0% found this document useful (0 votes)
96 views

DR Bandwidth

This document is a thesis submitted by Jeffrey Norwood Harrison to Macquarie University in fulfillment of the requirements for a Doctor of Philosophy degree. The thesis examines the dynamic range and bandwidth limitations of analog CMOS circuits through state-space analysis techniques. It includes chapters on calculating dynamic range using state-space equations, designing a 350MHz opamp-RC filter as a case study, and analyzing limitations on sigma-delta modulator performance.
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
96 views

DR Bandwidth

This document is a thesis submitted by Jeffrey Norwood Harrison to Macquarie University in fulfillment of the requirements for a Doctor of Philosophy degree. The thesis examines the dynamic range and bandwidth limitations of analog CMOS circuits through state-space analysis techniques. It includes chapters on calculating dynamic range using state-space equations, designing a 350MHz opamp-RC filter as a case study, and analyzing limitations on sigma-delta modulator performance.
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 205

DYNAMIC RANGE AND BANDWIDTH OF

ANALOG CMOS CIRCUITS

JEFFREY NORWOOD HARRISON


Bachelor of Engineering (Monash University, 1996)

A thesis submitted to Macquarie University in ful lment


of the requirements for the degree of Doctor of Philosophy

January 2002

Department of Electronics
Division of Information and Communication Sciences
Macquarie University, Sydney, Australia
CONTENTS

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

Statement of originality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

Acknowledgement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

Notation and Acronyms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

1 Introduction 17
1.1 Dynamic range and bandwidth . . . . . . . . . . . . . . . . . . . . . . . . . . 17

1.2 Aim . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

1.3 Approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

1.3.1 Analysis techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

1.4 Scope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

1.4.1 Pitfalls of roadmapping . . . . . . . . . . . . . . . . . . . . . . . . . . 22

1.4.2 Target application and frequency range . . . . . . . . . . . . . . . . . 23

1.5 Synopsis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

2 State-space calculation of dynamic range 29


2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

2.1.1 Four questions and the organisation of this chapter . . . . . . . . . . . 30

2.2 State-space equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

2.3 Maximum signal swing and some scaling considerations . . . . . . . . . . . . 34

2
2.3.1 Maximum signal swing . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

2.3.2 Basic points regarding power supply voltage scaling . . . . . . . . . . 34

2.4 Noise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

2.4.1 Thermal noise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

2.5 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

2.5.1 LC ladder lter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

2.5.2 Feedback ampli er using a Miller-compensated opamp . . . . . . . . . 41

2.5.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

2.6 The dynamic range limitation . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

2.6.1 Derivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

2.6.2 Interpretation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

2.6.3 Application to lter dynamic range . . . . . . . . . . . . . . . . . . . . 48

2.6.4 Basic example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

2.7 802.11a receiver example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

2.7.1 Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

2.7.2 Calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

2.7.3 Comments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

2.8 State-space realisation theory . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

2.8.1 Controllability and observability Gramians . . . . . . . . . . . . . . . 53

2.8.2 Similarity transformation . . . . . . . . . . . . . . . . . . . . . . . . . 54

2.8.3 Product of the Gramians . . . . . . . . . . . . . . . . . . . . . . . . . 55

2.9 An optimal realisation based on the dynamic range limitation . . . . . . . . . 57

3
2.9.1 Singular value decomposition . . . . . . . . . . . . . . . . . . . . . . . 58

2.9.2 Numerical realisation algorithm . . . . . . . . . . . . . . . . . . . . . . 61

2.9.3 Properties of the optimal realisation . . . . . . . . . . . . . . . . . . . 62

2.9.4 Example of frequency-weighted input-normal realisation . . . . . . . . 65

2.10 Transfer function optimisation . . . . . . . . . . . . . . . . . . . . . . . . . . 67

2.10.1 Ampli er transfer function . . . . . . . . . . . . . . . . . . . . . . . . 68

2.10.2 What is feedback? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

2.11 Links to other literature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

2.11.1 The Hankel operator . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

2.11.2 De ning a frequency-weighted Hankel operator . . . . . . . . . . . . . 76

2.11.3 Controllability and observability operators . . . . . . . . . . . . . . . . 76

2.11.4 Optimum dynamic range realisation . . . . . . . . . . . . . . . . . . . 77

2.11.5 Sensitivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

2.11.6 Sensitivity theory and other related previous work . . . . . . . . . . . 80

2.11.7 Energy storage, sensitivity and group delay . . . . . . . . . . . . . . . 81

2.12 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

3 A 350MHz opamp-RC lter 83


3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

3.1.1 Organisation of the chapter . . . . . . . . . . . . . . . . . . . . . . . . 84

3.2 Choice of transfer function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84

3.2.1 Transfer function optimisation . . . . . . . . . . . . . . . . . . . . . . 85

4
3.2.2 Graphical examination of dynamic range limitations . . . . . . . . . . 86

3.2.3 Relationship with existing theory . . . . . . . . . . . . . . . . . . . . . 86

3.3 Realisation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88

3.3.1 Unsuitability of the frequency-weighted input-normal realisation . . . 88

3.3.2 Merit of lossless realisations . . . . . . . . . . . . . . . . . . . . . . . . 89

3.3.3 Lossless realisation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

3.3.4 Filter taps - practical details . . . . . . . . . . . . . . . . . . . . . . . 92

3.3.5 Replica oscillator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

3.4 Opamp design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96

3.4.1 Choice of opamps as the active element . . . . . . . . . . . . . . . . . 96

3.4.2 Fundamental limitations on opamp bandwidth . . . . . . . . . . . . . 97

3.4.3 E ect of nite opamp gain-bandwidth . . . . . . . . . . . . . . . . . . 98

3.4.4 Gain stage principles . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101

3.4.5 Compensation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104

3.4.6 Opamp circuit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106

3.5 Measured performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106

3.5.1 Frequency response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106

3.5.2 Dynamic range . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

3.6 Additional simulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112

3.7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116

4 Analytic limitations on sigma-delta modulator performance 118

5
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118

4.2 The sigma-delta modulator . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118

4.2.1 Trade-o between resolution and oversampling ratio . . . . . . . . . . 120

4.2.2 A subjective comment regarding use of \fundamental limitations" . . 121

4.3 Noise transfer function analysis . . . . . . . . . . . . . . . . . . . . . . . . . . 122

4.3.1 Quantisation noise spectrum . . . . . . . . . . . . . . . . . . . . . . . 123

4.3.2 Stability and maximum input signal . . . . . . . . . . . . . . . . . . . 124

4.3.3 Applicability of the NTF analysis to the modulator investigated . . . 127

4.4 Integral limitations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128

4.4.1 Links with Bode's sensitivity integral . . . . . . . . . . . . . . . . . . . 128

4.4.2 Interpretation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130

4.4.3 Derivation of integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . 136

4.4.4 E ect of excess delay . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139

4.4.5 Excess delay with unstable loop lters . . . . . . . . . . . . . . . . . . 141

4.5 Connections with other theory . . . . . . . . . . . . . . . . . . . . . . . . . . 143

4.5.1 Linear predictor interpretation . . . . . . . . . . . . . . . . . . . . . . 144

4.5.2 Links with other system theory . . . . . . . . . . . . . . . . . . . . . . 146

4.6 Design implications of the integral limitations . . . . . . . . . . . . . . . . . . 148

4.6.1 Multiple-loop modulators . . . . . . . . . . . . . . . . . . . . . . . . . 148

4.6.2 Multiple-bit quantisers . . . . . . . . . . . . . . . . . . . . . . . . . . . 150

4.6.3 Compensation for excess loop delay . . . . . . . . . . . . . . . . . . . 151

4.7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154

6
5 A 520MHz sigma-delta ADC with a FIR DAC loop lter 155
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155

5.1.1 Argument for FIR DAC loop lters . . . . . . . . . . . . . . . . . . . . 155

5.1.2 FIR DAC loop lter . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158

5.2 Choice of noise transfer function . . . . . . . . . . . . . . . . . . . . . . . . . 159

5.3 Architecture details . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160

5.3.1 Error ampli er . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160

5.3.2 Latency and rst-tap problems . . . . . . . . . . . . . . . . . . . . . . 163

5.4 Circuit and layout details . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164

5.4.1 Error ampli er . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164

5.4.2 Quantiser . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167

5.4.3 Shift register and DAC . . . . . . . . . . . . . . . . . . . . . . . . . . . 167

5.5 Test circuit board and testing . . . . . . . . . . . . . . . . . . . . . . . . . . . 168

5.5.1 Data and clock inputs . . . . . . . . . . . . . . . . . . . . . . . . . . . 169

5.5.2 Digital output . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170

5.5.3 Power rails . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171

5.5.4 Reference currents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171

5.6 Measured performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172

5.7 Alternative architectures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174

5.7.1 Bandpass option . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176

5.7.2 Digital lter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176

5.7.3 Continuous-time loop lter versus FIR DAC . . . . . . . . . . . . . . . 180

7
5.8 Sigma-delta versus Nyquist-rate converters at 20-200MHz bandwidth . . . . . 181

5.8.1 Performance comparison . . . . . . . . . . . . . . . . . . . . . . . . . . 181

5.8.2 Limitations of sigma-delta ADC's and Nyquist-rate ADC's . . . . . . 182

5.9 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187

6 Conclusion and suggestions for future work 189


6.1 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189

6.2 Suggestions for future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191

6.2.1 Circuit design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191

6.2.2 Circuit theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192

Bibliography 193

8
ABSTRACT
This thesis studies the dynamic range of CMOS analog circuits, with particular emphasis
on the interaction between dynamic range and frequency response. Dynamic range has
become a more signi cant issue for analog circuits recently due to the increased market for
battery-powered products such as mobile telephones. The low power supply voltage of the
deep-submicron CMOS processes used for system-on-a-chip integration also constrains
dynamic range.

Previous authors such as Vittoz have investigated the dynamic range limitations of
simple ampli er stages and the like. However, for more complicated analog circuits such as
active lters and sigma-delta analog to digital converters, this is only part of the problem.
The other part is how the dynamic range of a complicated circuit relates to that of the
sub-blocks from which it is built.

An example familiar to many analog designers is that the dynamic range of a high-Q
bandpass lter is less than that of the integrator sub-blocks from which it is built. Previous
authors have been able to show a 1 / Q dependence of the dynamic range for second-order
systems. A more general relationship between the transfer function H (j!) and dynamic
range is derived here. It is based on a quantity [H (j!1 ) ; H (j!2 )]=j (!2 ; !1 ) which is a
new frequency-domain representation of the so-called Hankel operator of state-space system
realisation theory.

An opamp-RC lter was designed and fabricated in a deep-submicron CMOS process to


assess the usefulness of this theory. The opamp-RC lter achieves state-of-the-art
performance, with a remarkably high cuto frequency, 350 MHz, and a signal swing of
0.5 Vp;p for {40 dB distortion. Further simulations show that it should be practical to push
these speci cations to 800 MHz and 2 Vp;p .

Like lters, sigma-delta modulators have a non-trivial dependence between frequency


response and dynamic range. An integral limitation by Bode is used to investigate the
trade-o between oversampling factor and signal to noise ratio. The analysis has a common
feature with the lter work, namely that the quantisation error is represented as additive
noise in the same way that thermal noise in a lter is. That is, a noise transfer function

9
analysis is used. A new integral limitation is derived which shows the detrimental e ects of
excess delay in the sigma-delta modulator feedback loop. The analysis also shows that large
performance gains are possible by using multiple-bit quantisers, provided that noise transfer
functions with high out-of-band gain are used.

Based on the predictions of the theory, a sigma-delta modulator with a four-bit quantiser
and a high out-of-band noise gain was constructed. It uses a novel topology in which the
loop ltering is implemented in a nite-impulse-response digital to analog converter. This
allows a sample rate of 520 Ms=s, beyond the limits of switched-capacitor technology, to be
used. The fabricated sigma-delta analog to digital converter has a limited signal to noise
and distortion ratio of 40 dB, but achieves it over the very wide bandwidth of 60 MHz. The
potential of wideband sigma-delta modulators to compete with pipelined and other
Nyquist-rate approaches is assessed.

The general conclusion from this work is that achievable dynamic range depends heavily
on the bandwidth of the sub-blocks making up a circuit such as lter or analog to digital
converter, not just their dynamic range.

10
STATEMENT OF ORIGINALITY
I certify that the work contained in this thesis is all my own work except where
acknowledged and that the thesis has not been presented to any other University or
Institution.

Candidate Signature:

11
ACKNOWLEDGEMENT
I express gratitude to my primary supervisor, Prof. Neil Weste. The most important of his
contributions are the opportunity to work for Radiata (now part of Cisco Systems, Inc.),
the provision of design tools and funding for fabrication runs which has allowed me to learn
far more in four years than I would otherwise have, and non-technical skills which I have
picked up by observing him. Neil's basic message is to the e ect that the shortest distance
between two points is a straight line, in both design and teamwork.

I thank Cadence Design Systems, Inc. for supplying Macquarie University with the
design tools used throughout this work, and Cisco Systems, Inc. for fabrication run space
for the lter in chapter 3.

The ideas developed in this thesis have more distant origins. The interest in achieving
high feedback over wide bandwidths essentially comes from knowing the work of Ed Cherry
at Monash University, where I was an undergraduate. The appreciation of Bode's work
comes from Greg Cambrell and Ed Cherry at Monash. The work in chapter 2 has a long
and tangled history, but the little-known work \Finite State Predictors for Gaussian
Sequences" by Justesen (1978) provided the key observations that there is a natural state
space for a given lter transfer function, and that norms of the state space might have
frequency-domain expressions.

Important contributions come from friends and colleagues, notably Marie


Wintrebert-Fouquet at Macquarie, Rodney Chandler and Andrew Adams at Radiata, and
Katrina from Toongabbie. I thank my system administrators Rienzie Jayasekara and
particularly Chris Corcoran, who have greatly increased my productivity by shielding me
from irrelevant computer knowledge. Dave Barkin and the math library sta were
particularly helpful during my time at Stanford.

12
NOTATION AND ACRONYMS

Notation
Where possible, an upper-case quantity is the Laplace transform of the lower-case
time-domain quantity eg. a system has impulse response h(t) and transfer function H (s).
Vector transfer functions are bold, eg. F (s). The list below omits symbols which are used
within a single subsection, as di erent symbols are often given the same name eg. K or M .
A conjugate transpose of A: A = amn + jbmn ! A = anm ; jbnm
AT transpose of A
A complex conjugate of A
jAj absolute value (magnitude) of A
jjA(ej )jj1 1-norm of the discrete-time frequency response, which is the peak gain
jjajj1 P
1-norm in the time-domain, which is jai j
i
1.44/0.24 transistor with gate width 1.44 m and gate length 0.24 m
log logarithm to base e
tr
P
trace (the sum of the diagonal entries, which equals i )
i
HA(s) transfer function H (s) after A is changed to A + A, section 2.11.5
Vp;p volts peak to peak

Notes on matrix algebra


A useful way to think of a matrix expression such as
y = ABC x
is as a block diagram in which the input signal x successively passes through blocks C , B
and A to give the output y. That is, in matrix notation the signal ow is from right to left.

The conjugate transpose A is the natural generalisation of the complex conjugate of a


scalar a. AA and A A are similar to jaj2 , but check whether the result is a scalar, or a
covariance matrix like those in statistical signal processing. The norm
p p
jjAjj2 = tr(AA ) = tr(AA) is analogous to jaj. Recall that (AB ) = B A. Some
books also use A for the adjoint of A, which is de ned by an expression involving the inner
13
product. The conjugate transpose is the adjoint for a matrix, but for transfer functions it
needs to be checked whether the adjoint is being taken at a spot frequency or for the
operator over all frequencies.

