PLL Thesis Mansuri Ucla 2003
PLL Thesis Mansuri Ucla 2003
PLL Thesis Mansuri Ucla 2003
Los Angeles
in Electrical Engineering
by
Mozhgan Mansuri
2003
The dissertation of Mozhgan Mansuri is approved.
_________________________________________
Majid Sarrafzadeh
_________________________________________
Mau-Chung Frank Chang
_________________________________________
Behzad Razavi
_________________________________________
Chih-Kong Ken Yang, Committee Chair
ii
Dedication
To my parents
iii
Table of Contents
1. Introduction..................................................................................................................1
1.1 Motivation...............................................................................................................2
1.2 Organization............................................................................................................6
iv
2.5.1 Device Electronic Noise .............................................................................22
2.5.2 Supply or Substrate Noise...........................................................................23
2.5.3 Noise Sensitivity Metric .............................................................................23
2.6 Summary ...............................................................................................................24
v
5.3 Loop Filter ............................................................................................................92
5.3.1 Proposed Loop Filter Design ......................................................................94
5.4 Phase-Frequency Detector ....................................................................................97
5.4.1 Conventional PFD Design ..........................................................................97
5.4.2 Pass-Transistor PFD Design .....................................................................100
5.4.3 Latch-Based PFD Design..........................................................................101
5.4.4 Simulated Transfer Curve of PFDs...........................................................104
5.5 Measurement Results ..........................................................................................106
5.6 PLL Performance Comparison ...........................................................................112
5.7 Summary .............................................................................................................114
7. Conclusion ................................................................................................................130
Appendices......................................................................................................................135
Bibliography ...................................................................................................................144
vi
List of Figures
vii
z = 0.65 (3) f-3dB = 11.4%fref, z = 1.63)..................................................36
Figure 3.8: Long-term jitter (due to VCO noise) as a function of: (a) loop bandwidth,
(b) loop damping factor .............................................................................37
Figure 3.9: Comparison of long-term jitter (due to VCO noise) in: (a) 2nd, 3rd order
loop (b) without loop delay and c) with loop delay ...................................39
Figure 3.10: PLL bandwidth (at minimum jitter) as a function of 3rd pole frequency and
PLL loop delay...........................................................................................40
Figure 3.11: Output clock jitter (due to input clock noise) behavior vs. input clock jitter
behavior .....................................................................................................42
Figure 3.12: Output to input jitter ratio behavior of a 2nd-order loop as a function of: (a)
loop bandwidth, (b) loop damping factor ..................................................43
Figure 3.13: Comparison of long-term jitter (due to white noise at PLL input) in: (a) 2nd,
3rd order loop (b) without loop delay and (c) with loop delay..................44
Figure 3.14: An adaptive bandwidth PLL with tunable loop parameters ......................45
Figure 3.15: Die photograph of the PLL ........................................................................46
Figure 3.16: Measurement technique in time domain, referenced to reference clock ...47
Figure 3.17: Measured and calculated tracking jitter as wz is reduced in constant KLoop
....................................................................................................................48
Figure 3.18: Measurement technique for calculating PLL loop transfer function .........50
Figure 3.19: Measured PLL loop transfer function (@ 700MHz reference clock) at a con-
stant ICPintegral (constant KLoop) ...........................................................50
Figure 3.20: Measurement technique in time domain, referenced to output clock ........51
Figure 3.21: Measured and calculated short-term jitter (@ 700MHz reference clock) for
four different loop parameters ...................................................................51
Figure 3.22: Output jitter (due to input clock noise) behavior for three different PLL loop
parameters: (a) measurement results, (b) analytical results ((1) Input jitter
(2) z = 0.2, f-3dB = 39MHz (3) z = 0.65, f-3dB = 45MHz (4) z = 1.63, f-3dB
= 80MHz)...................................................................................................52
Figure 4.1: The PLL block diagram with VCO and input noise ..................................56
Figure 4.2: Loop transfer functions from VCO and input clock noise to the PLL output
....................................................................................................................57
Figure 4.3: Behavior of output clock jitter due to VCO noise for various loop parame-
ters: (a) 3-D, (b) contour ............................................................................59
Figure 4.4: Behavior of output clock jitter due to input noise for various loop parame-
ters: (a) 3-D, (b) contour ............................................................................60
Figure 4.5: Behavior of output clock jitter due to both VCO and input noise for various
loop parameters: (a) 3-D, (b) contour ........................................................62
Figure 4.6: A PLL architecture with adjustable loop parameters using adjustable R and
viii
ICP .............................................................................................................64
Figure 4.7: Jitter measurement with a flash TDC architecture.....................................65
Figure 4.8: Jitter measurement with a dead-zone window establishment ....................66
Figure 4.9: PLL die photograph ...................................................................................69
Figure 4.10: Test setup for the jitter measurement and optimization.............................70
Figure 4.11: (a) Measured percentage hits distribution for one set of PLL loop parameters
for N=500 and N=5000, (b) standard deviation of measured percentage hits
....................................................................................................................71
Figure 4.12: Jitter measurement contours (due to VCO noise) for all loop parameters
with (a) constant dead-zone width and measuring hits (percentage), (b) con-
stant 4% measured hits and measuring dead-zone width ..........................73
Figure 4.13: Jitter measurement contours (due to input noise) for all loop parameters with
constant 4% measured hits and measuring dead-zone width.....................75
Figure 4.14: Flow chart of jitter minimization algorithm ..............................................77
Figure 4.15: Measured minimum jitter due to the sum of VCO and input noise for (a)
3000hits, (b) 300hits ..................................................................................78
Figure 5.1: The proposed PLL architecture..................................................................84
Figure 5.2: Power-supply regulated VCO ....................................................................85
Figure 5.3: VCO with a feedback cascode using OTA ................................................86
Figure 5.4: Voltage-controlled oscillator with a noise-canceling circuit .....................87
Figure 5.5: Quadrature pseudo-differential current-controlled oscillator (CCO) ........88
Figure 5.6: Simulated V-I converter gain characteristic across process corners..........89
Figure 5.7: VCCO response of V-I converter to -10% VDD step inserted at t=2ns ....91
Figure 5.8: Conventional loop filter .............................................................................93
Figure 5.9: Implementing the PLL stabilizing zero with two charge-pump currents and
a regulator ..................................................................................................94
Figure 5.10: Proposed loop filter architecture................................................................94
Figure 5.11: Charge-pump current circuit ......................................................................95
Figure 5.12: Loop stabilizing zero with a 4-bit controller (n=4)....................................95
Figure 5.13: (a) Linear PFD architecture, (b) PFD state diagram ..................................97
Figure 5.14: (a) Ideal PFD characteristic. (b) Nonideal linear PFD characteristic. (c) PFD
nonideal behavior due to nonzero reset delay............................................99
Figure 5.15: Pass-transistor DFF PFD architecture......................................................101
Figure 5.16: (a) Behavior of a latch-based PFD, including the description of the nonideal
behavior origin. (b) characteristic of a latch-based PFD .........................102
Figure 5.17: Latch-based PFD architecture..................................................................103
Figure 5.18: Characteristics of three PFDs at 435MHz ...............................................105
Figure 5.19: Simulated frequency acquisition..............................................................105
ix
Figure 5.20: PLL and clock buffer die photograph ......................................................106
Figure 5.21: Measured and simulated VCO gain .........................................................107
Figure 5.22: PLL output jitter histogram at 1GHz .......................................................107
Figure 5.23: Measured sensitivity of VCO output clock frequency to static and dynamic
supply noise .............................................................................................109
Figure 5.24: Die photograph of three different PFDs implemented in a PLL..............110
Figure 5.25: Measured frequency acquisition ..............................................................111
Figure 6.1 (a) Ideal compensation of supply-induced inverter delay variation, (b) pro-
posed compensator inverter .....................................................................118
Figure 6.2 (a) Delay variation of compensated inverter due to VSG variation, (b) delay
sensitivity of compensator circuit, normalized to delay sensitivity of an in-
verter ........................................................................................................120
Figure 6.3 Behavior of normalized delay sensitivity of compensator circuit due to VSG
(VDD) variation as a function of: (a) PMOS capacitor, (b) PMOS resistor..
..................................................................................................................121
Figure 6.4 Supply-induced delay variation of: (1) uncompensated inverter, (2) com-
pensated inverter with inverter’s VDD held constant and (3) compensated
inverter .....................................................................................................122
Figure 6.5 Bias circuit generating Vgap....................................................................123
Figure 6.6 Sensitivity of supply-induced delay variation of compensated inverter due
to Vgap offset...........................................................................................124
Figure 6.7 Delay variation of compensated clock buffer over temperature as VDD var-
ies £ ±10% ...............................................................................................125
Figure 6.8 Delay variation of compensated clock buffer across the corners as VDD var-
ies £ ±10% ...............................................................................................126
Figure 6.9 Five stages of fanout of four (FO-4) compensated inverters (n=5) .........127
Figure 6.10 Measured supply-induced delay variation of uncompensated (--) and com-
pensated clock buffer ...............................................................................128
x
List of Tables
Table 3.1: Tracking jitter (in ps) for different loop parameters (fref = 700MHz) ......48
Table 5.1: PFDs performance summary ...................................................................105
Table 5.2: PLL performance summary (1)................................................................106
Table 5.3: PLL performance summary (2)................................................................107
Table A.1: Comparison of estimated tracking jitter (by 2nd-order analysis) with mea-
sured tracking jitter (fref = 700MHz) ......................................................131
xi
Acknowledgments
During my study and research at UCLA, I have been extremely blessed by God to
meet and collaborate with so many people that were so supportive and helpful in this
research.
I would like to deeply thank my advisor, professor Ken Yang, for his continuos
support, encouragement and help. He has been my best research advisor and it has been a
privilege collaborating and working with him these past four years. He has been source of
ideas and knowledge, yet, his wisdom allowed me to direct my research successfully.
I would also like to thank professor Behzad Razavi for his support and useful
technical discussions. I would like to extend my appreciation to him, professor Frank
Chang and professor Majid Sarrafzadeh for serving on my committee and providing me
with their fruitful comments.
xii
It has been a pleasure to work with so many talented people in UCLA. I wish to
thank, in particular, Siamak Modjtahedi, who generously provided me with his help and
useful discussions, Jackie Wong and Hamid Hatamkhani, with whom I collaborated in the
design of low-power links, and Ali Hadiashar, who helped me with the development of
run-time algorithm for jitter optimization. I would also like to thank Dean Liu for his
collaboration and great help on the design of phase-frequency detectors.
Also, I am greatly thankful to my friends for their constant support and friendship.
I would like to thank, in particular, Hamid Rafati, Esmaeil Heidari, Rahim Bagheri, Ali
Karimi, Omid Oliaei, Alireza Razzaghi, Vladimir Stojanovic, Saeed Chehrazi and Pejman
Kalkhoran for countless discussions.
I wish to thank National semiconductor, Intel corporation and UCMicro 02-102 for
fabrication and their support. Also I would like to thank Makoto Murata for his great help
in wire bonding and Dorothy Tarkington for her wonderful help in purchasing the lab
equipment and components.
xiii
VITA
M. Mansuri and CK.K. Yang, “A Low-Power Low-Jitter Adaptive Bandwidth PLL and
Clock Buffer,” Submitted for publication, IEEE, Journal of Solid-State Circuits, Novem-
ber 2003
M. Mansuri, A. Hadiashar, and CK.K. Yang, “Methodology for On-chip Adaptive Jitter
Minimization in Phase-Locked Loops,” Submitted for publication, IEEE, Journal of
Transactions on Circuits and Systems II, November 2003
xiv
KL.J. Wong, M. Mansuri, H. Hatamkhani and CK.K. Yang, “A 27-mW 3.6-Gb/s I/O
Transceiver,” Proceedings of Symposium on VLSI Circuits, pp. 99-102, Japan, June 2003
M. Mansuri and CK.K. Yang, “A Low-Power Low-Jitter Adaptive Bandwidth PLL and
Clock Buffer,” ISSCC Digest of Technical Papers, pp. 430-431, San Francisco, CA, Feb-
ruary 2003
M. Mansuri and CK.K. Yang, “Jitter Optimization Based on Phase-Locked Loop Design
Parameters,” IEEE, Journal of Solid-State Circuits, vol. 37, no. 11, pp. 1375-1382,
November 2002
M. Mansuri and CK.K. Yang, “Jitter Optimization Based on Phase-Locked Loop Design
Parameters,” ISSCC Digest of Technical Papers, pp. 138-139, San Francisco, CA, Febru-
ary 2002
xv
ABSTRACT OF THE DISSERTATION
by
Mozhgan Mansuri
Phase locked-loops (PLLs) are widely used to generate well-timed on-chip clocks
Switching activity in large digital systems introduces power supply or substrate noise
xvi
Power dissipated by PLLs is often a small fraction of total active power. However,
during sleep modes where the PLL must remain in lock, it can be a significant fraction of
dissipated power. Also, for some applications such as high speed parallel links and
distributed synchronous clocking, multiple PLLs are employed to minimize the timing
uncertainty. Therefore, demand for low-power PLLs has been increasing. The low-power
This research describes the design of a fully-integrated low-jitter PLL for low-
power applications. To achieve the low-jitter performance, this work proposes jitter
At the system level, this work investigates the effects of PLL design parameters,
such as bandwidth and peaking in the frequency response, on timing jitter of PLL output
clock. The analysis includes several common noise sources in a PLL and develops an
intuition for selecting design parameters to obtain minimum output jitter based on the
dominant noise source. The proposed PLL is equipped with digitally-controllable loop
parameters that independently adjusts the loop parameters. Based on jitter analysis, a
methodology for on-chip adaptive jitter minimization in PLLs is developed. The proposed
method measures the output jitter and adjusts the PLL loop parameters toward minimizing
the jitter by a closed loop control system. The experimental results verify the success of
the proposed method in minimizing jitter to within 5ps of the minimum long-term peak-
to-peak jitter.
xvii
At the circuit level, two new supply rejection techniques for VCOs and clock
buffers are developed. Both methods demonstrate the delay sensitivity of ≤0.1%-delay/%-
VDD due to both static and dynamic supply noise. While the jitter performance is
comparable with prior state-of-art work, the proposed VCO and clock buffer consume less
power with smaller area than previous designs. The VCO is designed to operate over a
wide frequency range and has a linear voltage-to-frequency gain. The PLL is designed
with scaling loop parameters that track over a 10x frequency range of the VCO and allow
the adaptive loop bandwidth. The PLL is implemented in 0.25-µm CMOS technology and
xviii
Chapter 1
Introduction
synchronize between functional units and between ICs. Clock frequencies and data rates
have been increasing with each generation of processing technology and processor
architecture. Figure 1.1 shows the clock frequency versus technology generation
according to 2002 ITRS1. Within these digital systems, well-timed clocks are generated
with phase-locked loops (PLLs) and then distributed on-chip with clock buffers. The rapid
increase of the systems’ clock frequency poses challenges in generating and distributing
the clock with low uncertainty and low power. This research presents innovative
techniques at both system and circuit levels that minimize the clock timing uncertainty
1
7
1
130 120 110 100 90 80 70
Technology (nm)
1.1 Motivation
A PLL is essentially a feedback loop that locks the on-chip clock phase to that of
an input clock or signal. Because the on-chip clock toggles a large capacitive load, a series
of clock buffers efficiently increases the drive strength of the PLL output to drive the load.
High-performance PLLs and clock buffers are widely used within a digital system for two
For clock generation, since off-chip reference frequencies are limited by the
maximum frequency of a crystal frequency reference1, a PLL receives the reference clock
and multiplies the frequency to the multi-gigahertz operating frequency. The high-
2
frequency clock is then driven to all parts of the chip. Timing recovery pertains to the data
communication between chips. As data rates increase to satisfy the increase in on-chip
processing rate, the phase relationship between the input data and the on-chip clock is not
fixed. To reliably receive the high-speed data, a PLL locks the clock phase that samples
maintain proper synchronization, large timing uncertainty would result in lower frequency
of operation. Jitter is due to both intrinsic random noise (i.e. thermal noise and flicker
switching activity introduces power-supply or substrate noise which perturbs the PLL
elements and clock buffers. Supply or substrate noise is the dominant source of jitter in
these systems. This research focusses on the design of the most sensitive blocks in a PLL
and clock buffer with high immunity to supply/substrate noise. The research also
represents a powerful noise-filtering technique that minimizes jitter through adjusting the
key loop parameters of a PLL based on the dominant noise source in the PLL.
Power dissipated by PLLs is often a small fraction of the total active power. However, it
can be a significant fraction of the power dissipated in the sleep mode where the PLL must
remain in lock. Also, as operating clock frequency of digital systems is increasing, the
systems become less tolerable to clock skew. There is an increasing demand for using
parallel links [8]-[10] and distributed synchronous clocking [1]-[7]. In both applications,
3
multiple PLLs are employed to reduce the timing uncertainty across the entire system with
CKref PLL0
data0
PLL1
data1
dataN
PLLM
ref
increase the bandwidth, the architecture utilizes a set of parallel data signals. The
synchronization is achieved through transmitting a reference clock with the parallel data
signals. In the receiver, the on-chip clock is locally generated by multiple PLLs from the
transmitted clock to recover the data. Locally distributed PLLs reduces the timing
clock is distributed to many locations on the chip over a tree-like or grid-like network
(Figure 1.3-(a) or (b)) with repeaters at necessary intervals. These networks are passive
because it does nothing to reduce the uncertainty of the clock delivered to the sequential
elements. As the clock frequency goes up, the number of required repeaters increases and
shielding the interconnect segments becomes more difficult; thus, the timing uncertainty
inevitably increases. Skew compensation [13]-[14] is used to reduce the delay mismatches
4
introduced during fabrication. However, this technique does not suppress jitter. A possible
Driver
Root
Leaf
(a) (b)
Figure 1.3 Clock distribution networks: (a) trees, (b) grids
signal at multiple nodes across the chip (Figure 1.4). Phase detectors (PDs) at boundaries
produce error signals to adjust frequency of the node PLL. Within the tree, the clocks will
be driven as sinusoidal signals without intermediate buffering; thus, the clocks at each
terminal have a small swing due to resistive losses. With locally generated clocks, there
are no full swing clock lines to couple in jitter. Also, since the clock is generated at each
node, jitter does not accumulate with distance from the clock source.
Master Local
PLL
PLL Clock
Region PD
5
Since many of these phase-locking systems are required to be integrated within a
single chip, the overall power and area overhead of a single phase-locking circuit are key
similar components as a PLL. The power and area constraints make the design of a low-
jitter PLL even more challenging due to the trade-off between low-jitter and low-power
(and low-area) design techniques. This research presents new filtering techniques in the
design of PLL components and loop parameters to overcome the low-power and low-area
constraints. The proposed filtering techniques minimize the clock timing uncertainty
1.2 Organization
This thesis is composed of seven chapters. The functioning and components of a
phase-locked loop (PLL) are described in Chapter 2. Then, the two common PLL
architectures, delay-line based PLL (DLL) and oscillator-based PLL (PLL), are discussed
and compared. The noise and power constraints associated with the design of a PLL are
the next subject of the chapter. Noise minimization techniques at both system and circuit
At the system level, the timing jitter of the PLL output clock is minimized by
proper design of PLL loop parameters, such as bandwidth and peaking in the frequency
response. The jitter minimization relies on the fact that a PLL is a closed-loop system and
filters each noise source in the PLL based on the transfer function from the correspondent
noise source to the PLL output. For instance, a high-bandwidth PLL can track the phase of
6
a low-noise input clock and filter out voltage-controlled oscillator (VCO) noise.
