An Efficient Jitter Measurement Technique
An Efficient Jitter Measurement Technique
Volume 51
123
Dr. Dongwoo Hong Prof. Kwang-Ting Cheng
Broadcom Corporation University of California
5300 California Ave. Santa Barbara College of Engineering
Irvine CA 92617 Dept. Electrical & Computer Engineering
USA Santa Barbara CA 93106-9560
dwhong@broadcom.com USA
timcheng@ece.ucsb.edu
Springer
c Science+Business Media B.V. 2010
No part of this work may be reproduced, stored in a retrieval system, or transmitted in any form or by
any means, electronic, mechanical, photocopying, microfilming, recording or otherwise, without written
permission from the Publisher, with the exception of any material supplied specifically for the purpose
of being entered and executed on a computer system, for exclusive use by the purchaser of the work.
With the increasing demand for higher data bandwidth, communication systems’
data rates have reached the multi-gigahertz range and even beyond. Advances in
semiconductor technologies have accelerated the adoption of high-speed serial in-
terfaces, such as PCI-Express, Serial-ATA, and XAUI, in order to mitigate the high
pin-count and the data-channel skewing problems. However, with the increasing
number of I/O pins and greater data rates, significant challenges arise for testing
high-speed interfaces in terms of test cost and quality, especially in high volume
manufacturing (HVM) environments.
This dissertation proposes novel testing techniques for high-speed interfaces,
which can significantly reduce testing costs while maintaining high test coverage.
The primary focus is on efficient testing methods for jitter and bit-error-rate (BER),
which are widely used to represent the quality of a communication system. The
reader is assumed to have a basic understanding of high-speed I/O and its test
methodologies.
There are eight chapters in this dissertation. Chapter 1 gives a general introduc-
tion to high-speed I/O systems and describes commonly used testing methods along
with their limitations.
Chapter 2 presents an efficient jitter measurement technique using fast compara-
tor sampling. The comparator under-sampling technique is discussed, followed by
the efficient jitter calculation algorithm. To validate the accuracy of the technique,
the experimental setup and results using a high-speed sampling prototype and vari-
ous types of measurement instruments are presented.
The BER estimation technique for high-speed serial links that incorporate the
linear clock and data recovery (CDR) circuit is described in Chapter 3. The jitter
transfer characteristics of the linear CDR loop are analyzed based on the conven-
tional phase-locked loop (PLL) theory. This chapter then describes how the input
jitter and the CDR circuit’s internal jitter affect the recovered clock jitter and the
dependency of the BER on the characteristics of the CDR circuit. The BER estima-
tion technique is extended to the serial links which incorporate the bang-bang (BB)
CDR circuit. This method is explained in Chapter 4. Due to the highly non-linear
characteristics of the BB CDR loop, the jitter transfer function of the loop strongly
varies depending on the input jitter magnitude. Thus, the loop’s dependency on the
jitter magnitude is fully characterized for accurate BER estimation.
vii
viii Preface
Chapter 5 first reviews the basic concepts of the timing margining test, which is
a widely adopted design-for-test (DFT) technique. Then, this chapter describes four
possible gaps that the timing margining test might have compared to conventional
jitter testing. Chapter 6 deals with the total jitter (TJ) estimation technique for im-
proving the quality of the timing margining test, which can accurately predict the TJ
at a very low BER level using only the information from the higher BER region. The
limitations of the existing TJ estimation method, which is based on the dual-Dirac
model, is described, and then a high-order polynomial fitting technique is proposed
to overcome the limitations.
A two-tone testing method for continuous-time adaptive equalizers is described
in Chapter 7. This chapter starts with an introduction to the continuous-time adaptive
equalizers. Then, the proposed two-tone testing method is described followed by the
transistor-level implementation details of the technique.
Chapter 8 concludes the dissertation by summarizing the results of the research
conducted and discussing possible future research areas.
This research could not be accomplished without guidance and support of pro-
fessors, colleagues, family members, and friends. First of all, I would like to thank
my advisor, Professor Tim Cheng, for his endless support and great advice dur-
ing the challenging periods. I feel grateful and privileged to have worked with
him. I would also like to thank a number of professors and professionals who re-
viewed the dissertation and gave me helpful feedback: Professor Forrest D. Brewer,
Professor Stephen I. Long, Professor Patrick Yue, and Dr. Mike P. Li.
I would like to acknowledge Teradyne for providing test equipment and facilities
for experiments. Especially, I would like to thank Cameron Dryden, Michael Panis,
Jacob Scherb, and Wolfgang Maichen for their assistance and feedback. I would
like to thank Anne Meixner, Benoit Provost, James Jaussi, and Bryan Casper of
Intel. Their great technical expertise always inspired me and helped me to solve
several challenging problems on a number of occasions.
I would like thank Linda Dailey Paulson for her attentive help in improving my
writing skill throughout the writing process of this dissertation. I would like to ex-
press my sincere gratitude towards my parents for their support and encouragement
during this project. Finally, I give my warmest thanks to my wife, Joomi Kim. Her
love and continual encouragement have been a great source of strength that allowed
me to complete this work.
Contents
1 Introduction .. . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . 1
1.1 Overview of High-Speed Serial Links .. . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . 1
1.1.1 High-Speed Serial Link System . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . 1
1.1.2 Testing High-Speed Serial Links . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . 2
1.2 Challenges in Testing High-Speed Serial Links .. . . . . . . . . . . . .. . . . . . . . . . . . 3
1.3 Contributions of the Dissertation . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . 4
ix
x Contents
8 Conclusions . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . 89
References .. . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . 95
Chapter 1
Introduction
High-speed serial links are composed of a transmitter (TX) and a receiver (RX)
communicating over a channel. Figure 1.1 shows the typical block diagram of a
transceiver for high-speed serial links. Due to the limited number of I/O pins in a
chip and density constraints on the number of wires between the chips, the links
usually convert parallel data to serial one using a serializer before transmitting the
data. In the receiver side, this serial data is reconverted to the original parallel data
using a deserializer. A clock and data recovery (CDR) circuit in the receiver extracts
the clock information from the data to synchronize the receiver with the incoming
data because, in serial communication systems, the clock signal is embedded in the
data. Thus, the CDR circuit plays a critical role in determining the quality of serial
communication systems, including influencing metrics such as bit error rate (BER).
As the data rates continue to increase and approach speed of multi-gigabits/second
the signal is distorted by the bandwidth limitation of the channel. In order to com-
pensate for channel loss, a pre-emphasis at the TX and an equalizer at the RX are
implemented in the system. In addition, a simple pattern generator and an error de-
tector are found in most transceiver designs for testing purposes.
D. Hong and K.-T. Cheng, Efficient Test Methodologies for High-Speed Serial Links, 1
Lecture Notes in Electrical Engineering 51, DOI 10.1007/978-90-481-3443-4 1,
c Springer Science+Business Media B.V. 2010
2 1 Introduction
Data Clock TX
Data from Core Pre-emphasis
Serialize
Pattern
generator
Pass/ RX Equalizer
Fail Error
detector Deserialize
Data to Core
CDR
BER and jitter have been widely used as measurements to ensure the performance
of high-speed communication systems. The BER represents the ratio of the number
of bits received in error to the total number of bits transmitted. Jitter is defined as the
deviation of a signal event from its ideal position in time. Total jitter (TJ) is typically
divided into two categories – deterministic jitter (DJ) and random jitter (RJ). DJ is
further divided into periodic jitter (PJ) (also known as Sinusoidal Jitter (SJ)), data-
dependent jitter (DDJ), and bounded uncorrelated jitter (BUJ). The characteristics
and sources of different kinds of jitter are summarized as follows [1–3].
Random jitter (RJ) – RJ comes from device noise sources, which include shot
noise, flicker noise and thermal noise. It is commonly modeled by the Gaussian
distribution function.
Periodic jitter (PJ) – PJ is typically caused by external deterministic noise sources
coupling into a system, such as switching power supply noises or strong local RF
carriers. It may also caused by an unstable CDR circuit.
Data-dependent jitter (DDJ) – DDJ is correlated with the bit sequence in a data
stream. DDJ is mainly caused by the duty-cycle distortion (DCD) and inter-symbol
interference (ISI). The slew rate differences between the rising and falling signal
edges and the voltage offsets between the differential inputs generate the DCD,
while the bandwidth limitation of the transmission channel causes the ISI.
Bounded uncorrelated jitter (BUJ) – BUJ is typically due to coupling from
adjacent lines (i.e. crosstalk) or on-chip random logic switching. It is bounded,
and its characteristics depend on the data pattern, coupling signal, and coupling
mechanism.
Jitter or BER can be measured using a variety of methods, either using external
equipment or on-chip circuitry. The external instruments include oscilloscope, spec-
trum analyzer, time interval analyzer (TIA), and bit-error-rate tester (BERT) [1, 4].
A TIA measures many single-shot edge-to-edge time intervals on a non-continuous
and random basis. The statistics of these measurements can be used to perform
1.2 Challenges in Testing High-Speed Serial Links 3
the total jitter calculation. BERT is composed of a pattern generator and an error
detector. In order to obtain the amount of eye closure as a function of BER, the
BERT can vary the clock edge placement with respect to the data edge for BER
measurement, which is called the BERT scan technique. Various on-chip measure-
ment techniques have been proposed [5–9]. For jitter measurement, a ring oscillator
and a calibration circuit are used in [5]. This technique is capable of measuring jit-
ter with a resolution as low as a single gate delay in the given process. In [6, 7],
researchers have proposed novel ways of measuring jitter with a resolution below a
single gate delay either by applying an undersampling technique or a Vernier delay
line. In [8, 9], the BER or timing margin is measured by simply providing a loop-
back path from the transmitter to its own receiver. A transceiver includes a pattern
generator and an error-detector for BER measurement, and has features to control
the clock phase in order to adjust the point in the eye from which data samples are
taken for margining test.
With the increasing number of I/O pins and greater data rates, testing high-speed
interfaces has posed significant challenges in terms of test cost and quality.
In this section, we describe a few main challenges in testing these high-speed
communication links.
Currently available jitter measurement techniques either require expensive mea-
surement instruments or else they do not guarantee sufficient test quality. Although
jitter measurements using external instruments can be performed in the lab for char-
acterization, it is not appropriate for high volume production due to its cost and
scalability limitations. Some instruments, such as the oscilloscope and spectrum an-
alyzer, do not lend themselves to fast parallel testing of devices with a large number
of high-speed interfaces due to their hardware complexity. In addition, current on-
chip jitter measurement circuits do not have sufficient resolution [5, 7], or cannot
separate various jitter components [6].
BER measurement down to the 1012 level, which is required to ensure system
reliability for most multi-gigabit communication standards, is test time prohibitive.
In order to guarantee 1012 BER with 95% confidence level, at least 3 1012 bits
need to be captured without a single error. Even at 3 Gb/s data rate, it takes more than
15 min to capture that many bits. The jitter tolerance test is even more troublesome
than that because it has to measure the BER by sweeping periodic jitter (PJ) in
frequency and amplitude. Although the testing time can be reduced via extrapolation
from higher BERs using the BERT scan method, the results may not be accurate if
the extrapolation is from inappropriate BER levels.
Loopback-based margining test may not detect some failures in devices. With
the challenges of keeping test costs down and testing large numbers of serial I/Os
in a single chip, the timing margining test along with loopback configuration has
been widely adopted by companies [8, 9]. However, a simple loopback test that
4 1 Introduction
determines maximum and minimum limits on phase offset (i.e. timing margin) does
not provide adequate test coverage for analog performance variations [10]. Thus, it
might not fully replace either conventional jitter or BER testing.
While equalizer design has been studied for a long time and a number of novel ar-
chitectures have recently been developed to mitigate channel bandwidth limitation,
high-quality and cost-effective production test methods for these equalizers are not
yet well developed. The pre-emphasis block in the TX can be directly observed from
output pins; however, since equalizer outputs are typically embedded in the RX, they
may not be directly measured using external equipment unless additional pins are
added. In addition, measuring the eye-diagram on-chip to guarantee the operation
of equalizers in the multi-gigahertz range requires a significant amount of internal
circuitry.
In this dissertation, we propose novel testing techniques that address the previously
mentioned challenges in high-speed interface testing. The contributions of the re-
search can be categorized into four groups:
1. An efficient jitter measurement technique using fast comparator sampling
We have proposed an efficient jitter measurement technique using fast com-
parator sampling. This technique uses a simple sampler circuit and an efficient
analysis algorithm for jitter measurement. Due to its simplicity, small size,
and high bandwidth, it is applicable for either on-chip jitter measurement or
measurement using Automatic Test Equipment (ATE). The approach combines
partial measurements based on individual data edge regions, in contrast to more
common approaches that first accumulate data from multiple edge regions. This
enables the technique to effectively separate RJ from DDJ and low-frequency PJ.
2. BER estimation technique
We have proposed the BER estimation technique for high-speed serial links,
which utilizes the jitter spectral information extracted from the transmitted data
and some key characteristics of the CDR circuit in the receiver. In addition to
giving insight into both the behavior of the CDR loop and the contribution of
the jitter to the BER, the estimation technique can be used to accelerate the jitter
tolerance test by eliminating the conventional BER measurement process. We
propose two different versions: one for use with linear CDR circuits and the other
for non-linear CDR circuits. Experimental results comparing the estimated BER
and the measured BER demonstrate the high accuracy of the proposed technique.
3. Gap analysis in timing margining test and solution for reducing the gap
Timing margining test used in conjunction with loopback configuration has
become a popular design-for-test (DFT)-based test method for high speed
interfaces. The timing margining test can evaluate most of the transceiver’s
functionality without relying on very expensive high-speed and high-pin-count
ATE. However, does it adequately cover every type of jitter specification? To
1.3 Contributions of the Dissertation 5
answer this question, we first explore possible sources that might generate some
gaps between the timing margining test and the conventional jitter test. In order
to reduce the gap between two methods, we also propose an accurate total jitter
estimation technique. In this technique, instead of relying only on the timing
margin, a few BER measurements are taken at high BER levels (i.e. higher than
106 levels). Then, the timing margin at 1012 level is estimated via extrap-
olation using different fitting algorithms. We demonstrate its value for a more
efficient and accurate total jitter estimation at a very low BER level.
4. Two-tone test method for continuous-time adaptive equalizer
A novel test method for continuous-time adaptive equalizers is proposed. This
technique applies a simple two-sinusoidal-tone signal as a stimulus and includes
a root-mean-square (rms) value detector for testing. The advantages of the pro-
posed technique are as follows: (1) The test stimuli can be easily generated and
applied. (2) The test output is a DC signal and thus can be easily measured.
(3) The extra on-chip circuitry needed for supporting the technique requires only
a very small area overhead and does not result in any performance degradation.
To validate the technique, we used a recently published adaptive equalizer as
our test case and conducted both behavioral and transistor-level simulations.
Simulation results demonstrate that the technique is effective in detecting defects
in both the equalization filter and the adaptation loop, either of which might not
be easily detected using the conventional eye-diagram method.
Chapter 2
An Efficient Jitter Measurement Technique
As device data rates increase, comparator undersampling has advantages both for
ATE [13] and for on-chip testing [6] due to its simplicity, small size, and high
bandwidth.
D. Hong and K.-T. Cheng, Efficient Test Methodologies for High-Speed Serial Links, 7
Lecture Notes in Electrical Engineering 51, DOI 10.1007/978-90-481-3443-4 2,
c Springer Science+Business Media B.V. 2010
8 2 An Efficient Jitter Measurement Technique
DUT Differential AT E
Buffer
High-Speed Sample Memory
Output Latch
DATA
VT
Comparator ADDR
Data
Clock
ΔT
2ΔT
Resolution (ΔT) = 1ps
Data
Clock …
Comparator
00001011001111…1110110010000…0000101100111…
Samples
Fig. 2.2 Comparator data acquisition: (a) real-time data sampling, (b) successive data samples
Figure 2.1 is a block diagram of our comparator sampler. Differential data from
a device transmitter is buffered and compared against a threshold, which represents
the edge transition voltage.