Greek and Script Symbols


C Controllability operator
O Observability operator
H Hankel operator
i
p
eigenvalues; i (K  W ) are the Hankel singular values
(s) input signal weight, section 2.8.1
(s) output noise weight, section 2.8.1
! angular frequency (rad/s)

14
Roman symbols
A; B; C; D matrices of state-space realisation
Cgd transistor gain-drain capacitance
Cgs transistor gate-source capacitance
Cds transistor drain-source capacitance
e noise added to states
E expected value
E (z) added quantisation noise in sigma-delta modulator
fmax maximum frequency of oscillation
fs sampling (clock) frequency
fT transition frequency (the frequency where current gain jy21 j falls to unity)
F (s) gain to states (column vector), section 2.2
G(s) noise gain from noise added to states (row vector), section 2.2
gm transconductance, eg. of a transistor. Also gm in gm-C lter
H (s) a generic transfer function
H (z ) a generic discrete-time transfer function,
but also the noise transfer function in chapter 4
K controllability Gramian, section 2.8.1
L(s), L(z) open-loop gain
R(z) sigma-delta modulator signal input
s frequency variable from Laplace transform
T a transformation of the state space
u state-space system input
VDD power supply voltage
Vsat drain to source saturation voltage of a transistor
VT threshold voltage of a transistor
W observability Gramian, section 2.8.1
x state (column vector)
y state-space system output
z discrete-time frequency variable from Z-transform

15
Acronyms
ADC analog to digital converter
BNC British naval connector
CMOS complementary metal-oxide-semiconductor (the term MOS
is still used despite the transistors not having metal gates)
DAC digital to analog converter
DSP digital signal processing
FFT fast Fourier transform
FIR nite impulse response
IF intermediate frequency (in a radio)
IP3 third-order intercept point (Razavi 1998, p. 19)
I/Q in-phase / quadrature signals
LAN local area network
LC inductor and capacitor
nMOS n-channel metal-oxide-semiconductor transistor
NTF noise transfer function
pMOS p-channel metal-oxide-semiconductor transistor
OSR oversampling ratio (Nyquist frequency / useful bandwidth)
PCB printed circuit board
Q quality factor (energy storage / energy loss per cycle)
QAM quadrature amplitude modulation
ROM read-only memory
RMS root mean square
SAW surface acoustic wave (passive lter)
SFDR spurii-free dynamic range (output amplitude / largest distortion spur)
SNR signal to noise ratio
SNDR signal to noise and distortion ratio (SINAD)
SRAM static random-access memory
T/H track and hold

16
1 INTRODUCTION

Dynamic range is the maximum signal a circuit can handle without overload divided by
the noise or other signal processing error. Dynamic range has become a more signi cant
analog design issue with the rise in commercial importance of battery-powered devices such
as mobile telephones. This work investigates ways to optimise the dynamic range of active
lters and sigma-delta analog to digital converters.

This chapter introduces the particular aspects of the dynamic range problem which were
studied, and the methods used to study them. It also provides a short introduction to
relevant existing literature. This is not a literature survey { existing results are introduced
when required in the body of this work.

1.1 DYNAMIC RANGE AND BANDWIDTH


There seem to be two reasons why dynamic range has received attention lately. One is the
desire to increase the battery life of mobile telephones, laptop computers and the like. In
the past it has been possible to achieve satisfactory performance by increasing power supply
current and lowering impedances to reduce the thermal noise voltage. In active lters,
dynamic range could normally be made less of a problem than sensitivity. If battery life and
hence power consumption is an issue this is no longer acceptable; Nokia will not buy your
chip.

The other reason for a focus on dynamic range has to do with integration. The cost of a
product can potentially be reduced by combining the circuitry in fewer integrated circuits.
An ultimate aim is to include almost all circuitry on a single \system-on-a-chip" and to use
as simple a fabrication process as possible. Now the speed of digital circuits can be
increased by reducing the transistor gate length, and the power consumption can be
reduced by decreasing the power supply voltage VDD . Shorter gate length transistors
require lower VDD to avoid gate oxide breakdown or hot electron degradation. Reducing the

17
power supply voltage reduces voltage swing, and hence it would be expected that analog
dynamic range would decrease. This is basically true, although the reduction in dynamic
range is less than expected as noise voltage also falls.

Fig. 1.1 shows typical transistor VDD over a range of process generations. We can plot
VDD against some other quantity; the author has chosen to plot it against a measure of
transistor speed, peak fT . Given that VDD already decreased from 10 { 20 V in early analog
CMOS designs to the 3.3 V popular for analog circuits today, why are we concerned that it
will fall further? It is largely because VDD has fallen to the point where transistor
saturation voltage and threshold voltage are a signi cant fraction of power supply voltage,
requiring changes to circuit techniques. Dynamic range depends on signal swing as a
fraction of supply voltage, and this reduces if a xed proportion of the supply voltage is lost
as a saturation or threshold voltage. The question is then whether the increased speed of
the transistors can in some way compensate. The conventional answer to this is \not
much"; the author is more optimistic.

1.2 AIM
It seems natural to try to \design-in" good dynamic range, and to do so in as systematic a
manner as possible. At the level of individual transistor stages, the limitations on dynamic
range such as maximum signal swing, distortion, thermal noise and icker noise, are
reasonably well known. There are fundamental trade-o s between dynamic range and power
consumption; see for example (Vittoz 1994), (Castello & Gray 1985) and (Annema 1999).
However, what seems to be less clear is how the dynamic range of a system relates to that
of its sub-circuits. This is particularly true for mathematically complicated circuits such as
active lters and sigma-delta modulators. How does a sigma-delta modulator achieve higher
resolution than its quantiser? Why do narrowband bandpass lters always seem to have
poor dynamic range? This work aims to answer these questions.

A contemporary designer will usually answer the rst question by saying that the
sigma-delta modulator trades o bandwidth for resolution { it uses oversampling and noise
shaping. If such a tradeo exists between bandwidth and dynamic range (and we shall see

18
8
1
10 5
Power supply voltage (V)

2 1.2 0.8

0.5 0.35
0.25

0.18

0.13
0
10 0.09
0.07

0.05
0 1 2
10 10 10
Transition frequency (GHz)

Figure 1.1: Power supply voltage VDD versus peak transition frequency fT for past, present and
roadmapped processes. Labels are feature size (m). Symbols indicate source: square { fT from long
channel model in Wong & Salama (1983); circle { Kakumu et. al. (1990) with fT estimated from
inverter delay; diamond { SPICE models from the foundry used by the author; triangle { predictions
in Morifuji et. al. (1999).

19
that it does), it is reasonable to ask how much bandwidth needs to be traded o . Where is
the optimum point on g. 1.1; a deep submicron process with wide bandwidth and lower
signal swings, or an older process? Obviously, the answer is likely to depend on the circuit
involved and its frequency range of operation, so this work does not attempt to give a
de nitive answer. Rather, the aim is to develop theories of dynamic range which include
bandwidth considerations.

1.3 APPROACH
This work has two basic lines of investigation:

1. Design of circuits with high dynamic range. As we want to design systems-on-a-chip in


deep submicron CMOS, the circuit design concentrates on deep submicron processes.

2. Determination of fundamental limitations on the achievable dynamic range. This is in


some sense the complement { designing a circuit shows that a given level of
performance is achievable; nding a fundamental limitation shows that a given level of
performance is not achievable.

(There is a third possibility, namely providing a non-constructive proof that a circuit with a
given performance exists. The author cannot recall any such results in circuit theory.)

The author notes that the second method varies widely in appeal, and urges those
readers for whom it is not a natural way of thinking to at least bear with it and see what
comes from it. More will be said on this in chapter 4.

1.3.1 Analysis techniques


As dynamic range is de ned by maximum signal swing and noise, we need a way to
calculate the maximum internal signal swing, and the e ect of internal noise on the output.

20
Additive noise
The theoretical work here analyses the error which limits dynamic range as an additive
\noise", whether the source of such noise is quantisation error, thermal noise or distortion.
The gain from this noise to the output is a key quantity. This same approach is used for
both lters (in chapter 2, using the noise gain vector G(s)) and sigma-delta modulators (in
chapter 4, using the noise transfer function H (z )).

This linear analysis is widely applicable as any distortion or parameter variation can be
considered as the original signal plus a \noise". Other approaches can give more
information if the \noise" has a consistent structure. Distortion cancellation and calibration
become possible, and problems such as tones in sigma-delta modulators need to be
considered. However, quantisation noise and particularly thermal noise are often fairly
white, so the linear analysis does not miss any information. For distortion this is less true,
but the author still considers that much can be learned by considering a transistor as a
linear two-port with an additive drain current distortion \noise".

Analog signal processing and causality


A distinctive feature of analog signal processing is that it is done in real time. Circuits are
of course causal, so only past inputs are available. This statement seems banal, but it is
important because the key derivations in chapters 2 and 4 rely on the causality property. If
the output at all past and future times was available, like when running a batch digital
signal processing task, the dynamic range limitations would be entirely di erent.

Another notable feature of analog signal processing is that storing signals is dicult.
Feedback cannot be used to correct for inaccurate storage. Compare this with a digital
lter where signals can be stored exactly and (rounding) error is only in multipliers and the
like. This restriction, together with causality, means that a good lter or ADC works on
signals in the recent past. In general, the more recent the signal the better, but at some
point transistor speed becomes a limiting factor. This idea is present in the way we consider
an additive noise to each integrator in chapter 2 and in the way we consider excess
sigma-delta modulator loop delay in chapters 4 and 5.

21
1.4 SCOPE
The work tries to identify new circuit techniques. It is not a general \roadmapping"
consideration of analog circuit performance versus process feature size, despite the fact that
the \fundamental limitations" derived have relevance in this context. The circuit types
considered are also limited.

1.4.1 Pitfalls of roadmapping


This work does not attempt to analyse in detail the e ects of scaling on analog CMOS
circuits. The following example helps illustrate why:

A scaling of  = 4 would reduce the minimum channel length to 2 m and


the circuit area by approximately a factor of 16 while still keeping second-order
e ects within bounds. Any further scaling would lead (as will be seen in the
following discussion) to unacceptable degradation in analog performance. (Wong
& Salama 1983)

The \unacceptable degradation" mentioned is a combination of higher power


consumption for a given dynamic range and lower opamp gain due to the increased output
conductance of short-channel devices. Today, the process generation at which further
scaling is considered detrimental seems to be around 0.35 m (Annema 1999) (Buss 1999).
The same reasons are given. Attempts to nd a point at which scaling becomes detrimental
seem to be too dependent on assumptions. To be quantitative, such analyses must use
today's circuit techniques, which are suited to today's processes. The aim of this work is
rather to identify tomorrow's circuit techniques.

It should be noted that the general studies of analog scaling do serve the valuable
function of identifying \On Growth and Form" (Thompson 1917) quality factors and gures
of merit. These simplify design optimisation by making the design variables orthogonal.
Some papers on analog scaling which the author has found useful are (Vittoz 1993),
(Vittoz 1994), (Castello & Gray 1985), (Groenewold 1992), (Annema 1999) and (Bult 2000).

22
1.4.2 Target application and frequency range
This thesis concentrates on communications circuits with bandwidths of 10 MHz { 1 GHz.
The Radiata (now Cisco) / Macquarie University / CSIRO 5 GHz 50 Mbit=s wireless LAN
(Skellern et al. 1997) was often the target application in mind. This has several implications:

 Flicker noise is not discussed as thermal noise dominates at these bandwidths.


 The emphasis is on continuous-time circuits. Partly this is because of the clock-rate
limitations of switched-capacitor circuits. Partly it is because continuous-time circuits
can allow larger dynamic range because they do not \fold down" thermal noise into
the passband. Partly it is because switched-capacitor circuits have already been
researched by many competent researchers.
 The bandwidth is a little high to be investigating class-B and rail-to-rail opamps.
\High signal swing" at these bandwidths generally means a signi cant fraction of
rail-to-rail { the aim is to avoid circuits which are limited to swings of the order of
ID =gm .

The consideration of RF CMOS was beyond the scope of this thesis. Active lters and
ADC's do actually present more of a dynamic range bottleneck than the RF signal path.
VCO phase noise and digital noise coupling are big RF dynamic range issues, but they are
large problems in themselves.

\Dynamic range" will mean the ratio of maximum signal to the noise at the same time.
The other popular use of \dynamic range" which is connected with automatic gain control
(AGC), companding and log-domain lters (that is, circuits which can have small signal
and small noise at one time, and large signal and large noise at another) will not be
discussed. The statement \dynamic range will degrade with CMOS scaling" is obviously
not about limitations of AGC circuits.

Current-mode and switched-current circuits are not considered. Only a few


state-of-the-art circuits could be designed in the time available, and the author's guess was
that feedback, voltage-mode, polysilicon resistors and routing metal capacitors were the
way to get optimum dynamic range. This work tries to demonstrate that such circuits allow
23
performance equal to that achieved in older CMOS processes; the question of whether
current-mode circuits can also achieve such performance is left open.

1.5 SYNOPSIS
The body of this work consists of a chapter of active lter dynamic range theory, a chapter
describing an active lter, a chapter of sigma-delta modulator dynamic range theory, and a
chapter describing a practical sigma-delta modulator, in that order.

State-space calculation of dynamic range


Chapter 2 describes how to calculate the dynamic range of a system such as an active lter,
using its state-space representation. If the reader has seen the work of Groenewold (1991),
they should think of the chapter as developing that paradigm. This work is to be presented
as

Harrison, J. & Weste, N. (2002), A limitation on active lter dynamic range,


accepted for `IEEE Int. Symp. Circuits Syst.'

An earlier version of this work has been presented as

Harrison, J. & Weste, N. (2001), Energy storage and Gramians of ladder lter
realisations, in `IEEE Int. Symp. Circuits Syst.', Vol. 1, pp. 29{32.

However, the results in that paper are superseded by the dynamic range limitation
described in chapter 2 and in (Harrison & Weste 2002).

The state-space treatment of dynamic range originates in work done on nite-wordlength


e ects in digital lters in the 1970's. Storing a digital signal in a nite number of bits is
clearly a limitation on the dynamic range. Both digital lters and analog active lters based
on integrators have very convenient state-space descriptions. There are an in nite number
of di erent states-space realisations for a given transfer function. The question is then
which one is the best in terms of dynamic range. A reasonable answer to this was found by
24
Mullis & Roberts (1976) and Hwang (1977). This theory involves certain invariants known
as Hankel singular values or second-order modes.

The Hankel singular values also appeared in the related work of the Russian operator
theorists (Adamyan, Arov & Krein 1971). This work was taken up by the control systems
community to use for model reduction { the reduction of the dimension of the state-space
model of a system. Model reduction is an extreme case of nite wordlength realisation {
some states are given no bits at all! Chapter 2 uses a frequency-weighted version of the
dynamic range optimisation / model reduction theory, which is due to Enns (1984) and
Thiele (1986). The developments in control systems and digital ltering largely proceeded
independently.

The state-space theory was applied to analog active lters as it developed. The
state-space description itself has a long history; see for example Anderson & Vongpanitlerd
(1973). The dynamic range theory (with its Hankel singular values, Gramians, gains to
states and noise gains) was only used more recently, by Snelgrove & Sedra (1986), Johns &
Sedra (1987) and Groenewold (1991). The main reason for this is probably that the
\optimal" realisations turn out to be impractical. Because of this, the main use of the
theory is for analysis and conceptual understanding. Chapter 2 continues the development
in this direction. With Groenewold's work and this work, the theory is reaching the point
where it will become the 21st century textbook introduction to lter design. The rst half
of chapter 2 is similar to Groenewold's and Snelgrove's development; the second half
continues it with a new limitation on dynamic range, a new frequency-domain presentation
of the optimal realisation theory, and an application to feedback ampli ers.

A 350MHz opamp-RC lter


Chapter 3 presents an opamp-RC lter built in a 0.18 m standard CMOS process. The
lter is notable in that it has a very high cuto frequency (up to 350 MHz) for an
opamp-RC architecture, and large signal swing (0.5 Vp;p di erential for {40 dB THD)
considering the low VDD (1.8 V) and frequency of operation.

Most of the prior art { opamp-RC lters, state-space simulation of LC ladders,

25
feedforward and two-pole compensation of opamps { was in place by the early 1970's. For
example, see (Anderson & Vongpanitlerd 1973) and (Roberge 1975). Why research
opamp-RC lters in 2001 then?

Much of the research in the intervening period involved developing active lter
architectures compatible with CMOS processes. MOSFET's were used in place of resistors
to allow continuous tuning (MOSFET-C lters), and transconductor-C (gm-C) lters were
developed to allow grounded MOS capacitors and supposedly to allow operation at higher
frequencies. The number of papers published on transconductors is extraordinary.
Switched-capacitor lters were of course also developed. However, circumstances have
gradually changed to make an on-chip opamp-RC lter practical:

 Even digital CMOS processes have reasonable polysilicon resistors (non-silicided


polysilicon is used in ESD protection).

 Tuning (to overcome process variations) can use resistors and CMOS switches rather
than MOSFET's. In the past, such a switched tuning made tuning loops a little more
awkward. However, tuning loops are often digital today and no such problem exists.

 With the increasing number of routing metal layers, sandwich capacitors with little
bottom-plate capacitance can be built in standard digital processes. Grounded
capacitors are thus no longer necessary.

The nal consideration is to make an opamp with sucient bandwidth, which is the
subject of much of chapter 3. Note that opamp-RC lters potentially have an advantage
over gm-C lters with respect to dynamic range. This is both because feedback circuits
tend to allow larger signal swing with acceptable distortion, and because the fundamental
limitations on integrator dynamic range are about 4 dB better (Moreira & Silva 2001)
(Groenewold 1992).

Analytic limitations on sigma-delta modulator performance


Chapter 4 derives fundamental limitations involving the tradeo between sigma-delta
modulator oversampling ratio and resolution. It has been presented as:
26
Harrison, J. & Weste, N. (2000a), Analytic limitations on sigma-delta modulator
performance, in `IEEE Int. Symp. Circuits Syst.', Vol. 3, pp. 746{749.