Conversely, a low-bandwidth PLL filters a noisy input clock. The goal is to explore an
intuition for selecting design parameters to obtain the minimum output jitter based on the
dominant noise source. Chapter 3 reviews jitter definitions and major timing jitter sources
in a PLL. The relationship between the jitter, the power spectral density of each noise
source and the correspondent PLL noise transfer function is extracted next. Based on the
extracted equations, the sensitivity of jitter to PLL bandwidth and peaking in loop
frequency response is derived. Finally, a PLL with tunable loop parameters is used to
The proper design of PLL loop parameters for minimum output jitter performance
requires knowledge of the dominant noise source in the PLL. For many systems, the
magnitude of the noise sources are not well known which makes the design of loop
minimization in PLLs. The algorithm functions during system operation and minimizes
jitter as noise source conditions vary. The chapter shows that since the total jitter has only
minimum. The chapter, then, describes the circuit components necessary that dynamically
explores designs of low-noise PLL components. Although both device noise and supply/
substrate noise are present, supply/substrate noise is the dominant noise source in digital
systems which perturbs the most sensitive blocks such as voltage-controlled oscillators
7
(VCOs) and clock buffers. To achieve a high-noise performance requires design of VCOs
Design of the PLL components are discussed in Chapter 5, starting with the design
of a VCO. The new noise filtering technique is presented that achieves similar noise
performance with improved power and area performance comparing with state-of-the-art
designs. The chapter presents a self-biased charge-pump current and loop filter, next, that
allows the PLL to operate over a wide frequency range with an adaptive bandwidth in a
introduced next that has lower power consumption and larger lock-in range than
conventional PFDs.
Clock buffers with improved supply sensitivity of buffer elements are introduced
in Chapter 6. The design goal is to compensate the supply-induced delay variation with an
improved dynamic behavior while introducing minimum power, area and delay overhead.
The noise performance of the compensated buffer is verified with experimental results.
8
Chapter 2
frequency synthesizer. The basic concept of phase locking has remained the same since its
invention in the 1930s [20]. However, design and implementation of PLLs continue to be
consumption and area become more stringent. A large part of this research focuses on the
challenges and trade-off behind the design of such a PLL, this chapter provides a brief
Section 2.1 provides an overview of a PLL system and briefly discusses the basic
concept of phase locking. PLL components for charge-pump PLLs are discussed in
Section 2.2. Section 2.3 discusses and compares the two possible PLL architectures: (1)
delay-line based PLL and (2) oscillator-based PLL. Study of loop characteristics and loop
9
parameters is the subject of Section 2.4. This section provides a simple analysis of the
The noise sources present in digital systems are discussed in Section 2.5. The
chapter concludes with a summary of design goals and issues involved in the design of
feedback system that sets fixed phase relationship between its output clock phase and the
phase of a reference clock. A PLL tracks the phase changes that are within the bandwidth
of the PLL. A PLL also multiplies a low-frequency reference clock, CKref, to produce a
φref, CKref
φout, CKout
Phase error Low-Pass Filter Oscillator
Detector
φfeedback, CKfeedback
Frequency Divider
:N
produces an error output signal based on the phase difference between the phase of the
feedback clock and the phase of the reference clock. Over time, small frequency
differences accumulate as an increasing phase error. The difference or error signal is low-
10
pass filtered and drives the oscillator. The filtered error signal acts as a control signal
(voltage or current) of the oscillator and adjusts the frequency of oscillation to align
φfeedback with φref. The frequency of oscillation is divided down to the feedback clock by a
frequency divider. The phase is locked when the feedback clock has a constant phase error
and the same frequency as the reference clock. Because the feedback clock is a divided
version of the oscillator’s clock frequency, the frequency of oscillation is N times the
reference clock.
current, (3) loop filter, (4) voltage-controlled oscillator, and (5) frequency divider. The
Divider
:N
UP ICP
Reference PD/PFD Output
VCO
Clock DN Clock
ICP R
CCP C1
Charge-Pump
Loop Filter
11
2.2.1 Voltage-Controlled Oscillator (VCO)
An oscillator is an autonomous system that generates a periodic output without any
input. A CMOS ring oscillator shown in Figure 2.3 is an example of an oscillator. So that
Vctrl (or Ictrl)
CKout
the phase of a PLL is adjustable, the frequency of oscillation must be tunable. In the
example of an inverter ring oscillator, the frequency could easily be adjusted with
controlling the supply (voltage or current) of inverters. The slope of frequency versus
control signal curve at the oscillation frequency is called voltage-to-frequency (or current-
φ VCO = ∫ KVCO ⋅ Vctrl ⋅ dt . In other words, the VCO in the frequency domain (s-domain),
φ VCO K VCO
is modeled as ------------ ( s ) = ------------- . Ideally, for the linear analysis to apply over a large
V ctrl s
frequency range, KVCO, needs to be relatively constant.
from tens to a few hundreds of MHz. On the other hand, VCOs for clocking and parallel
link applications operate at a few GHz or even ten GHz. For proper functioning of the
12
phase detector or phase-frequency detector, discussed in the next section, a frequency
divider divides down the VCO frequency to the frequency of the reference clock.
and produces an error signal that is proportional to the phase difference. In the presence of
a large frequency difference, a pure phase detector does not always generate the correct
direction of phase error. Phase error accumulates rapidly and can oscillate between phase
error of >180oand <180o from cycle to cycle. The average phase detector output contains
little frequency information and no valuable phase information. Since the phase detector is
insensitive to frequency difference at the input, upon start-up when the oscillator’s
frequency divided by N1 is far from the reference frequency, the PLL may fail to lock. The
problem is known as an inadequate acquisition range of the PLL. To remedy the problem,
a phase-frequency detector (PFD) is used that can detect both phase and frequency
differences. Figure 2.4 conceptually demonstrates the operation of a PFD for two cases:
(a) the two input signals have the same frequency, and (b) one input has higher frequency
than another input. In both cases, the DC contents of PFD’s outputs, UP and DN, provide
13
Ref Ref
Ref UP CK CK
PFD
CK DN UP UP
DN DN
(a) (b)
Figure 2.4: Operation of a PFD: (a) fref=fCK, φref#φCk and (b) fref>fCK
DN outputs of PFD as shown in Figure 2.2. The charge-pump injects the charge into or out
of the loop filter capacitor (CCP). The combination of charge-pump and CCP is an
integrator that generates the average of UP (or DN) pulses. This average voltage adjusts
the frequency of the subsequent oscillator circuit. Since the VCO introduces another
integrator, the loop gain of a charge-pump PLL has two poles at origin; thus, the closed-
loop system is unstable. To stabilize the system, a zero, ωz = 1/RCCP, is introduced in the
The PFD, charge pump and filter are often modeled with a linear continuous-time
model. In reality, the PFD acts as a pulse modulator system and drives the charge-pump
for the duration of pulse width which is equal to PFD input phase difference, ∆φ. The
actual phase response is not linear because phase is cyclical. Furthermore, the phase
14
However, a linear continuous-time approximation is often used to model the
stability of an operating point. The error due to approximation is negligible if the PLL
bandwidth is 1/10th or smaller than the reference clock frequency [79]. The reference
frequency determines the rate that PFD output is refreshed. With a linear approximation,
V ctrl I CP
Vctrl is equal to: ---------- ( s ) = -------- ⋅ F ( s ) where F(s) is the transfer function of the loop filter
∆φ 2π
1
and is equal to: F ( s ) = ------------ ⋅ ( 1 + RC CP s ) , ignoring C1 in Figure 2.2.
C CP s
PLL or a delay-locked loop (DLL). A DLL is similar to a PLL except that a variable delay
line replaces the oscillator [21]. Thus, phase is the only state variable in a DLL while both
phase and frequency are the state variables in a PLL. The basic DLL building blocks are
shown in Figure 2.5, similar to that of a PLL. A phase detector (PD) measures the phase
delay_in
Delay Line delay_out
difference between the reference clock and the delay-line output. The error signal is low-
15
pass filtered to produce the control signal that adjusts the delay of the delay line. Note that
the delay-line input can be a separate external clock instead of the CKref.
To eliminate the phase offset in a DLL, the filter is an integrator. DLL with only a
single pole is unconditionally stable. Only at loop bandwidths close to the reference
frequency, where the loop delay and the sampling nature of the PD degrade phase margin,
error before correcting the error because the output phase is an integration of the
frequency change. In contrast, a DLL does not accumulate the phase error and corrects the
Although, a simple loop characteristic of a DLL is desirable, a DLL has its own
limitations. First, for clock generation, only one input clock is available so the clock is
used as the input to the delay line as well as the phase detector. Therefore, any high-
frequency jitter at the reference clock directly passes through the delay line to the DLL
to any phase variations in a reference clock. Secondly, it is not as easy to multiply the
reference frequency [65]-[66] as a PLL. Third, delay lines usually have a finite delay
range. The limited delay range causes the loop to not lock properly. In contrast, a PLL can
filter out a noisy reference clock by lowering the PLL bandwidth. A PLL can achieve a
wide frequency range, provided that the VCO is designed to operate over a wide range.
The output frequency can be any frequency different from the reference clock frequency.
The advantages of a PLL over a DLL motivates us to focus on a design of a PLL in this
16
research. Nevertheless, the circuits and jitter reduction techniques discussed in following
chapters are applicable to DLLs because PLL and DLL architectures share many similar
presentation of each loop element, discussed in Section 2.2, is depicted within each block
:N
UP ICP
Vctrl VCO
PD/PFD φout
φref KVCO /s
KPD DN
ICP R
CCP C1
ICP/2π F(s)
gain, F(s) is the loop filter transfer function and KVCO is the conversion gain of the VCO.
The open-loop transfer function for a second-order PLL (ignoring C1 in the loop filter) is
equal to:
I CP K VCO
H open ( s ) = K PFD ⋅ --------------------- ⋅ ( 1 + RC CP s ) ⋅ ------------
- ‹2.1›
2π ⋅ C CP s
2
17
This transfer function has two poles at origin and one compensating zero that guarantees
the closed-loop stability. Including the third pole, the open-loop transfer function is equal
to:
I CP K VCO
H open ( s ) = K PFD ⋅ --------------------------------------- ⋅ ( 1 + RC CP s ) ⋅ -------------------------------------------------------
- ‹2.2›
2π ⋅ ( C CP + C 1 ) 2
s ⋅ [1 + R(C C )s ]
CP 1
The magnitude and phase of the open-loop transfer functions for a second and
1 1
third-order PLL are shown in Figure 2.7. ω z = -------------- and ω p3 = ------------------------------ indicate
RC CP R ( C CP C 1 )
|Hopen(s)|.1/N
|Hopen(s)|.1/N
40dB/dec 40dB/dec
20dB/dec 20dB/dec
ωp3
0dB
ωz ω 0dB ω
ωc ωz ωc
40dB/dec
Hopen(s) Hopen(s)
ω ω
-90O -90O
-180O -180O
(a) (b)
Figure 2.7: Magnitude and phase of the open-loop transfer function for (a) a second-
order PLL, (b) a third-order PLL
the zero and third pole frequency, respectively. ωc is the open-loop unity gain frequency.
18
order PLL is stable. The zero locus for an ideal second-order loop is not critical for
To understand the effect of the zero and other PLL parameters on the closed-loop
behavior of the PLL, the closed-loop transfer function of a PLL from input phase to output
phase is calculated:
φ out H open ( s )
--------- ( s ) = H closed ( s ) = ---------------------------------------------
- ‹2.3›
φ in 1 + H open ( s ) ⋅ 1 ⁄ N
φ out K Loop ⋅ ( 1 + RC CP s )
--------- ( s ) = -----------------------------------------------------------------------------------
- ‹2.4›
φ in 2
s + ( K Loop ⁄ N )RC CP s + K Loop ⁄ N
The closed-loop transfer function from the input phase to the output phase (Equation 2.4)
is a low-pass filter. This low-pass behavior of a PLL is desirable because it rejects input
noise frequencies higher than the PLL bandwidth. Similarly, the closed-loop transfer
function from the VCO control voltage, Vctrl, to the output phase is calculated:
φ out K VCO ⋅ s
---------- ( s ) = -----------------------------------------------------------------------------------
- ‹2.6›
V ctrl 2
s + ( K Loop ⁄ N )RC CP s + K Loop ⁄ N
This closed-loop transfer function is a band-pass filter. This band-pass filter rejects
Filtering out noise sources by the closed-loop behavior of the PLL forms the
baseline for jitter analysis discussed in Chapter 3. Noise of the PLL’s output clock can be
19
optimally filtered by adjusting the loop bandwidth and peaking in frequency response
based on the dominant noise source. The loop bandwidth and peaking are adjustable by
K Loop
The natural frequency, ωn, and damping factor, ζ1, are equal to ω n = -------------
- and
N
ωn
ζ = ------------- , respectively. Natural frequency is proportional to square-root of the loop gain.
2 ⋅ ωz
Since KPFD, KVCO and CCP are typically design constant parameters, the natural
frequency (typically through the loop filter resistor, R) and charge-pump current, ζ and ωn
can be adjusted. In other words, the bandwidth and peaking in frequency response are
adjustable by varying ωz and ICP. The closed-loop frequency response for different values
of ωz in constant ICP are shown in Figure 2.8-(a). As ωz decreases the loop bandwidth
For a third-order PLL with sampling/feedback delay, decreasing the zero frequency
increases the bandwidth. However, the peaking in frequency response increases because
of the phase margin degradation due to the third pole and delay. The phase margin (PM)
for a third-order PLL with loop delay of tdelay can be approximated with [79]:
ωc ωc 360
o
PM = atan ------ – atan --------- – ----------- ⋅ ω c ⋅ t delay ‹2.7›
ω z ω p3 2π
2 2 2 K Loop K Loop
s + 2ζω n s + ω n ≡ s + -------------- RCs + --------------
N N
‹2.8›
20
The closed-loop frequency response of a third-order PLL for different values of ωz in
Magnitude (dB) 0
10
(a)
ωz
−2
10 −4 −2 0
10 10 10
Magnitude (dB)
0
10
(b) ωz
−2
10 −4 −2 0
10 10 10
Frequency/fref
Figure 2.8: Closed-loop frequency response of: (a) an ideal second-order PLL, (b) a
sampling third-order PLL
generate an output clock with minimum timing uncertainty. The timing uncertainty arises
Device mismatches causes a static phase shift (or skew) in the PLL output clock
from its desired phase. Skew can be minimized with a careful layout and increasing the
device size [11]-[12]. Skew is generally less critical than jitter because, due to its static
21
nature, the system can compensate for the static errors [13]-[14]. Dynamic noise causes a
random phase shift (or jitter) in the PLL output clock. The noise sources in a PLL are (1)
device electronic noise such as thermal noise or flicker noise and (2) power-supply or
substrate noise.
clock timing. Numerous studies provide models that predict the jitter due to device noise.
Most of these studies ([22]-[33]) focus on the modeling and prediction of jitter (or phase
noise) due to VCOs. A few studies discuss the effect of noise in other PLL blocks such as
PDs ([34]-[35]) and frequency dividers ([36]-[38]) on the PLL output jitter.
The previous studies also provide some guidance to reduce jitter. Some
architectures demonstrate an improved jitter performance over the others. For example,
resonant circuit-based VCOs (or harmonic oscillators) exhibit less jitter than relaxation
oscillators (such as ring oscillators) [24]-[25]. The jitter due to device electronic noise
components ([22], [26] and [30]-[32]). Therefore, there is a trade-off between power
consumption and jitter performance. For instance, Hajimiri in [26] demonstrates that the
jitter of a ring oscillator with a constant frequency decreases as the number of stages and
power increase.
22
2.5.2 Supply or Substrate Noise
Switching activities in digital systems introduces supply or substrate noise. The
supply or substrate noise perturbs the sensitive blocks in a PLL such as VCO and clock
VCO which changes the VCO operating frequency. The change in the oscillation
frequency of a VCO appears as a phase step in the input of the phase detector. The phase
error accumulates jitter until it is corrected by the PLL. Therefore, supply or substrate
noise causes jitter in a VCO which is persistent for the time duration equal to the time
For a clock buffer1, supply or substrate noise varies the delay and introduces a
phase shift at the output clock of the buffer. The impact of the supply voltage step for a
clock buffer is considerably shorter lived. However, clock buffers are designed for power
and area efficient capacitance driving and not supply rejection. The long chain of buffers
needed in modern processors causes a significant transient phase shift at the output.
with noise sensitivity metric. Noise sensitivity for a VCO is defined as a percentage of
VCO clock frequency (or period) variation per percentage of supply voltage (or substrate)
23
variation; %-fVCO/%-VDD. Similarly, noise sensitivity for a clock buffer is defined as a
percentage of the inverter’s delay variation per percentage of supply voltage (or substrate)
clock buffer is to minimize the noise sensitivity of these circuits to supply or substrate
noise. For most digital systems, the supply or substrate noise does not exceed ±10-15%
[49].
2.6 Summary
This chapter discussed the basic concept behind phase locking and in particular, a
PLL. The operation of each PLL component is briefly explained which provides a
framework to understand the design of a PLL as discussed in the following chapters. Two
main architectures to design a PLL were discussed. A DLL has a simpler loop
characteristic than a PLL and does not suffer from jitter accumulation presented in a PLL.
However, a DLL passes input clock noise while a PLL low-pass filters the input noise.
The frequency multiplication is easier in a PLL than a DLL. These two reasons motivate
The primary goal to design a PLL is to generate a low-jitter clock due to noise and
mismatches. This chapter discussed sources of noise. It also showed that there is a trade-
To reduce noise, this research first studies the effect of loop parameters in filtering
out noise sources in a PLL. Chapter 3 develops a simple yet accurate model that predicts
the output jitter and provides an intuition toward optimum loop parameter design for
24
minimum jitter. To further adaptively minimize the jitter, Chapter 4 discusses a
Supply or substrate noise is a dominant noise source in large digital systems. This
research presents innovative filtering techniques at circuit level that achieve the noise
performance comparable to prior work but with lower power and area. The design of such
clock buffer with minimum power, area and delay overhead is discussed in Chapter 6.
25
Chapter 3
Timing jitter has been the subject of numerous studies ([22]-[39]) which provide
many models to predict the jitter of individual blocks in a PLL, in particular, different
types of voltage controlled oscillators (VCOs) due to device noise and supply/substrate
noise. While most of previous work focuses on jitter study of individual blocks, there has
been done less work on modeling the overal jitter at PLL output clock ([22] and [43]-
[46]). This research extends the previous work by investigating the effect of PLL
parameters such as bandwidth and damping factor toward minimizing output clock jitter
The common design practice for systems with low-noise input clock is to
design the loop with the highest possible bandwidth to eliminate the effects of noise
26
sources at the output. Very low bandwidth and high damping factor are commonly used to
filter a noisy input clock with a clean oscillator within the PLL. By understanding the
sensitivity of jitter to loop parameters, we can refine these common practices in designing
low-jitter PLLs. Section 3.1 reviews the definitions of timing jitter. The brief study of the
previous work on jitter optimization is discussed in Section 3.2. The noise sources in a
PLL are the subject of the next section. Section 3.4 extracts the relationship between the
overall rms jitter at the PLL output clock, the power spectral density of each noise source
and the correspondent PLL noise transfer function. In Section 3.5, the sensitivity of jitter
to PLL damping factor and bandwidth is first derived for second-order loops and then
extended to third-order loops. The sensitivity of jitter to loop parameters is studied for all
primary noise sources in a PLL. Section 3.6 describes the design of a tunable PLL that is
used to minimize jitter and to verify our analysis. Finally, the experimental methods and
results that verify the jitter analysis are given in Section 3.7.
between the first cycle and mth cycle of the clock (Figure 3.1). Timing jitter can be
T
∆T = m.T 1
σ ∆T = ⋅ σ ∆φ
ω0
27
expressed in terms of phase jitter by σ ∆T = ( T ⁄ 2π ) ⋅ σ ∆φ = ( 1 ⁄ ω 0 )σ ∆φ where the
clock period, T, is 2π/ω0. Timing jitter is called short-term jitter for small ∆T and long-
term jitter as ∆T goes to infinity. The tracking jitter, σtr, is a commonly used metric for a
PLL output clock. It is measured as the phase difference between a clean reference clock
and the PLL output clock as shown in Figure 3.2. The tracking jitter is related to timing
σ ∆T → ∞
jitter by σ tr = ------------------- at very large ∆T as shown in [22].
2
PLL CKref
PLL CKout
σtr
which depends on the VCO design. For the case of a first-order PLL with bandwidth of f-
3dB, the long-term jitter of the output clock due to VCO noise is calculated in [22] as
1
σ ∆T → ∞ = σ T = κ ------------------- . The first-order loop roughly approximates an overdamped
2πf – 3 dB
28
second-order PLL. The short-term jitter of the first-order PLL is calculated in [40].