Figure 2.2 illustrates the comparator data acquisition. In Fig. 2.2a, the clock pe-
riod is set at 1ps greater than the period of a repeating pattern, so each clock samples
2.2 Random Jitter Measurement 9
the pattern 1ps later than the previous sample. The clock acts as a strobe walking
across the pattern with 1ps resolution.
Figure 2.2b illustrates the reconstructed data that appear to have been sampled
very densely with each sample separated by 1ps. If the data contains jitter, the result
in memory is a generally non-monotonic transition region between all 0s and all
1s (gray region). We define the transition region for a zero-to-one transition as the
first-one-to-last-zero, and vice versa.
A “1” (“0”) sample indicates that the buffered data transitioned high (low) at a
time before the specific clock edge that took the sample. Since jittered edge positions
are stochastic, the likelihood of an edge occurring in a particular position can be
described using statistical functions. If samples were taken at the same position
multiple times, the density of ones would represent the probability that a rising edge
occurred prior to that effective position.
The distribution of increasing probabilities could be described as a cumulative
distribution function (CDF) of the jitter. The derivative of the CDF yields the prob-
ability distribution function (PDF), which in this case describes the probability that
an edge occurred in a specific region. Since random jitter can be defined as the prob-
ability of an edge occurring over a range of positions, we can analyze the PDF to
determine the edge’s jitter.
It should be noted that in our method, we only sample each effective position
once. This results in a CDF that is only a crude estimation of the edge’s actual
CDF. In fact, it appears little like a CDF, which would be monotonic. Despite this,
if the sample resolution is small enough and the latch bandwidth high enough, this
estimate is good enough to produce accurate measurements of RJ, as will be shown,
although non-Gaussian jitter sources complicate the matter [14, 15].
The mean and standard deviation are subsequently calculated from each transi-
tion region’s PDF. The mean edge positions can be used to estimate DDJ. Rms jitter
is calculated by an aggregate estimate of each region’s standard deviation. The de-
tails of how each region’s mean and standard deviation is calculated and the validity
of this technique are described in the following sections.
Looking at an individual edge region, i, RJ for that edge equals its standard
deviation:
RJEdge .i/ D ¢Edge .i/
However, most practical measurements will deduce RJ from multiple edge mea-
surements. The RJ component must be measured accurately to avoid overstating TJ,
which is often used to differentiate good devices from bad. Given a system with
only deterministic and Gaussian random jitter, TJ is estimated by [1]:
TJ D DJ C .14:069 RJ/
10 2 An Efficient Jitter Measurement Technique
Using the binary data for each edge transition region, our approach is to calculate the
standard deviation of each individual region instead of first accumulating data for
multiple regions. We then calculate an aggregate estimate. This approach is insen-
sitive to timing disturbances occurring between edges, since the measurement does
not accumulate timing errors from one edge to the next, unlike cumulative-edge
techniques, such as a histogram measurement performed by a traditional scope.
Also, calculating the standard deviation does not require edge distributions to be
Gaussian or symmetric.
Our RJ measurement for N edge transition regions is computed according to:
s
¢12 C ¢22 C ¢32 C : : : C ¢N2
RJ D
N
where ¢j is the standard deviation of the jth pattern edge. The optimal way to com-
bine the measurement values depends on the nature of the data itself. In practice,
the approach is selected primarily for computational efficiency and robustness to
outliers, given an expected distribution.
When a sigma is calculated using many sample points per transition region, the
root-mean-square average works well for combining the individual sigmas. How-
ever, it is not unusual to have fewer than a dozen points per region, causing the
variances to follow a chi-squared distribution with a pronounced skew, rather than
one that is nearly Gaussian. The longer tail on the high side of the distribution, along
with the root-mean-square average (which more heavily weights large values) over-
estimates the true value of sigma. Our proposed method should work well given 20
or more samples per transition region. Other methods of combining the individual
sigmas may work better in practice, given fewer points.
Using information sampled at a fixed sample resolution t, the mean and stan-
dard deviation for each edge transition are calculated as follows:
1. Compute discrete derivatives (Œi) between successive samples as a discrete
approximation of the PDF.
2. Compute the first and second statistical moments (M1 and M2) of the difference
values for each edge region, given by:
X X
M1 D .Œi .i C 0:5// D .M1i /
Edge region Edge region
2.2 Random Jitter Measurement 11
X X
M2 D .Œi .i C 0:5/2 C jŒij =12/ D .M2i /
Edge region Edge region
Table 2.1 gives an example of the steps necessary to calculate the mean and stan-
dard deviation for a single edge transition. In calculating M1, 0.5 is added to each
sample number on the assumption that all information within the histogram bin is
located in the center of the bin. This reduces the square sum by 1/12, compared to
a uniform distribution in each bin, which must be compensated in computing the
second moment. In every jitter region, the same steps are applied to derive the mean
and standard deviation individually.
This technique generally yields different results than one superimposing data
from multiple edges, because it filters out variations in the mean positions of each
edge. This effect is illustrated in Fig. 2.3. This also greatly reduces low-frequency
PJ and DDJ components in the data from appearing in the RJ measurement, since
the same roll-off affects all the above components similarly. If a device’s RJ has
a substantial low-frequency component of concern, then the proposed approach
would be disadvantageous. However, since Clock Data Recovery (CDR) circuits
12 2 An Efficient Jitter Measurement Technique
Jitter
Region 1
Jitter
Region 2
Jitter
Region 3
Jitter
Region 4
Note that as with all sampling techniques, this technique is subject to aliasing
effects. An example of how aliasing can affect jitter measurement is as follows:
Sampling frequency D 10 MHz
PJ D 100 MHz
Every time a strobe occurs, it will hit the same phase of PJ; thus, the PJ will
not be observed. The data rate is irrelevant for the above example. However, if one
considers the following example, where
Data rate D 1 Gb=s (nominal, ignoring sample resolution)
Bits=pattern D 30
Pattern rate 30 MHz (ignoring sample resolution)
In this case, the PJ would affect the DUT signal, even though the measurement
would not capture it, due to the relative synchronicity of the PJ and the sampling
frequency.
First, data with only injected RJ was analyzed. Calculations were performed us-
ing the proposed method and a conventional histogram method that does not adjust
mean edge positions, for comparison purposes. Figure 2.4 illustrates the simu-
lation results. In each case, 20 iterations are simulated and averaged to reduce
measurement variability.
In the absence of injected PJ or DDJ, both methods measure the pure RJ ac-
curately, as expected. If we inject PJ as shown in the next section, the differences
increase.
where RJrms is the rms value of RJ, and PJrms is the rms value of PJ.
7
6
Measured RJ (ps)
5
4
3
2
1
0
0 1 2 3 4 5 6
Injected RJ (ps)
Histogram (control) Proposed Method
K is a scalar value relating the random jitter rms value to peak-to-peak jitter am-
plitude. K is a weak function of both the sample resolution granularity relative to
the sigma of the distribution and the number of edges captured during the measure-
ment. Our simulations suggest that for 100–1,000 edges captured and 5–50 sample
steps per sigma, 2:5 < K < 3:5. For K D 2:8; AJA D 17:66ps. For Case 2, the
characteristic frequency is:
(Note: Since the errors between the injected and the measured RJ are represented
in Fig. 2.5, the values are close to 4:69 2 D 2:69 ps). The proposed technique is
accurate to within 0.5 ps range up to Fc =3 (10 MHz). Above this frequency, the PJ
component appears in the RJ measurement.
A 20-bit pattern was used for this simulation. Figure 2.6 shows the data pattern and
how much DDJ is injected on each edge transition. The DDJ applied to each edge
was chosen arbitrarily, not based on a specific channel model. As before, a 2 ps rms
value of RJ is injected.
3
Error (ps)
0
1 10 100
Injected PJ Freq. (MHz)
Histogram (control) Proposed Method
0 0 0 0 1 1 0 0 1 1 0 0 1 0 1 0 1 1 1 1
Table 2.2 shows simulated RJ measurement results for both methods. As can be
seen, measured RJ from the proposed method is not affected by DDJ – it returns the
same value as Case 1 (i.e., when only RJ is injected). On the other hand, the cumu-
lative histogram overstates the measurement substantially (to be fair, a traditional
scope could be set up to make this measurement correctly).
First, we injected RJ using the same conditions as for the simulation and measured
it using both the oscilloscope and fast sampling prototype. Figure 2.8 illustrates the
results.
2.3 Experimental Results 17
RJ PJ
Source Source
D
Pattern Digital
Generator D Oscilloscope
Triggering clock
(6.4GHz)
Fast
Clock Sampler
Source Sampling clock
Jitter
Analysis
7
6
Measured RJ (ps)
5
4
3
2
1
0
0 1 2 3 4 5 6
Injected RJ (ps)
Scope Prototype
In most cases, the error between the scope and the prototype measurements are
less than 0.5 ps, and the prototype generally overstates the result. The scope result
is plotted as the ideal, although it is difficult to know precisely how much jitter is
delivered to the experimental setup. We suspect that the measurement differences
are due to the way the aggregate estimate is computed, but that close correlation
between the techniques could be achieved with additional calibration.
Next, we injected PJ C RJ. The rms value of RJ is fixed at 2 ps, and we injected
different PJ frequencies, one at a time, with the same peak-to-peak amplitude as the
18 2 An Efficient Jitter Measurement Technique
3
Error (ps)
0
1 10 100
Injected PJ Freq. (MHz)
Scope Prototype
simulation (i.e. 12 ps). Then, RJ was measured using the fast sampler and the scope.
Figure 2.9 illustrates the experimental results. The error represents the difference
between the injected RJ and the measured RJ.
As expected, RJ measured by the scope overstates the result due to PJ, regardless
of frequency. However, the proposed method measures RJ much more accurately,
within 0.5 ps error, up to Fc =4. The proposed method overstates the result above
about FC =2, as PJ elongates the individual edge transition regions.
2.4 Summary
A random jitter measurement technique using a simple sampler circuit and effi-
cient analysis algorithm has been demonstrated. The technique has advantages for
ATE applications testing multiple high-speed serial interfaces. Simulation and mea-
surement results are compared between the proposed technique and the cumulative
histogram based technique, and the proposed technique demonstrably measure RJ
more accurately in the presence of DDJ and low-frequency PJ, up to 25% of a
deduced characteristic frequency.
High-frequency PJ is shown to contaminate the RJ measurement results. Further
work should be to develop a method of separating PJ from the measured RJ, and
to further quantify the effect of sample resolution and total number of samples on
measurement accuracy.
Chapter 3
BER Estimation for Linear Clock and Data
Recovery Circuit
D. Hong and K.-T. Cheng, Efficient Test Methodologies for High-Speed Serial Links, 19
Lecture Notes in Electrical Engineering 51, DOI 10.1007/978-90-481-3443-4 3,
c Springer Science+Business Media B.V. 2010
20 3 BER Estimation for Linear Clock and Data Recovery Circuit
In principle, if the frequency of input jitter is relatively low, the CDR circuit can
track the jitter perfectly, and thus no bit errors will occur. However, if the input
jitter varies rapidly, the CDR circuit may not track the jitter, and some errors will
occur. On the other hand, the CDR circuit has an opposite reaction to the CDR
circuit’s internal noise. That is, the high frequency component of the internal noise
is transparent (and thus will be directly added) to the recovered clock instead of
being filtered out by the CDR circuit. In addition, the phase response of the CDR
circuit, which determines the timing response in clock recovery, has a very strong
correlation to the BER. If the jitter frequency falls into the range in which the phase
delay is non-zero, the CDR circuit introduces some timing delay to the recovered
clock which will, in turn, contribute to the BER. Within a specific frequency range,
this timing delay can cause a significant increase in the BER. Detailed descriptions
of how the magnitude and the phase responses of the CDR circuit affect the BER
will be described in the following sections.
In the next section, we first summarize the relationship between the BER and
the rms value of the RJ. Section 3.2 gives details of the BER variations, which
depend on the jitter spectrum and a CDR circuit’s jitter transfer characteristics.
Section 3.3 extends the BER estimation by including the noise of the CDR circuit it-
self. Section 3.4 presents the experimental setup and results for the validation of the
proposed technique. Section 3.5 summarizes the chapter and discusses future work
for further improving the technique.
If only RJ is present in the transmitted data, the BER can be easily estimated [1].
Since the CDR circuit can track only the low frequency components of the jitter, the
3.1 BER Analysis with Random Jitter 21
1T
Data
Clock T/2
RJ
Ideal
sampling
point
Error Error
Fig. 3.2 The input jitter and the recovered clock jitter
rapidly varying RJ cannot be tracked. Figure 3.2 shows the jitter in the transmitted
data (2.5 Gbps) with only the RJ component and the recovered clock jitter produced
by the CDR circuit (from simulation). As observed, the recovered clock does not
track input jitter at all (i.e. the jitter of the recovered clock is close to zero whereas
RJ in the transmitted data is significant). Thus, errors occur when the input jitter
exceeds plus or minus half of a Unit Interval T/2 (where T is denoted as one Unit
Interval (UI)).
22 3 BER Estimation for Linear Clock and Data Recovery Circuit
T=2
xD
¢RJ
The Q-function is multiplied by 2 because the error occurs on both sides of the data
transition (i.e. when jitter is greater than the threshold or less than .1/ threshold).
In addition, the BER is linearly proportional to the transition density (¡T ) of the data
[27]. For a clock-like pattern (i.e. a periodic “1010” stream), the transition density
is equal to 1.
The estimation becomes much more complex when both PJ and RJ components
are present. Table 3.1 shows the simulation results of the BER with different com-
binations of RJ and PJ (the details of the simulation setup will be explained in
Section 3.4). In this experiment, we injected the PJ into the transmitted data at
four different frequencies 50 KHz, 1 MHz, 10 MHz, and 100 MHz – with the same
amplitude (0.1 T). In addition, we injected RJ with an rms value of T/7.8. We then
compared the resulting BER for the following three cases: (1) only PJ is injected;
(2) only RJ is injected; and (3) both PJ and RJ are injected. As shown in Table 3.1,
if the frequency of the injected PJ is low (say, at 50 KHz), the BER for Case (3)
(i.e. containing both PJ and RJ) is the same as the BER for Case (2) (i.e. containing
only RJ). However, when the frequency of the PJ increases, the BER of these two
cases becomes different although the presence of the PJ alone (i.e. Case (1)) does
not cause any error. In the following sections, we will give a thorough analysis to
explain how the frequency of the PJ affects the BER.
As discussed earlier, the CDR circuit can tolerate low frequency components of the
input jitter. This characteristic results in an increase of BER when the frequency
of the PJ increases. However, this alone cannot completely explain the phenomena
observed in Table 3.1. To understand these effects more precisely, we have to take
into account the phase response of the CDR circuit in addition to its magnitude
response.
A CDR circuit is commonly implemented using the architecture of a Phase-
Locked Loop (PLL). The basic block diagram of the CDR circuit under such
architecture is shown in Fig. 3.3.
The open- and closed-loop transfer functions of the CDR circuit can be expressed
by Eqs. 3.3 and 3.4, respectively [28],
2Ÿ¨n C ¨n2
HO .s/ D 2Ÿ 3
(3.3)
m¨ n
s C .m C 1/s2
™o .s/ 2Ÿ¨n C ¨n2
HC .s/ D D 2Ÿ (3.4)
™i .s/ m ¨n s3 C .m C 1/s2 C 2Ÿ¨n s C ¨n2
where m is the capacitance ratio, is the damping ratio, and !n is the natural fre-
quency. These parameters can be expressed by the following equations:
C2
mD
C1
IP KVCO /s
θi θo
Phase Charge
VCO
Detector Pump
R C2
C1
Ip KVCO
¨n2 D
2 C1
IP RKVCO
2٬n D
2
Note that Eqs. 3.3 and 3.4 are valid only if the data follow a clock-like pattern.