The results draw on two theories. One is the noise transfer function (NTF) linearised
analysis of sigma-delta modulators. The basic approach of the NTF analysis is to regard
quantisation error as an additive noise, and to calculate the noise transfer function H (z )
from this noise source to the output. Analysing quantisation error in this way is classical,
and the NTF analysis seems to have developed over time. An important more recent aspect
of this theory is that the stability of sigma-delta modulators can be roughly assessed by the
maximum of the NTF jjH (ej )jj1 . (This maximum of course occurs out-of-band; the aim is
to make the in-band NTF small.) An example of such work is Schreier (1993). Excessive
NTF's make the sigma-delta modulator overload from its own quantisation noise.

The other theory is that of integral sensitivity limitations, which basically starts with
Bode (1945). These state that while the sensitivity of a control system can be made very
small at a spot frequency, it cannot be made very small over a wide bandwidth. The
de nition of sensitivity in control theory is the same as that of the NTF, 1=(1 + L(z )) where
L(z) is the loop gain. The discrete-time version of Bode's result and another less well
known result from linear prediction theory are used here. The application of the Bode
integral to sigma-delta modulators has occurred to previous authors such as Gerzon &
Craven (1989); the best exposition elsewhere in the literature seems to be Nawrocki,
Goldberg & Sandler (1999). (This was published after the author had re-derived and
submitted similar results, although reviewers suggest that the ideas have been around for a
while and were known to researchers such as Magrath and Sandler.)

The Bode integral seems to be regarded as somewhat obscure and academic by circuit
designers, but it is considered one of the key results of robust control theory:

It is natural that Bode's integral formula should have a central place in any
theory of complex systems, as it was the rst result to focus completely on
robustness trade-o s, in this case imposed by causality. [J. C. Doyle, in
(Antsaklis 2000)]

27
A 520MHz sigma-delta ADC with a FIR DAC loop lter
Chapter 5 presents a sigma-delta ADC using a novel architecture in which the loop lter is
implemented in the DAC. The ADC clocks at 520 MHz, faster than that possible with
switched-capacitor techniques. This work has been presented as

Harrison, J. & Weste, N. (2000b), A multi-bit sigma-delta ADC with an FIR


DAC loop lter, in `Proc. IEEJ Int. Analog VLSI Workshop', pp. 26{31.

The history of sigma-delta ADC's has two eras. In early work, a wide variety of
architectures were tried. Eventually the successful combination of loop lter in the forward
path and switched-capacitor implementation became established. Later work concentrated
on improving the switched-capacitor sigma-delta modulator, although some diversity has
returned with investigation of continuous-time loop lters and mismatch shaping. The
dividing line between the two eras can be taken as (Candy 1985). The work presented in
chapter 5 is peculiar in that the loop lter is in the feedback path, and the implementation
is not switched-capacitor (although it is discrete-time). It in fact owes a lot to earlier work
such as (Tewksbury & Hallock 1978) and (Spang & Schultheiss 1962).

28
2 STATE-SPACE CALCULATION OF DYNAMIC
RANGE

2.1 INTRODUCTION
The dynamic range of a signal-processing building block is the maximum signal which can
be processed without overload, divided by the processing error. This de nition can be
applied to a wide variety of circuit building blocks. Indeed, it is useful to do so because
similar state-space theory can be used. In this work, the idea will be used in three cases:

1. An active lter or integrator in an active lter has a dynamic range limited by the
maximum signal swing, distortion and thermal noise.

2. A quantiser, be it a stand-alone ADC or the quantiser in a sigma-delta modulator, has


a dynamic range limited by the number of bits. Basically we are talking about a
dynamic range of (6n + 2) decibels here, where n is the number of bits. (Of course,
comparator o sets and the like can degrade the dynamic range.)

3. A transistor stage can be considered to have a dynamic range which is the ratio of
maximum signal to distortion plus noise. An output stage with a drain current which
swings from 0.1 mA to 0.9 mA might have a total harmonic distortion around 5%,
giving a dynamic range of 26 dB.

The reader may consider any one of these examples, particularly the last, an abuse of the
term \dynamic range". However, the point of this chapter (and indeed much of this thesis)
is that the three can be analysed in the same way, and that it is conceptually useful to do
so. The way they are analysed is to consider all noise and distortion as an additive noise on
top of the correct signal. This work is limited to studying those systems which are linear
apart from the additive noise. Some name for the signal processing accuracy has to be used;
perhaps an alternative to dynamic range might be peak signal to noise and distortion ratio,

29
but this is cumbersome. The processing error will be generically termed \noise"; remember
that it may be quantisation error or distortion as well as thermal noise.

It is fairly straightforward to calculate the dynamic range of sub-blocks such as


opamp-RC or gm-C integrators (Groenewold 1992) (Moreira & Verhoeven 1998)
(Groenewold 1991), quantisers, switched-capacitor stages (Dias, Palmisano, O'Leary &
Maloberti 1992) (Castello & Gray 1985), nite-wordlength digital multipliers (Rabiner &
Gold 1975) (Roberts & Mullis 1987), ampli er stages (Annema 1999) (Bult 2000) and the
like. Some of these results are in this chapter. (The term \sub-blocks" is used here as a
circuit such as an active lter or sigma-delta ADC is still only a building block in a
complete system such as a radio receiver.) The considerations for improving the dynamic
range of such sub-blocks are also well understood { for an ampli er stage, increase the
transistor widths and supply current to lower the noise; for a quantiser or digital circuit,
add additional bits. There are always trade-o s with power consumption.

However, this is only part of the problem. The dynamic range of an active lter is not
simply the dynamic range of its integrators. For instance, a high-Q bandpass lter has less
dynamic range than the integrators it is built from (Groenewold 1991) (Abidi 1992).
Similarly, a negative-feedback ampli er can have more dynamic range than the stages it is
built from. Consider an audio power ampli er { the class-B output stage might have 1%
distortion, or a dynamic range of 40 dB if distortion is counted as \noise". However,
feedback might reduce the distortion to 0.01%, or 80 dB dynamic range. (The input stage of
the audio power ampli er need not have 80 dB dynamic range either { it does not have
large output.)

2.1.1 Four questions and the organisation of this chapter


Dynamic range can be analysed by considering noise and distortion to be an additive signal.
We want to analyse circuits composed of sub-blocks. An archetypal example of such a
problem is an active lter where the sub-blocks are integrators. To nd the dynamic range,
there seem to be four natural questions:

1. What is the maximum signal swing of the sub-blocks? For the lter archetype this is
30
the maximum integrator output swing for acceptable distortion.

2. How large is the additive noise at the point where it is generated? We are talking here
about calculating the thermal noise of an integrator, the quantisation error of a
quantiser, or the noise and distortion of a transistor stage.

3. How does the signal swing in the sub-block relate to the input signal amplitude? We
have restricted ourselves to systems which are linear (apart from the additive noise),
so we are interested in the gain from the input to internal nodes. For active lters we
will de ne a vector gain to states F(s).

4. What is the gain from this noise source to the output? For a lter, how does lter
noise relate to integrator noise? We will de ne a vector noise gain G(s) from noise
added to the lter states, to the output.

Section 2.3 brie y considers question 1, the maximum signal swing in analog circuits.
Section 2.4 brie y considers question 2 for one noise source which ultimately limits
performance, thermal noise. These problems seem relatively well understood, and the
author does not have a large contribution to make. Results are mostly reviewed only to the
extent that they will be used later.

The bulk of this chapter is about questions 3 and 4. To analyse a system some structure
needs to be assumed. In this chapter the well-known state-space equations are used as the
structure. (In chapter 4, we use another well-known structure, that of a feedback loop.)
The state-space equations are introduced in section 2.2. (This section is put before the
discussions of maximum signal swing and thermal noise, so that noise can be given a
state-space description immediately.) Section 2.5 contains two examples of giving systems a
state-space description { an LC lter and an opamp.

Section 2.6 derives a key limitation on the achievable dynamic range of a state-space
system as a function of its frequency response. Section 2.7 gives an example of combining
this dynamic range limitation with practical maximum signal and thermal noise results.
The example is a lter needed for a particular wireless LAN architecture. Sections 2.8 and
2.9 use this result to derive a state-space lter realisation with certain optimal properties

31
with respect to dynamic range. The work in these sections appears to be new. For
comparison, existing related work is given at the end of the chapter in section 2.11.

The archetypal example considered in this chapter is the active lter composed of
integrators. To show that the state-space theory is conceptually useful in a wider sphere,
section 2.10 continues the application of it to opamps started in section 2.5.2 by discussing
how feedback ampli ers are essentially \predicted" by the state-space theory.

2.2 STATE-SPACE EQUATIONS


A versatile formalism for studying how the dynamic range of a lter relates to the dynamic
range of its integrators uses the state-space equations. This approach was rst applied to
digital lters (where nite-wordlength multipliers constrain dynamic range) in the 1970's
(Mullis & Roberts 1976) (Hwang 1977) (Roberts & Mullis 1987). Later it was also applied
to analog active lters (Groenewold 1991) (Johns, Snelgrove & Sedra 1989). Other useful
references are Thiele (1986) and Lutz & Hakimi (1988).

The state-space theory can also be usefully be applied to transistor stages. The
\integrators" are the drain current of one stage owing into the gate capacitance of the
next. The t is not quite as good as with opamp-RC or gm-C integrators or digital delays
(z ;1 ). However, the idea allows a deeper understanding of feedback, as will be discussed
later in this chapter.

The internal structure of a system (such as a lter made up of integrators) can be


speci ed by the state-space equations
dx = Ax + Bu + e (2.1)
dt
y = C x + Du (2.2)
where u is the input, y is the output, and x is a vector of internal states. One element
which may be less familiar is e, which represents noise input to the states. In an opamp-RC
lter like that in the next chapter, the state x is the set of integrator output voltages, and e
is a combination of noise current from the tap resistors and opamp noise. Anderson &
Vongpanitlerd (1973) and recent undergraduate control theory textbooks such as
32
(Ogata 1997) or (Kuo 1995) contain further introductory material on the state-space
equations. Filters not made up of integrators can also be given state-space representations
using capacitor voltages and inductor currents as states.

In the Laplace-transform domain (with zero initial conditions),

sX (s) = AX (s) + BU (s) + E(s)


Y (s) = C X (s) + DU (s);
which may be solved to give

X (s) = (sI ; A);1 BU (s) + (sI ; A);1 E (s)


Y (s) = C X (s) + DU (s)
= (C (sI ; A);1 B + D)U (s) + C (sI ; A);1 E (s):

These expressions will be frequently used, so we de ne the quantities

F(s) = (sI ; A);1 B (2.3)


G(s) = C (sI ; A);1 (2.4)
H (s) = C (sI ; A);1 B + D (2.5)

so that

X (s) = F (s)U (s) + (sI ; A);1 E (s)


Y (s) = H (s)U (s) + G(s)E (s):
H (s) is the familiar input-output transfer function. F (s) is the gain vector from input to
states. If F (s) is large, then the integrators in a lter will overload with small inputs. G(s)
is the gain from the noise added to the states E (s) to the output. If G(s) is large, then the
lter will have much output noise. The aim of much of this chapter is to see how small F (s)
and G(s) can be made, while still realising the desired lter transfer function H (s).

A comment about the dimensions and units of some of these quantities should avoid
possible confusion. F (s), x and B are column vectors, assuming there is one input. G(s)
and C are row vectors, assuming there is one output. A is a matrix and D is a scalar. The
units are particularly unexpected. Given the symmetry between F (s) and G(s), it might be
33
expected that they have the same units. They do not. Suppose that the input, state and
output are all voltages. F (s) maps the input voltage to the state voltages, so it is
dimensionless. The transfer function H (s) = C F (s) + D is dimensionless as it is a ratio of
voltages, so C and D are dimensionless. For (sI ; A);1 to be dimensionally consistent, A
must have units of frequency like s. Thus (sI ; A);1 has units of time. Now
G(s) = C (sI ; A);1 , so it has units of time. (G(s) is the confusing one.) The noise output
p p
G(s)E (s) is in V/ Hz, so E (s) is in V Hz. (The noise E (s) is an integrator input rather
p
than output, so it is not measured in V/ Hz.) If some of the signals are current-mode, the
units will change although the peculiar time dependences remain.

2.3 MAXIMUM SIGNAL SWING AND SOME SCALING


CONSIDERATIONS
The maximum signal swing of the integrators themselves will now be discussed. The
material has been covered previously by many authors (Moreira & Verhoeven 1998)
(Annema 1999) (Bult 2000) (Buss 1999), so only a brief introduction is given here.

2.3.1 Maximum signal swing


Maximum signal swing is basically limited by the power supply voltage. Opamp-based and
transconductor-based lters use similar ampli ers. The output stage is usually either a
cascoded stage or a simple (uncascoded) common-source stage. For a cascoded output
stage, the output can swing within two saturation voltages Vsat of each supply rail, so the
swing (per side) is VDD ; 4Vsat . Without cascoding, the swing is VDD ; 2Vsat . It is dicult
to nd data on how Vsat has varied with scaling, but from experience it seems to have
remained constant at around 0.2 V.

2.3.2 Basic points regarding power supply voltage scaling


If dynamic range was optimised by maximising signal swing, the low VDD of the
deep-submicron CMOS processes favoured for digital circuits (and hence systems-on-a-chip)

34
would be a big problem. However, the situation is not that bad.

For a given power consumption, if VDD halves then the power supply current can double.
All impedances in the circuit can be divided by four. This halves the thermal noise voltage,
and the dynamic range is unchanged. When authors (Annema 1999) calculate that dynamic
range will fall with falling VDD , they are not just claiming that signal swing will fall; they
are claiming that signal swing as a percentage of the supply voltage will fall. This point is
of course well known to the experts, but in casual discussion (particularly with digital
designers, whose noise problems are rarely thermal) it is often missed.

Using the results in section 2.3.1, the signal swing as a fraction of VDD is thus
approximately 1 { 0.8 V / VDD for cascoded output stages and 1 { 0.4 V / VDD for
uncascoded output stages. Thus, without circuit changes dynamic range does degrade with
scaling for common opamp and transconductor circuits. Clearly, cascodes face severe
problems by the 0.13 m generation (VDD = 1.2{1.5 V). This trend seems to be visible
today with the 0.25{0.18 m generations { Miller compensated opamps with simple output
stages are seen more often than they were ve years ago.

For more formal derivations of these results, see (Castello & Gray 1985) for
switched-capacitor lters and (Groenewold 1991), (Groenewold 1992), (Vittoz 1993) and
(Vittoz 1994) for continuous-time lters. A representative simple result is that for a
continuous-time class-A stage which has a maximum signal swing of VDD ; V , the
minimum possible power consumption P is (Annema 1999)
P = 81kTf sig SNR
; V=VDD (2.6)
when outputting a single sinusoid of frequency fsig . As discussed above, a typical V is
0.8 V for cascoded stages, and 0.4 V for uncascoded stages. Note that the result is linear not
quadratic with signal swing { if the signal swing is reduced, the impedance levels can also
be reduced.

Using (2.6), the dynamic range loss for a VDD as low as 0.8 V is only 3 dB for simple
common-source stages. The popular opinion seems to be that higher VDD , say 3.3 V, is
preferable to avoid this degradation. The author's opinion is that the increased fT of
deep-submicron devices will allow feedback circuits to be used to achieve signal swings
35
closer to rail-to-rail with acceptable distortion. The lter in chapter 3 and the sigma-delta
modulator in chapter 5 are investigations of this e ect. A 3 dB loss is not a huge amount to
regain. It should be stressed that the aim of this work is not to argue this point, but to try
to improve the dynamic range whether it gets better or worse.

There are some other scaling considerations apart from output swing, which are brie y
mentioned below.

Capacitor area
Dividing all impedances by four multiplies all capacitances by four, so chip area may be
prohibitive. This is o set by being able to run switched-capacitor circuits with higher clock
rates and hence smaller capacitors for the same dynamic range (Dias, Palmisano, O'Leary
& Maloberti 1992, eq. (12)). The area capacitance of metal-insulator-metal and routing
metal capacitors is also greater in more modern processes (Semiconductor Industry
Association 1999).

Mismatch
Another argument made is that the limiting \noise" is mismatch rather than thermal noise.
However this has stayed roughly proportional to thermal noise (Pelgrom, Tuinhout &
Vertregt 1998, g. 4) at around 100kT .

Power supply noise


If SNR is limited by power supply noise, if voltages are halved and currents are doubled,
then power supply impedance has to fall by a factor of four for the same SNR. This and
other signal integrity issues are beyond the scope of this thesis. Veendrick (2001) makes the
interesting point that at the 0.1 m generation, VDD falls below two diode drops so CMOS
latch-up should not be possible as the parasitic SCR cannot latch. This allows the use of
non-epi substrates without conservative design rules to prevent latchup. Non-epi substrates
allow better signal isolation.

36
Biasing diculties
There are also other diculties at low VDD . Using a common-mode voltage of VDD =2, the
input di erential pair requires Vgs + Vsat  VDD =2. This becomes dicult below VDD =
1.8 V once process, temperature and VDD variations are taken into account. The input
common-mode voltage can be moved toward VDD or ground in some switched-capacitor
circuits however. There are also problems with CMOS switches; the switch at the output of
an opamp requires VDD > 2VT plus some overdrive. In a 0.18 m process, VT is 0.5 V
typical, 0.6 V at the slow corner, and close to 0.7 V at the slow corner at low temperature.
VDD is 1.8 V nominal and perhaps 1.8 V {10% = 1.62 V worst-case. This leaves too little
switch overdrive for decent speed.