Although, [40] conceptually discusses jitter in higher-order loops and for different noise
sources, it does not elaborate the impact of loop parameters on the output jitter. The
previous work in [42] investigates the effect of only loop bandwidth on jitter due to VCO
noise. Recently, the impact of the loop parameters on long-term jitter in an ideal second-
order PLL is studied [41]. While this con-current work achieves similar closed-form
equations for jitter as our analysis, it does not include higher-order effects of a PLL on
jitter.
In this work, we extend the jitter analysis to different noise sources and to any
second-order and third-order PLL loop parameters by including the delay and sampling
nature of the loop in the analysis. The main goal of this analysis is to provide a simple, yet
accurate model, to predict the short-term jitter as well as long-term jitter. The model
should also provide designers with some guidance for proper design of the loop
parameters for minimum jitter performance. First, we explain the primary noise sources in
(Vnin), VCO noise (VnVCO), and clock buffer noise (Vnbuf) as shown in Figure 3.3. Open
2
N Clk – in 2 en
loop noise psd of a clock source is equal to S φnin ( f ) = ------------------
- . Nin-CLK is K 0 ⋅ ------- [22]
f
2 2
where K0 ( Hz ⁄ V ) represents the gain of the clock source oscillator and en (V ⁄ Hz) is a
N Clk – in
white noise source. NClk-in is related to κ with κ = ----------------------- [22]. Being a clock source
ω in ⁄ 2π
29
as well, the VCO has a similar noise that can be characterized using Nvco to represent the
noise sources in the VCO1. For the buffer, open-loop noise psd is calculated by
N buf
S φn Buf ( f ) = ---------------------------
2 2
- where fBuf is the buffer 3-dB bandwidth (typically much larger
f ⁄ f buf + 1 2
2 en
than PLL loop bandwidth) and N buf = ( K delay ⋅ 2π ⋅ f VCO ) ⋅ ------- . Kdelay ( s ⁄ V ) represents
2
buffer delay variation to voltage noise. Multiplying Kdelay by clock frequency (fVCO)
The transfer functions from each noise source to the output of the PLL shape the
noise. For example, the loop transfer function from the input phase to the output phase is a
low-pass filter as seen from Equation 2.2. The lower the PLL loop bandwidth, the more
strongly the PLL rejects the input clock noise. Next section discusses and extracts the
relationship between the timing jitter at PLL output, each noise source and PLL loop
parameters.
1. VCO noise spectrum falls as 1/f2 for a bounded frequency range. At lower frequencies, it falls as 1/
3
f , and at higher frequencies, it flattens out. Since low-frequency noise is suppressed by the PLL, and high-
frequency noise is inconsequential to jitter (because it is so small), the 1/f2 approximation is a reasonable
assumption.
30
3.4 Jitter Calculation Model
The goal is to relate the timing jitter at the PLL output clock to each noise source.
As shown in Appendex A.1, the relationship between the timing jitter, σ∆T and noise
2 8 ∞ 2
σ ∆T = --------2- ∫ S φ ( f ) sin ( πf∆T ) df ‹3.1›
ω0 0
2 2 4 ∞
σ T = --------2- R φ ( 0 ) = --------2- ∫ S φ ( f ) df ‹3.2›
ω0 ω0 0
Figure 3.4 graphically depicts Equation 3.1 and as shown, reducing the area under the
phase noise psd lowers jitter at the output. The phase noise psd associated with each noise
source is shaped as each noise is filtered out by the loop transfer function of the PLL from
sin2(πf∆T) Sφ(f)
1/ ∆T
f
The filtering of the PLL on each input noise is included in the timing jitter by
replacing the noise psd in Equation 3.1 (or Equation 3.2) with closed-loop noise psd.
31
2
S φ ( f ) = S φn –c losed ( f ) = ∑ Sφn –o pen ( f ) ⋅
i
Hn i ( j2πf ) ‹3.3›
i
2
Hn i ( j2πf ) is the square magnitude of noise transfer function (NTF) from each input
φ out
phase noise to PLL output phase, i.e. --------- ( f ) = Hn i ( j2πf ) . Sφni−open(f) indicates the open-
φn i
loop phase noise of each noise source as calculated in Section 3.3.
Replacing the open-loop phase noise of each noise source, the total noise psd at the output
is given by:
Note that this analysis assumes white noise sources. The same analysis can be done for
2
en
colored noise sources (such as supply and substrate noise) by replacing ------- by
2 2
en 1
------- ⋅ -------------------------------
- where fnoise is the 3-dB bandwidth of the noise.
2 f 2 ⁄ f 2 noise + 1
The loop transfer function from the input phase to the output phase was calculated in
:N
VnVCO Vnbuf
Vnin
ICP
Input PD VCO Clock
KPD φout
Clock φnin KVCO /s φn Buffer φn
ICP R VCO buf
32
Section 2.4 (Equation 2.4). Similarly, the noise transfer functions from VCO1 and clock
buffer phase noise are calculated. The NTFs for three noise sources are2:
2
φ out K Loop RCs + K Loop 2ζω n s + ω n
Hn In ( s ) = ----------- = ------------------------------------------------------------- = --------------------------------------------
φn In 2 2 2
s + K Loop RCs + K Loop s + 2ζω n s + ω n
‹3.5›
φ out 2 2
s s
Hn VCO ( s ) = Hn buf ( s ) = ---------------------------- = ------------------------------------------------------------- = --------------------------------------------
φn VCO, buf 2 2 2
s + K Loop RCs + K Loop s + 2ζω n s + ω n
I CP
where K Loop = ----------- K PD K VCO , ω n = K loop , and ζ = K loop RC ⁄ 2 .
2πC
The NTFs for VCO and clock buffer noise are high-pass filters while the NTF for
input clock noise is a low-pass filter. Multiplying each noise source’s NTF with the
transfer function of the correspondent block provides the overall transfer function from
φ out RCs + 1
Tn In ( s ) = ----------- = ( K 0 ⋅ K Loop ) ⋅ --------------------------------------------------------------------
Vn In s ⋅ (s + K
2
RCs + K )
Loop Loop
φ out s
Tn VCO ( s ) = ---------------- = K VCO ⋅ --------------------------------------------------------
- ‹3.6›
Vn VCO 2
s +K RCs + K Loop Loop
φ out 1 s
2
Tn buf ( s ) = ------------- = -------------------------- ⋅ --------------------------------------------------------
-
Vn buf s ⁄ ω buf + 1 s + K 2
RCs + K Loop Loop
As seen from Equation 3.6, the overal loop transfer functions are low-pass filter,
band-pass filter and high-pass filter for input clock noise, VCO noise and clock buffer
noise, respectively. The overall transfer function for a clock buffer can be approximated as
1. For the VCO control voltage noise, the gain from the noise source to the VCO output phase is KVCO.
For power-supply noise, KVCO is substituted with the gain from supply noise to VCO output phase.
2. The loop multiplication factor is one.
33
ω buf
a high-pass filter because the buffer 3-dB bandwidth, f buf = ---------- , is typically much
2π
larger than the PLL bandwidth. Figure 3.6 demonstrates the overall transfer functions for
20
10
Noise transfer function (dB)
−10
−20
VCO noise
−30 (b)
−50 4 6 8 10
10 10
frequency (Hz)10 10
Figure 3.6: Loop transfer function from each noise source to PLL output
Equation 3.1. The noise transfer functions in Equation 3.4 are substituted from Equation
3.5.
34
3.5.1 Jitter due to VCO Noise
To study the effect of each noise source on jitter, we first consider the VCO noise
2 8 ∞ N VCO 2 2
σ ∆T = --------2- ∫ ------------
2
- ⋅ Hn VCO ( j2πf ) sin ( πf∆T ) df ‹3.7›
ω0 0 f
We first study the jitter due to VCO noise in an ideal second-order PLL.
By substituting the VCO NTF from Equation 3.5 into Equation 3.7:
4N VCO ∞ 2 2 2
2 s sin ( πf∆T )
2 ∫
σ ∆T = ----------------
- ---------------------------------------- ---------------------------
2
- df ‹3.8›
ω0 – ∞ s 2 + 2ζω s + ω 2 f
n n s = jω
2
4π N VCO ∞ ∆T ∆T 2
x t + ------- – x t – -------
2
σ ∆T -∫
= ---------------------- dt ‹3.9›
2 2 2
ω0 –∞
s
where x(t) is inverse Fourier transform of --------------------------------------- . For damping factors
2
-
2
s + 2ζω n s + ω n s = jω
smaller and larger than one, the jitter expression is as follows (Appendex A.3):
– ζω n ∆T
1 e sin ( ω d ∆T + θ ) cos ( ω d ∆T )
2 ------------
2ζω - + ----------------------- ⋅ ------------------------------------- – ---------------------------- ζ<1
4π N VCO n 2(1 – ζ )
2 ωn ζω n
2
σ ∆T = ------------------------- ⋅
2
‹3.10›
ω0 2 2
1
------------ – a∆T 2αβ α – b∆T 2αβ β
-– e ------------ + ------ – e ------------ + ------ ζ≥1
2ζω n a + b a a + b b
κ2
2 2 2 –a
where ω d = ω n ⋅ 1 – ζ , cos θ = 1 – ζ , a, b = ζω n −
+ ω n ⋅ ζ – 1 , α = ------------
b–a
b
and β = ------------ .
b–a
35
Figure 3.7-(a) shows the short-term jitter behavior for different damping factors.
The details of Figure 3.7-(b) is discussed in the next section. For ∆T of within a few
similarly to the time-domain step response of the PLL output phase with similar
dependence on the damping factor and bandwidth. The lower damping factor appears as
more peaking in short-term jitter. For small short-term jitter, damping factor should be
designed to be equal to or greater than one to avoid ringing in the jitter response.
5
0 0
0 50 100 0 50 100
1.5
5 1
(2)
0.5
0 0
0 50 100 0 50 100
5 1.5
1 (3)
0.5
0 0
0 50 0 50
Figure 3.7: Short-term jitter behavior with different f-3dB and ζ due to (a) VCO and
(b) clock buffering noise. ((1) f-3dB = 5.5% fref, ζ = 0.2 (2) f-3dB = 6.4% fref, ζ = 0.65 (3)
f-3dB = 11.4%fref, ζ = 1.63)
1
At large ∆T, long-term jitter converges to final value of κ ⋅ ------------- . Note that this
2ζω n
result is similar to the result derived in [41]. The sensitivity of jitter to loop parameters can
36
while ζ is constant results in Figure 3.8-(a) in which jitter is reduced proportional to
1
---------------- . Figure 3.8-(b) illustrates the effects of varying ζ (or peaking in the frequency
f –3 dB
response) with constant f-3dB. In the plot, fn is adjusted to maintain the same f-3dB while
sweeping ζ. For ζ less than one (or greater peaking in frequency response), long-term jitter
1
is proportional to ------- , but the sensitivity reduces as ζ increases. For ζ greater than 2 with
ζ
constant loop bandwidth, long-term jitter is relatively constant, independent of ζ value.
4
Constant ζ
Normalized RMS Jitter
2 (a)
1
0
0 5 10 15 20 f-3dB (%fref)
2
0.4 2 4 fn (%fref), ζ=1
Constant f-3dB
1.5
(b)
1
0.5
0 1 2 3 4 5 ζ
2.81.6 1.15 1.05 1.013 1.01 Peak(%)
Figure 3.8: Long-term jitter (due to VCO noise) as a function of: (a) loop bandwidth,
(b) loop damping factor
So far we investigated the effect of VCO noise using an ideal second-order PLL
without considering the effects of the third-order pole or the inherent loop delay in a
sampled system. In many PLLs, a 3rd-order pole is often included to filter control voltage
ripple. For high loop bandwidths, this pole degrades the phase margin and causes peaking
37
in the frequency response. A similar frequency response peaking occurs when accounting
for the delay in the feedback loop and the sampled-nature of the loop. These non-idealities
can be taken into account using Equation 3.2 with a more accurate NTF.
the output long-term jitter as bandwidth is increased for a second-order loop (curve-a),
third-order loop without loop delay (curve-b), and third-order loop with loop delay (curve-
c). In the plot, the 3rd-order pole is kept constant while the zero frequency is decreased
which simultaneously increases the open-loop cross-over frequency, ωc, and the damping
factor. The plots on the right illustrate the loop frequency responses for a 2nd-order, 3rd-
order PLL without and with loop delay as zero frequency (ωz) is decreased. Curve-a
shows the anticipated decrease in jitter due to the higher bandwidth and damping factor. In
curve-b, as the loop bandwidth nears the 3rd-order pole, the peaking in frequency
response increases due to phase margin degradation. Thus jitter is roughly flattened at
bandwidths higher than 3rd pole due to the opposing effect of peaking and bandwidth on
jitter. Accounting for loop delay (curve-c), the jitter increases at high bandwidth due to the
additional peaking in the NTF from more phase margin degradation1. A minimum exists
and is modestly flat over a significant range of loop parameter variations. This implies that
a loop designed near this minimum has an output jitter that is relatively insensitive to the
parameter variations that may be due to process, voltage and temperature (PVT).
1. To the first order, using the loop delay accounts for the effect of the sampled system. The
measurement results of Section 3.7 matches the simulated results from this model better than that from a z-
domain model using impulse invariant transformation [80].
38
5 4
2nd-order PLL
)2 de 4
nd r 3rd-order PLL
-o +delay
rd
1 er 2
ωz
3rd pole
0 0
0 0.2 0.4 10
Figure 3.9: Comparison of long-term jitter (due to VCO noise) in: (a) 2nd, 3rd order
loop (b) without loop delay and c) with loop delay
Analysis of the minimum indicates that it depends on all four variables (loop gain,
zero frequency, 3rd-order pole frequency, and loop delay) because each contribute to
phase margin degradation (Equation 2.7). The analytical results show that jitter is
minimum with PM between 30o and 45o. Consequently, the PLL bandwidth at minimum
jitter reduces as 3rd-pole frequency decreases or loop delay increases as shown in Figure
3.10. This result counters common practice of designing with large phase margins and
damping factor of 1 ⁄ 2 .
39
45
1/3 Tref
35
/2 Tref
delay = 1
30
25
Figure 3.10: PLL bandwidth (at minimum jitter) as a function of 3rd pole frequency
and PLL loop delay
Noise from the buffering and the input clock can be similarly analyzed using the
corresponding closed-loop noise psds. Similar to the VCO noise, we first analyze the jitter
behavior in an ideal second-order PLL. The final equations are summarized in Appendex
A.3 and Appendex A.4, respectively. Then, the jitter analysis is extended to a third-order
PLL taking into account the delay and sampling nature of the loop.
to VCO noise except for small ∆T where jitter is increased sharply due to the high-pass
filtering of the buffer NTF. Figure 3.7-(b) illustrates the output jitter for different ∆T with
40
To compare buffer noise magnitude with VCO noise, the jitter values are extracted
from Equation 3.10 and Equation A.8 (Appendex A.4) for ∆T → ∞ . The ratio of the
2 2 2 2 2
σ Buf ( N Buf ⁄ ω 0 ) ⋅ ω Buf m ⋅ K delay ⋅ ω 0 ⋅ ω Buf ⋅ e n buf ⁄ 2
------------
2
- ≈ -----------------------------------------------------------------------
2 2
- = -------------------------------------------------------------------------------
2 2 2
- ‹3.11›
σ VCO ( 4π ⋅ N VCO ⁄ ω 0 ) ⋅ ( 1 ⁄ 2ζω n ) 4π ⋅ K VCO ⋅ ( 1 ⁄ 2ζω n ) ⋅ e n VCO ⁄ 2
where m is the number of buffer stages. For a ring oscillator with the same delay elements
–1
as the buffering, the KVCO can be expressed in terms of Kdelay, K VCO = K delay ⋅ ----------------2-
2n ⋅ t d
where n is the number of stages in ring oscillator VCO and td is the delay of each stage.
2
σ Buf mζω n
- ≈ --------------
------------ ‹3.12›
σ VCO nf osc
2
With ωn=0.2fosc and ζ=1, in order for the noise contribution of the buffer to be less than
that of the VCO, either m<5n or the VCO element must have 5x lower noise sensitivity
than the buffer elements. With lower loop bandwidths, buffer noise contribution decreases
proportionally.
When accounting for the effect of the PLL filtering on a noisy input clock, the
analytical results1 for a 2nd-order PLL show that the output clock timing jitter is
41
suppressed at small ∆T and asymptotically approaches a value, κ 1 ⁄ ( 2ζω n ) , greater
than the input jitter at large ∆T. The shape and final value depend on the bandwidth and
the damping factor. Figure 3.11 illustrates the behavior of output clock jitter for different
damping factors with constant bandwidth. The figure also includes the behavior of input
−7
x 10
4
ζ=0.5
(Output Clock Jitter/κ)2
3 ζ=1.2
0
0 50 100 150 200
Figure 3.11: Output clock jitter (due to input clock noise) behavior vs. input clock
jitter behavior
clock jitter. The ∆T at which the jitter exceeds the input jitter (the crossover time, ∆Tcr) is
larger for higher damping factors and lower bandwidths. For most clock source PLLs,
jitter of the overall system is suppressed as long as ∆Tcr is longer than the response time of
any subsequent PLLs locking to the output clock. The jitter analysis due to noisy input
clock not only confirms common practice but also elaborates the roles of bandwidth and
damping factor on the output jitter. Figure 3.12-(a) shows how the output jitter (at ∆T=100
output jitter (at ∆T=100 cycles) is reduced as damping factor is increased for two different
42
bandwidths. Similar to VCO noise analysis, output jitter is roughly constant for damping
factor greater than 2. For instance, for output jitter to be less than 0.1 input jitter at ∆T>
100 cycles, the PLL should be designed with a damping factor greater than 2 and
100
Constant ζ
Output/Input jitter (%)
50 (a)
0 −1
10 f-3dB(%fref)
0 1 2
10 10 10
12 80
Constant f-3dB
10 70
f-3dB = 0.002% fref (b)
8 60
f-3dB = 0.1% fref
6 −1 50
10 10
0
10
1
ζ
Figure 3.12: Output to input jitter ratio behavior of a 2nd-order loop as a function
of: (a) loop bandwidth, (b) loop damping factor
To investigate the effects of the loop non-idealities, the jitter (due to input clock
noise) of an ideal 2nd-order loop is compared to that of a 3rd-order PLL with loop delay.
To better show the comparison, we assume white noise at PLL input phase instead of 1/f2
noise (of a noisy input clock). Figure 3.13 illustrates the output long-term jitter while the
zero frequency is decreased which simultaneously increases the loop cross-over frequency
and the damping factor. Jitter decreases initially for all three curves due to the lower
frequency-response peaking where the bandwidth changes only slightly. As the zero
43
frequency decreases further, the bandwidth increases causing jitter to increase. At
bandwidths close to 3rd pole, the peaking is increased due to phase margin degradation
which results in more jitter increase in curve-b compared with curve-a. When accounting
for loop delay (curve-c), additional peaking in the NTF from more phase margin
20
Output RMS Jitter (ps)
15
y
la
de
r+
de
10
or
d-
3r
o rd e r
b ) 3r d-
c)
5 o rd e r
a ) 2n d-
3rd pole
0
0 10 20 30
Figure 3.13: Comparison of long-term jitter (due to white noise at PLL input) in: (a)
2nd, 3rd order loop (b) without loop delay and (c) with loop delay
the noise from within the loop. A high-bandwidth PLL can track the phase of a low-noise
input clock and filter out VCO and clock buffer noise. Conversely, a low-bandwidth PLL
filters a noisy input clock while it is transparent to VCO and clock buffer noise. We design
a PLL with adjustable loop bandwidth and peaking in frequency response to verify the
44
results in the previous section. The parameters can be adjusted by varying the loop
One possible architecture [52] is shown in Figure 3.14. This PLL has an adaptive
CPintegral 1/gmReg
Regulator
PFD + Clock
CCP -
VCO Buffer
CKref CKout
C1
d10 d1n
CPproportional
d20 d2n
bandwidth with tunable loop parameters. The design employs two digitally controllable
charge pump currents in the proportional and integral paths to adjust ωz and Kloop:
ω = 1 ⁄ --------------
1 I CPproportional 1
z - ⋅ ---------------------------------- ⋅ C CP + --------------- ⋅ C 2
gm Reg I CPintegral gm Reg
‹3.13›
K VCO ⋅ I CPintegral
K Loop = ------------------------------------------
N ⋅ C CP
While the proportional charge-pump current varies the zero locus only, sweeping the
integral charge-pump current changes both the zero and the open loop gain. Varying any
of the two charge-pump currents does not vary the position of the PLL third-order pole.