If random data are applied, the transfer function will be different because the phase
detector (PD) updates the phase error only when a transition occurs in the data. Thus,
if the data have a lower transition density, the PD gain will be reduced proportion-
ally. Since the transition density of a pseudo-random bit stream (PRBS) pattern,
commonly used as a general purpose test pattern, is half of a clock-like pattern’s
transition density [29], the open-loop and closed-loop transfer functions for a PRBS
pattern could be expressed by Eqs. 3.5 and 3.6.
1
HO PRBS .s/ D HO .s/ (3.5)
2
2Ÿ¨n C ¨n2
HC PRBS .s/ D 4Ÿ 3
(3.6)
m¨ n
s C 2.m C 1/s2 C 2Ÿ¨n s C ¨n2
The solid lines in Fig. 3.4 show the magnitude and the phase responses of the
CDR circuit derived from Eqs. 3.4 and 3.6. The parameters are set to m D 0:005,
¨n D 2 2 105 rad=s, and Ÿ D 5, which result in a 2 MHz loop bandwidth and less
than 0.1 dB jitter peaking for the clock-like pattern. Note that jitter peaking could
cause degradation to the BER, if its value is large. For our analysis, the jitter peaking
was chosen to be at a level low enough so that it will not cause any non-trivial BER
degradation. To validate these transfer functions and to observe their dependency
on the transition density, a behavioral model of the CDR circuit was designed for
MATLAB simulation. The differences of the magnitude and the phase between the
injected PJ and the recovered clock jitter were computed for both clock-like and
PRBS patterns. The circles and the stars on the lines in Fig. 3.4 represent the simu-
lation results for the clock-like and the PRBS patterns, respectively. As shown, the
simulation results matched the responses derived from the equations very well for
both patterns.
To analyze the BER, which depends on the jitter frequency, we divide the fre-
quency into four regions based on the magnitude and phase responses as shown in
Fig. 3.4:
1. Region 1 (clock-like: 0–70 KHz, PRBS: 0–50 KHz) where the magnitude gain is
1, and the phase curve is flat (i.e. 0ı ).
2. Region 2 (clock-like: 70 KHz–2 MHz, PRBS: 50–700 KHz) where the magnitude
gain is 1, and the phase curve has a non-zero slope.
3. Region 3 (clock-like: 2–40 MHz, PRBS: 700 KHz–30 MHz) where the magni-
tude gain is less than 1, and the phase curve has a non-zero slope.
4. Region 4 (clock-like: 40 MHz and greater, PRBS: 30 MHz and greater) where the
magnitude gain is negligible.
3.2 BER Analysis with Random Jitter and Periodic Jitter 25
Magnitude
10
1010-TF
0 PRBS-TF
1010-Sim
PRBS-Sim
–10
–20
Gain (dB)
–30
–40
–50
–60
–70
–80
103 104 105 106 107 108 109
Frequency(Hz)
Phase
0
1010-TF
–20 PRBS-TF
1010-Sim
–40 PRBS-Sim
Phase (degree)
–60
–80
–100
–120
–140
–160
–180
103 104 105 106 107 108 109
Frequency(Hz)
Clock-like– R1 R2 R3 R4
PRBS – R1 R2 R3 R4
In the following section, we first describe the modification to the conventional dual-
Dirac model for the BER estimation. Then, we describe the method for estimating
the BER in each region.
One popular approach to the BER estimation is the use of the dual-Dirac model [14].
To calculate the total jitter distribution, it approximates the DJ as two Dirac delta
functions and convolves them with the RJ. Figure 3.5 represents this convolution of
the DJ and the RJ.
26 3 BER Estimation for Linear Clock and Data Recovery Circuit
Error Error
sRJ
* =
As discussed in Section 3.1, an error will occur when the total jitter exceeds plus
or minus T/2 without considering the clock recovery function (as shown in Fig. 3.5).
Thus, in this case, the threshold value x for the Q-function would be
T=2 ADJ
xD :
¢RJ
Note that, compared to Eq. 3.2, the mean position of the Gaussian distribution is
shifted by the amount of ADJ . Therefore, the BER can be estimated as:
T=2 ADJ
BER D 2 ¡T Q : (3.7)
¢RJ
In our case, both PJ and RJ are considered for the BER estimation. However, be-
cause the PJ distribution cannot be properly modeled as two Dirac delta functions,
we make some modifications to Eq. 3.7. We use two new variables – the effective
mean shifting .Aeff / and the effective rms value .¢eff / – to replace ADJ and ¢RJ . The
details of the derivation of these two new variables, which are based on the ratio of
the amount of RJ to the amount of PJ, are given in Appendix A. With the variable
replacement, the BER becomes:
T=2 Aeff
BER D 2 ¡T Q : (3.8)
¢eff
Now we take into account the receiver’s clock recovery capability for BER
estimation. Each region in the jitter transfer function of the CDR circuit, specified in
Section 3.2.1, has distinct magnitude and phase responses to the input jitter. There-
fore, we derive different equations for computing Aeff and ¢eff for different regions
and use the resulting values to estimate the BER. We first consider the clock-like
pattern, and then extend the analysis for the PRBS pattern.
3.2 BER Analysis with Random Jitter and Periodic Jitter 27
When the input data is a clock-like pattern, the transition density is one. Thus, the
BER will be:
T=2 Aeff
BERCLK D 2 Q : (3.9)
¢eff
1. Region 1
If the RJ and the PJ in Region 1 are present in the data, the PJ will be perfectly
tracked whereas the rapidly varying RJ cannot be tracked by the CDR circuit.
Therefore, only the RJ contributes to the BER.
Suppose that the total jitter is composed of a PJ component s.t/ D a1 sin.¨t/
and an RJ component n(t) with an rms value of ¢RJ . Errors occur when the input
jitter exceeds the error boundaries, which are the recovered clock jitter plus or
minus T/2. Because the CDR circuit can completely track the PJ, the recovered
clock jitter will be the same as s(t). That is, errors occur when
The left hand sides represent the total input jitter, and the right hand sides
represent the error boundaries. After simplification, we arrive at the following
inequality:
jn.t/j T =2:
Because only the RJ affects the BER, the values of Aeff and ¢eff for Equation (3.9)
will be:
Aeff D 0
¢eff D ¢RJ : (3.10)
2. Region 2
Since the phase response of the CDR circuit has a non-zero slope, the PJ in
this region is tracked by the CDR circuit with certain delay introduced into the
recovered clock. This time delay also shifts the error boundaries, thus increasing
the BER. Figure 3.6 shows the input jitter, the delayed recovered clock jitter, and
the boundaries at which the errors occur.
The time delay introduced into the recovered clock jitter depends on the slope
of the phase curve. According to [30], the time delay at each frequency ¨ equals
the negative of the slope of the phase at ¨. That is:
d
£.¨/ D f†HC .j¨/g (3.11)
d¨
Fig. 3.6 The input jitter and the recovered clock jitter for Region 2
If the CDR circuit tracks the input PJ (i.e. a1 sin.¨t// with some delay t0 .t0
can be obtained from Eq. (3.11) at a known frequency), the recovered clock jitter
can be represented as a1 sin.¨.t t0 //. Therefore, errors occur when
Aeff D 0
p
¢eff D a1 2 .1 cos.¨t0 // C ¢RJ 2 : (3.14)
3.2 BER Analysis with Random Jitter and Periodic Jitter 29
Note that ¢eff is based on Eq. A.1 in Appendix A, and the variance of a sinusoidal
term with amplitude A is A2 =2. Therefore, the BER will be calculated simply by
substituting Eq. 3.14 for Eq. 3.9.
ii. When ¢RJ =¢PJeff < 1
Aeff and ¢eff for this case are:
!
¢RJ p
Aeff D 1 0:9 p a1 2 2 cos.¨t0 /
a1 1 cos.¨t0 /
p
¢eff D .1 C 0:842 / ¢RJ 2 : (3.15)
They are derived based on Eqs. A.2 and A.3 in Appendix A. These values can
then be applied to the BER estimation.
3. Region 3
In this region, the phase curve has a non-zero slope, and the magnitude gain is
less than one. Therefore, the recovered clock jitter has a certain delay and smaller
amplitude. In addition, since the time delay is significant (it could be as much
as half of the period of the PJ when the phase response of this region is near
180ı), the input PJ and the recovered clock jitter can be out of phase. That is,
when the input PJ has a maximum (minimum) value, the recovered clock jitter
can come close to a minimum (maximum) value. Since the error boundaries are
determined by the recovered clock jitter, this out-of-phase phenomenon brings
the error boundary curves closer to the input jitter, as indicated in Fig. 3.7. This
can make the BER of Region 3 worse than those of other regions (including
Fig. 3.7 The input jitter and the recovered clock jitter for Region 3
30 3 BER Estimation for Linear Clock and Data Recovery Circuit
Region 4, which will be discussed later), when the same amount of PJ and RJ
are present. This explains an interesting phenomenon: the CDR circuit’s tracking
ability may sometimes cause a worse BER than the case in which the CDR circuit
cannot track the jitter at all (i.e. Region 3 has a higher BER than that of Region 4,
as shown in Table 3.1).
If we denote the amplitude of the recovered clock jitter as a2 , which is
a2 D Gain a1 (Gain is obtained from Eq. 3.4), then the error boundaries can be
expressed as:
Combining these two inequalities and further simplifying them result in the
following inequality:
ˇp ˇ
ˇ ˇ
ˇ a1 2 C a2 2 2a1 a2 cos.¨t0 / sin.¨t C ‚2 / C n.t/ˇ T=2
Aeff D 0
s
a1 2 a2 2 a2
¢eff D .1 C 2 2 cos.¨t0 // C ¢RJ 2 : (3.16)
2 a1 a1
4. Region 4
Since the magnitude gain is negligible in this region, the PJ component is not
tracked at all, as shown in Fig. 3.8. As neither PJ nor RJ components are tracked
by the CDR circuit, the error boundaries can be expressed as:
Fig. 3.8 The input jitter and the recovered clock jitter for Region 4
For this region, the effective PJ is the same as the injected PJ.
i. When ¢RJ =¢PJeff > 1
Aeff D 0
a1 2
¢eff 2 D C ¢RJ 2 : (3.18)
2
They can then be used for Eq. 3.9 to estimate the BER.
ii. When ¢RJ =¢PJeff < 1
Aeff and ¢eff for Eq. 3.9 will be:
¢RJ
Aeff D 1 0:9 p a1
a1 = 2
p
¢eff D .1 C 0:842 / ¢RJ 2 : (3.19)
For all cases, we have derived Aeff and ¢eff for Eq. 3.9 to estimate the BER. As the
frequency of the PJ changes, the values of Aeff and ¢eff will be changed, so is the
BER. The above analysis shows how the frequency component of jitter affects
the BER corresponding to the CDR circuit’s jitter transfer characteristics.
32 3 BER Estimation for Linear Clock and Data Recovery Circuit
For the PRBS pattern, the equations derived for Aeff and ¢eff in each region do not
change. The only differences are: (1) The time delay (t0 ) and the magnitude of the
recovered clock jitter in Region 3 (a2 ) will be calculated using Eq. 3.6 instead of
Eq. 3.4. (2) The transition density (¡T ) is 1/2. Thus the BER equation becomes:
T=2 Aeff
BERPRBS DQ : (3.20)
¢eff
The jitter transfer analysis can be extended to consider the intrinsic noise of the CDR
circuit. The main noise source of the CDR circuit is from the Voltage Controlled
Oscillator (VCO), whose frequency- and time-domain properties can be specified
as phase noise and jitter, respectively [31]. The CDR circuit behaves as a high-pass
filter for the VCO noise in contrast to the low-pass filtering behavior for the input
noise [23, 32, 33]. The transfer function between VCO phase noise (™n / and the
output phase .™o / is
2Ÿ 3
™o .s/ m¨ s C .m C 1/s2
Hn .s/ D D 1 H.s/ D 2Ÿ n
(3.21)
™n .s/ m ¨n s3 C .m C 1/s2 C 2Ÿ¨n s C ¨n 2
where m, Ÿ, and ¨n are the same as Eq. 3.4. Figure 3.9 shows a CDR circuit’s jitter
transfer characteristic for the VCO noise.
Fig. 3.10 The input jitter and the recovered clock jitter with VCO noise
The VCO noise is often assumed to be a white Gaussian [34]. As white noise
has a significant high-frequency component, most of the VCO noise appears in the
recovered clock jitter, as shown in Fig. 3.10.
Suppose there is VCO noise n2 .t/ with an rms value of ¢VCO in the presence of
the input PJ a1 sin.¨t/ and RJ n1 .t/. As an example, if the input PJ is in Region 2,
by modifying Inequalities 3.12, the error boundaries can be expressed as:
By combining and simplifying the inequalities of Eq. 3.22, we arrive at the following
inequality:
ˇ p ˇ
ˇ ˇ
ˇa1 2 2 cos.¨t0 / sin.¨t C ‚1 / C n1 .t/ n2 .t/ˇ T=2:
From this inequality, the rms value of the effective RJ, which combines the effects
of the input RJ and the VCO noise, is calculated as:
p
¢RJ eff D ¢RJ 2 C ¢VCO 2 : (3.23)
Note that these two noise sources are assumed to be independent. This ¢RJ eff , instead
of ¢RJ , can then be applied to Eqs. 3.14 and 3.15 to estimate the BER for Region 2.
For the other regions, the BER can be calculated by applying ¢RJ eff instead of ¢RJ
to Eqs. 3.10), and 3.16–3.19.
34 3 BER Estimation for Linear Clock and Data Recovery Circuit
A behavioral model of a CDR circuit was developed using MATLAB, which oper-
ates at 2.5 Gbps with loop parameters specified in Section 3.2.1. In order to reduce
the simulation time, we use the cycle-domain model of the CDR circuit [35]. Data
with various combinations of RJ and PJ were used as input to the CDR circuit. The
BER was then measured by comparing the timing information of the input data and
the recovered clock signal. Four different cases were simulated:
Case 1: Clock-like data with jitter satisfying the condition of ¢RJ =¢PJeff > 1;
Case 2: Clock-like data with jitter satisfying the condition of ¢RJ =¢PJeff < 1;
Case 3: PRBS with jitter satisfying the condition of ¢RJ =¢PJeff > 1; and
Case 4: PRBS with jitter satisfying the condition of ¢RJ =¢PJeff < 1
In each case, the amounts of PJ and RJ are fixed, and only the frequency of PJ is
varied to cover all four regions.
The main difficulty with this validation process is the excessively long simulation
time required to capture enough samples to measure a low BER. The measured BER
approaches the actual BER as the total number of compared bits approaches infinity.
However, if we measure the BER by comparing a finite number of bits, the value
will vary depending on the size of the compared bits. The BER can be modeled
using a Binomial distribution because there are only two possible outcomes (i.e.
error or no error). In most cases, as we are dealing with a low BER and a large
number of compared bits, the Poisson distribution can further approximate the error
distribution [36].
Suppose ’ is the expected number of errors at a given BER and Nbits is the num-
ber of compared bits, then ’ will be expressed as:
’ D BER Nbits :
Then, the number of errors (Nerr ) has the distribution following the PDF of:
’Neff ’
Ppoisson .Nerr / D e :
Nerr Š
a b
2.0E-06 1.0E-01
1.8E-06 1.0E-02
1.6E-06 1.0E-03
BER
BER
1.4E-06 1.0E-04
1.2E-06 1.0E-05
8.0E-07 1.0E-07
50 KHz 1 MHz 10 MHz 100 MHz 50 KHz 1 MHz 10 MHz 100 MHz
PJ Freq. PJ Freq.
c d
1.2E-06 1.0E-01
1.0E-02
1.0E-06
1.0E-03
8.0E-07
BER
BER
1.0E-04
6.0E-07
1.0E-05
4.0E-07
1.0E-06
Est. BER Sim. BER Est. BER Sim. BER
2.0E-07 1.0E-07
50 KHz 400 KHz 5 MHz 100 MHz 50 KHz 400 KHz 5 MHz 100 MHz
PJ Freq. PJ Freq.