There are two process solutions for these problems (Semiconductor Industry
Association 1999). One is low-threshold or zero-threshold transistors. The large foundries
o er these, but do not seem to comprehend the importance of having minimum-length
zero-threshold devices. The other is to use the higher voltage (often 3.3 V) digital I/O
transistors for analog circuits (Buss 1999). This has merit. The possibility of using a
combination of fast low-voltage transistors and slower high-voltage transistors also deserves
attention. If, say, 1.8 V and 3.3 V rails are available, one option is to use 1.8 V opamps with
3.3 V nMOS-only switches.

There are also circuit design solutions. For switched-capacitor circuits, switched opamps,
clock boosting and alternative switch con gurations which remove the switch at the output
can be used (Bult 2000). If switched tuning is used in continuous-time circuits, similar
techniques are possible.

2.4 NOISE
The state-space analysis is applicable to a variety of noise sources. It was rst used for
digital lters, where the noise source is rounding errors due to nite wordlength, often in
multipliers. For analog circuits, distortion can be included in e as well as noise.

Discussion in this thesis centres on frequencies of 10 MHz to 1 GHz. As icker noise

37
Cf

Gi Gf
x

u iN

Figure 2.1: Opamp-RC lter for noise calculation. Signals can be vectors.

corners in CMOS transistors are rarely above 1 MHz (Martin, Archer, Boulin, Frei, Ng &
Yan 1997), icker noise was not a signi cant issue.

2.4.1 Thermal noise


For opamp-RC and switched-capacitor circuits, thermal noise is easy to calculate as opamp
contributions are usually not dominant. There are two useful ways to consider thermal
p
noise. One is to use the resistor noise voltage 4kTRf . The other is the thermodynamic
consideration that for linear passive circuits there is 12 kT of average noise energy per degree
of freedom (Pecseli 2000, p. 42). This leads rapidly to the switched-capacitor noise voltage
p
of kT=C using 21 kT = 12 CV 2 . As switched-capacitor circuits are not used in this thesis
(apart from one sampling switch in a less critical position in the sigma-delta modulator in
chapter 5), switched-capacitor circuits will not be considered further. It would be
interesting to combine the results on switched-capacitor dynamic range in (Castello &
Gray 1985) with the original discrete-time version (Mullis & Roberts 1976) (Hwang 1977) of
the state-space theory.

For an opamp-RC lter ( g. 2.1) with ideal (noiseless, in nite gain) opamps, the rst
state equation (2.1) is

Cf ddxt + Gf x + Giu + iN = 0
dx = ;C ;1G x ; C ;1 G u ; C ;1 i
dt f f f i f N
= Ax + Bu + e

38
The noise current iN over a bandwidth f depends on the sum of the conductances at the
summing node (Groenewold 1991) (Pecseli 2000):
E (iN iN ) = 4kTGt f
The symbol * means conjugate transpose. E () is expectation. The conductance Gt \seen"
by the summing node is a diagonal matrix with elements (Groenewold 1991)
X
gt(ii) = jgi(i) j + jgf (ij) j:
j
The absolute values are used because in di erential circuits negative taps are possible by
swapping negative and positive inputs. The notation here is somewhat awkward; gf (ij )
means the ij th element of the Gf in g. 2.1 etc.

The state-space noise spectral density is therefore


E(j!)E (j!) = 4kTCf;1 GtCf;:
(Of course, Cf is real so Cf; = Cf;T . Inverse and transpose commute:
Cf;T = (CfT );1 = (Cf;1)T .)
For approximate calculations, it is usually simpler to consider an average integrator
capacitance Cav and average total conductance at the opamp input Gav , and take the noise
input at each state as
p
e (j!)
i RMS ' 4kTGav : Cav (2.7)

In addition to this, there is opamp thermal noise. However, with sensible design this is
usually smaller (Groenewold 1992). For example, in the lter in the next chapter, the
opamp thermal noise adds only 0.9 dB to the noise of an integrator. (For the opamp noise
contribution to be small, the opamp rst stage transconductance gm1 must be large
compared to the tap conductances Gf and Gi .)

2.5 EXAMPLES
To illustrate the state-space theory, we consider two examples. One is an LC ladder lter,
or an active simulation of it. Active simulations such as gyrator or leapfrog structures
39
1 L2 L4

+
2u C1 C3 C5 1 y
-

Figure 2.2: 5th order low-pass LC lter

inherit the state-space equations of the LC prototype they were based on (Johns
et al. 1989). Such \lossless" structures are known to have low sensitivity and high dynamic
range (Orchard, Temes & Cataltepe 1985). The other example is an ampli er stage using a
Miller-compensated opamp. This is a less traditional use of the state-space theory. It is
important because much of this thesis is about how feedback a ects dynamic range. It is
also a step towards the author's (currently unrealised) goal of synthesising opamps in the
same way that lters are synthesised.

2.5.1 LC ladder lter


The state-space equations for an LC ladder lter are readily derived using Kircho 's laws
and the constitutive relations (Anderson & Vongpanitlerd 1973, p. 147). The states are
taken as the capacitor voltages and inductor currents:
h iT
x = vC1 iL2 vC3 iL4 vC5

40
The coecient matrices are:
2 1 1 3
66 C C ; ; 0 0 0 7
66 L1 0 ; L1 0 0 777
1 1

66 2
7 2

A = 66 0 C1 0 ; C1 0 777
66 0 0 1 0 ; 1 77
3 3

4 L L 54 4

0 0 0 1 ; C1
C
223 5 5

1 66 C 77
66 0 77
B =
66 77
66 0 77
66 0 77
4 5
0
h i
C = 1 0 0 0 0
hi
D = 0
For the Butterworth response with 1 rad=s {3 dB frequency used, C1 = C5 = 0:618,
L2 = L4 = 1:618 and C3 = 2.

2.5.2 Feedback ampli er using a Miller-compensated opamp


The circuit investigated is shown in g. 2.3. This might typically be used as an
intermediate frequency or baseband ampli er in a radio receiver. C1 and C2 represent
transistor Cgs of the rst and second stages (and Cds and parasitics capacitances etc.). The
small-signal equivalent circuit is for the di erential-mode signal.

Transistor Cgd is omitted as it complicates the analysis. For the rst stage, gm1 is
reversed in sign to represent the left-hand-side opamp input being connected to the
right-hand-side opamp output and vice versa. This is dicult to analyse if Cgd is present.
There are also diculties relating to writing state equations for loops of capacitors. It can
be done using nodal analysis and converting (sC + G) into C (sI + C ;1G), but the results
are more dicult to interpret as a state no longer represent the voltage of a single capacitor.

The state-space equations were written using the states x1 to x4 as labelled on g. 2.3.
The component values used were:
41
bias
R5

R1

C4 R4 R4 C4

R1
cmfb
R5 cmref
R3 C3 C3 R3

R5

R1 R4 + x4 -
x2 x3 = y
x1
u C1 C4 C3 R3
C2

gm1.x1 gm2.x2

Figure 2.3: Feedback ampli er, opamp used, and small-signal equivalent circuit for the state-space
analysis.

42
Component Value Description
R1 5k
input resistor
R5 15k
feedback resistor
R3 5k
load resistance and gds of output stage
C3 200fF load, routing, second stage Cds , C4 bottom-plate
gm1 0.7mS rst stage
gm2 2mS second stage
C1 80fF Cgs of rst stage, routing parasitics
C2 80fF Cgs of second stage and Cds of rst stage
C4 100fF Miller compensation
R4 1k
removes Miller comp. right-half-plane zero

The values used are fairly realistic for a wideband IF ampli er stage. The transistors are
18/0.36 for the input stage and 18/0.18 for the output stage. The input stage uses around
100 A per transistor and the output stage uses around 250 A per transistor. The load is a
similar ampli er stage. The stage gain is 3 (9.5 dB).

Recall that the noise gain G(j!) is not dimensionless. For an otherwise identical system,
if the bandwidth is doubled, then the noise gains halve. To allow the feedback ampli er to
be compared with the LC ladder lter, all of the capacitances in the feedback ampli er were
multiplied by a factor of 2:5  109 , to bring the feedback ampli er bandwidth down from
400 MHz to 1 rad=s.

2.5.3 Results
Fig. 2.4 shows the transfer function H (j!) = [h1 ], gains to states F (j!) = [f1 f2    ]T , and
noise gains G(j!) = [g1 g2    ] for both the LC ladder and feedback ampli er examples.

LC lter
All elements of F (j!) are 0 dB at DC, as the current through the inductors is unity and the
voltage across the capacitors is unity. The noise gains G(j!) have di erent DC values

43
20 20
h h
Transfer function H (dB)

Transfer function H (dB)


10 10

0 0

10 10

20 20

30 30

40 40
1 0 1 1 0 1
10 10 10 10 10 10
Frequency (rad/s) Frequency (rad/s)
20 20
f1 f1
10 10
Gain to states F (dB)

Gain to states F (dB)

f2 f2
f3 f3
0 f4 0 f4
f5
10 10

20 20

30 30

40 40
1 0 1 1 0 1
10 10 10 10 10 10
Frequency (rad/s) Frequency (rad/s)
20 20
g1 g1
10 10
Noise gain G (dB(s))

Noise gain G (dB(s))

g2 g2
g3 g3
0 g4 0 g4
g5
10 10

20 20

30 30

40 40
1 0 1 1 0 1
10 10 10 10 10 10
Frequency (rad/s) Frequency (rad/s)

Figure 2.4: Transfer function H (j!), gain to states F (j!) and noise gains G(j!) of the LC lter (left)
and opamp (right) examples.

44
because they depend on the component value.

The signal levels here are unscaled; if this lter was used in a radio receiver it would be
desirable to scale down f1 to prevent the rst state overloading. This would increase g1 by
the same factor. The scaling depends on the application; for instance the unscaled levels
here would be good for a transmit lter designed to handle energy mostly concentrated in
the passband.

Feedback ampli er
Think of the elements of F (j!) as the AC voltages in the circuit for 1 V AC input. f1 is the
error voltage, which is small at low frequencies where the opamp gain is high. f2 is the gate
voltage of the second stage which is also small at low frequencies. f3 is the output, and is
therefore equal to H (j!). f4 is the voltage across the compensation capacitor, which is
similar to the output.

The noise gains G(j!) also make sense. The highest gain is from noise injected into the
opamp input, g1 . Fortunately transistor distortion currents do not ow into this node. The
noise gain from the output of the rst stage is g2 , and is somewhat lower. Distortion
currents from the rst stage ow into this node, so it is fortunate that the signal swings
associated with the rst stage (f1 and f2) are small and hence should have low distortion
component. The noise gain g3 from distortion currents injected into the output node is very
small at low frequencies. This is the feedback working { it is important that g3 be small as
the output stage is likely to generate signi cant distortion current. g4 represents the gain
from noise in the compensation capacitor, and is quite large. Being a passive component,
hopefully the compensation capacitor should generate little noise and distortion.

In the introduction (section 2.1) it was claimed that the components in a feedback
ampli er can have less dynamic range than that required of the output. It can be veri ed
that jf1 jjg1 j, jf2 jjg2 j and jf3 jjg3 j are all signi cantly less than one at low frequencies where
the feedback \works". The LC lter does not show this behaviour. Thus the dynamic range
required of the components really is reduced by a feedback ampli er.

However, note that jf4 jjg4 j in the feedback ampli er is large. Thus the compensation

45
capacitor (whose voltage is state x4 ) requires high dynamic range. This should be achievable
by using a linear capacitor such as a metal-insulator-metal or routing metal structure.

Do circuits exist in which all states do not require high dynamic range? The theory
which follows provides some information.

2.6 THE DYNAMIC RANGE LIMITATION


This section derives a key relationship, (2.8), which the author calls the \dynamic range
limitation". This name is used as the result limits the lter dynamic range achievable with
a given integrator dynamic range and lter transfer function.

2.6.1 Derivation
The following identity will be used:

(s1 I ; A);1 ; (s2 I ; A);1 = (s2 ; s1 )(s1 I ; A);1 (s2 I ; A);1

The proof is straightforward:

M = (s1I ; A);1 ; (s2 I ; A);1


(s1 I ; A)M (s2 I ; A) = (s2 I ; A) ; (s1 I ; A)
(s1 I ; A)M (s2 I ; A) = (s2 ; s1 )I
M = (s2 ; s1)(s1 I ; A);1 (s2 I ; A);1
This is known as the resolvent identity (the expression (sI ; A);1 is called the resolvent).
Here s1 and s2 are scalars; think of them as two frequencies.

Now using M in the above proof,

CMB = C (s1I ; A);1 B ; C (s2I ; A);1 B


= C (s1I ; A);1 B + D ; C (s2 I ; A);1 B ; D
= H (s1 ) ; H (s2 )

46
but also

CMB = (s2 ; s1)C (s1 I ; A);1 (s2I ; A);1 B


= (s2 ; s1 )G(s1 )F (s2 )

so
H (s1 ) ; H (s2 ) = G(s )F (s ) (2.8)
s2 ; s1 1 2

Eq. (2.8) will be called the \dynamic range limitation" in the discussion which follows.
It relates the transfer function H at two frequencies s1 and s2 to the inner product of the
noise gain G(s1 ) at one frequency and the gain to states F (s2 ) at the other.

2.6.2 Interpretation
The dynamic range limitation (2.8) states that if the transfer function changes rapidly with
frequency, then at least one state must have a sizable product of the gain to state and noise
gain. That is, at least one state must require sizable dynamic range. The intuitive
explanation for why this is necessary is that if H (s) varies with frequency then past signals
must be stored, and they must be stored with dynamic range similar to that required of the
output.

The quantity G(s1 )F (s2 ) is a map from the s-plane to the H (s)-plane. If two points s1
and s2 are close together in the s-plane, but far apart in the H (s)-plane, then G(s1 )F (s2 )
has to be large. The map acts on the set of chords rather than on individual points.

If a lter was built from resonators with a transfer function 1=K (s) rather than
integrators with a function 1=s, then the quantity in the dynamic range limitation would
become [H (s1 ) ; H (s2 )]=[K (s2 ) ; K (s1 )]. Thus the best possible dynamic range depends on
the relative quality of the lter and the resonators it is built from. The Leeson (1966)
equation for oscillator phase noise is also a result of this nature.

47
2.6.3 Application to lter dynamic range
A radio receiver is an archetypal example of a circuit requiring high dynamic range. Usually
the greatest problem is to receive a small in-band signal in the face of large out-of-band
interferers.

Let s1 = j!P be a frequency in the receiver passband, and s2 = j!S be the frequency of
a large stopband interferer. Assume that jH (j!P )j > kP , and jH (j!S )j < kS . Then
H (j! ) ; H (j! ) k ; k
P S P S
j (!S ; !P )  !S ; !P
so
k ; k
jG(j!P )F (j!S )j  !P ; !S
S P
or
X k ; k
jfi(j!S )jjgi (j!P )j  !P ; !S (2.9)
i S P

If the input signal is large at the blocking frequency !S , it is important to have small
gains to states fi (j!S ) to prevent the states overloading. It is also important to have low
output noise in the passband, which requires small noise gains gi (j!P ). Eq. (2.9) shows that
this is only possible if there is sucient lter transition bandwidth j!S ; !P j. (jkP ; kS j
cannot be changed without more passband droop or much less stopband attenuation.)

2.6.4 Basic example


The following example gives an indication of whether (2.9) is tight enough to be useful.

We use the 5th order Butterworth lter analysed earlier in this chapter. It is {3 dB at
1 rad=s, and {40 dB at 2.5 rad=s. We will take these as !P and !S respectively. The
stopband gains to states and passband noise gains are as follows:

48
i fi (j!S ) gi (j!P ) fi (j!S )gi(j!P )
1 1:2296 ;52 0:2196 135 0:2696 173
2 0:3216 ;142 0:6736 167 0:2166 51
3 0:0686 128 1:1446 ;135 0:0786 ;97
4 0:0196 43 1:3276 ;60 0:0256 103
5 0:0106 ;14 0:4896 18 0:0056 ;33

This results in
X
fi(j!S )gi (j!P ) = 0:4776 45
i
H (j!P ) ; H (j!S ) = 0:4776 45
X j (!S ; !P )
jfi(j!S )jjgi (j!P )j = 0:592
i
kP ; kS
!S ; !P = 0:465

As expected, the dynamic range limitation (2.8) is obeyed. What is more remarkable is how
close (2.9) is to equality. This illustrates the well-known fact (Orchard et al. 1985) that LC
ladders and active simulations thereof are \good" lters with close to optimum dynamic
range.

2.7 802.11A RECEIVER EXAMPLE


The following example puts together the results derived so far.

2.7.1 Problem
An 802.11a 5 GHz wireless LAN must receive a {65 dBm 54 Mbit=s signal in the face of a
{50 dBm alternate adjacent blocking signal (IEEE Std. 802.11a 1999). The alternate
adjacent channel starts 20 MHz from the passband edge.