45
3.7 Experimental Methods and Results
The adaptive bandwidth PLL clock generator with tunable loop parameters (shown
in Figure 3.14) is designed and fabricated in 0.25-µm CMOS technology. The PLL die
photogragh is shown in Figure 3.15 where the area overhead due to digital controller logic
Digital
Controller
Logic
Loop Filter
CP1 Reg
CP2 VCO
PFD CLK
Buf
ps) produces the reference clock and the design uses only a few buffer stages in the
To verify the presence of minimum tracking jitter due to VCO noise, the integral
charge pump current is kept constant (i.e. KLoop = constant) while the proportional charge
46
pump current is swept (i.e. ωz is decreased). For each value of ICPproportional, the rms
tracking jitter of PLL output clock is measured based on the configuration of Figure 3.16.
Digital Oscope
Reference Output
clock
clock
PLL Input
Trigger
Table 3.1 summarizes some of the results at reference clock frequency of 700MHz
Figure 3.17-(a) and (b) show the measured and calculated jitter for one set of
measurements repeated for two reference clock frequencies. As seen in the figure, the
measured jitter corresponds closely with the analytical results and there is a minimum
jitter with a low sensitivity to loop parameter variations. For example, ±20% of bandwidth
variation increases jitter by less than 5%. In each set of measurements, jitter initially
decreases because the peaking decreases (or ζ grows linearly) with ICPproportional and the f-
3dB increases with the decreasing zero frequency (fn is held constant). As ICP2 increases,
the cross-over frequency approaches the third-order pole and degrades the phase margin.
Jitter reaches a relatively flat minimum before increasing due to the loop delay
(approximately 0.47ns).
47
Increasing reference clock frequency from 700MHz to 1.1GHz in our adaptive
bandwidth PLL, effectively measures the result of changing the loop’s feedback delay
from 1/3 to 1/2 of the reference clock period. The bandwidth at minimum jitter is reduced
Measurement
Analytical results
2
0 5 26% 10 15
3
fref = 1.1 GHz
2 (b)
1
0 5 12% 10 15
5
tdelay1 > tdelay2 t delay
1
tdelay2
(c)
0 26% 10
0 512% 15 ICPproportional/ICPintegral
33 100 180 250 f-3dB (MHz)
48
Table 3.1: Tracking jitter (in ps) for different loop parameters (fref = 700MHz)
The short-term jitter sensitivity to PLL loop parameters is also verified. The short-
term jitter is calculated with the analytical model. The time domain figure of merit of the
and peaking used for the model are first calculated through circuit simulations and then
verified with direct measurements. The test setup that measures the loop parameters is
49
shown in Figure 3.18. A radio frequency (RF) signal is added to the input clock. The
Digital Oscope
Pulse Generator
Clock + jitter Output
Ref clock clock
PLL Input
Trigger
RF generator
(jitter source)
Figure 3.18: Measurement technique for calculating PLL loop transfer function
output clock jitter is measured over different RF frequencies. The measured PLL loop
transfer functions with their effective f-3dB and effective peaking (Appendex A.6) are
shown in Figure 3.19 for four different values of ICPproportional with constant ICPintegral.
3 f-3dB
(MHz) Peak ζ ICP2
|H(s)| (Loop Transfer Function)
1.5
0.5
0 0 2
10 10
Input RF Frequency (MHz)
Figure 3.19: Measured PLL loop transfer function (@ 700MHz reference clock) at a
constant ICPintegral (constant KLoop)
50
The rms jitter is measured over different time interval (∆T) for each of the four
shown in Figure 3.20. The dummy delay in the test setup is critical to compensate for the
Digital Oscope
Reference Output Dummy
clock clock Trigger
PLL Input
Delay Trigger
σ∆T
T
∆T
triggering delay of an oscilloscope. Figure 3.21 shows the measured and calculated short-
Measurement
Analytical results
10
5 (a)
f-3dB = 39 MHz, Peak = 2.8% (ζ = 0.2)
0
Output RMS Jitter (ps)
4 (b)
f-3dB = 45 MHz, Peak = 1.26% (ζ = 0.65)
2
0 50 100 150
4
3 (c)
f-3dB = 80 MHz, Peak = 1.07% (ζ = 1.63)
2
0 10 20 30 40 50 60 70
5
(d)
f-3dB = 320 MHz, Peak = 2.4% (ζ = 0.3)
0
0 5 10 15 20 25
Number of Cycles of CKref (∆T/Tref)
Figure 3.21: Measured and calculated short-term jitter (@ 700MHz reference clock)
for four different loop parameters
51
term jitter. A slight timing shift between predicted and measured jitter is present because
of time uncertainty due to the delay of input trigger and dummy trigger delay at the input
of oscilloscope.
at 700MHz as a reference clock of the PLL. A white noise source is injected to the control
voltage of the free running VCO so that the input clock noise is the dominant noise source.
(a) (b)
200 200
Output RMS Jitter2 (ps2)
150 150
) (4)
) (1
50 (1 (4) 50
0 0
0 5 0 5
Number of Cycles of CKref (∆T/Tref)
Figure 3.22: Output jitter (due to input clock noise) behavior for three different PLL
loop parameters: (a) measurement results, (b) analytical results ((1) Input jitter (2) ζ
= 0.2, f-3dB = 39MHz (3) ζ = 0.65, f-3dB = 45MHz (4) ζ = 1.63, f-3dB = 80MHz)
As the baseline measurement, we measure the rms jitter of this reference input over
different time interval (∆T) based on the self-referenced technique (Figure 3.20). We also
52
measure the PLL output rms jitter while varying ∆T for three different loop parameters.
The measurement results in Figure 3.22-(a) demonstrate the same behavior to the
3.8 Summary
This chapter investigates the role of PLL loop parameters on timing jitter. Several
common noise sources have been included in the analysis. We develop an intuition for
designing low-jitter PLLs both by deriving a closed-form solution for a second-order loop
and by plotting the jitter sensitivity to various loop parameters for higher-order loops. One
possible PLL architecture with digitally-controllable loop parameters is designed that can
optimize jitter performance. Furthermore, the loop serves as a test bench to verify our
analysis.
The analysis shows a simple expression for long-term jitter due to VCO and
buffering noise to the damping factor and natural frequency. We derive an expression that
relates the jitter contribution of clock buffering (in the feedback) and VCO to the same
parameters. We validate the common design practice of using high loop bandwidth to
reduce VCO-induced jitter. However, to minimize jitter, we find that accounting for the
loop delay in the phase margin is critical. Interestingly, this minimum is very insensitive to
PVT and parameter variations making such a design robust. For applications that require
small short-term jitter (i.e. short distance links and block to block interconnect), an
underdamped loop can result in much higher short-term rms jitter. For applications that
filters input jitter, our modeling shows that very low bandwidths (0.002% fosc) are
53
necessary to reduce noise by a factor of 10 while a damping factor greater than 2 is
sufficient.
The result of jitter analysis extracted in this chapter can be applied to the optimum
design of PLL loop parameters to minimize the PLL output jitter. The jitter optimization
requires a well-known knowledge about the noise sources in a PLL. Since the noise
sources are not predetermined, the preliminary design of loop parameters does not
performance, the loop parameters of a PLL should be tuned for a minimum output jitter in
real system noise conditions. The next chapter presents a methodology for on-chip jitter
minimization and verifies the accuracy of the method in converging to the minimum jitter
54
Chapter 4
The previous chapter shows that the output jitter of a PLL depends strongly on the
magnitude and frequency response of the noise sources and the loop parameters. For many
systems, the loop design is complicated because the magnitude of the noise sources is not
well known; a noisier clock reference may be used or larger on-chip switching noise may
be present. Jitter can still be minimized under various noise conditions if jitter can be
dynamically measured with an on-chip noise measuring circuit and the loop parameters
can be adapted with a programmable loop filter. This chapter investigates the
methodology and accuracy of jitter minimization that occurs during system operation and
Section 4.1 reviews the relationship between the minimum jitter and the loop
parameters for two noise sources, input clock noise and internal VCO noise, as extracted
55
in Chapter 3. It is observed that the total jitter due to the two noise sources has only one
minimum that is global for a range of loop parameters that the PLL is stable. This result
leads to the gradient-descent algorithm described in Section 4.3. The circuit components
needed to dynamically minimize jitter is described in Section 4.2. Several of the existing
circuits that can measure jitter both for clocking and for data recovery are discussed.
Section 4.3 discusses the algorithms that converge to the minimum jitter during active
system operation. Because jitter is a stochastic process, any on-chip measurements are
subject to errors depending on the amount of averaging. The section illustrates the
performance of the convergence as related to the amount of jitter information. The chapter
4.1 Overview
Previous chapter discussed the relationship between minimum jitter due to each
noise source and PLL loop parameters. The block diagram of a PLL with two primary
noise sources, input clock and internal VCO noise, is shown in Figure 4.1. Although this
:N
VnVCO
Vnin
Lowpass φout
Input PFD Filter VCO
Clock
φVCO + φnVCO
φin + φnin
Figure 4.1: The PLL block diagram with VCO and input noise
56
work only considers these two noise sources, the results can be extended to other noise
sources in a PLL.
Each of the noise sources is shaped by the loop transfer function from the
corresponding noise voltage source to the output phase. Figure 4.2 illustrates the filter
10
Noise transfer function (dB)
0
Input clock noise
(Vnin)
−10
−20
−50 4 6 8 10
10 10 10 10
frequency (Hz)
Figure 4.2: Loop transfer functions from VCO and input clock noise to the PLL
output
As seen in Figure 4.2, the loop transfer function for the VCO noise (VnVCO) is a
band-pass filter that suppresses the VCO noise within the PLL bandwidth. In a second-
order PLL, the long-term rms jitter due to VCO noise1 is calculated as:
1. VCO phase noise is assumed to fall as 1/f2. The long-term jitter is calculated from Equation 3.10
when ∆T goes to infinity
57
N VCO 1 -
σ rms = ----------------- ⋅ ------------ ‹4.1›
f0 2ζω n
where f0 is the VCO frequency, ζ is the PLL damping factor and ωn is the PLL natural
frequency. The two loop parameters that can be easily tuned in a charge-pump PLL are the
PLL zero frequency, ωz, and the PLL loop gain, Kloop. Sweeping ωz and Kloop effectively
changes the PLL bandwidth and peaking in the PLL frequency response. Substituting ζ
N VCO 1
σ rms = ----------------- ⋅ ------------------------ ‹4.2›
f0 –1
( ω z ) ⋅ K Loop
Based on Equation 4.2, the relationship in a second-order PLL between the VCO-induced
jitter and the loop parameters, (ωz)-1 and Kloop, can be shown to be convex and hence has
only a global minimum without local minima. The jitter behavior as a function of (ωz)-1
and Kloop, in a third-order sampling PLL is graphically shown in Figure 4.3. The plot
includes the higher-order pole and sampling/feedback delay and still maintains the
convexity. The minimum jitter, as shown with the contours in Figure 4.3-(a), occurs at a
high loop bandwidth with low peaking in the PLL frequency response. As the bandwidth
is further increased, the phase margin degrades which increases the peaking and
58
25
15
(a)
10
5
15
10
1
K
15
x 10
l oo 5 0.5
p - 1
−7
x 10
0 0 (ω z)
x 10
15 Min jitter
10
9
10
10
8.7
8.4
11
9
11
9
8.7
7
Kloop
9
13
10
10
11
6
(b)
11
5
10
4
11
13
11
3
20
156
18
1
1 2 3 4 5 6 7 8
(ωz)-1
−8
x 10
Figure 4.3: Behavior of output clock jitter due to VCO noise for various loop
parameters: (a) 3-D, (b) contour
In contrast with the VCO noise, the loop transfer function for the input clock noise
is a low-pass filter that suppresses the input noise outside the loop bandwidth. The long-
term jitter due to input clock noise1 in a second-order PLL is equal to:
1. The input clock noise is assumed to be white, i.e. S φn – in ( f ) = N Clk – in . The detail of long-term
jitter calculation is given in Appendix A.5.2.
59
N Clk – in –1 1
σ rms = ----------------------- ⋅ K loop ⋅ ( ω z ) + --------------- ‹4.3›
2πf 0 (ω )
–1
z
14
12
rms jitter (ps)
10
(a)
8
6 15
4 10
15
0 p
K loo
0.2 5 x 10
0.4 0.6
−7 (ωz ) -1 0.8 1 0
x 10
15
x 10
10
9
8
11
9
12
7
8
10
6
11
9
7
8
Kloop
(b) 6 10
5
7
5.
9
6
5 8
5. 7
4 5
4.9
6
3 7
5.5
(ωz )-1 x 10
Figure 4.4: Behavior of output clock jitter due to input noise for various loop
parameters: (a) 3-D, (b) contour
60
It can be shown that the relationship in Equation 4.3 is not convex1. Including the
higher-order pole and sampling/feedback delay, Figure 4.4 plots the output jitter due to
only input-clock noise for a third-order sampling PLL. As seen in the figure, no local
minimums exist. The concavity of the surface is not very apparent and only occurs when
the phase margin is small (<30o). Such small phase margin is an unlikely operating point
due to possible loop instability. Similar to VCO noise, a single minimum exists except that
The total jitter is the sum of the jitter variances due to both VCO and input noise
sources. Figure 4.5-(a) shows the total jitter when two noise sources are comparable. As
one noise source becomes dominant, the minimum point of the contour shown in Figure
4.5-(b) moves toward the minimum for that particular noise source (Figure 4.3-(b) or
Figure 4.4-(b)).
Although, the jitter function due to the input noise is not entirely convex, it is shown in
Appendix A.8 that the total jitter in a second-order PLL has one global minimum without
any local minima. Simulation results for various ratios of VCO and input noise sources
show that the single minimum holds even when including a higher-order pole and
algorithm of Section 4.3 when the loop parameters are dynamically adjusted to achieve
minimum jitter.
61
30
25
rms jitter (ps)
(a) 20
15
15
10 10
0 15
p x 10
0.5 5 K loo
−7 (ωz ) -1 0
x 10 1
15
x 10
10
20
14.8
14.8
VCO noise
15.5
9
16
15.5
16
22
18
24
7
20
Kloop
22
14.
(b)
14.8.5
18
6
16 5
14.2
1516
18
15
.
14.2
5 20
Input noise
4 18
14
16
15
.8
22
14
20
3
.5
.8
18
16
1 2 3 4 5 6 7 8
(ωz)-1 x 10
−8
Figure 4.5: Behavior of output clock jitter due to both VCO and input noise for
various loop parameters: (a) 3-D, (b) contour
62
4.2 Jitter Detection Circuits and Architectures
To dynamically minimize jitter at the PLL output during system-operation, the
design requires three elements: 1) a PLL that has appropriately adjustable loop
parameters, 2) an on-chip jitter measuring that can compare the jitter between
measurements, and 3) an algorithm that adjusts the loop parameters to minimize jitter
based on the on-chip measurements. The first two are discussed in this section.
significantly impact jitter are the loop bandwidth and peaking in the frequency response.
They can be adjusted by varying the loop stabilizing zero and the open loop gain. One
possible PLL architecture is the one used in Chapter 3 to verify the jitter analysis (Figure
3.14). While the proportional charge-pump current varies the zero locus only, sweeping
the integral charge-pump current changes both the zero and the open loop gain.
In the second configuration shown in Figure 4.6, ωz and Kloop, are independently
adjustable by varying the loop stabilizing resistor (R) and charge pump current (ICP),
respectively:
ω z = 1 ⁄ ( R ⋅ C CP )
K VCO ⋅ I CP ‹4.4›
K ≅ -------------------------
-
loop N ⋅ C CP
63
In this configuration, third-pole does not move as ICP or R are changed.
:N
CP
ICP Vctrl
PFD VCO CKout
CKref ICP
R C1
d0 dm-1 CCP
cn-1
c1
c0
Up
d2 d1 d0
Controller
Figure 4.6: A PLL architecture with adjustable loop parameters using adjustable R
and ICP
measurements. The design permits 4-bits of digital adjustment for resistor that varies the
zero position by more than 10x (from 0.1 to 1.6 rad/sec). In this implementation of the
adjustable resistor, the resistance steps with non-linear digital quantization levels1. The
design also permits 3-bits of digital adjustment for charge-pump current that varies the
loop gain by 5x (from 2e15 to 10e15 (rad/sec)2) with a linear quantization level.
64
4.2.2 On-chip Jitter Measurement Architectures
The on-chip jitter measurement circuit depends on the application. This section
first describes the approach for a data recovery system. Possible approaches for on-chip
1) Data-Recovery Applications:
In data-recovery applications, clocks sample not only the center of the data eye to
recover the data pattern but also the data transitions to determine phase information. The
goal of the PLL is to track the data jitter while rejecting the noise from the VCO. By
correlating the sampling clock with the data transitions, the loop minimizes the phase error
In [67], a flash time-to-digital converter (TDC) measures the data jitter with the sampling
clock. This technique requires significant number of arbiters and on-chip buffering of the
Data
Buffer
Clock
Arbiter
Decoder Logic
65
Another technique demonstrated by [68]-[70] uses a dead-zone phase detector to
measure the jitter. Figure 4.8 illustrates the basic concept of the jitter measurement. A
D1 D2 D3
Histogram DCK
XCKL XCKR
sampling the data in the middle of the eye. The transition sampling clocks, XCKL and
XCKR, are programmed to track the left and right edges of the data eye and adjust the
dead-zone width, WDZ, accordingly. The design in [70] uses only one data-transition
sampler to construct the dead-zone window (WDZ) by alternating the edge sampling clock
position. The data transition outside the window is detected when the value of data
sampled by the transition sampling clock is equal to that sampled by the data sampling
clock. The magnitude of jitter is estimated by comparing the number of data transitions
outside the dead-zone for a given total count of data transitions. The window size is
adjusted when the number of transitions (measured hits) outside the zone is greater or less
than predetermined bounds to avoid saturating the counters. A similar method can adjust
66
the width of the dead-zone window until the number of measured hits is roughly a fixed
percentage of the total hits. This effectively directly measures the width of the jitter
histogram. The dead-zone technique is the measuring jitter circuit that is mimicked in the
next section.
large digital system’s clock. A similar architecture to [67] has been shown in [71] where
an array of phase detectors compares consecutive clock edges and measures the cycle-to-
cycle jitter. However, it is important to note that cycle-to-cycle jitter can not be minimized
through adapting the loop parameters. As shown in Section 3.5, cycle-to-cycle jitter is
primarily determined by the noise characteristics of the VCO alone and not the PLL loop
parameters. Adjusting loop parameters may result in large long-term jitter or an unstable
loop.
In the event that the long-term tracking jitter is important, a circuit that
accumulates phase over multiple cycles is necessary. The design of the accumulation
circuit is very challenging because it must strongly reject supply and substrate noise. A
simple delay line that spans multiple cycles is not adequate because a multi-cycle on-chip
delay line would likely introduce a significant noise floor to the measurement. Integrator
techniques similar to that used by Wavecrest SIA-3000 would suffer similar issues on-
chip. The design of this challenging circuit is left as future work and not addressed in this
work.
67
4.3 Jitter Minimization Algorithms and Measurements
Due to the stochastic nature of jitter, the measurement accuracy is a function of the
number of the samples. In addition, the jitter measurement circuit itself introduces some
noise. After describing the measurement setup, this section discusses the sensitivity of the
measurement to the total number of samples. Next, two jitter minimization algorithms are
0.25-µm CMOS technology. The chip die photograph is shown in Figure 4.9. To
proxy of the on-chip dead-zone phase detector circuit to measure the jitter of the output
clock. None of the features of the scope such as rms or p2p jitter information is used.