Fig. 3.11 BER simulation results for (a) Case 1, (b) Case 2, (c) Case 3, and (d) Case 4 responses,
(b) for measuring phase responses
Figure 3.11 shows the simulation results for each case. Each graph plots the
estimated BER and the simulated BER. For the clock-like pattern, four different fre-
quencies of PJ were injected in each case, which are 50 KHz for Region 1, 1 MHz
for Region 2, 10 MHz for Region 3, and 100 MHz for Region 4. The rms value of
RJ is fixed at T/9.8, and the peak-to-peak amplitude of the PJ was fixed at 0.06T for
Case 1 and 0.6T for Case 2. For the PRBS pattern, we chose 50 KHz for Region 1,
400 KHz for Region 2, 5 MHz for Region 3, and 100 MHz for Region 4. The ampli-
tude of the PJ was set at 0.07T for Case 3 and 0.7T for Case 4. The rms value of RJ
was fixed at T/9.8. The simulation results indicate that for all cases, the difference
between the estimated BER and the simulated BER is very small.
We further conducted the experiments using the MAXIM 3873A CDR circuit,
which operates at 2.488 Gbps with 2 MHz bandwidth and less than 0.1 dB jitter
36 3 BER Estimation for Linear Clock and Data Recovery Circuit
peaking [37]. This chip was chosen because its specifications are very close to those
used in our simulation setup. We used Synthesys Research’s BERTScope for jitter
injection and for BER measurement.
We first characterize the CDR circuit’s jitter transfer function using the experimental
setup shown in Fig. 3.12. For measuring the jitter transfer function, the BERTScope
a
BERT
Pattern Error
Generator Detector
CLK_O CLK_I
+ −
MAX 3873A
Recovered
DATA_I +
DATA_O Clock
−
+
+
− CLK_O
−
b
Scope BERT
Pattern Error
DATA1 DATA 2
Generator Detector
+ − + −
CLK_O CL K_I
+ −
DATA_O DATA_I
Clock
+ − + −
with PJ
Recovered
Clock
MAX 3873A
Data with PJ
+ DATA_I
DATA_O
−
+ +
CLK_O −
−
Fig. 3.12 Setup for characterizing the CDR circuit: (a) for measuring magnitude responses, (b) for
measuring phase responses
3.4 Experimental Results 37
has the ability to sweep the frequency of the injected PJ and, in turn, measure
the magnitude difference between the injected PJ and the recovered clock jitter.
However, it does not support phase difference measurements. Thus, Tektronix’s Os-
cilloscope (TDS 6124C) equipped with jitter analysis software was used for the
phase measurement. Their jitter analysis package can measure the Time Interval Er-
ror (TIE) of two inputs [38]. Thus, we compare the timing difference between the
TIE of the Pattern Generator’s clock with PJ and the TIE of the recovered clock
to measure the phase responses of the CDR circuit. Fig. 3.12(b) shows the phase
measurement setup.
Figure 3.13 shows the jitter transfer characteristics of the CDR circuit. Both the
clock-like pattern and the 27 1 PRBS pattern were used and their magnitude and
the phase responses were measured.
We measured the magnitude responses of two extra patterns – periodic 1100 and
11110000 patterns, which have lower transition densities – to assess the effect of
the data’s transition density on the jitter transfer characteristics. The results show
–5
–10
Gain (dB)
–15
–20
–25
–30
–35
10 100 1000 10000 100000
PJ Freq.(KHz)
1010 1100 11110000 PRBS
0
–20
Phase (Degree)
–40
–60
–80
–100
–120
–140
–160
–180
10 100 1000 10000 100000
Freq.(KHz)
1010 PRBS
Fig. 3.13 Measurement results for the CDR circuit’s jitter transfer characteristics
38 3 BER Estimation for Linear Clock and Data Recovery Circuit
that the PRBS pattern and the 1100 pattern, whose transition density is equal and
is half that of the clock-like pattern, have very similar transfer curves. In the high-
frequency range, the PRBS pattern has a higher gain than that of the periodic 1100
pattern. This difference is because the DDJ leaks into the recovered clock jitter for
the case of the PRBS pattern. This phenomenon does not make any difference in
the low frequency range because the recovered clock jitter is much greater than the
DDJ in this range. In addition, the periodic 11110000 pattern, which has a transition
density of 1/4, has an even lower bandwidth. These results validated our finding that
the jitter transfer characteristic has a strong correlation with the transition density,
which was analyzed in Section 3.2.1.
After characterizing the CDR circuit, we measured the BER by injecting both PJ
and RJ into the data. Figure 3.14 shows the experimental setup. The BERTScope
can inject the PJ up to 2.2T, from 1 KHZ to 10 MHz, and 0.5T, from 10 to 80 MHz.
In addition, it can inject RJ up to 0.5T at 1012 BER level, which means the rms
value of maximum injectable RJ is limited to 0.036T (0:5T=14:069) [39]. Due
to this limitation that only a small amount of RJ can be injected, we cannot af-
ford to conduct the BER measurement experiments for Cases 1 and 3 (as the BER
level would be significantly lower than 1012 ). Therefore, the experiments were
conducted only for Cases 2 and 4.
We injected PJ at eight different frequencies, ranging from 50 KHz to 80 MHz,
to cover all four frequency regions. We fixed the rms value of RJ at 0.036T for all
cases. For the clock-like pattern, we injected PJ with two different peak-to-peak
amplitudes, 0.5T and 0.45T, and sweeping the PJ’s frequency. For the PRBS pat-
tern, we injected PJ with 0.5T peak-to-peak amplitude and sweeping its frequency
BERT
Pattern Error
Generator Detector
CLK_O CLK_I Recovered
+ − Clock
Data with
PJ + RJ DATA_O DATA_I
+ − + −
MAX 3873A
DATA_I DATA_O +
+ −
Fig. 3.14 Experimental +
− CLK_O
setup for BER measurements −
3.4 Experimental Results 39
a b
1.0E-05 Mea. BER Est. BER 1.0E-07 Mea. BER Est. BER
1.0E-06
1.0E-08
1.0E-07
BER
BER
1.0E-08 1.0E-09
1.0E-09 1.0E-10
1.0E-10
1.0E-11
1.0E-11
1.0E-12 1.0E-12
0.01 0.1 1 10 100 0.01 0.1 1 10 100
PJ Freq (MHz) PJ Freq (MHz)
c
1.0E-07 Mea. BER Est. BER
1.0E-08
BER
1.0E-09
1.0E-10
1.0E-11
1.0E-12
0.01 0.1 1 10 100
PJ Freq (MHz)
Fig. 3.15 BER measurement results for (a) clock-like pattern w/ 0.5T PJ, (b) clock-like pattern
w/ 0.45T PJ, and (c) PRBS pattern w/ 0.5T PJ
too. Figure 3.15 shows the comparison between the measurement results and the es-
timation results which are calculated based on Eqs. 3.9 and 3.20. We made minor
necessary modifications in our estimation: we used the effective rms value of RJ
(Eqs. 3.23) instead of using the rms value of the input RJ to incorporate the CDR
circuit’s intrinsic noise, and we included the extra DJ value caused by the cable into
Aeff for the PRBS pattern.
In the low frequency range, the expected BER is too low to be measured (i.e. less
than 1012 ) within a reasonable amount of time due to the CDR circuit’s tracking
ability. Thus, if no error is captured within a time limit (in our experiment, the
limit was set to 7 min 1012 b/(2.5 Gbps)), the BER is represented as 1012 as in
Fig. 3.15. For these cases, the estimated BER is indeed much lower than 1012 . The
results clearly show that the measured BER and the estimated BER match very well
in almost all cases. When the measured BER level is around 1012 , the difference
is larger than those of other cases as shown in Fig. 3.15b. We suspect that this is
because the injected RJ does not have the unbounded tails in real applications – thus
the Gaussian model for the RJ may not accurate in the very low probability region.
Further investigation is needed to more accurately model the RJ in this region.
The measurement results also clearly indicate that the BER peaks in Region 3
and drops when the PJ frequency reaches Region 4. This validates our analysis that
40 3 BER Estimation for Linear Clock and Data Recovery Circuit
the out of phase phenomenon between the injected PJ and the recovered clock jitter
could cause significant degradation in BER performance. Note that this increase in
BER is not caused by jitter peaking. This is evident because the device has a less
than 0.1dB jitter peaking and the magnitude gain of the jitter transfer function in
this frequency range (10 MHz) is about 15 dB as shown in Fig. 3.13.
Jitter has been used for measuring the quality of high-speed serial links, and sev-
eral standards specify the jitter performance for these links. However, jitter alone
would not fully reflect the overall system performance. The spectral information of
the jitter and the magnitude and phase responses of the CDR circuit’s jitter transfer
function should be jointly considered to determine the system performance. In this
chapter, we show how the frequencies of the input jitter and the CDR circuit’s in-
ternal jitter affect the recovered clock jitter and the dependency of the BER on the
characteristics of the CDR circuit. The experimental results demonstrate the validity
of our analysis and the roles of these parameters in determining the BER.
As the data rates continue to increase, the DDJ, caused by inter-symbol inter-
ference, becomes more significant due to the bandwidth limitation of the electrical
wires [40]. Various equalization techniques have been proposed to compensate for
the DDJ [41–43]. We will extend our work to incorporate the DDJ and some key
parameters extracted from the equalizer for the BER estimation.
Chapter 4
BER Estimation for Non-linear Clock
and Data Recovery Circuit
The block diagram of a typical BB CDR circuit is shown in Fig. 4.1 [44]. It consists
of a phase detector, a charge pump, a loop filter and a voltage-controlled oscillator
D. Hong and K.-T. Cheng, Efficient Test Methodologies for High-Speed Serial Links, 41
Lecture Notes in Electrical Engineering 51, DOI 10.1007/978-90-481-3443-4 4,
c Springer Science+Business Media B.V. 2010
42 4 BER Estimation for Non-linear Clock and Data Recovery Circuit
KVCO/s
θin Up
IP θout
BB PD Charge
VCO
Pump
Down R
–θth
–ICP
(VCO). The only difference between a linear and a BB CDR circuit is the PD oper-
ation. The bang-bang PD discards the magnitude information of the phase error and
only measures its polarity while a linear PD’s output magnitude is proportional to
the phase error.
Jitter transfer is defined as the ratio of a CDR circuit’s output jitter to the input
jitter, which is a function of the input jitter frequency. Since the binary PD in a BB
CDR circuit exhibits non-linear behavior to the input jitter, conventional linear PLL
theory cannot be used for analysis. Figure 4.2 shows the characteristic of the BB
phase detector to the input jitter, in D in;p cos w t. It operates in the linear region
for a small phase error whereas operates in the non-linear region for a large phase
error. The ideal characteristic of a BB phase detector should not have any linear
operation region. However, the average PD gain is smoothed by the metastability of
flipflops in the PD and random jitter in the input data and the VCO, and, thus, has a
finite slope across a narrow range of the phase error [44].
4.1 Jitter Analysis for BB CDR Circuits 43
θin,p
θin
t
ICP
IP
t
–ICP
θout,p
θout
t
Tθ /4
At a low jitter frequency, the PD can still track the input jitter closely, and the gain
is close to 1. As the jitter frequency increases, so does the phase error. This results in
a larger current from the charge pump flowing into the loop filter. Since the available
current beyond the linear region (i.e. ICP ) is constant, large and fast variation of the
input jitter results in slewing. The slewing phenomenon is illustrated in Fig. 4.3. For
illustration, we assume an extreme case in which the phase error changes the polarity
in every half cycle of the jitter. The charge pump current, IP , alternates between CICP
and ICP as a result. Because the loop filter capacitor is typically large, the output
of the loop filter tracks IP R. The output phase thus follows a triangular waveform as
shown in Fig. 4.3. The peak value of the output phase occurs after integration of the
control voltage for a duration of T™ =4 .T™ D 2 =w™ /; that is,
KVCO IP RT™
™out;p D :
4
Thus, the jitter transfer function can be represented as
ˇ ˇ
ˇ ™out;p ˇ KVCO IP R
ˇ ˇ
ˇ ™ ˇ D 2™ w : (4.1)
in;p in;p ™
Equation 4.1 expresses the dependence of the jitter transfer upon the input jitter
magnitude, ™in;p . In addition, the equation also reveals a 20 dB=dec slope in the
slewing region. As w™ decreases, slewing eventually vanishes and the jitter trans-
fer approaches unity, as shown in Fig. 4.4. Extrapolation of the linear and slewing
regions yields an approximate value for the 3 dB bandwidth of the jitter transfer,
which represents the boundary between the linear operation region and the slewing
region:
KVCO IP R
w3dB D (4.2)
2™in;p
44 4 BER Estimation for Non-linear Clock and Data Recovery Circuit
Slewing
(slope: –20dB/dec)
w–3dB wθ
θin,p
θin θout
t0 t1
Δθmax
Jitter tolerance is the process used to determine the magnitude of input PJ at which
a CDR loop starts to introduce errors. As the phase error, ™in ™out , approaches
(i.e. half of the unit interval [UI]), the BER increases rapidly. In this sub-section, we
summarize the relationship between the input PJ and the maximum phase error. It is
important to recognize that a BB CDR loop must slew if it incurs errors.
Figure 4.6 shows an example of the slewing for which the maximum phase error,
™max occurs at t1 . The phase error at t0 is used to approximate the maximum phase
error because it is close enough to ™max and much simpler to calculate. If ™out slews
for most of the period, t0 is approximately equal to T™ =4. Denoting the input jitter
as ™in;p cos.¨™ t C •/, the maximum phase error could be approximated as:
q
ˇ ˇ 4w2™ ™in;p
2
2 K2 2 2
VCO IP R
ˇ ˇ
™max ™.t0 / D ˇ™in;p cos. C •/ˇ D (4.3)
2 ™max 2w™
The details of the derivation can be found in [44]. Based on Eq. 4.3, we can calcu-
late the maximum tolerable input jitter by expressing ™in;p in terms of w™ when the
maximum phase error, ™max , is equal to .
The behavior of a BB CDR loop is well understood when only PJ is present in the
data. However, the analysis becomes much more complex when RJ is also present.
In the last section, we have shown that the jitter transfer function of the BB CDR
loop varies with respect to the PJ magnitude. In the presence of RJ, even with a
46 4 BER Estimation for Non-linear Clock and Data Recovery Circuit
fixed PJ magnitude, the jitter transfer function will vary for different amounts of RJ.
In the following section, we discuss the variation of the jitter transfer function with
respect to the rms value of RJ, followed by the derivation of the BER estimation
technique.
If the only jitter component in the data is PJ, the peak-to-peak magnitude of the jitter
is fixed and, thus, the dependency of the jitter transfer on the jitter magnitude can
be defined. However, when RJ is present in the data, the peak-to-peak magnitude of
the jitter becomes difficult to define because the RJ is typically assumed to follow a
Gaussian distribution and thus has unbounded characteristic.