If a superheterodyne 5 GHz receiver is used, the intermediate frequency (IF) must be at


least, say, 300 MHz for the image to be suciently far away to lter out with a SAW, LC or
49
microstrip lter. The next chapter shows that on-chip ltering at 300 MHz is certainly
possible, although the Q-tuning required for 20 MHz passband centred at 300 MHz would be
delicate. Is such a lter practical from a dynamic range perspective?

2.7.2 Calculations
Filter dynamic range
The 54 Mbit=s signal is multi-tone 64-QAM, which with white noise requires a signal to
noise ratio (SNR) of around 20 dB for 10;6 bit error rate. As there are coding losses and
are other noise sources such as the DAC, ADC and digital lters, it seems sensible to aim
for an SNR of at least 30 dB for the lter.

The blocking signal is {50 dBm, but multi-tone modulation produces a high
peak-to-average-power ratio. Some margin for this is needed, say that overload should not
occur below {44 dBm. The lter dynamic range therefore needs to be 30 dB + ({44 dB {
{65 dB) = 51 dB.

Signal swing
The best signal swing which could be hoped for is the VDD ; 2Vsat per side mentioned in
section 2.3.1 for an uncascoded common-source output stage. In a 0.18 m process this is
typically 1.4 Vp;p per side, or 2.8 Vp;p di erential. In practice, at high frequencies it is
very dicult to achieve this with acceptable distortion. The lter in the next chapter
achieves {40 dB distortion at 0.5 Vp;p di erential signal swing. The extensions in section
3.6 show that 2 Vp;p appears achievable. We will assume a 1 Vp;p maximum di erential
signal swing (0.25 V peak per side).

Filter components
Assume that the power budget is around 20 mW. Assume there are three pairs of
integrators; 2{4 would be typical. This is 3.3 mW per opamp, or 1.9 mA for VDD = 1.8 V. If
the output stage current is 0.6 mA per side, a current swing of 0.4 mA peak per side seems
achievable. The voltage swing is 0.25 V peak per side, so the load on an opamp should be

50
0.25 V / 0.4 mA = 625
. This load is typically one capacitor and two or three resistors, so
a unit resistor should be around 2 k
.

To achieve a 300 MHz pole frequency with a 2 k


resistor, the integrator capacitors need
to be 250 fF.

Dynamic range limitation


Assume that the input, state and output signals are similar in magnitude. Then the gains
to states fi (j!S ) should not exceed 0 dB in magnitude. Assume that jH (j!P )j = 0:9
({1 dB) and jH (j!S )j = 0:03 ({30 dB). Then from (2.9),

X k ; k
jfi(j!S )jjgi (j!P )j  !P ; !S
i S P
X 0:9 ; 0:03
jgi(j!P )j  2  20  106 rad/s = 6:9  10;9 s
i

Integrator noise
Using (2.7), the noise input is
p4kTG p
Ei(j!)RMS ' C av 10;23  300  1  10;3 = 16 VpHz
= 4  1:38 250
av  10;15
assuming that the average conductance Gav connected to each opamp summing node is
p p p
1 mS. This is per side; the di erential noise is 2  16 V Hz = 23 V Hz.

Filter noise
P
As output noise is proportional to jgi j2 but the dynamic range limitation constrains
P jg j, minimum noise occurs if equali noise is produced by all six poles. As
i
Pi jg (! )j = 6:9  10;9 s, each jg (! )j is 1:15  10;9 s. In this case, the output noise
i P i P
i
contribution from each integrator is
p p
1:15  10;9 s  23 V Hz = 26 nV= Hz
Hence the total noise is
p p p
26 nV= Hz  6 = 65 nV= Hz
51
Over a 20 MHz bandwidth, this integrates to 290 V RMS.

In practice, the dynamic range is limited by the rst pair of integrators and the noise is
not so evenly averaged over the six poles. Based on the example in section 2.5.1 and actual
calculations, active simulations of LC ladders are likely to achieve around 700 V RMS
noise. The di erential signal swing is 0.35 V RMS, so the dynamic range achieved is 54 dB.
This is just above the minimum required.

2.7.3 Comments
Clearly such an architecture is marginal from a dynamic range perspective. Only thermal
noise from the resistors has been considered; on top of this is distortion and noise from the
opamps. Against this, a signal swing of 2 Vp;p di erential might be possible, and the power
budget could be increased a little. (There is a lot of broadband digital signal processing in a
54 Mbit=s wireless LAN, so 20 mW is unlikely to be a large component of the total power
consumption.)

To achieve higher dynamic range, a Q-enhanced LC lter could be used. This cuts the
power consumption for a given integrator dynamic range, as energy is stored for multiple
cycles rather than being taken from the power supply each cycle. The thermal noise
currents can also be lower. However, the frequency variation of inductor loss needs to be
carefully considered when designing the Q tuning, and the negative resistance to cancel the
losses still has to be linear for swings of a signi cant fraction of rail-to-rail.

2.8 STATE-SPACE REALISATION THEORY


The dynamic range limitation is useful in its own right. However, what is surprising is that
it can also be used to construct lter realisations with dynamic range which is in some sense
optimum. These \optimal" realisations were rst found in the 1970's and 1980's using a
time-domain and state-space approach. However, the new frequency-domain approach
presented here seems simpler and more comprehensible. (It should be noted that the
original algorithms are more computationally ecient though.)

52
In state-space terms, \realisation" means nding a set of matrices (A; B; C; D) which
give the correct transfer function H (s) = C (sI ; A);1 B + D. The key problem is that there
are di erent sets (A; B; C; D) which realise the same transfer function. Moreover, these
have di erent dynamic range properties, as the gains to states F (s) and noise gains G(s)
are di erent. The problem we are trying to solve is to nd a good set. Section 2.8.2
considers what sets exist, and section 2.8.1 considers what is a good one.

2.8.1 Controllability and observability Gramians


It turns out to be important to consider the gains to states F (s) and noise gains G(s) over
all frequency rather than at spot frequencies. Enns (1984) and Thiele (1986) have de ned
the frequency-weighted controllability Gramian matrix
Z1
K  = 21 [F (j!)(j!)][F (j!)(j!)] d! (2.10)
;1
and the frequency-weighted observability Gramian matrix
Z1
W = 12 [ (j!)G(j!)] [ (j!)G(j!)] d!: (2.11)
;1
(The symbol * is conjugate transpose.) The way to understand what these quantities are is
to look at their de nitions. K  is a weighted integral over frequency of the power gain to
states. The clearest way to think of this is that the diagonal element kii of K  is the signal
power in the i'th state for an input signal (s). The o -diagonal terms kij represent
correlations between the signals in the states. (K  and W are both n  n matrices for a
system with n states.) Similarly, W is a weighted integral of the noise power gain. The
noise output is weighted by an output lter (s). The diagonal elements wii represent
weighted output noise power gains, and the o -diagonal elements wij represent
frequency-weighted correlations between the noise gains from di erent noise inputs ei and
ej .
To optimise the dynamic range for input signal (s) and output noise weighting (s),
we want to make the kii and wii small. A brute-force way of doing this would be to
Pn
numerical minimise kii wii over the space of all realisations (A; B; C; D) which give a
i=1

53
transfer function H (s) = C (sI ; A);1 B + D which meets the passband ripple and stopband
attenuation speci cations. The main problem with doing this is that the optimisation
problem is too hard. For a fth-order lter, (A; B; C; D) has 36 parameters. This is quite
practical for a convex optimisation, but the passband ripple and stopband attenuation
speci cations appear to give a non-convex problem. Nevertheless, this approach should not
be ruled out, as a careful choice of a sparse (A; B; C; D) could reduce the dimensionality. For
example, (A; B; C; D) could correspond to the elements of an LC ladder. The optimisation
could also tweak an existing realisation within a convex region. The other reason why such
an approach is less theoretically interesting is that it o ers very little design insight.

Because of these problems, the most successful approach to lter design has been to split
the problem into choosing a transfer function and choosing a realisation of that transfer
function.

2.8.2 Similarity transformation


Di erent realisations (A; B; C; D) can be parameterised by substituting a transformed state

v = Tx

so that
dx = Ax + Bu
dt
T ;1 dvt = AT ;1v + Bu
d
dv = TAT ;1 v + TBu
dt
and

y = C x + Du
y = CT ;1v + Du:
Thus the substitution (A; B; C; D) (TAT ;1 ; TB; CT ;1 ; D) leaves the transfer function
unchanged. This is known as a similarity transformation. (Note that we do not include e in
the derivation, as the noise input is likely to be di erent in the transformed system.)

54
The gain to states and noise gain transform as follows:

Fv (j!) = (sI ; TAT ;1 );1 TB = T (sI ; A);1 B = T Fx (j!) (2.12)

Gv (j!) = CT ;1 (sI ; TAT ;1 );1 = C (sI ; A);1 T ;1 = Gx (j!)T ;1 (2.13)

Here Fv (j!) and Gv (j!) are the gain to states and noise gain for the transformed state
variable v, and Fx (j!) and Gx (j!) are the gain to states and noise gain for the original
state variable x.

Note that

Gv (j!1 )Fv (j!2 ) = Gx (j!1 )T ;1 T Fx (j!2 ) = Gx (j!1 )Fx (j!2 ):

If this did not hold, the dynamic range limitation could not be true.

The main signi cance of the similarity transformation is that it describes all minimal
state-space realisations with a given transfer function (Anderson & Vongpanitlerd 1973, pp.
101{103). (\Minimal" basically means having the same number of integrators as the order
of the transfer function. There is no evidence that non-minimal realisations allow superior
dynamic range, although it has not been proved impossible.)

2.8.3 Product of the Gramians


We could go on to calculate how K  and W transform under a similarity transformation.
The answer is not particularly interesting; they do not stay the same. What is more
interesting is how the product K  W transforms. From the de nitions (2.10) and (2.11) of

55
the frequency-weighted Gramians,

KvWv
Z1 Z1
= 21 [Fv (j!)(j!)][Fv (j!)(j!)] d!  21 [ (j!)Gv (j!)] [ (j!)Gv (j!)] d!
;1 ;1
Z1 Z1
= 41 2 [Fv (j!2 )(j!2 )][Fv (j!2 )(j!2 )] [ (j!1 )Gv (j!1 )] [ (j!1 )Gv (j!1 )]d!1 d!2
;1 ;1
Z1 Z1
= 41 2 Fv (j!2 )(j!2 ) (j!2 )Fv  (j!2 )Gv  (j!1 )  (j!1 ) (j!1 )Gv (j!1 )d!1 d!2
;1 ;1
Z1 Z1
= 41 2 Fv (j!2 )(j!2 ) (j!2 )[Gv (j!1 )Fv (j!2 )]  (j!1 ) (j!1 )Gv (j!1 )d!1 d!2
;1 ;1
Z1 Z1
= 41 2 T Fx(j!2 )(j!2 ) (j!2 )[Gx(j!1 )Fx (j!2 )] (j!1 ) (j!1 )Gx (j!1 )T ;1d!1d!2
;1 ;1
= TK W T ;1 :
x x

For any non-singular transformation M , the eigenvalues of TMT ;1 are the same as
those of M . (As an example, recall that A also becomes TAT ;1 under a similarity
transform. We know that this leaves the transfer function unchanged, so the poles are
unchanged. The poles of the system are the eigenvalues of A.) Thus the eigenvalues of
K W are invariant under a similarity transformation, and hence are the same for all
minimal realisations of a given transfer function. These eigenvalues i (K  W ) are
important, and (their square roots) are called frequency-weighted Hankel singular values.
The reason for this name will be apparent by the end of the chapter.
p
The Hankel singular values i (K  W ) are important because they restrict how small
the powers in the states kii and the noise power gains wii can be made. The matrices K 
and W cannot be both small if K  W is big. The diagonal elements kii and wii cannot
be small if K  or W are big, because like other covariance matrices, K  and W are
positive-de nite (and symmetric). There is rigorous theory concerning which norms to
de ne \big" and \small" in (Mullis & Roberts 1976) (Hwang 1977). The results become
somewhat messy because the norms are not quite the ones we would like, so we restrict
ourselves to an intuitive understanding.

56
2.9 AN OPTIMAL REALISATION BASED ON THE DYNAMIC RANGE
LIMITATION
The di erence between a good realisation of a lter transfer function and a bad realisation
seems somewhat mysterious at rst. By a good realisation we mean one with high dynamic
range and low sensitivity to component variations { fortunately these two properties are
compatible. Actually, it is reasonably straight-forward. A good realisation stores signals in
its states which are as orthogonal as possible, and it stores signals which are relevant to the
output. The situation is similar to conditioning of matrices { a transformation which
diagonalises the matrix improves conditioning as results are no longer produced by
subtracting large, nearly equal quantities. A good realisation \diagonalises" the gains to
states and noise gains. The Gramians K  and W become diagonal or diagonally
dominant. In terms of the discussion in the previous section, this is the way to minimise the
powers in the states kii and weighted noise output wii for a given transfer function and
hence given Hankel singular values.

The rst procedures for doing this were discovered in the mid-1970's (Mullis &
Roberts 1976) (Hwang 1977). These involved nding a similarity transformation T which
simultaneously diagonalises K  and W . (This work was done for digital lters, which are
discrete-time. However, the continuous-time theory is essentially the same.) Independently,
mathematicians developed similar techniques (Adamyan et al. 1971). The original work was
for weightings (s) = (s) = I , but it was generalised to the frequency-weighted case by
Enns (1984) and Thiele (1986).

A new frequency-domain solution to the problem is presented here. It is a direct


approach, in contrast to most current techniques which are based on transforming an
existing realisation. Direct approaches do already exist, developed by Zeiger & McEwen
(1974) and Kung (1978) from the work of Ho & Kalman (1966). However, they seem not to
have been extended to the frequency-weighted case.

The frequency-domain solution shows the signi cance of the dynamic range limitation.
However, it should be noted that the transformation-based approaches are often more
computationally ecient, so the best approach may be to understand the frequency-domain

57
solution but to use the traditional algorithms (Enns 1984) (Thiele 1986) (Anderson &
Moore 1989, problem 10.2-1). Previous approaches are discussed in detail in section 2.11.4
at the end of the chapter.

2.9.1 Singular value decomposition


The singular value decomposition of a matrix K is (Golub & Van Loan 1996, p. 69)
K = USV ;
where U and V are unitary matrices and S is a diagonal matrix with real, non-negative
entries i . The i are called singular values. They are almost always ordered from largest to
smallest, 1 > 2 >    , and the matrix S is arranged as
2 3
66 1 0 0   77
6 0 2 0   77
S = 666 7:
64 0 0 3   775
   
A unitary matrix is like a rotation { distances are unchanged by a unitary transformation.
A unitary matrix U obeys UU  = U  U = I . The singular value decomposition is widely
used in numerical linear algebra as the \rotate { stretch { rotate" sequence avoids
ill-conditioning. It is also useful for approximation techniques such as least squares, as small
singular values can be set to zero to approximate a matrix by one of lower rank.

For transfer functions, the equivalent decomposition is


K (j!1; j!2 ) = U (j!1 ) S V (j!2 );
or expressing this as a sum,
X
K (j!1 ; j!2 ) = iui (j!1 )vi(j!2 ):
i
The conditions for U (j!) and V (j!) to be unitary are chosen as
1 Z1 Z1
2 U (j!)U  (j!) d! = 1 2 U  (j!)U (j!) d! = I (2.14)
;1 ;1
1 Z1 Z1
2 V (j!)V  (j!) d! = 21 V  (j!)V (j!) d! = I: (2.15)
;1 ;1
58
One should think of !1 and !2 as an in nite-dimensional analog of matrix subscripts. In
general, S is a diagonal operator, although for certain K (j!1 ; j!2 ) it is nite-dimensional
and hence a diagonal matrix.

Singular value decomposition of the dynamic range limitation


Suppose we apply the singular value decomposition to the dynamic range limitation
multiplied by input weight (j!) and output weight (j!):

(j!1 ) H (j! 1 ) ; H (j!2 ) (j! ) = (j! )G(j! )F (j! )(j! ) = U (j! ) S V  (j! )
j (! ; ! ) 2 1 1 2 2 1 2
2 1
(2.16)
Then we have the lter realisation

F (j!)(j!) = V  (j!) (2.17)


(j!)G(j!) = U (j!) S: (2.18)

Why do this? Recall that F (j!) is the gain to states and (j!) can be thought of as the
input signal to the lter. By setting F (j!)(j!) equal to a unitary quantity V  (j!), we
are guaranteeing that there will be unit average power in each state. This is a sensible
scaling of signal level. There are further considerations to the e ect that the gains to states
and noise gains are uncorrelated, which are discussed in section 2.9.3. As the reader may
have guessed from their name, the elements i of the diagonal matrix S turn out to be the
p
Hankel singular values i (K  W ). Showing this is also deferred to section 2.9.3.