Instead, only the histogram data is downloaded to the computer through the GPIB port. By
percentage of the total number of transitions (total hits), the measurement replicates that of
a dead-zone phase detector. The number of measured hits (or percentage) is an indication
of the jitter magnitude1. The dead-zone width is adjusted when the number of measured
hits (outside the dead-zone) exceeds 1-10% of the total hits. The histogram can also model
1. The total jitter for a data-recovery system is the sum of the data jitter and sampling clock jitter.
68
the histogram can directly determine the width of dead-zone window such that the number
D/A
D/A
PLL controlling
controlling
resistor
CP current
Figure 4.10 shows the measurement setup. The PLL loop parameters (ICP and R)
are changed by D/A converters controllable by a data-acquisition board. The digital scope
69
PLL TestChip 3
D/A (ICP) Data Acquisition
CK D/A (R) Board
in 4 Computer
CKout (1 GHz)
GPIB “C”
Interface program
Digital Scope
Trigger
Pulse
Generator 250 MHz
Figure 4.10: Test setup for the jitter measurement and optimization
sensitivity of the measurement to the total number of hits. The inherent randomness of
jitter results in some measurement error that will degrade the performance of the jitter
minimization.
zone window varies between measurements. The percentage forms a distribution where
the standard deviation of the distribution is inversely proportional to the total hits, N, in
the histogram. Figure 4.11-(a) illustrates four distributions of the percentage of hits
outside the dead-zone. The curves represent two values of total hits (N=500 and N=5000)
70
and two dead-zone positions (WDZ=4σ and WDZ=5σ where σ is the jitter standard
deviation). The additional shaded lines illustrate the impact on the measurement when
0.04
N=5000
WDZ=5σ
0.03
N=5000
Distribution
WDZ=4σ
(a) 0.02
N=500
N=500
0.01 WDZ=5σ
WDZ=4σ
0
0.02 0.04 0.06 0.08 0.1
#hits (% total)
2
1.8
Standard deviation (%)
1.5
(b) 1
0.79
0.5
300
0
0 1000 5000 10000 15000
#hits (N)
Figure 4.11: (a) Measured percentage hits distribution for one set of PLL loop
parameters for N=500 and N=5000, (b) standard deviation of measured percentage
hits
there is a ∆WDZ of 0.1σ. Figure 4.11-(b) shows the measured standard deviation of the
measured hits as a function of N. Increasing the number of hits from 300 to 1000 reduces
the standard deviation of the measured percentage from 1.8% to 0.79%. With very large
71
number of hits and at the cost of more hardware and time, the jitter measurement
uncertainty can be reduced such that noise of the jitter measurement circuit dominates the
uncertainty.
VCO and input noise has only one global minimum without any local minima for a range
of loop parameters that the PLL is stable. Jitter can be dynamically minimized by an
algorithm that descends the gradient. However, the jitter measurement uncertainty can
degrade the performance of the descent algorithm or cause the algorithm to fail. To
understand how the uncertainty affects the algorithm, a table-lookup method is first
measuring the jitter for all values of ICP and R during a system calibration, the results in
the table can be compared to find the global minimum. The method adapts to the jitter
environment only during explicit calibration periods. Figure 4.12 shows a table-lookup
measurement only due to VCO noise as input clock is supplied from a clean signal
generator with long-term rms jitter less than 1ps. Figure 4.12-(a) on the left illustrates a
contour of the measured hits (as a percentage of total hits) for each loop parameter setting
with the large total hits, N=30khits. The minimum jitter, shown in the figure, is in
agreement with the absolute minimum from simulation. Reducing the number of total hits
72
results in greater measurement uncertainty. The minimum value, from measuring all table
values once, may deviate from the absolute minimum. The contours, overlaid in the figure
on the right, indicate the range of possible minima for smaller number of hits (N=300 and
N=3000). As expected the contour for N=300 is larger than N=3000. Figure 4.12-(b)
15
Min jitter
x 10
6.5 15
20
15
3 3.5
2.5 x 10
9
60
9
4
15
5
3
2.
6
3.5
9
40
9
N=3000 hits
3
6
4
15
15
6
20
80
5.5 3.5 N=300 hits
64
60
3.5
4
6
15 40 15
5 20 5
Kloop
Kloop
4 4
20
9
9
9
40
4.5
(a)
6
15
15
6
15 4
4
15
6
6
9 9 9
3.5 15
9 3
20 9
3 15 15
60
40
2.5
2 4 6 8
15
-1 −8
1 2 3 4
20
5 6 7 8
(ωz) x 10
-1
(ωz)
−8
x 10
Min jitter
15
x 10
6.5 16.3
23
20
18
17. 15
30
1 40 x 10
40
23
30
18
30
6 17.1
18
30
N=3000 hits
23
20
23
18
6
20
N=300 hits
20
5.5
18
5
18
Kloop
30
5
20
30
Kloop
(b)
23
30
20
23
23
23
4.5
40
30
20
4 4
60
20
23
23 23
3.5 23
23 30
23 3
3 30
80
40 30
2.5 2 4 6 8
1 2 3 4 5
30
6 7 8
(ωz)-1 x 10
−8
(ωz)-1 x 10
−8
Figure 4.12: Jitter measurement contours (due to VCO noise) for all loop
parameters with (a) constant dead-zone width and measuring hits (percentage), (b)
constant 4% measured hits and measuring dead-zone width
73
illustrates the same measurement by finding the width of the dead-zone with 4% of the
hits outside the zone. The contours represent actual measured jitter in picoseconds. The
impact of N is also overlaid in the figure on the right. Notice that the methods yield
essentially the same results. The added uncertainty for using an N>3000hits gives
Similar measurements are made to show the impact of the combined input and
VCO noise for all measurements. In the test setup, the VCO noise is mainly due to the
thermal noise whereas the input noise is adjustable. Figure 4.13 illustrates the contours (in
ps) for a large number of hits (N=30khits). The figure illustrates the case that the input
jitter is dominant. As the input noise is reduced, the minimum jitter moves upward toward
the minimum jitter point shown in Figure 4.12, in agreement with the simulation results.
The contours, overlaid in the figure on the right, illustrates the range of possible
minimums for smaller number of hits (N=3000 and N=300). The added uncertainty for
using N=300 is 9ps (20% minimum jitter). For N>3000, the added uncertainty is <5ps
(<10% minimum jitter). It should be noted that the local minimum seen in Figure 4.13 is
due to the errors of measuring dead-zone width. The local minima appears where the long-
term jitter difference between neighboring loop parameters is less than the measurement
error. The measured long-term rms jitter does not show any local minima.
74
15
x 10
6.5
84 Flat range of jitter x 10
15
66
84
6
90
66
84
66 6
5.5
66
84
5
5
Kloop
59
84
4.5 90
84
N=300 hits
59
66
66
84
66
56
84
66
59
56
4 4 59
56
52
56 84 90 N=3000 hits
8
3.5
52
66 4
66 59
59 3 52
59
56
3 Min jitter 56 66
56
84
559
66
52
52 50
N=3000 hits 6
84 0
59
59
66
2.5
56
56 2 4 6 8
9
47 50 52 −8
1 2 3 4 5 6 7 8 x 10
(ωz )-1 x 10
−8
Figure 4.13: Jitter measurement contours (due to input noise) for all loop
parameters with constant 4% measured hits and measuring dead-zone width
It is important to note that the variable range of the loop parameters is bounded.
With excessively large adjustment range, the loop may become unstable and lose lock at
the extreme values. If the loss of lock occurs during calibration, it can be detected by
observing the PLL control voltage and the PLL can be forced to reset. However, the range
that descends the gradient. This allows dynamic jitter minimization during run-time with
uncertainty at relatively flat regions of the jitter surface causes difficulty for the algorithm
75
to converge to the minimum. Proper initialization can improve the performance. One
viable choice is to use the table-lookup results from system calibration as the starting
point. In another option, the algorithm could be initialized with lower loop gains (Kloop =
2.5e15 in Figure 4.13). If the loop is dominated by input noise, the initialized value is
close to the optimum. If instead VCO noise dominates, the steeper slope of the surface at
lower loop gains (as shown in Figure 4.12) allows a rapid descent to the correct minimum
Figure 4.14 shows the flow chart for a descent jitter minimization algorithm. First,
the PLL is initialized to the starting values of the loop parameters (R[n], ICP[m]) and the
output clock jitter is measured. The width of the dead-zone is also initialized. Based on the
nearest neighbor measurements, the algorithm chooses the direction of descent for the first
loop parameter (R). The first loop parameter (R) is swept until the minimum jitter is found
while keeping the second parameter (ICP) constant. Then the algorithm chooses the
direction of descent for the second loop parameter (ICP) starting from (R[k], ICP[m]). The
minimum jitter for the second parameter is found for a fixed first parameter. The algorithm
repeats alternating between the two loop parameters. Several flags are used to keep track
Since the loop-parameter adjustments are digitally quantized, the curvature of the
jitter function, in particular as a function (ωz)-1, may be large enough such that the descent
gradient needs to be diagonal. The algorithm is designed to check diagonal neighbors once
76
a minimum is reached. An alternative is to add two more digital bits to reduce the non-
Initialize PLL
Is current jitter no
larger than previous? Store jitter into prev.
no
yes Increment
End of parameter parameter
sweeping?
yes
Set direction flag
yes
Count Min
reset flags
Similar to the jitter optimization with a table method, the algorithm converges to a
range of possible loop-parameter settings. Figure 4.15 shows the histogram of the
77
movement of the algorithm. The z-axis indicates the number of times the algorithm lands
at each loop setting. For N=3000 hits, the minimum jitter mostly occurs at the global
minimum while for N=1000 the minimum jitter moves over several loop settings as the
algorithm runs. The method converges to the minimum jitter that is higher than the
absolute jitter by <10% for N=3000 (or <20% results for N=300).
100 100
50 50
0
0 1
1
0.5 0.5
( (ω 2
x 10 ωz -1 0 8
2 −7 4
6 4 x 10 z ) -1 0 8 6
−7
Kloop 15
x 10 Kloop x 10
15
Figure 4.15: Measured minimum jitter due to the sum of VCO and input noise for
(a) 3000hits, (b) 300hits
parameters of the PLL for minimum jitter. The jitter analysis and measurements reveal
several key considerations when implementing the algorithm. First, for a rapid
convergence, the starting point of the algorithm is important. Calibrating the system upon
startup with a table of measurements would produce a near minimum initial point. As long
as the noise conditions change slowly, the algorithm will safely adapt. Second, the long-
78
term jitter magnitude must be measured and not cycle-to-cycle jitter. Adapting loop
parameters based on short-term jitter may result in an unstable loop. The paper shows that
applications. However, the measurement uncertainty limits the performance. Due to the
uncertainty, an algorithm would result in an operating point that wanders over a region.
Choosing a total number of hits >3000 produces reasonable results of <5ps of added jitter
from the uncertainty. The implication of a large number of hits is that ≥12-bit accumulator
A third issue in implementation is that the PLL needs to have adaptable loop
parameters and careful implementation is needed. Varying the loop parameters could
cause static phase offsets which would shift the dead-zone window and cause
proportional to the step-size. The injected charge must be sufficiently small so that the
loop can track and not lose phase lock. To ensure that the jitter measurements circuits
collect steady-state jitter information, a delay is needed between changing the loop
parameters and collecting hits. The waiting period corresponds to the time needed for the
loop to settle well within the measurement uncertainty at the worst-case parameter
settings. Since the tracking depends on the loop bandwidth, an intelligent implementation
would adapt the waiting period based on the digital loop-parameter settings.
79
An implementation using digitally-programmable loop parameters gives the
greatest flexibility in the design of the algorithm. A fourth issue to consider is the range
and the digital quantization of the loop parameter. The range must be bounded by the
ability of the loop to remain in lock especially if the algorithm operates when the system is
active. The quantization or resolution of the parameter adjustment has a similar constraint.
Typically, large quantization steps results in long waiting period and the risk of losing
lock. The resolution of 4-bits and 3-bits of the design shown in this paper, that provided
10x and 3x range for resistor and charge-pump current, is sufficient. However, the 4-bit
resolution for resistor is not fine enough to avoid being trapped in a false minima.
Although the jitter function is relatively flat over the minimum, the algorithm may become
stuck in the descent. The previously shown algorithm checks diagonals to alleviate the
problem. A more robust and less complex solution is to increase the resolution by two bits.
4.5 Summary
This chapter demonstrated a run-time technique that minimizes jitter at a PLL
output clock. This work addresses the considerations in the design of the PLL and the on-
chip jitter measuring circuit. Based on design considerations and jitter analysis results, an
the loop parameters and not knowing noise conditions a priori, jitter can considerably be
higher than the minimum. This work shows that jitter of a PLL can be minimized to within
80
This chapter concludes the jitter minimization method based on the PLL loop
designed with high immunity to noise. Due to switching activities in large digital systems,
power-supply or substrate noise, in particular, are of concern. The next two chapters
presents innovative circuit techniques in implementing PLL components and clock buffers
81
Chapter 5
substrate noise perturb the most sensitive blocks in a PLL such as voltage-controlled
oscillators (VCOs) which can significantly degrade the jitter performance of the PLL.
Prior state-of-the-art designs implement VCOs with high immunity to supply or substrate
noise with the cost of power and area. This research focuses on a new filtering technique
in the design of a VCO. The primary goal is to achieve similar noise performance as prior
focuses on the design of PLL that operates over a wide frequency range with adaptive
1. Wide operating frequency range allows to test a microprocessor or implement multi-rate links
82
techniques in the design of the loop filter. The loop filter is also designed with digitally
and Chapter 4. This work also addresses the limitation of the conventional PFD. It
proposes new circuit techniques for design of high-performance PFDs that achieve larger
Section 5.1 demonstrates the proposed charge-pump PLL architecture. Section 5.2
discusses the design of a low-power VCO with high immunity to noise. The design of a
self-biased loop filter will be discussed in Section 5.3. The design of high-performance
phase-frequency detectors (PFDs) are introduced in Section 5.4. The measurement results
are discussed in Section 5.5. The chapter concludes with summary performance of the
produces the VCO control voltage. The VCO is composed of a voltage to current (V-I)
signal of the VCO passes through a low-to-full swing (L-F) amplifier and feeds back to
83
:N
ICP
Vctrl L-F
PFD V-I CCO
Input ICP Amp Output
Ref. R Noise- C Clock
Clock CCP Canceling
Circuit
The primary design goals for the proposed PLL are: 1) to achieve high supply/
substrate noise rejection with adding a noise-canceling circuit to the VCO, 2) low power
and low area, and 3) to operate over a wide frequency range with an adaptive bandwidth.
because any noise coupled into the VCO control voltage is directly translated to the
change in the oscillation frequency. The change in frequency appears as a phase error
which is persistent for the time duration equal to the time constant of the PLL. The jitter
accumulation issue becomes more sever for lower loop bandwidths or higher loop
frequency multiplications. To remedy the problem, design of a VCO with high immunity
84
In high-performance digital systems, CMOS delay buffers are typically used to
implement voltage-controlled oscillators (VCOs) due to their wide tuning range, portable
design and relaxed supply headroom requirement. However, they have high noise
sensitivity to their control voltage (or VDD); 1%-delay/1%-VDD. The next two sections
discuss several techniques that improve the noise performance of CMOS buffers. First, the
advantages and drawbacks of prior design techniques are discussed. The design of the
filter the supply voltage using either a passive or active filter [51]-[55]. Designs in [51]-
[52] employ voltage regulators to filter out supply noise (Figure 5.2). Filtering a high-
VDD
Vctrl Cfilter
Regulator
frequency supply noise requires a supply coupling capacitor (Cfilter) [51] that shunts the
noise. The capacitor can occupy large area. Alternatively, a high-bandwidth regulator [52]
configuration that boosts the output resistance and rejects noise. Similarly, [54]-[55] use a
85
feedback cascode to boost the output resistance of the V-I converter circuit. Figure 5.3
transconductance amplifier (OTA) [54]. Although supply regulation and feedback cascode
techniques rejects the supply noise significantly, they typically consumes significant
amount of power to supply the VCO and clock buffer. A second technique is through
VDD
OTA C
Vctrl Cfilter
R
improving the supply sensitivity of VCO elements. A common design strategy employs
differential topologies. Differential VCOs and clock buffers ([39], [56]-[57]) demonstrate
filtering techniques, the differential elements consume significant power, especially in the
The next section discusses the design of the VCO with a new filtering technique
that reduces supply/substrate noise with less power consumption and area than prior
designs.
86
5.2.2 Proposed VCO Design
The four primary goals in design of the VCO are: 1) high static and dynamic
power-supply noise rejection ratio (PSRR), 2) low power and low area, 3) wide operating
frequency range, and 4) linear gain for the entire range of the control voltage (Vctrl).
Figure 5.4 shows the proposed VCO design. To achieve a wide operating
VDD
Mp1
Mp2 Mp5
Wp Wp Wp
VDD α
VDD
I0 Mp4 α>β>1
Mn3 Rout
φ0
ISF
Mp3
φ90
Vctrl VCCO
Mn1 Mn2 CCO φ180
Icomp ICCO
IDrv Source φ270
follower
Mn4 Mn5 C
Wn β.Wn
Feedback cascode
Noise-canceling circuit
V-I Converter
frequency range, the design uses a CMOS inverter ring oscillator with controllable supply.
Figure 5.5 shows the current-controlled oscillator (CCO) circuit composing of four stages
elements to enable the VCO to run faster at a given Vctrl. The CCO produces quadrature
87
clock phases, making the design suitable for applications such as clock/data recovery
ICCO
ICCO
inp+ inp-
o- o+
inn+ inn-
The V-I converter circuit, transistors Mn1, Mp1-Mp3, converts the control voltage
to current (IDrv) that drives the CCO and controls the frequency of CCO output signal. To
maintain linear VCO conversion gain (KVCO), Mp1-Mp3 are designed with large widths
for minimum overdrive voltage. The minimum overdrive voltage of PMOS transistors
guarantees the linear KVCO due to the fact that Mn1 stays in saturation for almost the entire
range of control voltage VCO, VTn1 ≤ Vctrl ≤ VDD, where VTn1 is the threshold voltage of
Mn1. However, at a Vctrl that is near VDD, Mn1 enters triode region which reduces the
conversion gain and saturates KVCO. To compensate for the gain drop at high Vctrl, the
88
circuit uses a source follower transistor (Mn3). Source follower is off for Vctrl - VCCO <
VTn3 and gradually turns on at high Vctrl which injects current (ISF) and compensates for
IDrv drop.