In order to observe the RJ’s effects on jitter transfer, we obtained jitter transfer
gains through simulation by injecting both PJ and RJ. The rms value of the RJ is
fixed at 0.1 UI. Three different peak-to-peak magnitudes of the PJ – 0.4, 0.6, and
0.8 UI – are injected with the RJ. Figure 4.7 shows the simulation results. The simu-
lated jitter transfer gains (symbols in Fig. 4.7) were compared to the predicted jitter
transfer functions (black dotted lines in Fig. 4.7) based on the analysis described in
the last section. There are two distinctive discrepancies between them: the 3 dB
bandwidth and the slope in the slewing region. The 3 dB bandwidth derived from
simulation is lower than the prediction because the addition of the RJ results in an
analytical relationship between the slope and the jitter magnitude, we empirically
derive their relationship from the simulation results as follows:
" 14:5 PJ
5:5 C 0:5 log2 0:1
, if Slope > 18:5
PJ CRJ RJ
SlopeD
18.5, otherwise
A bit error occurs when the magnitude of the phase error between the input jitter
and the recovered clock jitter exceeds half of a UI. Thus, we can estimate the BER
once we know the distribution of the phase error between the input jitter and the
recovered clock jitter.
When only PJ is present in the data, the phase error distribution can be approx-
imated using a Uniform distribution with a peak-to-peak value of 2 ™max . ™max
can be calculated from Eq. 4.3. This approximation correctly reflects the bounded
property of the phase error distribution and is analytically convenient for BER
estimation. When both RJ and PJ are present, the phase error distribution can be
approximated by the convolution of a Uniform and a Gaussian distribution because
most of RJ components are not tracked by the CDR circuit. Since the addition of
the RJ changes the jitter transfer function, we need to modify Eq. 4.3 for calculating
the magnitude of the Uniform distribution. To incorporate the bandwidth variation,
we simply substitute ™in;p in Eq. 4.3 by ™in;TJ . For the slope variation, we introduce a
4.3 Experimental Setup and Results 49
gain compensating factor that takes into account the gain difference in the slewing
region, which is introduced by the addition of the RJ:
SlopeC20
20
w™
Kslope .w™ / D
w3 dB
We include Kslope .w™ / in the denominator of Eq. 4.3 because the maximum phase
error is inversely proportional to the jitter transfer gain. As a result, the maximum
phase error becomes
q
4w2™ ™in;TJ
2
2 K2 2
VCO IP R2
™max TJ D (4.4)
2w™ Kslope .w™ /
Knowing the magnitude of the Uniform distribution (i.e. ™max TJ / and the rms
value of the Gaussian distribution (i.e. ¢RJ /, the PDF of the phase error can be cal-
culated by the convolution of these two distributions:
Z
1 1 xCATJ
2 =2¢ 2
P™ .x/ D p et RJ dt ;
2ATJ 2 ¢RJ xATJ
where ATJ is the maximum phase error in UI (i.e. ATJ D ™max TJ =.2 //. In turn,
the BER, which represents the probability that the phase error exceeds half of the
UI, would be:
" Z 0:5CATJ
1 1 2 2 2
2 .0:5CATJ /2 =2¢RJ
BER D 1 p .0:5 C ATJ / et =2¢RJ dt C ¢RJ e
2ATJ 2 ¢RJ 1
Z 0:5ATJ
2 2 2
2 .0:5ATJ /2 =2¢RJ
.0:5 ATJ / et =2¢RJ dt ¢RJ e (4.5)
1
In the experiment, PJ and RJ are injected into the transmitted data, which is then
applied to the input of the CDR circuit. The BER is measured by comparing the
timing information of the input data to the recovered clock signal. The amplitudes
and frequencies of the PJ and the rms values of the RJ were adjusted so that the
resulting BER fell within the desired BER levels. Due to the limited simulation
time, we could only afford to conduct simulations down to the 107 BER level.
For each case, we conducted the simulation to the point at which 1,000 errors were
captured. Then, we compared the error between the estimated BER (i.e. estimated
using Eq. 4.5) and the simulated BER (i.e. measured from simulation).
We chose two different rms values of RJ, 1/10 and 1/8 UI. The peak-to-peak PJ
magnitudes were chosen at 0.4 UI (i.e. Case 1) and 0.8 UI (i.e. Case 2) for the 1/10
UI RJ case, and 0.4 UI (i.e. Case 3) and 0.6 UI (i.e. Case 4) for the 1/8 UI RJ case.
These cases cover the BER range from 107 to 102 . Four different PJ frequencies
were chosen for each case, one of which was set to be located in the linear operation
region.
1.0E-04
1.0E-03
BER
BER
1.0E-05
1.0E-04
1.0E-06
1.0E-07 1.0E-05
4MHz 10MHz 15MHz 20MHz 4MHz 10MHz 15MHz 20MHz
PJ Freq. PJ Freq.
Est. BER w/o comp. Est. BER w/ comp. Sim. BER Est. BER w/o comp. Est. BER w/ comp. Sim. BER
1.0E-01 1.0E-02
Case 2 Case 4
1.0E-02
1.0E-03 1.0E-03
BER
BER
1.0E-04
1.0E-05 1.0E-04
1.0E-06
1.0E-07 1.0E-05
2MHz 6MHz 7MHz 8MHz 2MHz 8MHz 10MHz 12MHz
PJ Freq. PJ Freq.
Est. BER w/o comp. Est. BER w/ comp. Sim. BER Est. BER w/o comp. Est. BER w/ comp. Sim. BER
Figure 4.9 shows the simulation results of these cases. Each graph contains the
simulated BER and two estimated BERs: one is the estimated BER without com-
pensating the jitter transfer variations due to the RJ (i.e. Substituting Eq. 4.3 into
Eq. 4.5) and the other is the estimated BER with compensation of these effects (i.e.
Substituting Eq. 4.4 into Eq. 4.5). The simulation results indicate that for all cases,
the simulated BER and the estimated BER, after compensating the jitter transfer
variations, match pretty well (see Fig. 4.10).
4.4 Summary
In Chapter 3, we propose a BER estimation technique using the jitter spectrum in-
formation and the characteristics of the CDR circuit, which is implemented using a
linear PD. In this chapter, we extend the technique using the same set of informa-
tion for BB clock and data recovery circuits, which exhibit non-linear characteristics
with respect to the input jitter. We show the dependency of a BB CDR loop’s jitter
transfer function on the magnitudes of PJ and RJ. We further propose an analytical
technique to estimate the BER. The simulation results demonstrate the validity of
our analysis. We will further conduct hardware validation, using suitable measure-
ment equipment to assess the accuracy of the proposed estimation method.
Chapter 5
Gaps in Timing Margining Test
With the challenges of keeping testing costs down, DFT-based testing methods have
been pursued [6–9, 47, 48]. The timing margining test is one of the widely adopted
DFT methods by companies [8, 9]. Usually facilitated by a loopback on-die or off-
die, the idea is to assess the margin in a given I/O’s timing and make the pass/fail
decision. This test coupled with additional DFT testing methods (e.g. DFT specific
to clock data recovery circuitry as noted in [9]) enables computer product manu-
facturers to ignore the communications style testing, which involves expensive ATE
equipment and long testing times. Still, there continues to be one nagging question,
“Are we missing anything gross by relying heavily upon timing margining test?”
We decided to study this question and this chapter summarizes our results. The
next section gives a primer on timing margining. The core of the chapter explains
the testing method’s gap, which include random jitter, non-linear clock recovery
circuitry, jitter amplification, and duty cycle distortion.
Defect-based I/O screens using timing margining [8,47] methodology has been em-
braced by industries as an alternative to traditional functional timing test. The basic
goal of timing margining is to measure the amount of margin within a data eye.
For a source synchronous interface such as Intel’s Front-Side Data Bus, the distance
between the strobe and data eye edges are measured. Ideally, the strobe would be
centered on the data eye, and the right/left margin would be 1=2 UI (Unit Interval,
ideal data eye) each. In reality, because of jitter, circuit non-idealities, setup/hold
time, trace length mismatch, and clock recovery inaccuracy, these margins are re-
duced and unequal.
Figure 5.1 shows a simplified diagram of a typical timing margining implemen-
tation for a source-synchronous data bus. In normal mode, the data is fed from the
core and is transmitted to another agent. The strobe is also generated and the de-
lay cell is set to 0. In this mode the strobe should be centered in the data eye. In
timing margining mode, we use a loopback configuration which makes the strobe
latch its own data back to the inbound section. There, it is compared to the sent data
D. Hong and K.-T. Cheng, Efficient Test Methodologies for High-Speed Serial Links, 53
Lecture Notes in Electrical Engineering 51, DOI 10.1007/978-90-481-3443-4 5,
c Springer Science+Business Media B.V. 2010
54 5 Gaps in Timing Margining Test
Data clock
IOLB enable
Pattern
generator
Pass/ Data
Fail Error pad
detector
Data to core
Fig. 5.1 Timing margining general approach for source synchronous pins
and a pass/fail decision is made. By moving the strobe with respect to the data, the
delay cell allows a timing stress. In addition, this movement combined with mon-
itoring the pass/fail result, the margin between the original strobe position and the
data edge can be found. This assumes that the relationship between the strobe delay
setting and the actual delay is known and reliable.
The same concept can be applied to various single-ended I/O configurations such
as common-clock, DDR, and CMOS, as well as serial interfaces such as Serial ATA,
FBD, PCI Express. Most serial interfaces already have a phase interpolator (PI) in
the receiver circuitry, which can be re-used to perform timing stress on the received
data eye. Timing margining can also be applied in a non-loopback configuration by
sending data from one chip to another chip, if a pattern comparison protocol, such
as IBIST [49], has been implemented.
The timing margining test has been successfully deployed for testing Front Side Bus
at Intel [8, 47] and at other companies [9]. However, it is uncertain that the timing
margining test can completely replace the conventional jitter testing methods, such
as jitter measurement and jitter tolerance, for high-speed serial interfaces running at
multiple Gbps. In this section, we investigate possible sources that might generate
some gaps between the timing margining test and the conventional jitter test.
5.2 Gap Analysis in Timing Margining Test 55
Jitter is commonly divided into DJ and RJ. One of the motivations for separating
jitter into these classifications is to extrapolate system performance without direct,
time-consuming measurement. DJ is bounded and represented by a peak-to-peak
value, while RJ is unbounded and its peak-to-peak value highly depends on the
measurement time. Since RJ is best described by a Gaussian distribution, it is com-
monly represented by an rms value. Then, total jitter (TJ) at a certain BER level
is estimated using the peak-to-peak value of DJ (DJpp ) and the rms value of RJ
(RJrms ) as follows [27]:
the timing margining result. In order to cope with this problem, timing margining
can be performed at two or more different BER levels to separate the RJ and the DJ.
Non-linear behaviors of a system can also generate the gaps between the two testing
methods. For example, we can observe either no or a very small amount of eye
closure when the system has less than a certain amount of jitter. However, a slight
increase in jitter can suddenly close the eye due to the non-linear characteristics of
the system, instead of gradually closing it. In this situation, the timing margin is
not correlated with the jitter margin. An example of this non-linear behavior is a
PLL-based clock recovery circuit with a bang-bang phase detector (PD).
As discussed in Chapter 4, the bang-bang PD discards the magnitude information
of the timing error and only measures the polarity of the timing error. Since the
output of the bang-bang PD is constant regardless of the phase error, it can be greater
for small errors, or less for large errors than the output of the linear PD as shown in
Fig. 5.3. In other words, if the phase error is small (Ø1 in Fig. 5.3), the bang–bang
PD has greater gain than the linear PD. On the other hand, if the phase error is large
(Ø3 in Fig. 5.3), the bang–bang PD has less gain than the linear PD.
Since the effective gain of the bang–bang PD strongly depends on the phase error
(i.e. input jitter magnitude), the loop dynamics of the clock recovery circuit with
bang–bang PD also varies with the input jitter magnitude [44–46]. In other words,
5.2 Gap Analysis in Timing Margining Test 57
Ø1 Ø2 Ø3 PDin
Bang-bang
Linear
the loop bandwidth decreases as the input jitter increases because of the decrease in
the effective PD gain. In addition, jitter peaking suddenly increases when the jitter
magnitude exceeds a certain value because the PD output can have a long string of
either C1s or 1s, as shown in Fig. 5.4 [50]. This peaking causes severe resonances
at certain frequencies, and thus the system can be broken abruptly once the jitter
magnitude surpasses the threshold.
Bathtub curve simulations were conducted to validate this phenomenon. The
behavioral models of PLL-based clock recovery circuits are implemented using
MATLAB with either linear PD or bang–bang PD. The loop parameters were chosen
to reflect the jitter transfer characteristics shown in Fig. 5.4. Then, we injected sinu-
soidal jitter with 2 MHz frequency into the circuits, and plotted the bathtub curves
by gradually increasing the magnitude of the sinusoidal jitter. For the linear clock
recovery circuit, we varied the magnitude of the jitter from 3 to 5 UI with a 0.4
58 5 Gaps in Timing Margining Test
Fig. 5.5 Bathtub curve comparison: (a) for linear clock recovery circuit, (b) for non-linear clock
recovery circuit
UI step size. For the non-linear clock recovery circuit, the magnitude of the jitter is
varied from 4 to 6 UI with the same step size.
Figure 5.5 represents how the bathtub curves vary with the jitter magnitude for
both the linear and the non-linear clock recovery circuits. The non-linear case sud-
denly closes the eye when the jitter exceeds a certain level, while the other closes
the eye linearly as jitter increases.
These results indicate that for non-linear clock recovery circuits, the timing
margining test alone might not detect how close the system is to the breaking point.
In other words, even though the eye-opening passes within a high volume manufac-
turing (HVM) testing environment, the system might not work in real applications if
the jitter level is very close to the breaking point. Thus, jitter injection might be re-
quired on top of the timing margining test to address this problem, especially when
the system has a non-linear clock recovery unit.
5.2 Gap Analysis in Timing Margining Test 59
As a HVM testing environment is different from a platform environment, the test cri-
teria is usually determined more stringently with respect to the device specification
by having some guardbanding. Typical timing margining test conditions are cleaner
than those in the platform because there are no power/substrate noises from other
devices in the test environment and the loopback channel is shorter. If the jitter
caused by additional sources is simply added to the intrinsic jitter of the device un-
der test (DUT), guardbanding to the test specification can guarantee the operation of
the device in the platform. However, if jitter is amplified through some circuitry or
channel, the simple guardbanding might not always ensure the proper screening of
bad devices in HVM testing.
Jitter amplification mainly occurs due to the bandwidth limitation of devices (e.g.
clock buffer, delay cell, or channel). Specifically, the gain difference between the
signal and the jitter component causes jitter amplification. Frequency domain anal-
ysis can provide insight on jitter amplification. The clock waveform with jitter, ¥.t/,
can be modeled by the phase modulated sinusoid [51]:
When the magnitude of the jitter is small, the above equation can be approximated
as follows:
s.t/ D Ac cos.2 fc t/ C Ac sin.2 fc t/ ®.t/
If we denote the jitter as ®.t/ D Aj cos.2 fj t/, s(t) become
Ac Aj ˚
s.t/ D Ac cos.2 fc t/ C sin.2 .fc C fj /t/ C sin.2 .fc fj /t/ (5.2)
2
Figure 5.6 represents the frequency spectrum of the jitter components based on
Eq. 5.2. As can be seen, the low frequency component of the jitter is located close
fc Freq. fc Freq.
to the signal, while the high frequency component of the jitter is located relatively
farther away. Thus, when the signal passes through a bandwidth limited device, the
signal and the jitter component are attenuated differently. The high frequency jitter
in particular, which is located in the low frequency region in the spectrum, is not
attenuated as much as the signal component. This relative gain difference causes
jitter amplification, which can be expressed as Eq. 5.3 for a linear channel H(f) [52].
.H.fc C fj / C H.fc fj //
Jitter Amp D (5.3)
2H.fc /
There are several jitter sources in the system, such as power supply noise, thermal
noise, and the duty cycle distortion (DCD). The power supply noise, which usually
relates to package resonance frequency, is typically on the order of a hundred mega-
hertz. Since random jitter is distributed in the wide frequency range, most of the
jitter is attenuated with the signal. Among the jitter sources, DCD has the highest
frequency component, whose frequency is the same as the clock signal. Thus, the
DCD results in the greatest jitter amplification.