The values of the state-space matrices can be calculated. We use the property that

lim ( j!I ; A) ;1 = 1 I
!!1 j!
and the de nitions (2.3), (2.4) and (2.5) to calculate that

D = !lim
!1 H (j!) (2.19)
C = !lim
!1 j! G(j!)
(2.20)
B = !lim
!1 j! F (j!): (2.21)

59
Calculating A is a little less trivial. We use

F (j!) = (j!I ; A);1 B


(j!I ; A)F (j!) = B
AF (j!) = j!F (j!) ; B
AF (j!)(j!) = (j!F (j!) ; B )(j!)
and then using (2.17),

AV  (j!) = j!V (j!) ; B (j!):


We post-multiply by V (j!) and integrate over all frequency so that we can use (2.15):
Z1 Z1
A = 21 A V  (j!)V (j!) d! = 12 [j!V  (j!) ; B (j!)] V (j!) d! (2.22)
;1 ;1

The reader may have noticed that there is a big problem here, namely that the
realisation (A; B; C; D) is in nite-dimensional. However, it turns out that:

 If the transfer function being realised is rational and of order n, only n singular values
are non-zero.
 If only n singular values are non-zero, F (s) and G(s) are those of an n'th order
state-space system.
 If the transfer function closely approximates that of an nth order system, then all but
n singular values will be small. If these small singular values are set to zero, the
transfer function realised will be close to that desired. For weights (s) = (s) = 1, a
rigorous error bound is available, namely that the maximum transfer function error is
less than the sum of the singular values set to zero. For arbitrary weights, the
property is only known qualitatively.

These properties are dicult to establish. Indeed, the author is not even sure that they
have been proved for the frequency-weighted case. For the unweighted ((s) = (s) = 1)
case, they were proved by the Russian mathematicians Adamyan, Arov & Krein (1971).
Good (but by no means simple) descriptions are (Glover 1984) and (Zhou 1996, p. 159{).
60
The author would be surprised if anything went wrong with the derivation used here, as it
was found by transforming existing realisation theory into the frequency domain (see section
2.11) and including the frequency weights. Adding the weights should not cause problems as
the resulting frequency-weighted realisation is also well known. As engineering veri cation,
the author has tried the algorithm described below on many transfer functions (including
high-pass and bandpass, and non-rational weights) and found no unexpected problems.

2.9.2 Numerical realisation algorithm


Quite remarkably, the equations above are actually a practical way of calculating a
realisation. The transfer functions and operators have to be evaluated at a set of discrete
frequency points to reduce the problem to a nite number of dimensions. For a lter with a
cuto frequency around 1 rad=s, a good choice might be

!1 = ;10 : 0:1 : 10
!2 = ;9:95 : 0:1 : 9:95:
(This notation means that !1 = ;10; ;9:9; ;9:8;    ; 9:9; 10 etc.) We then calculate the
201  200 matrix

H = (j!1 ) H (j! 1 ) ; H (j!2 ) (j! ):


j (! ; ! ) 2
2 1
Note that we avoided !2 = ;10 : 0:1 : 10 to eliminate problems at !1 = !2 . A alternative
approach is to calculate (j!) dd(Hj!
(j!) (j!) for this diagonal.
)
Calculating the singular value decomposition of H is a standard problem in numerical
linear algebra (Golub & Van Loan 1996). In Matlab, the routines svd or svds can be used.
The routine svds produces only the n largest singular values and vectors, where n can be
chosen. The resulting U and V are rectangular. If these are used, the realisation produced
automatically only has order n. (For the example which follows in section 2.9.4, U is
200  6, V is 201  6, and S is 6  6.)

The matrices U , S and V which result are not quite what we need. Firstly, U and V are
only determined up to an arbitrary phase shift, which must be removed. This can be done

61
by calculating the unitary diagonal matrix
q
M = U T (j!)U (;j!)
and substituting
U UM ;1
V V M :
This substitution ensures that the realisation is real. Secondly, we need to normalise to
satisfy (2.14) and (2.15) rather than UU  = V V  = I . This is readily accomplished by
r
U 2 U
r !
V 2
! V
S ! S
2
where ! is the frequency step { here ! = 0:1.

F (j!) and G(j!) can then be calculated using (2.17) and (2.18). To determine D, C
and B using (2.19), (2.20) and (2.21), the author uses the approximations
D = 21 [H (;j 10) + H (j 10)]
C = 12 [;j 10 G(;j 10) + j 10 G(j 10)]
B = 12 [;j 9:95 F (;j 9:95) + j 9:95 F (j 9:95)] :
There are presumably more accurate ways of doing this calculation. The matrix A is
calculated using the Riemann sum to approximate the integral (2.22).

2.9.3 Properties of the optimal realisation


Using (2.17), (2.18), (2.14) and (2.15), and the de nition of the Gramians (2.10) and (2.11),
1 Z1
K = 2 [F (j!)(j!)][F (j!)(j!)] d!
;1
Z1
= 21 V  (j!)V (j!) d!
;1
= I
62
and
1 Z1
W = 2 [ (j!)G(j!)] [ (j!)G(j!)] d!
;1
Z1
= 21 S  U  (j!)U (j!)S d!
;1
= S2 S 3
 21 0 0   
66 7
66 0 22 0   777
= 66 7
4 6 0 0 32   77 5
   
:
The realisation will be called the frequency-weighted input-normal realisation. \Input
normal" means K  = I . In reading other literature, a little care should be taken as other
input-normal realisations which do not have diagonal W exist.

The frequency-weighted input-normal realisation has gains to states and noise gains
which are uncorrelated under the frequency weights (j!) and (j!). As the eigenvalues of
p
a diagonal matrix are the diagonal elements, the Hankel singular values i (K  W ) are
clearly the i . The realisation is optimal in the sense that it minimises a measure of
Q
dynamic range kii wii . Recall that kii is the frequency-weighted power gain to the ith
i
state, and wii is the frequency-weighted noise gain from the ith state to the output. For
digital lters, this can be shown to minimise the total number of bits required to store the
state to any given accuracy (Mullis & Roberts 1976). This fact will be used in an
interesting calculation in chapter 5.
P
For analog lters, the correct measure we want to minimise is probably wii subject to
kii  1 rather than Q kiiwii . However, the frequency-weighted input-normal realisation is
i
i
known to be fairly close to optimum in this measure also. It is possible to nd a true
P
minimum kii wii by brute-force optimisation, but the realisation does not have any other
i
conceptual signi cance.

There is a related realisation, called the frequency-weighted balanced realisation, which

63
has
2 3

66 1 0 0    77
6 0 2 0   77
K  = W = 666 7:
64 0 0 3   775
   
It was originally found by Enns (1984) and Thiele (1986). It is trivial to obtain this from
the input-normal realisation by diagonal scaling, so these authors were the rst to derive
the frequency-weighted input-normal realisation presented here. (They used a di erent
algorithm { see section 2.11.4.) The frequency-weighted balanced realisation is useful for
model reduction as states with small Hankel singular values i can be deleted without
a ecting the transfer function much.

The dynamic range limitation can be used to derive a frequency-domain formula for the
P
weighted output noise power wii achieved by the input-normal realisation. As K  = I ,
i
X
wii = tr K W : (2.23)
i
(The trace tr is the sum of the diagonal elements, which also equals the sum of the
eigenvalues.) From the de nitions of the Gramians,

tr K  W1
1 Z 1 Z1
= tr 2 [F (j!)(j!)][F (j!)(j!)] d!  2 [ (j!)G(j!)] [ (j!)G(j!)] d!

;1 ;1
1 Z Z tr[F (j! )(j! )][F (j! )(j! )] [ (j! )G(j! )] [ (j! )G(j! )] d! d! :
1 1
= 4 2 2 2 2 2 1 1 1 1 1 2
;1 ;1
Then using the commutivity of the trace, tr XY = tr Y X , with X = [F (j!2 )(j!2 )],

tr K  W
1 Z1 Z1 
= 4 2 tr [F (j!2 )(j!2 )] [ (j!1 )G(j!1 )] [ (j!1 )G(j!1 )][F (j!2 )(j!2 )] d!1 d!2
;1 ;1
Z1 Z1 
= 41 2 tr [ (j!1 )G(j!1 )F (j!2 )(j!2 )] [ (j!1 )G(j!1 )F (j!2 )(j!2 )]g d!1 d!2
;1 ;1
Z1 Z1  1 ) ; H (j!2 ) (j! )] [ (j! ) H (j!1 ) ; H (j!2 ) (j! )] d! d! :
= 41 2 tr [ (j!1 ) H (j!
j (! ; ! ) 2 1 j (! ; ! ) 2 1 2
;1 ;1
2 1 2 1

64
This is really just the dynamic range limitation as a weighted integral over input and
output frequency. This is important, because it shows that the dynamic range as limited by
the dynamic range limitation is in some sense achievable, namely by the frequency-weighted
input-normal realisation. For readers unfamiliar with matrix transfer functions, the
simpli cation for scalar (single-input, single-output) transfer functions may be easier to
comprehend:
1 Z1 Z1 H (j! ) ; H (j! ) 2
tr K  W = 42 j(j!2 )j j (j!1 )j j (!1 ; ! ) 2 d!1d!2
2 2
;1 ;1
2 1

2.9.4 Example of frequency-weighted input-normal realisation


As an example, we realise the same 1 rad=s low-pass 5th-order Butterworth transfer
function as in section 2.5.1. We use weightings
s)2
(s) = (1(1+ +0:25s)2
(s) = 1:
These are quite realistic for a radio receiver. The input weight (s) has two real zeros at
1 rad=s and two real poles at 4 rad=s. This represents an input signal which rises from 0 dB
in-band to +24 dB out-of-band. Output noise is considered equally important at all
frequencies.

As the transfer function is fth-order, only ve Hankel singular values should be


non-zero. However, we decide to keep six to investigate how accurate the algorithm is. The
magnitudes of the transfer function H (j!), gains to states F (j!) and noise gains G(j!) are
shown in g. 2.5. A wide frequency and amplitude range are shown, so as to show
imperfections.

The high-frequency error in the realised transfer function is due to the frequency grid
only extending up to 10 rad=s, leading to inaccurate values of B , C and D. Interestingly,
the number of points does not need to be large for an accurate transfer function { a 10  9
matrix gives the same transfer function as a 201  200 matrix. (The gains to states and
noise gains are badly suboptimal though.)

The F (j!) for states 1{5 can be seen to be sensibly scaled. The increased input
65
20
realised
Transfer function H (dB)

0 prototype
20

40

60

80

100
2 1 0 1 2
10 10 10 10 10
Frequency (rad/s)
20
f1
0
Gain to states F (dB)

f2
f3
20 f4
f5
40 f6
60

80

100
2 1 0 1 2
10 10 10 10 10
Frequency (rad/s)
20
g1
0
Noise gain G (dB(s))

g2
g3
20 g4
g5
40 g6
60

80

100
2 1 0 1 2
10 10 10 10 10
Frequency (rad/s)

Figure 2.5: Transfer function, gains to states and noise gains of the \optimal" realisation of a fth-
order Butterworth response. The sixth state (black) is only included for illustrative purposes.

66
weighting above 1 rad=s lowers the gains to states there, although the fallo in the transfer
function also contributes. The magnitudes of the noise gains G(j!) decrease in accordance
with the decreasing Hankel singular values as we move from g1 to g5 . The rst state has a
higher noise gain than in the LC ladder realisation. This is a consequence of the gains to
states being lower out-of-band than those of the LC ladder realisation. (Compare g. 2.5
and g. 2.4.)

The behaviour of the sixth state is interesting. We would expect the noise gain g6 to be
zero as the sixth frequency-weighted Hankel singular value should be zero, and indeed it is
quite small. It is not zero because of rounding and numerical routine inaccuracies, and
because of the two approximations in the algorithm { the nite number of frequency points,
and the calculation of B (and indirectly, A), C and D from values at 10 rad=s rather than
1. The error varies considerably from run to run, which suggests that the Matlab singular
value decomposition algorithm svds uses iteration with a random initial guess.
Approximating a rational transfer function by a higher-order rational transfer function
nearly always produces almost coincident pole-zero pairs, and this algorithm is no exception.
Here the pole-zero pair is very close to the origin, resulting in the low-frequency peak in f6 .

In this and other examples, the algorithm worked reliably when used with a sensible
matrix size and frequency range. With 201  200 matrices the calculations takes a few
seconds. It should be noted that existing matrix-based approaches (Zhou 1996, p. 163)
(Anderson & Moore 1989, problem 10.2-1) (Enns 1984) (Thiele 1986) are faster and more
accurate if state-space realisations of the transfer function and weights are available. This
other approach is brie y introduced in section 2.11.4.

2.10 TRANSFER FUNCTION OPTIMISATION


What transfer function optimises the dynamic range, subject to the dynamic range
limitation? In this thesis, we consider this for two cases, lters and ampli ers. Filters are
deferred to the next chapter. The brief summary for lters is that (2.9) seems to summarise
the situation { the achievable dynamic range is limited by how close the stopband is to the
passband and little else. There is surprisingly little di erence between, say, an Elliptic and

67
a Chebyshev response. (Sensitivity depends a lot more on transfer function than dynamic
range though, as will be shown in the next chapter.) Ampli ers are discussed in this section.

2.10.1 Ampli er transfer function


If a transfer function is constant, (2.8) is zero and the dynamic range is unlimited. For a
lter, obviously H (s) = c is useless. For an ampli er, H (s) = c with jcj > 1 would be nice,
but it is not possible because of the limited speed of the transistors. At frequencies above
fmax, the power gain must be less than 0 dB (Thornton, DeWitt, Chenette & Gray 1966, p.
114). We will investigate the example of an ampli er with a low-frequency power gain of
20 dB. The transfer function must therefore vary with frequency, and hence the dynamic
range limitation (2.8) must be non-zero.

In most applications, maximum signal and noise and distortion matter more at some
frequencies than others. The weights (s) and (s) and the Gramians were introduced to
deal with this problem, but for simplicity we work directly from the dynamic range
limitation in this section. We will assume that the aim is to maximise dynamic range over a
band starting at DC and extending up to a frequency !P . Transistor fmax and other
considerations (fT , load capacitance etc.) require that the gain of the ampli er fall to 0 dB
by !C .

According to (2.8), we should minimise the magnitude of [H (j!1 ) ; H (j!2 )]=j (!2 ; !1 )
over DC to !P . A straightforward way of doing this is to require H (j!) to be close to
constant over that frequency range. Note that H (j!) needs to be constant in phase as well
as magnitude { this is not the standard lter approximation problem. The target was
chosen as H (j!) = 10 + j 0. (The imaginary part has to be zero to be realisable at DC.)

The optimisation as described above tends to produce an H (j!) which has a large peak
between !P and !C . This is likely to be intolerable in a practical ampli er, so a constraint
on maximum jH (j!)j needs to be imposed. jH (j!)j < 14 was used here, limiting peaking to
3 dB. (It may be desirable to limit the magnitude of [H (j!1 ) ; H (j!2 )]=j (!2 ; !1 ) as well.
However, the peaking constraint alone gave reasonable results.)

68
The transfer function H (j!) also needs to be stable and causal. This is obvious but
mathematically important.

To summarise, the problem is:

Minimise (the maximum value of) jH (j!) ; 10j, 0  !  !P


for H (j!) stable and causal and subject to jH (j!)j < 14 for all ! and
jH (j!)j < 1 for ! > !C .

There are a variety of ways to solve this optimisation problem. It can be formulated as
trying to nd a causal approximation to the \centre"
8
>
<10; j!j < !P
C (j!) = >
:0; j!j  !P
within an error bound radius
8
>
> k; j!j < !P
>
<
R(j!) = >14; !P < j!j < !C
>
>
:1; j!j  !C
where we wish to minimise k. C (j!) is clearly non-causal as a lter with the same
magnitude response would have an associated phase lag in accordance with the Bode (1945)
gain-phase relationships. C (j!) as it stands has zero phase. In this formulation the
problem basically comes down to making a weighted (by R(j!)) causal approximation to
C (j!). It is discussed at length by Helton (1981). (The problem of making such causal
approximations is called the Nehari problem by mathematicians.) Fortunately, Helton &
Merino (1998) have written Mathematica procedures to solve the problem.

There are other solutions. Brute-force approximation, say over a space of FIR transfer
functions, can be quite practical. (The discrete-time FIR responses can be converted to
continuous-time by a bilinear transformation.) As we shall see, optimising the loop gain of a
feedback ampli er also basically solves the same problem. However, Merino and Heltons'
Mathematica routines seems to be faster and more consistent, so we use them. They also
give an indication of how close to optimum the approximation achieved is. The output of
69
these routines is the optimum frequency response H (j!) over a few hundred frequency
points.

Typical results H (j!) of the optimisation are shown on the left in g. 2.6. These (the
blue, green and red curves) are for !P = 1 rad=s and !C = 5, 10, 20 rad=s. It may occur to
the reader that the curves look like the closed-loop frequency responses of feedback
ampli ers. (This resemblance is pretty feeble at present; they looked more like feedback
ampli ers in earlier optimisation runs before Helton's superior algorithm was used.) They
are very at and have little phase shift in-band (as H (j!) ' 10 + j 0 for ! < !P ), and have
3 dB of (peculiar-shaped) peaking out-of-band (as constrained by jH (j!)j < 14). This
makes sense, for we suspect that feedback ampli ers have high dynamic range. The signal
transfer function of a high-order single-loop sigma-delta modulator also has a similar shape.