Figure 5.6 shows the simulated V-I converter gain characteristics for different
process corners. The proposed V-I converter achieves the linear gain that varies only by a
9
x 10
3.5
FF
3
VCO Clock Frequency (Hz)
2.5 FS
TT
2 SF
SS
1.5
0.5
0
0.5 1 1.5 2
Vctrl (V)
Figure 5.6: Simulated V-I converter gain characteristic across process corners
factor of less than 1.5 for almost the entire range of the control voltage (VTn1 ≤ Vctrl ≤
VDD). For instance, the KVCO varies between 1.15 and 1.7GHz/V at typical corner for
VTn1 ≤ Vctrl ≤ VDD. The slight variation of KVCO modestly impacts the loop dynamics. If
low-VT devices were available in the process technology, using one for the follower
would further improve the gain linearity at high VCO frequencies. The V-I converter in
[60] achieves a linear gain for the entire range of Vctrl (0 ≤ Vctrl ≤ VDD), slightly larger
89
than the range of this proposed V-I converter. However, the V-I converter in [60] suffers
from high power-supply noise sensitivity due to the coupling of Vctrl to both ground and
VDD. The gain linearity improvement technique proposed in this work resolves the
capacitor and output resistor (Rout) at VCCO forms the third pole of the PLL and filters the
high-frequency noise. The cascode current source that supplies IDrv uses a feedback circuit
(Mp4 and Mn2) to boost the output impedance [55]. The resulting supply noise sensitivity
is 0.2%-VCO frequency/1%-VDD because the finite output resistance of Mn1 causes IDrv
to vary with supply. An auxiliary noise-canceling circuit (Mp5, Mn4 and Mn5) is added to
compensate the residual variation of the output current (IDrv) due to supply noise. This
circuit generates a compensator current, Icomp, by mirroring a fraction of I0. Icomp is then
subtracted from IDrv. The current to the CCO is I CCO = I Drv – I comp for
V ctrl – V CCO < V Tn3 . The ideal supply noise cancellation occurs when IDrv variation is
equal to Icomp variation due to VDD noise, i.e. ∆I Drv = ∆I comp . In other words, when
designed to have a much worse supply sensitivity than the feedback cascode circuit that
generates IDrv. The noise-canceling circuit uses a single device without the feedback
cascode and with minimum channel length. The simulation result shows that Icomp is 4
∂I comp ⁄ ∂V DD
times more sensitive to VDD variation than IDrv, i.e. ---------------------------------- = 4 . By setting the
∂I Drv ⁄ ∂V DD
β 1
ratio of the mirroring, β/α, to the ratio of the supply sensitivity of the currents ( --- = --- ),
α 4
90
∆IDrv will be equal to ∆Icomp1. The power penalty to source the same ICCO for a given
To verify the noise performance of the proposed V/I converter, the dynamic
response of VCCO to supply noise is simulated. The curves (1) and (2) shown in Figure 5.7
demonstrate the VCCO response for the V/I converter without and with the noise-canceling
circuit, when a -10% VDD step with 100ps slew rate inserted at t=2ns. Adding the noise-
canceling circuit to the V-I converter improves the PSRR by 6dB for very high frequency
797
796
795
VCCO (mV)
(4)
794
(3)
w/ noise-canceling circuit:
(2)
793
792
791
(1) w/o noise-canceling circuit
790
2 4 6 8 10
Time (ns)
Figure 5.7: VCCO response of V-I converter to -10% VDD step inserted at t=2ns
noise. Increasing the slew rate of VDD step from 100ps to 1ns and 5ns (curves (3) and (4))
improves the dynamic PSRR of the V-I converter with noise-canceling circuit to 8dB and
1. Adjusting β/α alleviates any output impedance variation over the process corners. The
simulation results indicate that the proposed VCO maintains its noise rejection performance at the
process corners by adjusting the value of β/α, 3/16≤β/α≤1/4.
91
12dB, respectively. Also, the bandwidth of the feedback cascode current source of this
This bandwidth is larger than 20x the loop bandwidth of the PLL. For DC supply noise,
the PSRR is improved by more than 15dB. Equivalently, the supply sensitivity of VCO
frequency is improved from 0.2%-fVCO/1%-VDD (for the V-I converter without the noise-
circuit).
At very high frequencies, Mn1 enters triode region, which increases ∆IDrv beyond
the available ∆Icomp. Therefore, the supply sensitivity of the VCO degrades at high control
voltages similar to regulated VCOs and differential VCOs. At very low control voltages,
the supply sensitivity also degrades due to greater susceptibility of the CCO to noise.
While the VCO has an operating range of 200-2300MHz, the simulation results indicate
that the VCO achieves the supply noise rejection of ≤ 0.035%-fVCO/%-VDD over a smaller
pump injects the charge into or out of it. To stabilize the system, as discussed in Chapter 2,
a zero should be introduced by adding a resistor, R, in series with the loop filter capacitor.
Figure 5.8 shows the conventional loop filter, implemented with constant and linear RC.
To guarantee the loop stability under varying process or operating frequency, PLLs with
the conventional loop filter achieve a constant and relatively low bandwidth. The low
92
bandwidth results in a poor tracking jitter performance due to VCO noise as discussed in
Chapter 3.
ICP
To maximize the loop bandwidth over the operating frequency range requires that
the loop gain tracks the operating frequency. In order to maintain the loop stability, the
zero should also track the operating frequency such that the loop bandwidth scales with
the operating frequency in a constant phase margin. For a second-order PLL, damping
factor, ζ, and natural frequency, ωn, are calculated from Equation 2.4 and Equation 2.8:
ζ = 0.5 ⋅ R ⋅ K loop ⁄ N
2⋅ζ [4.1]
ω n = ------------------
R ⋅ C CP
Τhe designs proposed in [52], [55] and [57] employ self-biased techniques to
achieve an adaptive bandwidth PLL with a constant phase margin. These designs
93
implement the resistor through active components. Figure 5.9 shows an adaptive loop
filter [52] that uses two charge-pump currents to implement the resistor:
I CP – proportional 1
R = -------------------------------------- ⋅ ---------------
I CP – integral gm Reg
CPintegral 1/gmReg
Regulator
From PFD +
CCP Vctrl
-
CPproportional
Figure 5.9: Implementing the PLL stabilizing zero with two charge-pump currents
and a regulator
circuit, 2) loop stabilizing zero and 3) a third pole. The design is similar to [52], [57], and
[55] in that the loop characteristics track the VCO operating frequency such that the loop
ICP
Rout
Vctrl
From PFD V-I To CCO
ICP
R C
CCP Vint
94
The charge pump uses a similar structure as [52] where it is self-biased with the
VCO control voltage (Figure 5.11). Therefore, the charge-pump current scales with the
Up
(from PFD)
Vctrl
(to VCO) Vint
2W0
W0
Dn 2W0
(from PFD)
4W0
d2 d1 d0
Controller
PLL operating frequency. The series of a resistor and a capacitor forms the loop stabilizing
zero. The design implements the resistor and capacitor with a MOS channel resistance
[62] and a MOS capacitor, respectively, as shown in Figure 5.12. The MOS resistor is
Vctrl Controller
cn-1
c1
c0
Vctrl
R
R0/4 R0/2 2n-3.R0
CCP
R0 2.R0 2n-1.R0
95
biased by the VCO control voltage so that the loop zero scales with the PLL’s operating
frequency. The proposed circuit achieves the scalable zero with a modest improvement in
power and area upon the previous designs ([52] and [57]) that use an additional charge-
D/A converters in Figure 5.11 and Figure 5.12 adjust the charge-pump current and
MOS resistor to allow further loop-parameter adjustments to optimize jitter at the output
clock as discussed in Chapter 3 and Chapter 4. The area overhead due to a 3-bit controller
for the charge-pump current and a 4-bit controller for the loop filter resistor is negligible in
comparison with the overall charge-pump area and loop filter capacitor. The tunability of
the MOS resistor also provides an additional tuning to adjust the zero position for any
The switching activity of PFD produces ripple on the VCO control voltage at the
same rate as the reference clock frequency. The ripple modulates the VCO frequency
resulting in jitter at the output clock. This effect worsens with higher frequency
multiplication by the loop. The loop’s third pole (formed at the CCO input) filters out the
ripple. The third pole also tracks the PLL operating frequency because the output resistor
(Rout) scales with the oscillator’s frequency. With all primary loop parameters adapting to
the oscillator frequency, the loop operates with a wide frequency range with a constant
phase margin.
96
5.4 Phase-Frequency Detector
A common architecture for clock generation uses a phase-frequency detector
(PFD) for simultaneous phase and frequency acquisition. Generating high frequency clock
increases the difficulty of the design of the PFDs particularly for systems with a high input
5.4.1, the speed of the conventional NAND D-flip-flop phase-frequency detectors (PFDs)
limits the operating frequency and slows the frequency acquisition. This research proposes
CKref
Up=0
Up Dn=1
D Q CKout CKref
DFF Down State
R
Up=0
Reset Dn=0
CKref CKout
R Initial State
DFF Up State
Dn
D Q Up=1
Dn=0
CKout
(a) (b)
Figure 5.13: (a) Linear PFD architecture, (b) PFD state diagram
flops (DFFs) and its state diagram. This PFD generates an Up and a Dn signal that
97
switches the current of a charge pump. The DFFs are triggered by the inputs to the PFD.
Initially, both outputs are low. When one of the PFD inputs rises, the corresponding output
becomes HIGH. The state of FSM moves from an initial state to an Up or Down state. The
state is held until the second input goes high which in turn resets the circuit and returns the
The PFD’s characteristic is ideally linear for the entire range of input phase
differences from -2π to 2π (Figure 5.14-(a)). When the inputs differ in frequency, the
( T CK – T CK ) 1
phase difference changes each cycle by - .
2π ⋅ -----------------------------------------------
ref
max ( T CK , T CK )
out
On every clock cycle during
out ref
frequency acquisition, the phase difference steps across the PFD transfer curve from 0 to
+/-2π and repeats as the output clock cycle slips. The control voltage of voltage-controlled
oscillator (VCO) is pumped monotonically toward that of the desired frequency. As the
frequency error decreases, the sweep slows until the frequency difference is within the
lock-in range. Note that because phase roughly sweeps linearly and that the voltage is
integrated, the voltage accumulates quadratically between each slip of the clock cycle.
Once within the lock-in range, the cycle slipping stops and the phase is acquired, behaving
as a linear system.
However due to the delay of the reset path, the linear range is less than 4π (Figure
5.14-(b)). Figure 5.14-(c) illustrates the non-ideal behavior with the reference clock
(CKref) leading the output clock (CKout) causing an Up output. As the input phase
difference nears 2π, the next leading edge (CKref) arrives before the DFFs are reset due to
98
the finite reset delay. The reset overrides the new CKref edge and does not activate the Up
signal. The subsequent CKout edge causes a Dn signal. The effect appears as a negative
output for phase differences higher than 2π - ∆ where ∆ = 2π ⋅ t reset ⁄ T cyc which depends on
the reset path delay (treset) and the reference clock period (Tcyc). Note that treset is
determined by the delay of logic gates in the reset path and is not a function of input
frequency.
Vout Vout
-2π -2π
∆φ ∆φ
2π 2π
∆
(a) (b)
CKref
CKout
Up
Dn
Reset
(c)
Figure 5.14: (a) Ideal PFD characteristic. (b) Nonideal linear PFD characteristic. (c)
PFD nonideal behavior due to nonzero reset delay
During acquisition, the frequency will not monotonically approach lock-in range
because the non-ideal PFD gives the wrong information periodically. The acquisition
slows by how often the wrong information occurs which depends on ∆. At an input
99
frequency ( TCKref = 2 ⋅ t reset ) where ∆ equals π, the PFD outputs the wrong information half
the time and thereby fails to acquire frequency lock unconditionally. The maximum
1
operating frequency can be expressed as f ref ≤ ------------------- .
2 ⋅ t reset
A commonly used PFD design is one used in [72] using NAND-based latches to
build the D-flip-flops. The reset path includes one 2-input NAND, one 4-input NAND and
two 3-input NANDs. We characterize the reset delay by normalizing it with the delay of a
measures a delay of 5.3 FO-4 thereby limiting the maximum clock period to 10.6 FO-4.
The next two sections describe two proposed designs that significantly improve
minimize clock skew. As both outputs become HIGH, the slave is reset asynchronously
while the master is reset synchronously i.e., the reset is allowed only when the slave latch
is transparent. Synchronously resetting the master increases the operating range and also
reduces the power consumption. If the master latch is reset while it is transparent, then
there will be significant short-circuit current, resulting in more power. The synchronized
reset transistors (N1 and N4), must be at the bottom of the stack because “RST” is the late
arriving signal when the nodes “out” and “ref” are reset. The reset circuit shown in Figure
3 includes one pass transistor, one inverter and one NAND gate. In order to properly reset
100
the slave, the pass-transistor output should become HIGH before the master becomes
transparent. Hence, the NAND gate delay is counted twice in the delay path. The smaller
gates in the reset path as compared to NAND FF PFD reduces treset to 4.4 FO-4 and Tref by
Reset path
CKout P2
out
Up
P1 N3
N2
N1
RST
CKref P4
ref
Dn
P3 N6
N5
N4
which fundamentally changes the dependence on the reset delay. This is illustrated in
Figure 5.16-(a) with the same case as before. When CKref arrives during the reset, the
edge information propagates to the output as long as CKref pulse (Pulseref) is still HIGH
(level-sensitive) when the reset period ends. The PFD no longer loses the edge that arrives
101
during reset and does not output the wrong direction. However, since the PFD output
becomes active HIGH at the end of the reset (∆), the output pulse width would be constant
(2π-∆) for phase differences greater than 2π−∆. The characteristic is shown in Figure
5.16-(b). The input clock pulse widths (Win) should be designed to be slightly smaller than
CKref
∆φ ≥ 2π−δ
CKout
Up
Dn
Reset
treset
Condition for negative output voltage
(a)
Vout ∆
−2π −π π 2π
∆φ
δ
(b)
Figure 5.16: (a) Behavior of a latch-based PFD, including the description of the
nonideal behavior origin. (b) characteristic of a latch-based PFD
treset, otherwise the PFD would fail to lock at zero input phase difference. The PFD failure
is due to the fact that the input clock pulse that triggers the reset would activate the output
after the reset pulse ends for W in ≥ t reset . This design criteria results in a negative output
102
voltage for ∆φ ≥ 2π – δ as illustrated in Figure 5.16-(a) and (b). Note that this PFD has
faster acquisition rate compared to the first type (with the same operating frequency)
because it outputs less incorrect phase information. However, the PFD has a gain that
Figure 5.17 illustrates design of the latch-based PFD [74], using glitch latches. The
Reset path
P2 P1
Dn Up
N4 N1
Pulseref Pulseout
CKref RST CKout
N5 N2
Generates Pulsed Clock
D N6 N3 D
delay elements control the pulse width of the clocks. As shown in Figure 5.17, the reset
circuit includes two inverters and one NAND. The reset also traverses the circuit twice
because the reset should return HIGH. Therefore, treset delay is roughly 5.5 FO-4 and
contains three inverters and two NANDs. As the clock period is less than twice the pulse
width, the clock pulses from N2 (N5) and N3 (N6) are no longer constant width but reduce
with the period. Therefore δ is no longer constant and grows with increasing frequency.
1
The PFD fails as frequency approaches ------------
t reset
which is potentially twice that of the
previously proposed PFD for the same treset. Consequently the maximum frequency is
103
higher than the DFF-based designs despite longer treset. The higher performance is at a
cost of 3x the power as compared to the first proposed circuit due to DC current and extra
power consumption in the delay circuit. When the reset node and clock inputs are
simultaneously LOW and HIGH respectively, the DC current flows through N1, N2, N3
It should be noted that the first PFD design (Figure 5.15) can also be converted to
latch-based type PFD by adding a delay cell to the gate inputs of P1 and P3 transistors.
The delay allows P1 and N3 (P3 and N6) to both conduct briefly, behaving like a glitch
latch. This new design has the similar functionality as PFD in Figure 5.17 in terms of
proposed designs (Figure 5.15 and Figure 5.17) for reference clock of 435 MHz
PFDs, starting the VCO at 375 MHz and locking at 800 MHz. As expected, the PLL with
latch-based PFD has the fastest frequency acquisition among the three PFDs.
104
Vout
∆φ
-2 -1.6 -1.2 -0.8 -0.4 0.4 0.8 1.2 1.6 2 (× π)
1.28
1.24
VCO control voltage (v)
1.08
1.04
105
5.5 Measurement Results
The PLL and clock buffer1 have been designed and fabricated in a 0.25-µm CMOS
technology. As shown in the chip micrograph, Figure 5.20, the PLL core area is 0.028mm2
(120µm x 230µm).
120 µm
PLL
230 µm
Clock Buffer
The measured VCO operating frequency is 130-1600 MHz. Figure 5.21 depicts
the measured VCO gain indicating that the gain varies only between 0.9-1.35 GHz/V for
106
9
120 µm
x 10
2.5
0.5
0
0.5 1 1.5 2 2.5
VCO control voltage (V)
The input reference frequency generated by a signal generator is set to 250 MHz
and the loop multiplication factor is four. The long-term jitter performance of the PLL
output at 1 GHz is demonstrated in Figure 5.22. The jitter histogram measures the rms
RMS = 3.28 ps
P2P = 28.89 ps
107
jitter at 3.28 ps and P2P jitter at 28.89 ps (> 45 Khits) without the supply noise. The
measured power consumption is 10mW at 2.5-V supply and 1-GHz output clock
frequency.
To characterize the sensitivity of the VCO frequency to supply noise, both static
and dynamic VCO supply sensitivity measurements are performed. For static
measurement, the DC value of the supply is varied by ±10% and the frequency variation
results expressed in %-fVCO/%-VDD. The measurement results indicate that the VCO
achieves ≤ 0.03%-fVCO/1%-VDD at low frequency supply noise for 0.8 ≤ Vctrl ≤ 1.7 (in
terms of frequency, 300 MHz ≤ fVCO ≤ 1.4 GHz). At Vctrl greater than 1.7V, where the
noise-canceling circuit becomes less effective, the noise sensitivity increases to 0.25%-
overall jitter performance of the PLL to high frequency noise. A ±10% supply step with 1-
ns slew rate (the fastest possible on-chip frequency) is injected to the VCO supply and the
P2P jitter at PLL output clock is measured. Figure 5.23 demonstrates the measured long-
term P2P jitter expressed in terms of the percentage of the PLL output clock period, %-
TPLL. The measurement results indicate that the PLL achieves the jitter performance of ≤
0.1%−ΤPLL/1%−VDD step, with the VCO frequency varying from 800 ΜΗz to 1.4 GHz.
108
8
x 10 0.25 Static (VCO)
16
0.2 Dynamic (PLL)
%-fVCO 0.03
14
%-VDD 0.09
0.03
0
0.6 0.8 1 1.2 1.4 1.6 1.8 2
Vctrl (V)
Figure 5.23: Measured sensitivity of VCO output clock frequency to static and
dynamic supply noise
To verify the performance of the proposed PFDs, the three PFDs and PLL
proposed in [52] are fabricated in a 0.25-µm CMOS technology. The die photogragh is
shown in Figure 5.24. The first and second circuits show 18.5% and 41.7% improvements
109
Loop Filter
Integral CP OPamp
Proportional CP VCO
Latch-based PFD : N
PFD
Loop Filter
Integral CP OPamp
Proportional CP VCO
NAND DFF PFD : N
PFD
Loop Filter
Integral CP OPamp
Proportional CP VCO
Pass Transistor PFD : N
DFF PFD
The measured frequency acquisition time of PLLs are depicted in Figure 5.25 for
all three PFDs. To analyze the frequency acquisition, a reference clock of 1 GHz is
supplied while the VCO frequency is initially reset to 200 MHz. Sampling circuits
monitor the VCO control voltage as the PLL’s reset is disabled. The loop acquires lock
with a slightly underdamped behavior. The latch-based PFD has a 1.7x faster acquisition
110
rate than the NAND DFF PFD and is 1.4x faster than the pass transistor DFF PFD. Note
that the PFD with fast acquisition has larger lock-in range.
PLL reset
Table 5.1 summarizes the measured and simulated power consumption and speed
performance of each PFD. The power consumption is calculated for PFDs in the lock
mode for the reference clock of 500 MHz. The pass-transistor DFF PFD consumes the
111
Table 5.1: PFDs performance summary
with prior state-of-the-art designs. The first two designs by Sidiropoulos [52] and Ingino
[51] are examples of the regulated VCOs whereas the designs by Kaenel [54] and Ahn
[55] are examples of V-I converters with cascode current sources. The design by Maneatis
[57] is an example of the differential VCO and finally, the design by Minami [60] is an
example of a V-I converter with linear gain for the entire range of the VCO control
voltage. For a fair comparison, all designs are normalized to 0.25-µm technology with
2.5V supply by the use of scaling equations for the short-channel and long-channel
devices. The proposed design achieves the lowest power and area among all designs while
achieving comparable noise performance with the regulated VCO proposed in [52]. The
0.007%/1% by coupling the VDD to ground with a large capacitor of 1.2nF and with
higher power consumption. While the area and noise rejection performance of the
proposed PLL is comparable with [52], it consumes 43% less power than the design in
112
[52]. With a comparable power consumption, the proposed PLL achieves better dynamic
a. f, λ and V indicates the normalization with frequency, technology and voltage, respectively.
113
Table 5.3: PLL performance comparison (2)
5.7 Summary
To produce low-jitter clocks in noisy supply environments, we demonstrated an
effective supply rejection technique for the VCO. The proposed VCO achieves high
supply noise rejection comparable to that of a regulated supply VCO with lower power
consumption. The VCO operates over a wide operating frequency range and has a linear
114
voltage-to-frequency gain. The PLL design demonstrates scaling loop parameters with the
oscillator’s frequency that tracks over a 10x frequency range. The self-biased design
allows the PLL to operate over a wide frequency range with an adaptive loop bandwidth
and a constant phase margin. The proposed PFDs acquire frequency lock faster and
The generated on-chip clock by the PLL should be distributed through clock
buffers to the entire system with small uncertainty. Next chapter discusses the design of a
115
Chapter 6
One of the challenges in digital systems is the distribution of the generated on-chip
clock with a small uncertainty. Static CMOS inverters are traditionally used for clock
buffering due to their simplicity and drive capability with low power consumption.