In order to quantify the jitter amplification difference between the testing and the
platform environments, we compared the frequency responses of the tester board
channel and the platform channel as shown in Fig. 5.7. The tester board channel has
13 dB gain at 5 GHz, while the platform channel has 30 dB gain at the same fre-
quency. If we calculate the jitter amplification factor based on Eq. 5.3 for a 5 GHz
clock, the tester channel amplifies the DCD 2.4 times while the platform channel
amplifies it 16 times. This huge difference in jitter amplification can cause a signif-
icant gap in the timing margining between the two different environments. For the
Tester board
17dB Gain
difference Platform
at 5GHz
data loop-back channel, the jitter injection filter can be used to emulate the system
channel [22]. This can be extended to the clock channel for source-synchronous
interfaces to mimic the jitter amplification effect.
In general, since data is widely spread out through the frequency range from DC
to half of the data rate, the expected jitter amplification through the data would be
less than that of the clock. In addition, most of the deterministic jitter components
in the data, such as inter symbol interference (ISI) and DCD, will be captured in the
timing margining test because it directly closes the eye-opening. However, DCD in
the clock does not directly close the eye-opening and thus it might not be captured
by the timing margining test. We investigated the impact of DCD in the clock on the
eye-opening and the timing margining results.
If the clock has DCD, the clock recovery circuit in a receiver, which is com-
monly implemented by a delay-locked loop (DLL)/PLL, may or may not detect the
DCD depending on the architecture of the phase detector. Since many PD architec-
tures used in DLLs/PLLs are designed to be insensitive to the DCD [53, 54], the
DCD in the clock is directly transferred to the output of each delay cell. In order
to control the phase offset for the timing margining test, the output of each delay
cell is connected to the PI. Then, the PI chooses two adjacent phases to generate the
main clock by mixing the phases. The number of delay cells and the resolution of
the PI greatly depend on architectures. In our analysis, the DLL is assumed to have
four delay cells to generate eight different phases with 45ı separation as shown in
Fig. 5.8. In addition, the phase mixer is assumed to generate seven steps within a
45ı range, which gives a 6:4ı .D45ı =7/ phase resolution.
Vcontrol
CLKin
Delay Delay Delay Delay Delay
cell cell cell cell cell
CLKinb
MUX
PI
Phase Mixer
CLKout CLKoutb
0o
45o
90o
135o
180o
225o
270o
315o
When the clock has DCD, all eight phases from the delay cells have the same
amount of DCD because the phase detector does not detect the DCD. However, the
phase separations between adjacent phases are not all the same, as shown in Fig. 5.9.
In other words, the phase separations between true (complementary) signals are not
affected by the DCD. On the other hand, the phase separations between a true signal
and a complementary signal (e.g. 135ı and 180ı, 315ı and 0ı ) are affected by the
DCD. One of them (135ı and 180ı ) has a greater phase separation, while another
one (315ı and 0ı ) has less phase separation.
Therefore, if the PI chooses a true signal and a complementary signal for the
clock generation, the step size will be narrower/wider than the ideal. This non-ideal
step size can affect the timing margining results, as shown in Fig. 5.10. We analyzed
this effect for three different cases:
Case 1: The region which has wider phase separation is used to measure the
timing margin.
Case 2: The region which has narrower phase separation is used to measure
the timing margin.
Case 3: Non-ideal regions are located in the jittery region (i.e. only ideal
regions are used to measure the timing margin).
For Case 1 (shown in Fig. 5.10a), the timing margin result is decreased compared
to the case when the clock does not have the DCD, while it is increased for Case 2, as
shown in Fig. 5.10b. For Case 3 (shown in Fig. 5.10c), the timing margining result is
not affected by the DCD. Since the timing margining results can vary case by case,
the DCD in the clock may not be detected by the timing margining test.
In addition, the testing can give wider eye-opening results than the original eye-
opening. In some cases (i.e. Case 2), this might result in passing bad devices in
HVM testing. In order to address this problem, the duty cycle detection/correction
circuitry should be required to minimize the duty cycle error in the clock. This is
becoming more common place.
5.3 Summary and Future Work 63
a
Jitter
UI/4 UI/4 UI/4 UI/4
No DCD
(20 steps)
W/DCD
(17 steps)
No DCD
(20 steps)
W/DCD
(23 steps)
No DCD
(20 steps)
W/DCD
(20 steps)
Fig. 5.10 Comparisons of the timing margining measurement: (a) Case 1, (b) Case 2, and (c)
Case 3
Realizing that switching from one test method to another may result in gaps; we
explored possible misses by relying upon timing margining test alone. By compar-
ing against standard jitter measurements, we determined that RJ dominant systems
are not covered by a single timing margining measurement. However, we note that
64 5 Gaps in Timing Margining Test
computer products are dominated by DJ and we could augment the current testing
approach by taking two or more measurements for different BER levels. Depending
upon the type of data clock recovery circuitry, there are test vulnerabilities. Non-
linear CDR circuitry, which is based on bang-bang PLL implementations, can have
abrupt changes in data eyes with increasing jitter which would not be covered by
timing margin test. Hence, their inherent sensitivity would require additional test
coverage. While jitter injection is one means to address, an area for future work
is to develop a DFT-based methodology for non-linear CDR circuitries. With re-
spect to jitter amplification, the noted differences between the end-use system and
the test system results in potential gaps for which others have suggested techniques
[49]. Finally DCD can adversely affect the timing margining measurement itself.
A preventive measure such as duty cycle detection and correction circuitry could be
implemented.
Chapter 6
An Accurate Jitter Estimation Technique
One of the popular and efficient approaches for testing a high-speed transceiver
is based on the loopback principle. Even though the BER can be measured in the
loopback mode using a pattern generator and an error detector in a transceiver, only
the I/O performance margin (e.g. timing margin) is guaranteed in production testing
due to excessively long testing time for low BER measurement [8, 9]. However, as
we illustrated in Chapter 5, such a test might not detect some chips that fail the BER
specification when the RJ accounts for a non-trivial fraction of the TJ.
In this chapter, we introduce a new technique for estimating TJ, which can be
incorporated into the existing loopback-based test method without a significant in-
crease in testing time. An estimation technique based on the dual-Dirac model has
been widely used. This technique intends to quickly predict the TJ at low BER lev-
els based on jitter or BER measurement data taken at higher BER levels. However,
the technique has a few problems: (1) The estimation accuracy decreases as the de-
terministic jitter (DJ) characteristics become increasingly complex due to ISI and
crosstalk. (2) The estimation accuracy is extremely sensitive to the BER region
which the prediction is based upon. In order to cope with these problems, we pro-
pose a high-order polynomial fitting technique that can accurately estimate the TJ
down to a 1012 BER level using the information from a higher BER region.
In the next section, we first describe the DJ characteristics due to ISI and
crosstalk, which are the major contributors to the DJ for multi-Gbps links.
Section 6.2 presents the limitations of the estimation technique based on the dual-
Dirac model, followed by the description of the proposed technique. Section 6.3
concludes the chapter.
6.1 Characteristics of DJ
DJ can be divided into DDJ and PJ. PJ can be easily characterized by sinusoidal
functions. Modeling DDJ is much more difficult because it is highly data-pattern
dependent. In this section, we analyze the characteristics of DDJ by assuming the
application of commonly used PRBS patterns. ISI and crosstalk are two major
sources of DDJ. The bandwidth limitation of a channel generates significant ISI
D. Hong and K.-T. Cheng, Efficient Test Methodologies for High-Speed Serial Links, 65
Lecture Notes in Electrical Engineering 51, DOI 10.1007/978-90-481-3443-4 6,
c Springer Science+Business Media B.V. 2010
66 6 An Accurate Jitter Estimation Technique
due to an increased data rate. In addition, increasing the interconnect density makes
the crosstalk between lanes a significant contributor to DDJ. In the following, we
describe how to characterize both ISI- and crosstalk-induced jitter.
Channel bandwidth limitation causes ISI. Channel loss due to skin-effect and di-
electric loss causes signal dispersion, and thus the signal is distorted by its own
delayed versions. In other words, the signal is affected not only by the current bit
but also by a number of previous bits. For each signal transition, the previous bits
shift the signal amplitude, which, in turn, changes the relative time in which the
signal crosses a decision threshold. The amount of timing deviation is determined
by the data sequence and the channel bandwidth. The channel bandwidth affects the
memory depth of the channel, which determines the number of previous bits that
would have an impact on the threshold crossing time. If the bandwidth is low, more
bits affect the crossing time. As described in [55], if only the penultimate bit has a
non-negligible effect, there are two distinct sets of sequences of interest – 010 and
101, for which the signal transition would arrive early, and 001 and 110, for which
the signal transition would arrive late. Thus, there are only two delta functions in the
jitter histogram. If the signal speed increases and the relative channel bandwidth de-
creases, more previous bits affect the crossing time. Then, the DDJ jitter histogram
includes more delta functions. Since a PRBS pattern has a certain property, we can
easily derive the DDJ characteristic resulting from the ISI. Figure 6.1a illustrates
this property of a 4-bit PRBS pattern: for each memory depth, the PRBS pattern
has the same number of occurrences for each distinct set of sequences. Because of
this property, it results in the same probability for each delta function in the jitter
histogram. If the bandwidth limited channel causes a memory depth of n bits, the
jitter histogram would have 2n delta functions, each of which has an identical prob-
ability. Thus, ISI-induced jitter can be approximated as a uniform distribution, as
illustrated in Fig. 6.1b.
Crosstalk is the energy coupling from one trace to another, which could be caused
by either capacitive or inductive coupling between adjacent lanes. Figure 6.2a illus-
trates three different crosstalk modes between the coupled transmission lines [56]:
odd, superposition, and even modes. In the odd mode, the data transition in the ag-
gressor is opposite to that in the victim. In the even mode, the data transitions are
the same in both the aggressor and victim. In the superposition mode, the aggressor
does not have any data transition, and thus there is no crosstalk. If there is only one
6.1 Characteristics of DJ 67
a
4-bit PRBS pattern: 000111101011001
Memory depth Patterns # of occurrences
001 / 110 4
1
101 / 010 4
0001 / 1110 2
0101 / 1010 2
2
1001 / 0110 2
1101 / 0010 2
b
Probability
Jitter
Fig. 6.1 Characteristics of ISI induced jitter: (a) edge pattern probability for a 4-bit PRBS pattern,
(b) jitter distribution due to ISI resulting from an n-bit PRBS pattern
Victim 0 1 0 1 0 1
b
Probability
Jitter
Fig. 6.2 Characteristics of crosstalk induced jitter: (a) different modes of crosstalk, (b) jitter
distribution due to crosstalk
68 6 An Accurate Jitter Estimation Technique
aggressor line in a system, the PDF of the jitter due to crosstalk consists of three
delta functions, represented in [56] as:
1 1 1
PDFXtalk D •.JXtalk / C •.0/ C •.JXtalk /; (6.1)
4 2 4
where JXtalk is the jitter amount due to crosstalk. The probability of having no
crosstalk (i.e. the superposition mode) is twice the probability of getting speed-up
(the even mode) or slow-down (the odd mode). The analysis can be easily extended
for a case in which a victim trace is sandwiched between two aggressor traces. The
jitter PDF can be derived by the convolution of Eq. 6.1
1 1 3 1 1
PDFXtalk D •.2JXtalk / C •.JXtalk / C •.0/ C •.JXtalk / C •.2JXtalk /
16 4 8 4 16
The resulting PDF has five delta functions, as shown in Fig. 6.2b. In general, the
jitter histogram due to crosstalk has a higher probability around the middle (i.e. no
or little jitter) and the probability decreases as jitter increases [57]. Therefore, we
can approximate the jitter histogram as a triangular distribution.
The dual-Dirac model has been widely adopted to separate DJ and RJ for estimating
TJ at low BER levels. In this model, DJ is approximated by two delta functions and
RJ is assumed to have a Gaussian distribution that is represented by an rms value.
In order to easily separate DJ and RJ from the measured bathtub curve, the verti-
cal axis of the plot is changed from the BER to the Q-scale. The advantage of using
the Q-scale is that a Gaussian jitter distribution becomes a straight line in the Q
domain. Therefore, a simple first-order line fitting can separate RJ from DJ when
an appropriate fitting region is chosen. The relationship between the Q-function and
the BER can be represented as [27]:
p
Q.x/ D 2 erfc1 .BER.x//; (6.2)
Z1
1 t2
erfc.x/ D p exp. /dt
2 ¢ 2¢ 2
x
By measuring the BER at different positions in the eye and deriving their corre-
sponding Q values using Eq. 6.2, the bathtub curve can be plotted using the Q-scale.
6.2 Total Jitter Estimation 69
Figure 6.3 shows the bathtub curve in the Q-domain. The solid line gives the
distribution from an actual measurement. The dashed lines represent the estimation
results using the dual-Dirac model from two different fitting regions. As the fig-
ure indicates, the estimation accuracy is extremely sensitive to the choice of fitting
region. Choosing a lower BER region for fitting results in a more accurate estima-
tion. This is because we only use the tail of the distribution, which closely follows
the true Gaussian distribution for fitting. However, measuring error rates at low BER
levels demands longer testing time and, in turn, greater testing cost.
We conducted MATLAB simulations to find out the appropriate fitting region for
an accurate estimation of TJ at a 1012 BER level (i.e. Q D 7). However, generating
a bathtub curve down to the 1012 BER level is almost impossible in the simulation
environment due to the excessive simulation time. Instead, we derived closed form
equations for the BER distributions. We generated the distributions for three differ-
ent DJ cases: (1) Case 1: DJ has only an ISI component, and thus follows a uniform
distribution. (2) Case 2: DJ has only a crosstalk component, and thus follows a tri-
angular distribution. (3) Case 3: DJ has both ISI and crosstalk components, and thus
its distribution results from the convolution of uniform and triangular distributions.
Then, we applied the first-order line fitting by choosing different BER regions to
estimate the TJ, and compared the results with the actual TJ at a 1012 BER level.
The step size for generating the bathtub curve was set to 1/64 UI (the PI in a mod-
ern transceiver can generate 64 steps within a UI [9, 58]). We chose four different
BER ranges as the fitting regions 102 104 , 103 105 , 104 106 , and
105 107 . We conducted experiments for the three DJ cases. Figure 6.4 shows
the results when RJ is dominant (i.e., RJ has 0.05 UI rms, DJ has 0.07 UI peak-to-
peak). Figure 6.5 shows the results when DJ is dominant (i.e., RJ has 0.01 UI rms,
DJ has 0.35 UI peak-to-peak). The tables list the differences between the estimated
TJ and the actual TJ.
70 6 An Accurate Jitter Estimation Technique
Case 1 Case 2 Case 3
0 0 0
TJ TJ TJ
1e-2~1e-4 1e-2~1e-4 1e-2~1e-4
1 1e-3~1e-5 1 1e-3~1e-5
1 1e-3~1e-5
1e-4~1e-6 1e-4~1e-6 1e-4~1e-6
2 2 1e-5~1e-7 2 1e-5~1e-7
3 3 3
Q
Q
4 4 4
5 5 5
6 6 6
7 7 7
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
Time (UI) Time (UI) Time (UI)
3 3 3
Q
4 4 4
5 5 5
6 6 6
7 7 7
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
Time (UI) Time (UI) Time (UI)
When RJ is dominant, the estimation error for different fitting regions does not
vary too much. However, when DJ is dominant, the amount of error is very sensitive
to the fitting region. In addition, the amount of error in DJ Cases 2 and 3 is greater
than that of Case 1. Thus, for Case 1, measuring the BER down to the 105 level
is sufficient for estimating the TJ at the 1012 level (with less than 1% error). For
Cases 2 and 3, measuring the BER to at least a 107 level is required for sufficiently
accurate estimation.