If the responses were those of feedback ampli ers, it is reasonable to ask what the
open-loop gain would be. This can be calculated from

H (s) = 1 +LkL
(s) ) L(s) = H (s)
(s) 1 ; kH (s)
where L(s) is the opamp gain and k is the feedback factor. (\Opamp gain" just means
open-loop forward gain here.) As the ampli er has 20 dB gain, we use k = 0:1. This results
in the opamp gains L(s) on the right of g. 2.6. It is certainly conceivable that an
\optimum" feedback ampli er could have such an opamp gain. It may be easier to think of
rad/s as GHz { then these are opamps compensated for a closed-loop gain of 10 over 1 GHz
bandwidth, where the opamps have 5, 10, 20 GHz unity-gain bandwidth.

The phase margins are around 50 { remember that the loop gain falls to one where the
opamp gain is 20 dB as k = 0:1. This phase margin is primarily due to limiting peaking to
3 dB. More peaking allows more loop gain below !P at the expense of reduced phase
margin. For !C = 10 rad=s, doubling the peaking to 6 dB allows 77 dB opamp gain below
!P = 1 rad=s, as compared to 60 dB for the green curve. The at low frequency gain is
obviously the result of optimisation. The higher it is, the closer H (j!) is to 10 + j 0.
Gratuitous peaks in the low-frequency loop gain tend to get smoothed out as they degrade
the phase margin in accordance with the various Bode gain-phase relations (Bode 1945).
The responses are conditionally stable, albeit only just for the blue curve. This is normally

70
80 80

60 60
H(s) gain (dB)

L(s) gain (dB)


40 40

20 20

0 0

20 20
1 0 1 1 0 1
10 10 10 10 10 10
Frequency (rad/s) Frequency (rad/s)

100 100

0 0
H(s) phase (deg.)

L(s) phase (deg.)

100 100

200 200

300 300

400 400
1 0 1 1 0 1
10 10 10 10 10 10
Frequency (rad/s) Frequency (rad/s)

Figure 2.6: Optimised transfer functions H (s) and corresponding open-loop gains L(s). Blue, green,
red: !C = 5 rad=s, 10 rad=s, 20 rad=s. Cyan: one-pole rollo above !C = 10 rad=s.

71
avoided in ampli ers, but it is known to allow greater low-frequency loop gain (Oizumi &
Kimura 1957) and is used in high-order single-loop sigma-delta modulators.

The assumption that an ampli er can have a gain very close to unity at all frequencies
above !S may be suspect. For this reason, the cyan curve in g. 2.6 is included. In this
case, jH (j!)j was constrained to fall o at a one-pole rate at high frequency. It can be seen
that this reduces the in-band opamp gain L(j!) from 60 dB (the green curve) to 49 dB (the
cyan curve). Constraining gain to fall o at a two-pole rate causes a further deterioration to
around 40 dB low-frequency opamp gain, ie. 20 dB of feedback. It may be of some
signi cance that the very fast opamps in the next chapter have a gain which falls o at only
a one-pole rate at high frequencies, due to a feedforward stage. Perhaps other
\feedforward" techniques such as gain-boosted cascodes also have wide bandwidth because
they reduce the order of the high-frequency asymptotic gain rollo ?

2.10.2 What is feedback?


The optimisations in the section above and the feedback ampli er example in section 2.5.2
provide a distinctive view of feedback circuits. It seems that a feedback circuit \works" by
minimising the dynamic range required of its active components within a passband (the
frequencies where loop gain is high). The following comment from Bode (1945, p. 47)
indicates that this point has been known for a long time:

\: : : the usual conception of feedback includes two distinct ideas. The rst is
that of a loop transmission or return of voltage, and the second that of a
reduction in the e ects of variation in the tube characteristics. In normal
circuits these two are related by simple mathematical laws so that the term
\feedback" can refer generically to both.
In exceptional circuits, when the correlation between the two breaks down, the
rst idea is evidently the one which most nearly agrees with the usual physical
conception of feedback. It will therefore be taken as the basis for feedback in the
general case."

72
But what if we abandon the rst idea, and work entirely from the second? The
conventional de nition of loop gain using return di erence and signal ow graphs
(Bode 1945) (Mason 1954) is somewhat cumbersome. It would be better if the de nition
could be related back to nodal analysis or some other unambiguous circuit description. If
active circuit synthesis was suciently developed, could we synthesise \feedback" circuits
by minimising F (j!)G(j!) over the desired frequency range rather than by creating a
feedback path?

Compensation
The dynamic range limitation (2.8) shows that in a circuit with restricted bandwidth, at
least one state must be stored with decent dynamic range. Looking back to the feedback
ampli er example (section 2.5.2), we see that the state x4 , the voltage across the
compensation capacitor, serves this function. The other three states all have reduced
dynamic range requirements. Thus, instead of using a wide bandwidth and optimising the
dynamic range (as in the optimisation example, section 2.10), we also have the option of
making a distinction between states associated with transistors and hence distortion, and
states associated with passive components. The dynamic range only needs to be minimised
for the \transistor" states.

As compensation using linear capacitors is widely used, it would be useful to have an


optimisation method which could handle two classes of states { low dynamic range
\transistor" states and high dynamic range \capacitor" states. However, the author does
not have a method for doing this at present. If one is developed, transistor-level active
circuit synthesis becomes a real possibility.

2.11 LINKS TO OTHER LITERATURE


This chapter takes a distinctive frequency-domain approach to the state-space dynamic
range theory. The key is to use two frequency variables !1 and !2 and the dynamic range
limitation. The rst priority of this chapter is to present this theory without diversion, as a
combination of new concepts and new mathematics is usually dicult to understand.

73
However, the approach based on the dynamic range limitation did not come from
nowhere { to do this would be a task of similar magnitude to the original development of
state-space theory by Kalman, Bellman and others in the 1960's. Essentially, the author
came up with the dynamic range limitation, and some months later derived that it was the
Laplace transform of the so-called Hankel operator of time-domain realisation theory. This
existing theory was then used as a guide to developing the realisation in section 2.9. This
section introduces the existing theory so that new work can be distinguished from old.

2.11.1 The Hankel operator


The dynamic range limitation is the frequency-domain representation of the Hankel
operator. This does not seem to be existing knowledge. We need to start by mentioning
what an operator is.

We can think of an input or output signal as an in nite vector consisting of the values at
all the di erent times or frequencies. An operator is an in nite matrix which maps the
input to the output. (By operator we mean linear operator.) Mathematicians take a dislike
to the term \in nite matrix" as the key to operator theory is to separate the operators from
their matrix representation in a particular basis, and to consider nite-dimensional and
in nite-dimensional quantities in similar ways. A function is like a diagonal operator. A
time-domain convolution is a good example of a non-diagonal operator. A Fourier or
Laplace transform is also an operator. Note the analogy between a discrete Fourier
transform (DFT), which is a nite matrix, and the continuous Fourier transform which is
an in nite matrix or operator. As an aside, note that the Laplace or Fourier transform
diagonalises the convolution operator, giving the transfer function. This is why the
frequency domain is so useful for analysing linear time-invariant systems.

The Hankel operator H is the operator which maps past inputs to future outputs. A
standard reference on it is (Glover 1984). For a time-invariant system with impulse
response h(t),
Z0
(Hu)(t) = h(t ;  )u( )d; t > 0
;1

74
It is convenient to de ne the input to the Hankel operator as v(t) = u(;t), t > 0 rather
than u(t), t < 0 (Glover 1984). This is useful here as we are going to Laplace transform the
above expression, and the Laplace transform L is de ned as
Z1
V (s) = (Lv)(s) = e;st v(t)dt
0
which ignores t < 0. Thus we use
Z1
(Hv)(t) = h(t +  )v( )d; t > 0:
0
(The Hankel operator is still only de ned for output time t > 0; it is the input we have
time-reversed, not the output.)

We will need expressions for the impulse responses of the transfer function, gain to
states, and noise gain. These are readily derived from the state-space equations or from the
inverse Laplace transforms of H (s), F (s) and G(s) to be

h(t) = CeAtB + D(t)


f (t) = eAt B
g(t) = CeAt :
For simplicity we have assumed zero initial conditions. (t) is a unit impulse at time t = 0.
We assume D = 0 for now, although it seems possible to neglect this. Therefore
Z1
(Hv)(t) = CeA(t+ ) Bv( )d
0
Z1
= CeAteA Bv( )d
0
Z1
= CeAt eA Bv( )d
0
Z1
= g(t) f ( )v( )d:
0
This can be Laplace transformed to give
1 Z 1
(HV )(j!1 ) = 2 G(j!1 ) F (j!2 )V (;j!2 )d!2
;1
75
using the generalised Parseval's formula on the integral, and with !1 the frequency variable
for t, and !2 the frequency variable for  . (Doetsch (1974, p. 211) gives a good discussion of
Parseval's formula for Laplace transforms.) Then using the dynamic range limitation (2.8),
Z1 H (j!1 ) ; H (j!2 )
(HV )(j!1 ) = 21 j (!2 ; !1) V (;j!2 )d!2: (2.24)
;1
We see that the dynamic range limitation is the kernel of the Laplace-transformed Hankel
operator. This is why we were justi ed in calling the singular values of the dynamic range
limitation Hankel singular values.

2.11.2 De ning a frequency-weighted Hankel operator


In light of (2.24), it seems natural to de ne a frequency-weighted Hankel operator
1 Z 1
(HV )(j!1 ) = 2 (j!1 )G(j!1 ) F (j!2 )(j!2 )V (;j!2 )d!2 (2.25)
;1
Z1 1 ) ; H (j!2 ) (j! )V (;j! )d! :
= 21 (j!1 ) H (j!
j (! ; ! ) 2 2 2 (2.26)
;1
2 1
The next section shows that this quantity has the same properties as the unweighted
version.

2.11.3 Controllability and observability operators


Recall that the Hankel operator H maps past inputs to future outputs. The reason that it
is a useful quantity is that it can be factorised into a map from past input to current state
and a map from current state to future output (Glover 1984) (Zhou 1996). This step is
important for state-space realisation. These maps are known as the controllability operator
C and the observability operator O respectively. They are known to be given by (Zhou 1996)
Z1
C (v(t)) = eA Bv( )d
0
Z1
= f ( )v( )d
0
(Ox)(t) = CeAt x
= g(t)x
76
It is straightforward to Laplace transform these expressions, but the author has never
seen it done. Presumably, without the dynamic range limitation there is little reason to do
so. The result is
Z 1
C (V (j!)) = 21 F (j!)V (;j!)d!
;1
(OX )(j!) = G(j!)X :

In light of this, it is natural to de ne frequency-weighted controllability and observability


operators
1 Z 1
C (V (j!)) = 2 F (j!)(j!)V (;j!)d!
;1
(OX )(j!) = (j!)G(j!)X :

These expressions satisfy the same properties as the unweighted time-domain expressions,
which are (Zhou 1996)

H = OC
K  = CC 
W = O O:

2.11.4 Optimum dynamic range realisation


Section 2.9 described a new approach for calculating \optimal" realisations. Here we brie y
describe what existed before. There were two main approaches. One was based on
similarity transformation of an existing state-space realisation. This could deal with
frequency weightings. The other was closer to that in section 2.9, but used the singular
value decomposition of the Hankel operator in the time domain. This makes it more
dicult to apply frequency weighting, and the author has not seen a frequency-weighted
time-domain algorithm.

Similarity transform approach


We saw in section 2.8.2 that the di erent minimal state-space realisations of a transfer
function are parameterised by a similarity transformation T . The traditional approach to
77
optimum dynamic range realisation is to take any realisation, good or bad, and nd a T
which transforms it into an optimal one.

For any realisation, there exists a T which transforms it into one with the Gramians K 
and W diagonal. (The T of course depends on the realisation you start with.) The
existence of such a similarity transformation which simultaneously diagonalises two
positive-real symmetric matrices is a moderately well-known piece of linear algebra.

The desired transform T can be produced using a Cholesky decomposition


(upper-triangular R)

R R = W ; (2.27)

a Schur decomposition (unitary U , upper-triangular S )

USU  = RK R; (2.28)

and two square roots and an inverse,

T = S ;1=4 U R: (2.29)

This is taken from (Anderson & Moore 1989, problem 10.2-1). Thiele (1986) gives a similar
procedure. This T actually results in the frequency-weighted balanced realisation rather
than the frequency-weighted input-normal realisation, but this is easily corrected by
diagonally scaling the signal levels.

Time-domain Hankel singular value decomposition approach


That a state-space realisation could be derived by factoring the Hankel operator into
controllability and observability operators has been known for a long time. (In all practical
algorithms operators are represented by nite matrices.) The original algorithm is by Ho &
Kalman (1966). Ho and Kalman used a di erent matrix factorisation. The singular value
decomposition was later found to be convenient by Zeiger & McEwen (1974) and Kung
(1978).

Once the Hankel operator is factorised as H = OC , the matrices A, B , C and D are


uniquely determined and can be calculated. The time-domain procedures for doing this are
78
of a similar nature to the frequency-domain formulae in section 2.9.1, but involve matrix
shifts and the like.

In the time domain, the Hankel operator does not have an obvious frequency-weighted
generalisation. Thus previous authors do not seem to have been extended the Ho-Kalman /
singular value decomposition approach to the frequency-weighted case. This is the value of
the frequency-domain approach presented in this chapter.

2.11.5 Sensitivity
Sensitivity is how much the transfer function varies when a component value varies.
Dynamic range is closely linked to sensitivity. This is not surprising, because a component
with a changed value can always be represented as the original component plus a noise.
Deriving the change in transfer function is quite straightforward. For changes in B , C and
D,
HB (s) ; H (s) = [C (sI ; A);1 (B + B ) + D] ; [C (sI ; A);1 B + D]
= C (sI ; A);1 B
= G(s) B
HC (s) ; H (s) = [(C + C )(sI ; A);1 B + B ] ; [C (sI ; A);1B + D]
= C (sI ; A);1 B
= C F (s)
HD (s) ; H (s) = [C (sI ; A);1 B + D + D] ; [C (sI ; A);1 B + D]
= D:

For changes in A, we use a derivation remarkably like that for the dynamic range limitation:

M = (sI ; (A + A));1 ; (sI ; A);1


(sI ; (A + A))M (sI ; A) = (sI ; A) ; (sI ; (A + A))
= A
M = (sI ; (A + A));1 A(sI ; A);1

79
Then

CMB = [C (sI ; (A + A));1 B + D] ; [C (sI ; A);1 B + D]


= HA(s) ; H (s)

but also

CMB = C (sI ; (A + A));1 A(sI ; A);1B


= GA (s) A F (s):

Combining all these results for in nitesimal changes,

dH (s) = G(s) dA F (s) + G(s) dB + dC F (s) + dD:

This result is well known in a less elegant subscripted form. Other authors (Thiele 1986)
(Lutz & Hakimi 1988) (Snelgrove & Sedra 1986) calculate derivatives instead of using a
resolvent identity.

Note that whereas the dynamic range limitation uses the gain to states and noise gain at
two di erent frequencies, the sensitivity formulae use them only at one frequency. This is
because a parameter change produces an output \noise" at the same frequency as the input.
For this reason, there is more opportunity to optimise sensitivity than dynamic range by
appropriate choice of the transfer function. This consideration will be discussed further in
the next chapter.

2.11.6 Sensitivity theory and other related previous work


To the best of the author's knowledge, the dynamic range limitation (2.8) is new. The
closest result the author has seen is

G(s)F (s) = dHds(s) ;

derived by Snelgrove & Sedra (1986). This is a corollary of the dynamic range limitation for
s1 ! s2. Note that this is the case which is relevant for sensitivity analysis. Historically,
sensitivity has been a more important problem than dynamic range. There was great
interest in the sensitivity problem in the late 1960's and early 1970's, and it would not be
80
surprising if the more general dynamic range limitation was known in this context. The
author has read through the IEEE Transactions on Circuits and Systems for this period
and cannot nd the dynamic range limitation (even in sensitivity rather than state-space
notation).

The dynamic range limitation may not have been discovered by sensitivity theorists
because the resolvent identity was not widely known. However, the resolvent identity is
widely used in functional analysis (Dunford & Schwartz 1964, p. 568). Zames (1981)
mentions the use of resolvent identities for sensitivity analysis. It would not be at all
surprising if Zames had derived the dynamic range limitation too, considering his depth of
knowledge of functional analysis and control system robustness.

Livsic (1973, p. 29) uses a similar identity to derive some properties of lossless circuits,
and to develop a variant of Darlington synthesis. Livsic's S (!) and R(!) seem to be the
same as H (j!) and F (j!) here. Livsic does not use a noise gain G(j!).