1%-delay/1%-VDD. With long chains, this poor supply noise rejection of the inverter
could result in significant jitter. For example, in IBM S/390 microprocessor [76], the
generated on-chip clock is distributed through clock buffers to all latches in three levels of
H-like tree hierarchy. The total simulated delay is about 750ps. The amount of jitter with
10% supply noise is roughly 75ps which is 3% of clock period of 400MHz. As technology
scales down1 and clock frequency increases, jitter becomes a larger fraction of the clock
period and may cause system to fail. To reduce jitter, the delay sensitivity of the clock
1. The delay of interconnect is not scaling down as fast as technology does [77].
116
buffers should be improved. One method of improving the supply noise rejection is by
filtering, using a regulator or RC filter. These techniques require large capacitors and
hence area. This chapter introduces a compensator circuit added to the inverter that offsets
any supply-induced delay variation. This circuit technique supplements other methods of
equal delay sensitivity to the supply noise as an inverter (Figure 6.1-(a)). This noise
inverter such that as VDD drops, the capacitor value decreases to compensate for inverter
delay increase and vice versa. A simple circuit capable of the delay compensation is a
MOS resistor in series with a capacitor. Figure 6.1-(b) shows the clock buffer with the
compensator circuit, where the capacitor and resistor are implemented by PMOS
transistors1. The gate voltage of the PMOS resistor, Vgap, is set to a constant voltage with
respect to ground. As VDD drops, the source-gate voltage (VSG) of the PMOS resistor is
capacitive loading, which compensates for the increase in the inverter’s resistance.
1. If the PMOS resistor and capacitor are switched in the compensator circuit, the circuit is similar to
delay elements commonly used for variable-delay lines [63]. However, this configuration is undesirable for
the noise compensation because the VSG is decoupled from VDD by PMOS capacitor whereas the VSG in
Figure 6.1-(b) experiences the VDD noise directly.
117
Compensator circuit
VDD
+
VSG
Vgap - R
KT
Inv mp. C
. Co Cfilter
Delay Comp. Inv. C
VDD
CKin CKout
0.9 1 1.1 ×VDD Inv
Compensated inverter
(a) (b)
Figure 6.1 (a) Ideal compensation of supply-induced inverter delay variation, (b)
proposed compensator inverter
excellent dynamic behavior due to a very small time constant of the compensator circuit.
The compensated inverter can have a high power-supply rejection ratio (PSRR) for both
low and high supply noise frequencies. A regulator in comparison would have a larger
time constant with a large filter capacitor. Also, for most applications, the supply noise
does not exceed ±15% of supply voltage [49]. Thus, the extra capacitive loading
introduced by the compensator circuit would be within 10-15% of the inverter’s load and
does not change the fanout of the inverter significantly. Due to the small loading effect, the
delay and power overhead added by the compensator circuit are a small fraction of the
118
6.2 Design Implications
To achieve the high PSRR in the compensated inverter, the compensator circuit
should provide an inverse and equal delay variation (from VDD noise) as the delay
proportional to VDD, the delay variation of the compensator circuit varies non-linearly
with VDD. Figure 6.2-(a) shows the non-linear behavior of the delay variation of the
compensator circuit as a function of VSG. For minimum power, delay and area overhead,
the compensator circuit should be used over the range where it achieves the maximum
∂ V SG
delay sensitivity; Sensitivity = delay ⋅ -------------- . The maximum delay sensitivity
∂ V SG delay
occurs over the VSG range where the resistance of the PMOS device, R, is the most
sensitive to the VSG variation. V SG in Figure 6.2-(a) indicates the middle of the region
max
circuit approximates the delay sensitivity of the inverter. For VDD noise exceeding this
range, the delay compensation performance of the compensator circuit degrades. Figure
6.2-(b) shows the desired delay sensitivity of the compensator circuit (normalized to the
compensation. The voltage range where the normalized sensitivity curve crosses one is
when the compensator circuit compensates for the delay variation of the uncompensated
119
inverter due to VDD noise. The delay sensitivity and the compensating range of the
compensator circuit are adjusted through sizing the devices in the compensator circuit.
−10
x 10
3.5
1
Sensitivity
0.5
Comp. (b)
range
0
0 0.5 1 1.5 2 2.5 VSG (V)
Figure 6.2 (a) Delay variation of compensated inverter due to VSG variation, (b)
delay sensitivity of compensator circuit, normalized to delay sensitivity of an inverter
circuit (normalized to the delay sensitivity of the uncompensated inverter) as the PMOS
resistor and capacitor vary. Figure 6.3-(a) shows the delay sensitivity behavior as a
function of the capacitor while keeping the width of PMOS resistor constant. Using a
larger compensating capacitor as a fraction of the total capacitive load of the inverter
increases the delay variation and, hence, the delay sensitivity of the compensator circuit.
However, increasing the capacitor reduces the compensating range. Figure 6.3-(b) shows
similar curves, varying the PMOS resistor value in a constant capacitor value. Decreasing
120
the resistor value increases the delay sensitivity by introducing larger capacitive loading to
the inverter while reducing the compensating range. By proper adjustment of the resistor
and capacitor, both the maximum normalized delay sensitivity and the compensating
range can be set to one and peak-to-peak supply noise, respectively. Curve (2) in Figure
6.3-(a) or (b) is an example of the proper sizing that roughly results in the same delay
Increasing C
(1)
1
Normalized Delay Sensitivity
(2)
C(1)>C(2)>C(3) (a)
(3)
of Compensator Circuit
0.5
Constant WR
0 VSG (V)
0 0.5 1 1.5 2 2.5
Decreasing R
1
(1) R(1)<R(2)<R(3)
(2)
0.5 (3) (b)
0
Constant C
0 0.5 1 1.5 2 2.5 VSG (V)
compensated inverter (with R and C values of curve (2) in Figure 6.3) when VDD varies by
±10%. Curve(1) illustrates the supply-induced delay variation of the compensated inverter
while keeping VDD of the compensator circuit constant. This curve represents the delay
variation while keeping VDD of the inverter constant. This curve represents the delay
variation solely due to the compensator circuit. Curve(3) shows the overall delay variation
121
of the compensated inverter to VDD noise. Curve(3) is effectively an average of the first
two curves. The overall delay sensitivity of the compensated inverter is approximately
−10
x 10
4.4
3.8 (3)
3.6
3.4
3.2
2.3 2.4 2.5 2.6 2.7
VDD (V)
Figure 6.4 Supply-induced delay variation of: (1) uncompensated inverter, (2)
compensated inverter with inverter’s VDD held constant and (3) compensated
Although the delay sensitivity metric has been traditionally used to illustrate the
circuit noise performance, the overall delay variation of curve(3) in Figure 6.4 suggests
another useful metric. Since the delay may not change linearly with VDD variation, the
alternate metric is defined as the maximum percentage delay variation from its nominal
∆delay max ∆V DD
value ( ------------------------------- ⋅ 100 ) within the VDD noise range ( -------------- ⋅ 100 ). The maximum
delay nominal V DD
delay variation for curve(3) is 1.2% within ±10% VDD noise.
122
6.2.2 Bias Circuit for Vgap
The previous discussion of the delay compensation indicates that the bias circuit
for Vgap must be constant with respect to ground. Also, the optimum biasing point for the
Vgap (the middle of the voltage range with the maximum delay sensitivity) varies across
process corners, as PMOS devices become faster or slower. To maintain the high PSRR
across the corners, the biasing circuit for Vgap should track the variation of the PMOS
threshold such that Vgap is set to the middle of the compensating range. Therefore, the
desired Vgap should compose of a voltage that is independent of supply and PVT (a
bandgap reference) and a voltage that depends on the PMOS threshold voltage. Figure 6.5
shows a realization of the bias circuit. A diode connected PMOS transistor is biased with a
small current such that VSG ~ |VTp|. To generate Vgap, the |VTp| is subtracted from an
Bandgap Vbg
Circuit + (1+R2/R1).Vbg
- +
R1 R ~VTp
-
R2 Vgap = (1+R2/R1).Vbg - VTp
W1 W2
W1>>W2
123
The generated Vgap bias is distributed to the entire clock buffers. Due to the
coupling noise into Vgap, there is uncertainty in the Vgap voltage from buffer-to-buffer.
The deviation of Vgap from the middle of the compensating range decreases the effective
compensating range. Figure 6.6 shows the simulated delay variation of the compensated
inverter for the optimum Vgap, and ±100mV deviation from the optimum Vgap. The
maximum delay variation increases from 1.2% (within ±10% VDD noise) at the optimum
Vgap to 2% and 2.7% at 50mV and 100mV of the offset in the Vgap value, respectively.
The uncertainty in the Vgap can be reduced by minimizing the coupling capacitors with a
careful layout design. Also, the Vgap uncertainty can be significantly suppressed by
supplying the clock buffers with their own local Vgap bias generator with the cost of power
400
Vgap = 1.3V
1.35V
380 Vgap = 1.4V (optimum)
1.45V
Delay (ps)
360
1.5V
TT,
Un
340 c om
pen
sat
e d in
320 v er
te r
300
2.3 2.4 2.5 2.6 2.7
VDD (V)
124
6.2.3 Performance Sensitivity to PVT1
To characterize the performance of the delay compensating technique, the
compensated clock buffer is simulated over temperature and process variations. As the
temperature increases, the Vgap increases due to the negative sensitivity of VTp to the
the compensated clock buffer as the temperature varies between 0o to 125o. Increasing the
temperature from 25o to 125o increases the maximum delay variation from 1.2% to 2.4%
−10
x 10
4.4
TT, Temp = 125oC (Vgap = 1.5V)
Delay of Clock Buffer (s)
4.2
4
TT, Temp = 25oC (Vgap = 1.4V)
3.8
TT, U
3.4 nc o mp e
n sa t e
d inv
3.2 e rt e r
at 25 o
3
C
2.3 2.4 2.5 2.6 2.7
VDD (V)
Figure 6.7 Delay variation of compensated clock buffer over temperature as VDD
varies ≤ ±10%
Figure 6.8 shows the supply-induced delay variation across the process corners where
Vgap tracks the PMOS threshold variation. The maximum delay variation increases to
125
2.5% (within ±10% VDD noise) at fast NMOS corners in the worst case. The PSRR
degradation at fast NMOS corners is due to the fact that neither the compensated circuit
nor the Vgap voltage tracks the NMOS corner variation. To further improve the PSRR, a
series of an NMOS capacitor and resistor can be added to the compensator circuit.
4.5
Figure 6.8 Delay variation of compensated clock buffer across the corners as VDD
varies ≤ ±10%
Five stages of fanout-4 compensated inverters as shown in Figure 6.9 are used for
simulation. The optimum sizes of the PMOS resistor and capacitor are 0.5x and 3x the
PMOS transistor width size in the preceding inverter stage. The simulated power and
delay increase due to the compensator circuit (Vgap bias circuit is not included) are 25%
126
VDD VDD
Vgap R0 Rn-1
0.5.Wp0 0.5.4n-1.Wp0
Cfilter
C0 3.Wp0 Cn-1 3.4n-1.Wp0
VDD
VDD VDD
Wp0 4.Wp0 4n-1.Wp0
CKin CKout
Wn0=Wp0/2 4.Wn0 4n-1.Wn0
Inv0 Inv1 Invn-1 Cload = 4n.
(Wp0+Wn0)
Figure 6.9 Five stages of fanout of four (FO-4) compensated inverters (n=5)
VDD variations are measured. Five stages of FO-4 inverters and compensated inverters are
fabricated in 0.25-µm CMOS technology1. The compensator inverters includes the PMOS
compensator circuit only. For the measurement purpose, a separate power supply is used
to supply the Vgap instead of the bias generator shown in Figure 6.5. Vgap is held constant
as VDD noise is injected. The measurement results shown in Figure 6.10 indicates that the
3.8% within ±10% VDD noise for a slow corner device, which is 5x less than the
127
clock buffer is greater than the simulation results in a typical corner (1.2% within ±10%
VDD) due to not tracking the NMOS process variation and also the parasitic capacitances.
compensator circuit.
750
Delay of Clock Buffer (ps)
700
600
Unc V
omp ga
p= V
ens DD
550 ated
Inve
rter
500
450
0.9 0.95 1 1.05 1.1 × VDD (= 2.5V)
compensated clock buffer for Vgap values far from the optimum Vgap=1.45V. The
measured result at Vgap=0 shows an increased maximum delay variation to 5.7% (within
±10% VDD noise) and for Vgap=VDD, where the PMOS resistor is off, the maximum delay
variation becomes 22%, which is roughly the same as that of an inverter. The measured
power and delay overhead are 30% and 18%, slightly greater than simulation results due
to the parasitic capacitances. The area overhead, excluding decoupling capacitors, is 50%
128
as compared to inverters alone. The overhead numbers do not include the overhead due to
6.4 Summary
To distribute low-jitter generated on-chip clocks in noisy supply-noise
for the clock buffer. The proposed clock buffer achieves high supply-noise rejection with
an excellent dynamic behavior and with small area and power overhead. This technique
can supplement existing supply filtering using decoupling capacitors and supply-voltage
regulation. The design dissipates low power for its jitter performance and has low area
overhead.
129
Chapter 7
Conclusion
This dissertation has shown the generation and distribution of low-jitter on-chip
clocks for low-power applications in noisy supply environment. The major noise sources
in a PLL were discussed: VCO internal noise, clock buffer noise and input (reference)
clock noise. The performance of circuits to supply noise is characterized with noise
sensitivity metric; %-delay/%-VDD. This is a useful metric that reports the delay variation
of buffer element in percentage per percentage of supply variation rather than absolute
value. Therefore, it can conveniently be used to compare the noise performance of various
To generate and distribute a clock with small uncertainty requires to reduce the
noise sensitivity of the most sensitive blocks in a PLL, i.e., VCO and clock buffer. Priors
state-of-the-art designs regulate and filter the supply-voltage or use differential delay
consume large power to supply delay elements and occupy large area due to decoupling
130
capacitors. To overcome power and area issues associated with the prior designs, this
research proposes two new filtering techniques that effectively improves the noise
sensitivity of VCO and clock buffer with small power and area overhead. Furthermore,
both techniques demonstrate an excellent dynamic behavior with a faster response time
than the time constant of the PLL. The faster response time enables the VCO (or clock
buffer) to correct for errors introduced by high-frequency noise much faster than the loop
While the proposed filtering techniques are proved to reduce the jitter at PLL
of various noise sources with PLL closed-loop feedback system. Investigation to the
impact of PLL loop parameters on output jitter reveals that the loop parameter settings at
which minimum jitter occurs depends on the dominant noise source in a PLL. Therefore,
to achieve the minimum jitter performance based on the loop parameters requires
For most systems, the dominant noise source is not well-known. To minimize jitter
under different noise conditions, a run-time methodology measures the jitter on-chip and
adjusts the PLL loop parameters toward minimum jitter performance based on gradient-
descent algorithm. A dead-zone phase detection circuit suffices as a measuring circuit for
performance. The implication of a large number of hits is that ≥12-bit accumulator and
long measurement intervals are needed. The range of loop parameters must be bounded by
131
the ability of the loop to remain in lock especially if the algorithm operates when the
bandwidth, we used self-biased techniques in the design of the loop filter. Also, this
research addresses the drawback of the conventional PFD and proposes new circuit
techniques to design PFDs. The proposed PFDs consume lower power and achieve wider
The PLL and clock buffer were fabricated in 0.25-µm CMOS technology.
Experimental results indicate that both VCO and clock buffer demonstrate the delay
supply. The total power consumption of PLL is 10mW at 1GHz. Using the run-time
adaptive method of minimizing jitter for the PLL minimizes jitter to within 5ps of the
minimum peak-to-peak jitter as noise conditions are changed. The PLL demonstrates
scaling loop parameters with the oscillator’s frequency that tracks over a 10x frequency
range.
This research points to several areas of potential future work. Although the
performance of the circuits proposed in this thesis has continued to scale down with
technologies less than 90nm. Understanding and overcoming scaling limitations can be an
132
interesting area for future work. Design of PLL circuits might require innovative design
scaling.
implementing a compensated clock buffer. One might benefit form this technique due its
lower overhead and excellent dynamic behavior. Thus, another interesting area of research
is to extend the noise compensation concept to develop new circuit techniques with better
Noise sources in digital systems are not well known and also the noise conditions
However, this algorithm has its own implications such as error in on-chip jitter
measurement, required resolution and range for loop parameters that impacts the
convergence time are interesting subjects. Although a dead-zone phase detector suffices
for clock/data recovery applications as shown in this work, an appropriate on-chip circuit
Finally, supply or substrate noise are the most dominant noise sources in digital
1. From starting point of this research, the design has been fabricated in three process
technologies: 0.35µm [41], 0.25µm and 0.18µm
133
understanding of these noise sources. There are a few studies ([39] and [48]-[50]) that
focus on understanding and modeling of supply or substrate noise. However, their impact
on jitter distribution is not well understood, because supply or substrate noise are
deterministic rather than probabilistic and they vary from a system to another system.
134
Appendices
2 1 2
σ ∆T = -----2- ⋅ E { [ φ ( t + ∆T ) – φ ( t ) ] }
ω0
‹A.1›
1 2 2
= -----2- ⋅ { E [ φ ( t ) ] + E [ φ ( t + ∆T ) ] – 2 ⋅ E [ φ ( t ) ⋅ φ ( t + ∆T ) ] }
ω0
2 2
σ ∆T = -----2- [ R φ ( 0 ) – R φ ( ∆T ) ] ‹A.2›
ω0
Replacing autocorrelation with power spectral density (given by Khinchin theorem [82]),
∞
j2πft
Rφ ( t ) = ∫–∞ Sφ ( f )e df in Equation A.2, results in:
135
2 4 ∞ 2
σ ∆T = --------2- ∫ S φ ( f ) sin ( πf∆T ) df ‹A.3›
ω 0 –∞
Equation A.3 describes the relationship between the timing jitter and noise power spectral
density (psd), Sφ(f). As ∆T goes to infinity, timing jitter is calculated from Equation A.2:
2 2
σ ∆T = -----2- R φ ( 0 ) ‹A.4›
ω0
or,
2 2 ∞
σ ∆T → ∞ = --------2- ∫ S φ ( f ) df ‹A.5›
ω 0 –∞
the open-loop phase noise of each noise source, Sφni-open(f), with the square magnitude of
of PLL NTF from the correspondent phase noise to the PLL output phase1, Hni(j2πf):
2 4 ∞ 2 2
σ ∆T = -----2- ∫ S φni – open ( f ) Hn i ( j2πf ) sin ( πf∆T ) df ‹A.6›
ω0 –∞
1 ∞ 2
∞
2
To simplify the equation, Parseval’s relation is used, ------ ∫ Z ( ω ) dω = ∫– ∞ z ( t ) dt . To
2π –∞
do so, Z(ω) is expressed as:
∆T
Z ( ω ) = X ( ω ) ⋅ Y ( ω ) = H open ( jω ) ⋅ Hn i ( jω ) ⋅ jω ⋅ sin ω ------- ⁄ ω ‹A.7›
2
X(ω) Y(ω)
136
2
where S φni – open ( f ) = H open ( jω ) . z(t) is equal to convolution of x(t) and y(t). Since
1 ∆T 1 ∆T
y ( t ) = --- δ t + ------- – --- δ t – ------- where δ(t) represents dirac’s delta function,
2 2 2 2
1 ∆T 1 ∆T
z ( t ) = --- x t + ------- – --- x t – ------- where x(t) is the inverse fourier of (Hopen(jω).Hni(jω)).