6.2 Total Jitter Estimation 71
Our early analysis concluded that the TJ estimation using fourth-order polynomial
fitting results in very accurate results. In this section, we explore the number of
measurement points within the fitting region required to maintain the estimation
accuracy. In the earlier discussion, we assumed that the PI can generate 64 steps
within a UI. Thus, in the DJ dominant case, there are 24 points from the transition
72 6 An Accurate Jitter Estimation Technique
3 3
Q
Q
4 4
5 5
6 6
7 7
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
Time(UI) Time(UI)
edge to the 106 BER region. We reduced the number of BER measurements for
fitting to 12, 8, 6, 4, 3, and 2 by evenly skipping some PI steps. For each case, we
estimated the TJ using fourth-order fitting and compared the estimation error with
the original case which used all 24 points. Figure 6.7 shows the simulation results.
The results indicate that four points would be sufficient for estimating TJ in order to
achieve an error rate of less than 1.5%.
6.3 Summary
As BER requirements become more and more stringent, the industry demands new
testing methods, that are both efficient and accurate, to guarantee the product meets
the BER specification at a low test cost. In this chapter, we first explored the ap-
plicability of the dual-Dirac-model-based technique for estimating TJ to the widely
used loopback-based I/O margining test. Our conclusion is that the accuracy of the
former technique is highly sensitive to the choice of the fitting region. Also, as DJ
behaves more and more like a Gaussian distribution, this technique requires data to
be measured at lower BER levels to ensure reasonable accuracy. This requirement
implies longer testing time and, in turn, greater test cost. We propose the use of high
order polynomial fitting to improve both the efficiency and accuracy of the estima-
tion. This fitting strategy alleviates the need to measure the BER at a very low level,
6.3 Summary 73
and thus reduces the testing time. It also enables the use of all BER data points for
fitting to improve the estimation accuracy. The experimental results show that, for
a system affected by both RJ and DJ, fourth-order fitting using four BER measure-
ments, which are evenly spread from the transition region to the 106 BER region,
can accurately estimate TJ at the 1012 BER level.
Chapter 7
A Two-Tone Test Method for Continuous-Time
Adaptive Equalizers
With the increasing demand of higher bandwidth, the data rate of I/Os is approach-
ing the tens of gigahertz range. While the continuing advancement of process
technology enables an I/O chip to run at such frequencies, the bandwidth of the
communication channels, including cables and legacy backplanes, has become the
limiting factor.
The bandwidth limitation of the channel causes ISI. Various equalization tech-
niques, which multiply the inverse response of the channel to flatten out the overall
frequency response, have been developed to compensate for this channel effect. In
addition, the channel characteristics may not be known in advance and might be
time-variant [59]. To cope with such problems, several adaptation algorithms have
also been developed to adjust the overall system response depending on the channel
conditions.
The equalizer can be implemented either in the transmitter or in the receiver.
The implementation of the transmitter equalizer is relatively easier than that of the
receiver equalizer because the required Finite Impulse Response (FIR) filter deals
with the digital data at the transmitter side, rather than the received analog data at
the receiver side [60,61]. However, since channel information is not easily available
at the transmitter, it is difficult to apply the adaptive technique at the transmitter.
The approaches of equalization at the receiver can be divided into two categories:
discrete-time equalization and continuous-time equalization. A discrete-time equal-
izer, which is based on the FIR filter, can take advantage of various digital adaptive
algorithms [41,42,62,63]. However, since equalization is based on the samples cap-
tured by the receiver’s recovered clock, there exists a cross-dependence between the
equalizer and the clock recovery circuit. As the data rate increases, the power con-
sumption would increase dramatically due to the large number of taps implemented
in this type of equalizer [64]. On the other hand, a continuous-time equalizer does
not require a sampling clock and thus the equalizer would work independent of the
clock recovery circuit. Continuous-time equalizers have been investigated for low
power and high speed applications, and promising performance has been reported
[64–69].
While equalizer design has been studied for a long time and a number of novel
architectures have recently been developed, high-quality and cost-effective produc-
tion test methods for these equalizers are not yet well developed. The most popular
D. Hong and K.-T. Cheng, Efficient Test Methodologies for High-Speed Serial Links, 75
Lecture Notes in Electrical Engineering 51, DOI 10.1007/978-90-481-3443-4 7,
c Springer Science+Business Media B.V. 2010
76 7 A Two-Tone Test Method for Continuous-Time Adaptive Equalizers
Comparator
Equalization Output
filter
High pass
filter
Rectifier
− Σ +
冮 Integrator
for high-frequency loss due to the channel. The adaptive servo loop, which adjusts
the compensation gain of the equalization filter, determines the control voltage by
comparing the comparator’s input and the output signals. In practice, it is diffi-
cult to design a comparator that can generate a clean waveform for comparisons
at very high frequencies. Several new approaches for adaptation [67–69] have been
proposed to address this problem. These new methods use the power spectrum de-
rived from the output signal of the equalization filter for the adaptation, as shown
in Fig. 7.2. Since the power spectrum of a random signal can be described by a
sinc2 function, the high-frequency loss can be detected by comparing the power
densities of two different frequency ranges. Three different methods have been pro-
posed to compare the power spectrum of the random signal. Two band-pass filters
are used in [67] to compare the power of two specific frequencies. In [68], one
low-pass filter and one high-pass filter are used to compare the power between the
low-frequency and high-frequency portions of the signal. In [69], only one low-pass
filter is used and the entire signal’s power is compared to that of the low-frequency
portion of the signal. Figure 7.3 illustrates these three different power spectrum
comparison architectures. Our proposed test method is applicable to the modified
continuous-time adaptive equalizer (i.e. the type shown in Fig. 7.2) under any of
these three architectures.
Comparator
Equalization Output
filter
Power spectrum
comparison
Rectifier
− Σ +
a b c
The proposed test technique directly uses a two sinusoidal-tone signal as test stim-
ulus instead of a data pattern stressed through the channel as is commonly used
for validating and characterizing the equalizers. The frequency for one of the two
sinusoidal tones, denoted as fL , falls within the stop-band region; the frequency of
the other tone, denoted as fH , falls within the pass-band region. Figure 7.4 illustrates
the frequency response of an adaptive equalizer that can compensate for the chan-
nel loss up to ’0 dB, and the frequency bands of two sinusoidal tones. In order to
mimic the channel response, this technique repeatedly applies the two-tone signal
and gradually varies the magnitude ratio of fH and fL . Specifically, we gradually in-
crease the magnitude of fL while fixing the magnitude of fH . Such test stimuli mimic
different relative losses of the high-frequency components caused by the channel.
While this could also be accomplished by gradually reducing the magnitude of fH
and fixing the magnitude of fL , it is much easier in practical implementations to
adjust the magnitude of the low-frequency component.
With the test stimuli, we measure the rms value at the output of the equalizer.
The adaptive servo loop attempts to maintain the ratio of fL to fH to the expected
level based on the sinc2 function. Thus, as illustrated in Fig. 7.5, if the test stim-
ulus’s fL magnitude is within the range of R1 and, therefore, the magnitude ratio
of fL to fH is within the range of the equalizer’s maximum compensation gain (i.e.
GEQ M ax D 10˛0 =20 ), the rms value of the equalizer output should be a constant.
When the magnitude of fL is increased to the point that the magnitude ratio of fL
to fH exceeds the maximum compensation gain of the equalizer (as indicated in
Fig. 7.5, where the magnitude of fL is in the range of R2), the rms value at the
equalizer output will start to increase. Figure 7.6 illustrates how the rms value at the
Gain
(dB)
0
Control
voltage
-α0
log (f)
stop-band pass-band
Compensation Magnitude of
Gain fL is in R1 Constant rms
value
R2
R1
fL fH
Increased rms
fL fH value
fL fH
Magnitude of
fL is in R2
Input test tones Equalizer output
VExp
Compensable Uncompensated
equalizer output would vary with respect to the magnitude of the fL (with a fixed
magnitude of fH ). Therefore, we can test the equalizer’s maximum compensation
gain, which is a key design specification of the equalizer, by identifying the magni-
tude of fL when the rms value starts to deviate from the expected value, denoted as
AfL Max in Fig. 7.6. The adaptive loop also can be tested by the constant rms values
up to the AfL Max . Note that the conventional eye-diagram method may not easily
test these specifications because: (1) designing the test board channel to exactly
match the expected high-frequency loss is difficult. (2) A number of different chan-
nel lengths are required for testing the adaptive loop, and this may not be feasible
for production testing.
VC
LPF:
Adjustable
gain (GLPF)
Rectifier
Control Differential
Voltage power
(Vc) detector A B
+ Σ −
equalizer, since the rectifiers already exist in the architecture, we only need to add
an LPF to implement the rms detector.
There are two possible locations for measuring the rms value: either node A or
node B (denoted in Fig. 7.7). The rms measurement at node A would be the com-
bined rms value of fL and fH which is the same as the equalizer’s output rms value.
The rms measurement at node B would also result in a curve similar to that seen in
Fig. 7.6 because the magnitude of fH remains constant and only the magnitude of fL
varies.
We denote fL ’s magnitude after the equalization filter as AfL Out and fH ’s magni-
tude after the equalization filter as AfH . The adaptive servo loop attempts to make
the power of the two paths equal. That is:
Therefore, the expected ratio of AfH to AfL Out , denoted by KExp , becomes:
q
AfH
KExp D D G2LPF 1 (7.1)
AfL Out
Note that we modify both of the equations with respect to AfH using Eq. 7.1.
Then, the rms values for both cases are the same as expected because AfH is a
constant. Therefore, the expected rms value would be
q
VExp D GLPF AfH = 2G2LPF 2 (7.2)
In this region, since the equalizer cannot fully attenuate the low-frequency com-
ponent to the desired level, the rms value would be greater than the expected
level and would thus increase as AfL In increases. However, the increasing rate
would be different for these two locations. The increasing rate of the rms value
measured at B would be greater because it is multiplied by GLPF , as shown in
Eq. 7.3.
The magnitude of fL at which the rms value starts to deviate from the expected
value can also be represented using Eq. 7.1:
AfH
AfL Max D GEQ Max (7.4)
KExp
Therefore, by measuring the AfL Max , we can test both the equalization filter and the
adaptive servo loop. Any defects and/or errors that cause variations to GEQ Max and
KExp would be detected.
82 7 A Two-Tone Test Method for Continuous-Time Adaptive Equalizers
GEQ Max D 17 dB D 7
GLPF D 3:1
fH D 5 GHz and AfH D 80 mV
fL D 100 MHz and AfL In D 80 340 mV
Based on these parameters, the VExp and AfL Max can be calculated using Eqs. 7.3
and 7.4:
VRMS D 60 mV
AfL Max D 191 mV
Two sinusoidal signals are injected into the behavioral model of the equalizer and
the rms values are measured at A and B. For each location, the simulation is repeated
by gradually increasing the magnitude of the low-frequency component (AfL In /.
Figure 7.8 shows the simulation results for both cases. The rms values for both
cases are constant up to 180 mV and start to increase around 200 mV. The expected
AfL Max is about 190 mV. As the step size of AfL In in our experiment was 20 mV,
the rms value increases only slightly at 200 mV but increases rapidly starting from
220 mV. This result indicates that the proposed method could indeed identify the
device’s AfL Max fairly accurately. However, the increasing rate for measurement at
A is small the difference of the rms values between steps is only a few millivolts.
This difference would be too small to be detected in real applications, particularly
in the presence of various noise sources. For our test case, the expected ratio of AfH
to AfL Out is about 3. Therefore, the rms value of AfH , which is constant, would be
significantly larger than that of AfL Out . Thus, there would be a small variation to the
total rms value measured at A. Based on this observation, we implemented the rms
detector at B for transistor-level simulation. However, for designs with a small ratio
of AfH to AfL Out , the rms detector placed at A could be a viable solution, too.
We injected a couple of parametric faults into the equalizer for simulation and
examined the rms value at B to check whether the faults are detectable using this
technique. Specifically, we changed the maximum compensation gain of the equal-
izer to (1) 13 dB and (2) 10 dB from 17 dB. Note that defects/errors that cause the
equalizer to malfunction would cause changes to its maximum compensation gains.
Based on Eq. 7.4, if the maximum compensation gain is reduced, the rms value
would start deviating from the expected value at a lower magnitude. Figure 7.9
shows the simulation results of the three cases (fault-free circuit and faulty circuits
(1) and (2)). The maximum compensation gain’s variation can be easily identified
by detecting the change to AfL Max .
The continuous-time adaptive equalizer shown in Fig. 7.7 has been designed in a
0:13-m CMOS process. The design details of the tunable passive equalization filter
was reported in [69]. The die photo and the measured filter response are shown in
Fig. 7.10. The filter can compensate for a loss up to 17 dB at 5 GHz with a control
voltage ranging from 0.1 to 0.6 V.
The rms detector is implemented inside the differential power detector in the
servo loop (see Fig. 7.7). We reuse the rectifier in the servo loop and only add
a simple RC low-pass filter at the output of the rectifier. To avoid degrading the
Fig. 7.10 Die photo and the filter response: (a) chip micrograph of the 0:13-m filter prototype,
(b) measurement vs simulation results of the filter’s differential S21
7.4 Summary and Future Work 85
CL
βi1
2i βi1
Vb
Fig. 7.11 Schematic of the differential power detector with rms detector
servo loop’s performance, a current mirror is used to copy the output of the recti-
fier. Figure 7.11 shows the schematic of the differential power detector and the rms
detector.
Figure 7.12a shows the rms values measured before the simple RC LPF.
Figure 7.12b shows the DC output of the rms detector. For both figures, we com-
pare the nominal case with two defective cases. For the defective cases, the gain
of the servo loop’s LPF (GLPF ) was changed to 3.75 (Fault A) and 4.5 (Fault B),
respectively. This, in turn, also results in a different AfL Max . The rms values start to
deviate at a lower AfL In for the defective cases as expected from Eq. 7.4. Therefore,
any variation in the adaptive loop also can be easily identified by the change to
AfL Max .
Compared to the MATLAB simulation results, the transistor-level simulation
shows that the output rms values are not constant, even within the equalizer’s com-
pensation gain range. This is because the gain of the equalization filter at 5 GHz
slightly varies depending on the control voltage, which in turn contributes to the
variation in the expected rms value. However, since the difference in slope between
the two regions is obvious, we could still clearly identify AfL Max .
a
110
Nominal Fault A Fault B
100
RMS Value (mV)
90
80
70
60
50
80 100 120 140 160 180 200 220 240 260 280 300 320 340
Magnitude of the low frequency component (mV)
b
910
890 Nominal Fault A Fault B
870
850
Vout (mV)
830
810
790
770
750
730
710
80 100 120 140 160 180 200 220 240 260 280 300 320 340
Magnitude of the low frequency component (mV)
Fig. 7.12 Spectre simulation results: (a) measured rms value before the RC LPF, (b) DC output
of the rms detector
in the adaptive servo loop that might not be easily detectable by the conventional
eye-diagram method. We validated the idea by simulation using a recently proposed
continuous-time adaptive equalizer. The behavioral and the transistor-level simula-
tion results demonstrate the validity of our test technique.
In this chapter, we focused the discussion on testing a stand-alone continuous-
time adaptive equalizer. This technique could be extended for the adaptive
7.4 Summary and Future Work 87
This research has investigated efficient test methodologies for high-speed serial
links. We have shown in Chapter 2 that the comparator-based undersampling tech-
nique can efficiently measure the jitter using either ATE or on-chip circuitry due
to its simplicity and high bandwidth. The measurement results with the prototype
board demonstrate the accuracy of the technique compared to the conventional his-
togram method.