2 2 2 2
4 ⋅ 2π ∞ 1 ∆T ∆T 2
2
- ∫ --- x t + ------- – x t – ------- dt
σ ∆T = ------------- ‹A.8›
2 2
ω 0 –∞ 4
2
X ( ω ) = H open ( jω ) ⋅ Hn buf ( jω )
N VCO 2
s ‹A.9›
= ----------------- ⋅ --------------------------------------------------------
-
s 2
s + K Loop RCs + K Loop
s = jω
– ζω n t
e
x ( t ) = ------------------ ⋅ cos ( ω d t + θ ) ⋅ u ( t ) ‹A.10›
2
1–ζ
2 2
where ω d = ω n ⋅ 1 – ζ and cos θ = 1–ζ .
– ζω n ∆T
1 e sin ( ω d ∆T + θ ) cos ( ω d ∆T )
- ⋅ -----------------------------------
------------- + --------------------- - – ---------------------------- ζ<1
4π N VCO 2ζω n 2 ( 1 – ζ 2 )
2
2
ωn ζω n
σ ∆T = -----------------------
2
⋅ ‹A.11›
ω0 1 – a∆T 2αβ
2
α - – b∆T 2αβ β
2
------------- – e ------------ + ----- –e ------------ + ----- ζ≥1
2ζω n a + b a a + b b
2 –a b
where a, b = ζω n −
+ ω n ⋅ ζ – 1 , α = ------------ and β = ------------ .
b–a b–a
137
A.4 Relationship Between Output Jitter and Clock
Buffer Noise
For the clock buffer noise, X(ω) is calculated from Equation 3.4 and Equation 3.5:
X ( ω ) = H open ( jω ) ⋅ Hn buf ( jω )
N buf 2
s ‹A.12›
= -------------------------- ⋅ --------------------------------------------------------
-
s ⁄ ω buf + 1 s + K Loop RCs + K Loop
2
s = jω
2
–ωn 2ζω n –ζω t –ω t
x ( t ) = --------- ⋅ sin ( ω d t ) – ------------------ ⋅ cos ( ω d t + θ ) ⋅ e n + ω buf ⋅ e buf ⋅ u ( t ) ‹A.13›
ωd 1–ζ
2
2 2
where ω Buf = 2πf Buf , ζ<1, ω d = ω n ⋅ 1 – ζ and cos θ = 1–ζ .
1 – 12ζ
2
– ω Buf ∆T
ω Buf + ω n -------------------- – e ( ω Buf – 4ζω n )
2ζ
ζ<1
– ζω ∆T ω sin ( ω ∆T + 3θ – π ) ω cos ( ω ∆T ) 2ω sin ( ω ∆T + 2θ )
– e
n
–
n d
-------------------------------------------------------
- +
n d
----------------------------------
- –
n d
--------------------------------------------------
2 2
2(1 – ζ ) 2 ( 1 – ζ )ζ 1–ζ
2
N Buf
2
σ ∆T = ----------2- ⋅
2
‹A.14›
ω0 2
γ- υ
2 4ω Buf υ 4ω Buf γ – a∆T 2υω
2υγ- υ -
4υγ- -------------------- Buf
ω
Buf a + ---- + -----
- + ----------- + + -------------------- – e -------------------- + ----------- + -----
b a + b a + ω Buf b + ω Buf a + ω Buf a + b a
ζ≥1
– b∆T 2γω Buf 2υγ γ
2 – ω Buf ∆T 2υωBuf 2γω Buf
–e -------------------- + ------------ + ----- – e ω + -------------------- + --------------------
b + ω Buf a + b b Buf a + ω Buf b + ω Buf
2 2
2ζω n a – ω n – 2 ζω n b + ω n
where υ = -------------------------- and γ = -------------------------------- .
b–a b–a
138
X ( ω ) = H open ( jω ) ⋅ Hn in ( jω )
2
2ζω n s + ω n ‹A.15›
= H open ( s ) ⋅ --------------------------------------------------------
2
-
s + K Loop RCs + K Loop
s = jω
In the next two sections, we calculate the timing jitter for two different input noise psd.
N Clk – in
A.5.1 1/f2 noise, i.e. S φn in ( f ) = ------------------
2
-
f
2
Using S φnin ( f ) = H open ( jω ) , X(ω) is calculated:
2
2π ⋅ N Clk – in 2ζω n s + ω n
X ( ω ) = ---------------------------------- ⋅ --------------------------------------------------------
- ‹A.16›
s 2
s + K Loop RCs + K Loop
s = jω
– ζω n t
e
x ( t ) = 1 – ------------------ ⋅ cos ( ω d t + θ ) ⋅ u ( t ) ‹A.17›
1–ζ
2
– ζω n ∆T
1 sin ( ω d ∆T + θ ) cos ( ω d ∆T ) 2 sin ( ω d ∆T )
- + e------------------- ⋅ -----------------------------------
1 + ------------------------ - – --------------------------------
- – ------------------------------ ζ<1
2ζω n ⋅ ∆T ∆T 2 ( 1 – ζ )ω n 2
2 ( 1 – ζ )ζω n
2 ωd
2 2
σ ∆T = κ ⋅ ∆T ⋅ ‹A.18›
– a∆T 2 – b∆T 2
1 e 2α 2αβ α e 2β 2αβ β
1 + ------------------------- + ------------- ------- – ------------ – ------ + ------------- ------ – ------------ – ----- ζ≥1
2ζω n ⋅ ∆T ∆T a a + b a ∆T b a + b b
2
2ζω n s + ω n
X ( ω ) = N Clk – in ⋅ --------------------------------------------------------
2
- ‹A.19›
s + K Loop RCs + K Loop
s = jω
139
x(t) is calculated by taking the inverse fourier of Equation A.16:
– ζω t 2 – ζω t
e n ωn ⋅ e n
x( t) = N Clk – in ⋅ 2ζω n ⋅ -----------------
- ⋅ cos ( ω d t + θ ) + ------------------------- ⋅ sin ( ω d t ) ⋅ u ( t ) ‹A.20›
2 ωd
1–ζ
where ωd and θ are the same as Equation 3.10. The long-term timing jitter is calculated:
2 2 ⋅ N Clk – in ∞ 2
σ ∆T → ∞ = --------------------------
ω0
2 ∫ –∞
x ( t ) dt ‹A.21›
2
2 4ζ + 1
σ ∆T → ∞ = --------2- ⋅ N Clk – in ⋅ ω n ⋅ ------------------
2
‹A.22›
ω0 4ζ
contribution of different loop parameters, our analytical results and measurements have
found that tracking jitter due to VCO noise for a particular design can be easily estimated
by simply using the second-order equations. As shown in the jitter analysis of Section 3.4
and Section 3.5, tracking jitter (σtr) is the integral of the noise shaped by the frequency
response. The critical parameters that determine the jitter are the f-3dB and the peaking in
the NTF.
the equations for the second-order loop because the resulting frequency response can
differ greatly. To still use the equation, for a given frequency response, we find an
140
effective fn and effective ζ that result in the same bandwidth and peaking. Figure 3.8-(a)
and (b) shows the corresponding f-3dB for each value of fn, and the corresponding peaking
for each value of ζ. This method is verified by measuring the tracking jitter for the
different loop bandwidths and frequency-response peaking. Jitter is calculated for the
κ 1 - . Table A.1 compares the measured and
same parameters using σ tr = ------- ⋅ ------------
2 2ζω n
calculated jitter. By changing only one variable, we express the change in the jitter as a
ratio. The ratio can be directly predicted from Figure 3.8-(a) or (b). The small error
between measurement and predicted result is primarily due to the oscilloscope’s inherent
noise.
ratio
f-3dB Peak fn estimated rms measured
ζ (Figure 3.8(a),
(MHz) (%) (MHz) jitter (ps) rms jitter (ps)
(b))
141
A.7 Is Jitter due to Input Clock Noise Convex?
The long-term jitter due to input clock noise is given by Equation 4.4 (or
equivalently Equation A.22). To verify the convexity of jitter, the second derivite of jitter
2 –2
d σ rms N in ( ωz )
---------------- = –-----1- ⋅ ----------- - ⋅ --------------------------------------------------------------- <0
dK loop
2 4 ω0 3---
–1 1 2
( ω z ) ⋅ K loop + ---------------
–1
( ωz )
2 ‹A.23›
1 1
K loop – --------------- ---------------
2 ( ωz )
–2 –3
d σ rms – 1 N in ( ωz )
- = ------ ⋅ ------------ ⋅ --------------------------------------------------------------- + ---------------------------------------------------------------
--------------------
d(ω z) –1 2 4 ω0 3--- 1---
2 2
–1 1 –1 1
( ω z ) ⋅ K loop + ---------------
–1
( ω z ) ⋅ K loop + ---------------
–1
( ωz ) ( ωz )
As seen from Equation A.23, the second derivitive of jitter as a function of Kloop is always
2 2 2
σ tot = σ in + σ VCO
or,
2 N in –1 1 N VCO 1
σ tot = --------2- ⋅ ( ω z ) ⋅ K loop + ----------
–1
+ ------------- ⋅ ---------------------------
2 –1
- ‹A.24›
ω0 ωz f0 ω z ⋅ K loop
142
We take the first derivative of total jitter as a function of Kloop:
–1
N in ⋅ ω z N VCO
----------------------- – ----------------------------------------- -
2 –1 2 2
dσ tot 1 ω 0 ω z ⋅ K loop ⋅ f 0
---------------- = --- ⋅ -----------------------------------------------------------------------
- ‹A.25›
dK loop 2 σ tot
2π N VCO
K loop = ----------
–1
⋅ ------------- ‹A.26›
σ tot = min ωz N in
N in 1 - N VCO
- ⋅ K loop – ---------------
-------- ⋅ – ----------------------------------------------
2 –1 2 –1 2 2
dσ tot ω0 ( ωz ) ( ω z ) ⋅ K loop ⋅ f 0
------------------- = ------------------------------------------------------------------------------------------------------------- ‹A.27›
d ( ωz )
–1 σ tot
N VCO ( 2π ) 2
1 + ------------- ⋅ -------------
–1 N in K loop
ωz = -----------------------------------------
- ‹A.28›
K loop
143
Bibliography
April 2001
allel and Distributed Systems, vol. 6, no. 3, pp. 314-328, March 1995
chronous distributed oscillators”, ISSCC Dig. Tech. Papers., pp. 404-405, Feb.
1998
[4] E. Fayneh and E. Knoll, “Clock generation and distribution for Intel Banias mobile
144
[5] V. Gutnik, et al., “Active GHz clock network using distributed PLLs”, IEEE Jour-
nal of Solid-State Circuits, vol. 35, no. 11, pp. 1553-1560, November 2000
[7] K.L. Wong, et al., “Cascaded PLL Design for a 90nm CMOS high performance
[8] K-Y.K. Chang, et al., “A 0.4-4Gb/s CMOS quad transceiver cell using on-chip reg-
ulated dual-loop PLLs,” IEEE Journal of Solid-State Circuits, vol. 38, no. 5, pp.
[9] KL.J. Wong, M. Mansuri, H. Hatamkhani and CK.K. Yang, “A 27-mW 3.6-Gb/s I/
link with per pin skew compensation,” ISSCC Dig. Tech. Papers., pp. 255-256,
Feb. 2000
[11] S. Sauter, et al., “Effect of parameter variation at chip and wafer level on clock
Nov. 2002
145
[12] P. Zarkesh-Ha, et al., “Characterization and modeling of clock skew with process
1999
[13] H.-Y. Hsieh, et al., “Self-calibrating clock distribution with scheduled skews,”
[14] H. Sutoh, et al., “A clock distribution technique with an automatic skew compen-
[15] V. Gutnik, et al., “Embedded power supply for low-power DSP,” IEEE Transac-
tion on Very Large Scale Integration (VLSI) Systems, vol. 5, no. 4, pp. 425-435,
December 1997
[16] D. Duarte, et al., “A clock power model to evaluate impact of architectural and
December 2002
[17] D. Duarte, et al., “Impact of technology scaling in the clock system power,” IEEE
[18] J. Tschanz, et al., “Dynamic-sleep transistor and body bias for active leakage
power control of microprocessors,” ISSCC Dig. Tech. Papers., pp. 102-103, Feb.
2003
146
[19] V. Gutnik and A. Chandrakasan, “An efficient controller for variable supply-volt-
[20] B. Razavi, Design of Analog CMOS Integrated Circuits, McGraw Hill, 2001
[21] B. Razavi, Monolithic Phase-Locked Loops and Clock Recovery Circuits, IEEE
Press, 2003
[22] J.A. McNeill, “Jitter in ring oscillators,” IEEE Journal of Solid-State Circuits, vol.
1994
[25] A. Hajimiri, et al., “A general theory of phase noise in electrical oscillators,” IEEE
Journal of Solid-State Circuits, vol. 33, no. 2, Feb. 1998, pp. 179-194
[26] A. Hajimiri, et al., “Jitter and phase noise in ring oscillators,” IEEE Journal of
[27] R. Moore, “Phase noise and jitter”, Agilent Technologies, May 2001
147
[28] B. Razavi, “A study of phase noise in CMOS oscillators,” IEEE Journal of Solid-
ing theory and numerical methods for characterization,” IEEE Trans. Circuits Syst.
[30] T. C. Weigandt, B. Kim, and P. R. Gray, “Analysis of timing jitter in ring oscilla-
[31] B. Kim, T. C. Weigandt, and P. R. Gray, “PLL/DLL system noise analysis for low
jitter clock synthesizer design,” in Proc. ISCAS, May 1994, vol. 4, pp. 31-34
[33] X. Zhang, et al., “A theoretical and experimental study of the noise behavior of
subharmonically injection locked local oscillators,” IEEE Trans. on MTT, vol. 40,
1994
[35] V.F. Kroupa, “Noise properties of PLL systems,” IEEE Trans. Comm., vol. 30, no.
148
[36] D. Huffman, “Extremely low noise frequency dividers,” Microwave J., pp. 209-
Trans. Ultrasonic, Ferroelectrics and Freq. Cont., vol. 37, no. 4, pp. 295-304, July
1990
trosvyaz, Translated in: Telecomm. and Radio Eng., part 1, vol. 29, no. 2, pp. 52-5,
Feb. 1975
[39] F. Herzel, et al., “A study of oscillator jitter due to supply and substrate noise,”
IEEE Transactions on Circuits and Systems-II: Analog and Digital Signal Process-
one-chip optical receiver IC with 1:8 DEMUX,” IEEE Journal of Solid-State Cir-
[42] K. Lim, et al., “A low-noise phase-locked loop design by loop bandwidth optimi-
zation,” IEEE Journal of Solid-State Circuits, vol. 35, no. 6, pp. 807-815, June
2002
149
[43] A. Mehrotra, “Noise analysis of phase-locked loops,” IEEE Trans. Circuits Syst. I,
[45] D.C. Lee, “Analysis of jitter in phase-locked loops,” Circuits and Systems II: Ana-
log and Digital Signal Processing, IEEE Transactions, vol. 49, issue 11, pp. 704-
Parameters,” IEEE Journal of Solid-State Circuits, vol. 37, no. 11, pp. 1375-1382,
November 2002
[47] K.A. Jenkins and J.P. Eckhardt, “Measuring jitter and phase error in microproces-
sor phase-locked loops,” IEEE Design and Test of Computers, pp. 86-93, 2000
[48] P. Larsson, “Measurements and analysis of PLL jitter caused by digital switching
noise,” IEEE Journal of Solid-State Circuits, vol. 36, no. 7, pp. 1113-1119, July
2001
150
[50] A. Demir, “Phase noise and timing jitter in oscillators with colored-noise sources,”
IEEE Transaction on Circuits and Systems-I, vol. 49, no. 12, Dec. 2002
[51] J.M. Ingino, “A 4GHz 40dB PSRR PLL for an SOC Application,” ISSCC Dig.
[52] S. Sidiropoulos, et al., “Adaptive bandwidth DLL’s and PLL’s using regulated Sup-
ply CMOS buffers,” Proceedings of 2000 IEEE Symposium on VLSI Circuits, Dig.
[53] V. R. von Kaenel, et al., “A High-Speed, Low-Power Clock Generator for a Micro-
processor Application,” IEEE Journal of Solid-State Circuits, vol. 33, no. 11, pp.
[54] V. R. von Kaenel, et al., “A 320 MHz CMOS PLL for Microprocessor Clock Gen-
eration,” IEEE Journal of Solid-State Circuits, vol. 31, no. 11, pp. 1715-1722,
November 1996
[55] H. Ahn, et al., “A Low-Jitter 1.9-V CMOS PLL for UltraSPARC Microprocessor
Applications,” IEEE Journal of Solid-State Circuits, vol. 35, no. 3, pp. 450-454,
March 2000
[56] I.A. Young, et al., “A PLL clock generator with 5 to 110 MHz lock range for
pp. 1599-607
151
[57] J.G. Maneatis, et al., “Low-jitter process-independent DLL and PLL based on self-
biased techniques” IEEE Journal of Solid-State Circuits, Nov. 1996. vol.31, no.11,
p. 1723-32
[58] S-J Lee, et al., “A novel high-speed ring oscillator for multiphase clock generation
using negative skewed delay scheme,” IEEE Journal of Solid-State Circuits, vol.
[59] K.Y.K Chang, et al., “A 0.4-4Gb/s CMOS Quad Transceiver Cell using On-chip
[60] K. Minami, et al., “A 0.10mm CMOS, 1.2V, 2GHz Phase-Locked Loop with Gain
[61] M. Mansuri and C.-K. Yang, “A low-power low-jitter adaptive bandwidth PLL and
clock buffer,” ISSCC Dig. Tech. Papers., pp. 430-431, Feb. 2003
[62] P. Larsson, “A 2-1600-MHz CMOS Clock Recovery PLL with Low-Vdd Capabil-
ity,” IEEE Journal of Solid-State Circuits, vol. 34, no. 12, pp. 1951-1960, Decem-
ber 1999
[63] M.G. Johnson, et al., “A Variable Delay Line PLL for CPU-Coprocessor Synchro-
nization,” IEEE Journal of Solid-State Circuits, vol. 23, no. 5, pp. 1218-1223,
October 1988
152
[64] Y. Moon, et al., “An all-analog multiphase delay-locked loop using a replica delay
line for wide-range operation and low-jitter performance,” IEEE Journal of Solid-
[65] W.J. Dally, et al., “Clock multiplying delay-locked loop for data communication,”
US patent pending
ceedings of 2000 IEEE Symposium on VLSI Circuits, Dig. Tech. Papers, pp. 52-53,
June 2000
[68] B-J Lee et al., “A 2.5-10Gb/s CMOS transceiver with alternating edge sampling
phase detection for loop characteristic stabilization,” ISSCC Dig. Tech. Papers.,
with dead-zone phase detection for robust clock/data recovery,” ISSCC Dig. Tech.
153
[70] S-H Lee et al., “A 5Gb/s 0.25mm CMOS jitter-tolerant variable-interval oversam-
pling clock/data recovery circuit,” ISSCC Dig. Tech. Papers., pp. 256-257, Feb.
2002
[71] R. Kuppuswamy et al., “On-die clock jitter detector for high speed microproces-
sors,” Proceedings of 2001 IEEE Symposium on VLSI Circuits, Dig. Tech. Papers,
[73] H. Partovi et al., “Flow-through latch and edge-triggered flip-flop hybrid ele-
ments,” in IEEE Int. Solid-State Circuits Conf. Dig. Tech. Papers, San Francisco,
[74] H. Partovi et al., “Phase frequency detector having reduced blind spot,” US patent,
GSa/s Phase-Locked Loops,” IEEE J. Solid-State Circuits, vol. 37, no. 10, pp.
154
[77] J.A. Davis, et al., “Interconnect limits on gigascale integration (GSI) in the 21st
century,” Proceedings of IEEE, vol. 89, issue 3, pp. 305-324, March 2001
[80] J. Hein, et al., “Z-domain model for discrete-time PLL’s,” IEEE Transactions on
Circuits and Systems, Nov. 1988
[83] M. Rau, et al., “Clock/data recovery PLL using half-frequency clock,” IEEE
Journal of Solid-State Circuits, Jul. 1997, vol.32, no.7, pp. 1156-9
[84] SS. Sidiropoulos, M.A. Horowitz, “A semidigital dual delay-locked loop,” IEEE
Journal of Solid-State Circuits, Nov. 1997, vol.32, no.11, pp. 1683-92
155