To expedite the BER testing and correlate the measured jitter information to the
BER, we use the BER estimation techniques proposed in Chapters 3 and 4. These
techniques utilize the jitter spectral information and the characteristics of the CDR
circuit for BER estimation. The high-speed serial links that incorporate the linear
CDR circuit are first analyzed based on the linear PLL theory. Experimental results
comparing the estimated BER and the BERT-measured BER on a 2.5 Gbps com-
mercial CDR circuit demonstrate the high accuracy of the proposed technique. In
Chapter 4, the BER estimation technique is extended to the serial links with a BB
CDR circuit. Due to the heavily non-linear nature of the BB CDR circuit, the jitter
transfer characteristics are analyzed depending on the jitter magnitude. Then, the er-
ror occurring mechanism of the non-linear loop is analyzed for the BER estimation.
Loopback-based margining test has been widely adopted by industry because
it can test most I/O functionality without requiring expensive ATE. However, it
cannot fully replace conventional jitter test due to the random jitter, the non-linear
characteristics of a CDR circuit, jitter amplification, and the DCD in clock signal.
In order to cope with the RJ underestimation problem in timing margining test, we
propose the total jitter estimation technique in Chapter 6. The experimental results
show that fourth-order fitting using only four BER measurements, which are evenly
spread from the transition region to the 106 BER region, can accurately estimate
the timing margin at the 1012 level.
Adaptive equalizers have been widely used in most advanced transceivers to
compensate for channel loss. In Chapter 7, we propose a two-tone test method
for continuous-time adaptive equalizer, which can test both the equalizing filter
and adaptation loop with only a very small area overhead and no performance
degradation.
With an ever increasing demand for high bandwidth in communication systems,
recent transceivers include more functionality in their designs which results in more
D. Hong and K.-T. Cheng, Efficient Test Methodologies for High-Speed Serial Links, 89
Lecture Notes in Electrical Engineering 51, DOI 10.1007/978-90-481-3443-4 8,
c Springer Science+Business Media B.V. 2010
90 8 Conclusions
This appendix describes how to extract effective sinusoidal and random components
from the jitter histogram. The jitter histogram, which represents the jitter Probabil-
ity Density Function (PDF), is commonly used for characterizing the jitter. In order
to estimate the BER when both sinusoidal and random components are present and
contributing to errors, the PDF of the sum of these components should be known.
The PDF of the sum of two random variables is the convolution of the individ-
ual PDFs. However, there is no closed-form expression for the convolution of a
sinusoidal PDF and a Gaussian PDF. Therefore, we can approximate the total jitter
PDF using the double delta model. Figure A.1 shows the convolution of a sinusoidal
distribution and a Gaussian distribution. Figure A.1a illustrates the case when the
random term is dominant, and Fig. A.1b shows the case when the sinusoidal term is
dominant.
When RJ is dominant, the distribution follows the Gaussian distribution. On
the other hand, when the sinusoidal term is dominant, the jitter distribution has
two peaks.
Given a jitter histogram, in order to separate the jitter components, we can as-
sume the tail part of the given distribution is mostly determined by the random
component. Then, fitting it with a Gaussian curve would result in the rms value of
the random component, as illustrated in Fig. A.2. The distance between the means
of these two fitted Gaussians, 2Aeff , gives the effective peak-to-peak amplitude of
the PJ, which can be treated as the distance between two Dirac delta functions [14].
Note that this value is different from the injected peak-to-peak amplitude of the
PJ [27].
We ran MATLAB simulations to quantify the variations of Aeff and ¢eff as a
function of the ratio of the amount of RJ to the amount of the PJ. First, we computed
the convolution of a sinusoidal PDF and a Gaussian PDF by varying the rms value
of the random component with a fixed amount of PJ. Using the tail fitting algorithm
described in [14], we calculated the Aeff and ¢eff . Figure A.3 shows the simulation
results as a function of the ratio of the rms value of RJ, ¢RJ , to the rms value of PJ,
¢PJ . Figure A.3(a) shows the RJ rms error normalized with respect to ¢PJ , which is
defined as: q
¢RJ err D ¢eff 2 ¢RJ 2
91
92 A Extracting Effective PJ and RJ Components from Jitter Histogram
* σRJ = σeff
–APJ APJ
* σRJ =
–APJ APJ
Fig. A.1 The convolution of two PDFs: (a) when random distribution is dominant, (b) when
sinusoidal distribution is dominant
Fitted curve
σeff
Figure A.3(b) shows the ratio of Aeff to APJ with respect to the ratio of ¢RJ =¢PJ .
Based on the Fig. A.3, the characteristics of ¢eff and Aeff can divide into two cases
depending on the ¢RJ =¢PJ .
1. When ¢RJ =¢PJ > 1 (the case illustrated in Fig. A.1a)
In this case, the ratio of the RJ rms error to the rms value of PJ approaches one,
which means the effective rms value is increased by the amount of the rms value of
the PJ. In addition, Aeff is almost zero. Thus, the total jitter can be treated as pure
random jitter (i.e. Aeff D 0), which has the following rms value:
q
¢eff D ¢RJ 2 C ¢PJ 2 : (A.1)
a
1
0.8
σRJ_err
0.6
σPJ
0.4
0.2
0
0 0.5 1 1.5 2 2.5 3 3.5 4
σRJ / σPJ
b
1
0.9
0.8
0.7
0.6
Aeff
APJ
0.5
0.4
0.3
0.2
0.1
0
0 0.5 1 1.5 2 2.5 3 3.5 4
σRJ / σPJ
Fig. A.3 Simulation results of tail fitting: (a) RJ rms error, (b) effective PJ amplitude
Both of ¢RJ err =¢PJ and Aeff =APJ have an almost linear relationship with ¢RJ =¢PJ .
Thus, based on the two curves in Fig. A.3, we can easily derive the following equa-
tions to calculate Aeff and ¢eff :
¢RJ
Aeff D 1 0:9 APJ : (A.2)
¢PJ
q
¢eff D ¢RJ 2 C .0:84 ¢RJ /2 : (A.3)
References
95
96 References
20. Li M, Wilstrup J (Oct 2003) Paradigm shift for jitter and noise in design and test >1 gb=s
communication systems. In: Proceedings of International Conference on Computer Design,
pp 467–472
21. Ong CK et al. (Jan 2004) Jitter Spectral Extraction for Multi-gigahertz Signal. In: Proceedings
of Asia and South Pacific Design Automation Conference pp 298–303
22. Cai Y et al. (Jan–Feb 2002) Jitter testing for gigabit serial communication transceivers, IEEE
Des Test Comp. 19:66–74
23. Razavi B (2002) Design of integrated circuits for optical communications, McGraw-Hill,
New York
24. Lee K. et al. (Feb 1998) A Jitter-Tolerant 4.5 Gb/s CMOS interconnect for digital display. In:
Proceedings of International Solid-State Circuits Conference, pp 310–311
25. Chang K et al. (Feb 2005) Clocking and circuit design for a parallel i/o on a first-generation
CELL processor. In: Proceedings of International Solid-State Circuits Conference, pp 526–527
26. Leon-Garcia A (1994) Probability and random processes for electrical engineering. Addison
Wesley, Reading, MA
27. Agilent Technologies, Jitter Analysis: The dual-Dirac Model, RJ/DJ, and Q-Scale, White Pa-
per. Available at www.agilent.com
28. Jiang H et al. (Aug 2002) Optimal loop parameter design of charge pump PLLs for jitter
transfer characteristic optimization. In: Proceedings of Midwest Symposium on Circuits and
Systems, vol 1, pp 344–347
29. Maxim Integrated Products (2004) Pattern Creator/Converter software User Manual, Applica-
tion Note: HFAN-9.5.0, Available at www.maxim-ic.com
30. Oppenheim AV, Willsky AS (1983) Signals and systems. Prentice Hall, Englewood Cliffs, NJ
31. Hajimiri A (Feb 2001) Noise in phase-locked loops. In: Proceedings of Southwest Symposium
on Mixed-Signal Design, pp 1–6
32. Lee D (Nov 2002) Analysis of jitter in phase-locked loops. IEEE Trans Circuit Syst II
49(11):704–711
33. Moon U et al. (May 2002) Spectral analysis of time-domain phase jitter measurements. IEEE
Trans Circuit Syst II 49(5):321–327
34. McNeill J (June 1997) Jitter in ring oscillators. IEEE J Solid-State Circuit 32(6):870–979
35. James NK (Aug 2000) Cycle-domain simulation for phase-locked loops. In: Proceedings of
Southwest Symposium on Mixed-Signal Design. pp 77–82
36. Muller M et al. (Feb 2005) Total jitter measurement at low probability levels. Using Optimized
BERT Scan Method. Designcon
37. Maxim Integrated Products (2007), Low-Power, Compact 2.5 Gbps/2.7 Gbps Clock-Recovery
and Data-Retiming IC. Available at www.maxim-ic.com
38. Tektronix (2006) TDS JIT3-Jitter Analysis and Timing Software. Available at www.tek.com
39. SyntheSys Research Inc. (2006) BERTScope and BERTScopeS Signal Integrity Analyzers.
Available at www.bertscope.com
40. Stojanovic V, Horowitz M (Sept 2003) Modeling and Analysis of High-Speed Links. IEEE
proceedings of Custom Integrated Circuits Conference. pp 589–594
41. Jaussi JE et al. (Jan 2005) 8-Gb/s source-synchronous I/O link with adaptive receiver equaliza-
tion, offset cancellation, and clock de-skew. IEEE J Solid State Circuit 40(1):80–88
42. Stofanovic V et al. (April 2005) Autonomous dual mode (PAM2/4) serial link transceiver with
adaptive equalization and data recovery. IEEE J Solid State Circuit 40(4):1012–1026
43. Payne R et al. (Dec 2005) A 6.25-Gb/s binary transceiver in 0.13-um CMOS for serial
data transmission across high loss legacy backplane channels. IEEE J Solid State Circuit
40(12):2646–2657
44. Lee J et al. (Sept 2004) Analysis and modeling of bang–bang clock and data recovery circuits.
IEEE J Solid State Circuit 39(9):1571–1580
45. Walker RC (2003) Designing bang–bang plls for clock and data recovery in serial data trans-
mission systems. In: Razavi B (ed) Phase-locking in high-performance systems, IEEE Press,
New York, pp 34–45
46. Choi Y et al. (Nov 2003) Jitter transfer analysis of tracked oversampling techniques for multi-
gigabit clock and data recovery. IEEE Trans Circuit Syst II 50(11):775–783
References 97
47. Provost B et al. (Oct 2004) PAC IO loopback design for high speed /spl mu/processor IO test.
In: Proceedings of International Test Conference, pp 23–30
48. Sunter S, Roy A (Oct 2005) Structural tests for jitter tolerance in SerDes receivers. In: Pro-
ceedings of International Test Conference, pp 1–10
49. Nejedlo JJ (Oct 2003) IBISTTM (Interconnect Built-In Self-Test) architecture and methodology
for PCI Express: Intel’s next-generation test and validation methodology for performance IO.
In: Proceedings of International Test Conference, pp 784–793
50. Kim J (2002) Design of CMOS adaptive-supply serial links, Ph.D. thesis, Stanford University,
California
51. Shimanachi M (Oct 2001) An approach to consistent jitter modeling for various jitter aspects
and measurement methods. In: Proceedings of International Test Conference, pp 845–857
52. Balamurugan G, Shanbhag N (Aug 2003) Modeling and mitigation of jitter in high-speed
source-synchronous inter-chip communication systems. In: Proceedings of Asilomar Confer-
ence on Signals, Systems and Computers, pp 1681–1687
53. Sidiropoulos S, Horowitz MA, semidigital A (Nov 1997) Dual delay-locked loop. IEEE J Solid
State Circuit 32(11):1683–1692
54. Lee TH, Bulzacchelli JF (Dec 1992) A 155-MHz clock recovery delay- and phase-locked loop.
IEEE J Solid State Circuit 27(12):1736–1746
55. Buckwalter JF, Hajimiri A (Mar 2006) Analysis and equalization of data-dependent jitter. IEEE
J Solid State Circuit 41(3):607–620
56. Buckwalter JF, Hajimiri A (Mar 2006) Cancellation of Crosstalk Induced jitter. IEEE J Solid
State Circuit 41(3):621–632
57. Kuo A et al. (Oct 2005) Crosstalk bounded uncorrelated jitter (BUJ) for high-speed intercon-
nects. IEEE Trans Instrum Measure 54(5):1800–1810
58. Casper B et al. (Feb 2006) A 20 Gb/s forwarded clock transceiver in 90 nm CMOS. In: Pro-
ceedings of International Solid-State Circuits Conference, pp 263–272
59. Zerbe J (2005) Comparison of adaptive and non-adaptive equalization methods in high-
performance backplanes DesignCon
60. Liu J, Lin X (2004) Equalization in high-speed communication systems. IEEE Circuits Systems
Magazine, Second quarter, pp 4–17
61. Dally WJ, Poulton J (Jan–Feb 1997) Transmitter equalization for 4-Gbps signaling. IEEE
Microelec 17:48–56
62. Beukema T et al. (Dec 2005) A 6.4 Gb/s CMOS SerDes Core with feed-forward and decision-
feedback equalization. IEEE J Solid State Circuit 40(12):2633–2645
63. Kim J et al. (Feb 2005) A four-channel 3.125-Gb/s/ch CMOS serial-link transceiver with a
mixed-mode adaptive equalizer. IEEE J Solid State Circuit 40(2):462–471
64. Sun R et al. (Sept 2005) A low-power, 20-Gb/s continuous-time adaptive passive equalizer. In:
Proceedings of International Symposium on Circuits and Systems, pp 920–923
65. Choi J et al. (Mar 2004) A 0.18-um CMOS 3.5-Gb/s continuous-time adaptive cable equalizer
using enhanced low-frequency gain control method. IEEE J Solid State Circuit 39(3):419–425
66. Ahang G et al. (Feb 2004) A BiCMOS 10 Gb/s Adaptive cable equalizer. In: Proceedings of
International Solid-State Circuits Conference, pp 92–93
67. Maxim Integrated Products (2003) Designing a simple, wide-band and low-power equalizer
for FR4 copper link, DesignCon
68. Lee J (Feb 2006) A 20-Gb/s adaptive equalizer in 0.13 um CMOS technology. In: Proceedings
of International Solid-State Circuits Conference, pp 92–93
69. Sun R et al. (June 2006) A tunable passive filter for low-power high-speed equalizers. In:
Proceedings of VLSI Circuit Symposium, pp 198–201
70. Jaussi JE et al. (Dec 2003) An 8 Gb/s simultaneous bidirectional link with on-die waveform
capture. IEEE J Solid State Circuit 38(12):2111–2120
71. Tripp M et al. (Oct 2003) Elimination of traditional functional testing of interface timings at
Intel. In: Proceedings of International Test Conference, pp 1014–1022
72. Sun R (Dec 2005) A low-power, 20-Gb/s continuous-time adaptive passive equalizer, MS the-
sis, Carnegie Mellon University
98 References
73. Kromer C et al. (Dec 2006) A 25-Gb/s CDR in 90-nm CMOS for high-density interconnects.
IEEE J Solid State Circuit 41(12):2921–2929
74. Hidaka Y et al. (Feb 2007) A 4-Channel 3.1/10.3 Gb/s transceiver macro with a pattern-tolerant
adaptive equalizer. In: Proceedings of Solid-State Circuits Conference, pp 442–445
75. Momtaz A et al. (Apr 2007) A fully integrated 10-Gb/s receiver with adaptive optical dispersion
equalizer in 0.13-um CMOS. IEEE J Solid State Circuit 42(4):872–880
76. Harwood M et al. (Feb 2006) A 12.5-Gb/s SerDes in 65 nm CMOS using a baud-rate adc
with digital receiver equalization and clock recovery. In: Proceedings of Solid-State Circuits
Conference, pp 436–438
77. Palmer R et al. (Feb 2007) A 14 mW 6.25 Gb/s transceiver in 90 nm CMOS for serial chip-to-
chip communications. In: Proceedings of Solid-State Circuits Conference, pp 440–442