2.11.7 Energy storage, sensitivity and group delay


Some understanding of the relationships between transfer function and dynamic range can
be acquired by other means. In particular, the expressions which relate the stored energy in
a lossless network to the group delay (Kishi & Nakazawa 1963) (Smith 1965) constrain
F (j!) (Harrison & Weste 2001). Sensitivity has also been previously related to group delay
and stored energy (Kishi & Kida 1967) (Orchard et al. 1985) (Temes & Orchard 1973). It is
also possible to relate G(j!) to F (j!) using reciprocity (Harrison & Weste 2001). However,
the work presented here seems to give a more useful limitation, and is not limited to lossless
reciprocal systems.

2.12 CONCLUSION
This chapter has presented a means of calculating the dynamic range of an active lter.
The main bene t of the theory presented here is that it allows dynamic range and
sensitivity to be taken into account when choosing the lter transfer function. Of course,

81
previous approaches such as checking the maximum pole Q provide some insight, but the
dynamic range limitation and Gramians o er a rigorous approach which can cope with
transfer functions outside a designer's experience.

There is a natural state-space structure implicit once the transfer function is chosen. Its
dimensions are found from the singular values of the Hankel operator. In the
frequency-domain, the Hankel operator has kernel [H (j!1 ) ; H (j!2 )]=j (!2 ; !1 ), which is
also equal to the product of the gain to states F (j!2 ) and the noise gain from the states
G(j!1 ). The frequency-weighted input-normal form uses this natural structure to achieve
high dynamic range.

The state-space dynamic range theory also provides some insight into feedback ampli ers.
Feedback can be thought of as a combination of a at transfer function which allows high
dynamic range according to the dynamic range limitation, and a good realisation of this
transfer function. Opamps do not have quite as neat a state-space description as lters, so a
little more work is needed before they can be synthesised to optimise dynamic range.

82
3 A 350MHZ OPAMP-RC FILTER

3.1 INTRODUCTION
To further investigate dynamic range considerations for deep-submicron CMOS circuits, an
active lter was designed in a 0.18 m process. The transfer function is a fth-order Elliptic
low-pass which would be useful as an anti-alias lter (such as in a broadband wireless
LAN's baseband circuitry). The choice of the transfer function is a further example of the
use of the dynamic range limitation, and one of real practical merit.

It would be conceptually elegant to use the frequency-weighted input-normal realisation


discussed in the previous chapter. However, a more conventional realisation based on
simulating an LC ladder was used. This has advantages with respect to sparsity of the tap
matrix, as will be discussed in section 3.3.

The lter uses opamps as the active element. This is a somewhat radical choice for cuto
frequencies above 100 MHz; conventional wisdom would be to use a gm-C approach. An
opamp-RC approach was tried as feedback circuits seemed promising for achieving large
signal swings with low distortion. The theory in section 2.10 also suggested that feedback
ampli ers had not been pushed to their bandwidth limits.

Resistors are switched in or out to tune the lter. (Unlike switched-capacitor circuits,
continuous-time lters need to tune out the e ects of resistor and capacitor absolute
tolerance.) The general reason that resistors were preferred to a MOSFET-C approach is
the better linearity with high signal swings. In systems on a chip, having switched rather
than continuous tuning is not a problem as the tuning loop is usually digital and can easily
deal with switched elements.

83
3.1.1 Organisation of the chapter
The conventional design process for a lter involves three key choices { choice of transfer
function, choice of realisation, and choice of integrator (or resonator or biquad). Section 3.2
discusses choosing the transfer function. The dynamic range limitation is used. The section
summarises and uses some of the theory in the previous chapter, and represents a good
chance for the reader to further understand the use of that material. Section 3.3 discusses
the realisation. (Note that in lter design, \realisation" basically means the state-space
equations, not the transistor-level design.)

Section 3.4 describes the opamp. An opamp-based approach was chosen a-priori on the
grounds that gm-C lters have already received plenty of attention by many distinguished
authors. A comparison of the relative merits of the gm-C and opamp-RC approaches is
deferred to the results section (3.5). The opamp work is the major new contribution in this
chapter. The basic problem is to achieve sucient loop gain at high frequencies.

After the measured results on the fabricated 350 MHz lter, the chapter nishes with
some further simulations to see if even greater bandwidth and signal swing is possible. It is,
and as the fabricated lter is already state-of-the-art, the opamp-RC approach looks
promising indeed.

3.2 CHOICE OF TRANSFER FUNCTION


Choosing the lter transfer function is a trade-o between the passband and stopband
speci cations on one hand, and the dynamic range and sensitivity on the other. A basic
outcome of the previous chapter is that the transfer function H (j!) should be chosen to
minimise j[H (j!1 ) ; H (j!2 )]=j (!2 ; !1 )j. Some input signal frequencies !2 and output
noise frequencies !1 are more important than others, namely the !2 where the input signal
is large or the !1 where the output noise needs to be small. In section 2.11.5, sensitivity
was shown to depend on values along the line !1 = !2 , ie. on jdH (j!)=d(j!)j.

84
3.2.1 Transfer function optimisation
A logical approach would be to nd the transfer function H (j!) which minimises
Z1 Z1
tr K  W = 1 j (j! ) j 2 j (j!1 )j2 H (j!1 ) ; H (j!2 ) d!1 d!2
42 2 j (!2 ; !1)
;1 ;1
in accordance with the theory in section 2.9.3. The transfer function must of course be
constrained to meet the passband and stopband speci cations and to be causal and stable.
Recall that j(j!)j2 is the input power spectrum, and j (j!)j2 is the weight to be put on
the output noise. For an anti-alias lter in a wireless LAN receiver, j(j!)j2 might be
chosen as 0 dB in-band and +10 dB out-of-band to cope with 10 dB larger blocking signals.
The noise weight j (j!)j2 might be chosen as 0 dB in-band and for frequencies which the
ADC aliases down into the passband, and {10 dB for other frequencies. (It would probably
be unwise to go below {10 dB, as the weights a ect sensitivity versus frequency as well as
dynamic range.)

The most straightforward way to minimise tr K  W would be to numerically optimise


some parameterisation of H (s). This was tried for rational H (s) and for \FIR" H (s) with
(1 ; s)=(1 + s) in place of z ;1 as the delay element. The results were mixed. It seems fairly
clear that the optimum is a transfer function with a rounded passband corner and not more
stopband attenuation than is necessary. The results often look fairly similar to an
inverse-Chebyshev response or an Elliptic with low passband ripple, sometimes with the
stopband zeroes moved a little into the left half-plane. It is not possible to say much more
than this, because the optimisation tends to have local minima and re ects the
parameterisation chosen.

Values of tr K  W within 0.5 dB of the optimised results were possible by just choosing
an Elliptic response with 0.01{0.1 dB passband ripple, and stopband attenuation and order
to just meet the speci cation. Such a response, namely a 5th order Elliptic with 0.01 dB
passband ripple and {49 dB stopband, was chosen for the lter described in this chapter.

The optimal response shape is largely independent of the weightings j(j!)j2 and
j (j!)j2 .
Realising stopband zeroes can be cumbersome, so it was checked whether an all-pole
85
response such as a Chebyshev transfer function could achieve similar performance. For
weightings as above, the general result was that the best Chebyshev transfer functions had
tr K  W about 1.5 { 2 times that of the low-ripple Elliptic responses. This would result in
2{3 dB loss in dynamic range. The best Chebyshev responses had moderate ripple,
0.01{0.3dB.

3.2.2 Graphical examination of dynamic range limitations


To compare di erent transfer functions, it can be useful to plot the quantity in the dynamic
range limitation (the Hankel operator kernel) j[H (j!1 ) ; H (j!2 )]=j (!2 ; !1 )j. Fig. 3.1
shows this for an Elliptic and a Chebyshev transfer function with the same transition
bandwidth. The Elliptic response is that used in the lter described in this chapter. Both
lters are normalised to {3 dB at 1 rad=s and are {49 dB at 1.63 rad=s.

Recall that sensitivity is related to performance along the line !1 = !2 , as a parameter


variation generates a \noise" at the input frequency. Blocking performance is for passband
!2, and stopband !1. These regions are shown on g. 3.1. The two frequency responses
have almost identical dynamic range limits for a stopband blocking signal. This is a
consequence of the results in section 2.6.3. However, the Elliptic response has about 7 dB
lower passband sensitivity (no dark red region on the plots). (All of these results of course
assume a good realisation.)

It is worth spending time studying g. 3.1 in detail, as the plots bring together a lot of
the theory in the previous chapter in a way which is likely to be useful to a designer.

3.2.3 Relationship with existing theory


It is useful to relate to this new approach to choosing the transfer function using the
dynamic range limitation to existing work. It is known that minimising lter pole Q tends
to minimise sensitivity and maximise dynamic range (Abidi 1992) (Groenewold 1991). In
particular, the high-Q pole pair at the corner frequency of most lters causes problems with
peaking at that frequency. This is in consistent with the dynamic range limitation, because

86
Elliptic
4 30

3
20
Output noise frequency (rad/s)

2
10
1

0 0

1
10
2
20
3

4 30
4 2 0 2 4
Input signal frequency (rad/s)

Chebyshev
4 30

3
20
Output noise frequency (rad/s)

2
10
1

0 0

1
10
2
20
3

4 30
4 2 0 2 4
Input signal frequency (rad/s)


Figure 3.1: 20 log10 H (j!j(!1 )2;;H!1(j!
) versus input signal frequency !2 and output noise frequency !1
2)

for a 5th order, 0.01 dB ripple, {49 dB stopband Elliptic and a 7th order 0.2 dB ripple Chebyshev
(these have the same transition bandwidth). Sensitivity depends on the value along the diagonal line,
and blocking performance depends on the value in the boxes. Isopleths (colours) are 2 dB wide.

87
j[H (j!1 ) ; H (j!2 )]=j (!2 ; !1)j is usually largest for !1 and !2 near the corner frequency {
see the red regions in g. 3.1. (Note that the change in transfer function phase, ie. group
delay, contributes as well as the change in transfer function magnitude.)

As lter designers know the signi cance of pole Q, it is reasonable to ask whether the
dynamic range limitation makes any di erent predictions. Now some authors
(Premoli 1973) (Kim 1992) have suggested replacing the high-Q pole pair at a lter's corner
frequency by multiple lower-Q poles to decrease the sensitivity. Considerations based on
pole Q suggest this is worthwhile, but the dynamic range limitation rules out much
improvement from such techniques if the transfer function remains similar. The idea does
not seem to be used much in practice.

3.3 REALISATION
Having chosen the transfer function, we move on to the realisation.

3.3.1 Unsuitability of the frequency-weighted input-normal realisation


Suppose we wish to use the frequency-weighted input-normal realisation, which was shown
in the last chapter to theoretically give optimal dynamic range. Typical weights might be
8
>
<1; j!j < 2
j(j!)j2 = > (3.1)
:10; j!j  2
assuming blockers are 10 dB larger and start at ! = 2 rad=s, and
8
>
> 1; j!j  1
>
<
j (j!)j2 = >0:1; 1 < j!j < 2 (3.2)
>
>
:1; j!j  2
assuming that the ADC sample rate is three times the lter bandwidth, so noise above
2 rad=s folds into the passband etc. (In practice, j (j!)j2 = 1 everywhere gives almost
identical results. Increasing j (j!)j2 further may also decrease stopband sensitivity.)

88
The realisation which results is
2 3
66;0:0240 0:4060 ;0:0253 ;0:0257 ;0:005377
66;0:6525 ;0:1956 0:4146 0:0719 0:0229 77
66 7
A = 66;0:4458 ;0:9923 ;0:3734 ;0:3273 ;0:0675777
66 0:6989 0:6483 1:1722 ;0:4386 ;0:352677
4 5
;1:0441 ;1:1410 ;1:2403 1:7061 ;1:0441
2 3
66 ; 0 : 4618 77
66;0:400177
66 7
B = 66;0:3821777
66 0:3083 77
4 5
;0:4705
h i
C = ;0:1759 0:3960 ;0:3366 ;0:1244 ;0:0287
hi
D = 0

The problems with this realisation should be apparent. The dense tap matrices will require
lots of resistors. Moreover, small taps such as a11 require large resistors (100 k
), which will
be physically large and hence have high parasitic capacitance. (Sometimes they can be
rounded to zero without changing the transfer function much though.) Large taps such as
a54 will heavily load the relevant integrator, requiring the unit resistance to be increased to
prevent current limiting. This then results in increased thermal noise. Nevertheless, these
problems should not be considered insurmountable { the author used this approach once in
a 30 MHz lter. It is also possible to halve the number of non-zero coecients (Li,
Anderson, Gevers & Perkins 1992) or partition the system into lower-order sub- lters.
(Partitioning is sub-optimal though.)

3.3.2 Merit of lossless realisations


As the frequency-weighted input-normal realisation was not very suitable, a more
conventional realisation was used, namely a simulation of a doubly-terminated LC ladder.
The state-space equations describing an LC ladder are known to have good sensitivity
properties (Fettweis 1986) (Orchard et al. 1985). (These papers give a useful introduction

89
and refer back to most of the original work which dates from the 1960's.) The passband
sensitivity is low by the Orchard / Fettweis argument that changing the L or C values
cannot allow the gain of a doubly-terminated LC network to exceed 0 dB. Hence at
frequencies where the gain is already close to 0 dB, the rst-order sensitivity must be low to
prevent the gain exceeding 0 dB with a small parameter variation. The stopband sensitivity
of an LC ladder is good because the ladder structure does not require two nearly equal
quantities to cancel to achieve the stopband rejection. The stopband attenuation is
achieved progressively along the ladder.

We can quantitively compare the frequency-weighted input-normal realisation and the


LC ladder simulation using tr W with the diagonal elements kii (the powers in the states
for input spectrum j(j!)j2 ) of K  scaled to unity. For the weights (3.1) and (3.2) and the
Elliptic transfer function used, the result is tr W = 1:59 for the frequency-weighted
input-normal realisation and tr W = 1:02 for the LC ladder. The LC ladder is actually
better! This may appear to contradict the optimum nature of the frequency-weighted
input-normal realisation. However, it actually results from using a slightly di erent norm to
that in which the frequency-weighted input-normal realisation is optimal. The LC ladder
spreads the signal energy across the states better than the frequency-weighted input-normal
realisation. An \equal-wordlength optimal realisation" exists (Mullis & Roberts 1976) which
is better again than the LC ladder, but it is not discussed here as it has little connection to
the Hankel singular value theory, requires a dense tap matrix, and is only a little better.

In a lot of cases, an LC-ladder-based realisation achieves an output noise tr W close to


2 
3 tr K W . The author does not know of any theoretical justi cation for this. However, it
is very useful to know that the dynamic range achieved by a practical realisation, namely
the simulation of an LC ladder, usually correlates closely with the norm tr K  W which
can be calculated from the transfer function alone. This consideration was the basis of the
author's earlier paper (Harrison & Weste 2001).

90
C2 C4

R1

L2 L4
2u C1 C3 C5 R5 y

Figure 3.2: LC ladder used for state-space equations

3.3.3 Lossless realisation


Section 3.2 showed that transfer functions with nite stopband zeroes (notably low-ripple
Elliptic ones) were superior, particularly with respect to sensitivity at the passband edge.
However, this complicates the state-space realisation. The fth order LC ladder has ve
capacitors and two inductors, so a state consisting of all capacitor voltages and inductor
currents is not minimal. This problem was solved by using the descriptor state-space
equations (which have an extra matrix in front of ddxt ) in place of the state-space equations.
That is, the oating capacitors (C2 and C4 ) in the Elliptic LC ladder prototype were realised
as oating capacitors in the opamp-RC realisation. The descriptor state-space equations are

C ddxt = F x + Gu (3.3)
y = H x: (3.4)

The state is taken as


h iT
x = vC1 iL2 vC3 iL4 vC5 ;

91
from which is is straightforward to analyse the circuit and derive that
2 3
66 C1 + C2 0 ;C2 0 0 77
66 0 L2 0 0 0 77
C =
66 77
66 ;C2 0 C2 + C3 + C4 0 ;C4 77
66 0 0 0 L4 0 77
4 5
0 0 ;C4 0 C4 + C5
2 1 3
66 R ; ; 1 0 0 0 7
66 1 0 ;1 0 0 777
1

66 7
F = 66 0 1 0 ;1 0 777
66 0 0 1 0 ;1 77
4 5
0 0 0 1 ;R 1
h iT 5

G = 2 0 0 0 0
h i
H = 0 0 0 0 1 :

A similar matrix-based approach for gm-C lters is in Greer, Henderson, Ping & Sewell
(1991). As the matrix C is diagonally dominant, the dynamic range properties of the
descriptor equations are similar to the true state space equations. True state-space
equations can be obtained by algebraic manipulations, but the tap matrix then becomes
dense. (This is how K  W was calculated in section 3.3.2.) The LC element values were
taken from the tables in (Zverev 1967).

An important consideration is that F and H only have unit elements. (R1 and R5 are
1
.) The coecient of two in G is easily realised by two parallel elements. This means that
the lter can be tuned by adjusting a unit element.

3.3.4 Filter taps - practical details


Each element Cij of C corresponds to two capacitors, and each element of F , G and H
corresponds to two switchable unit resistors - see g. 3.3. (G uses two pairs of parallel unit
resistors as the tap coecient is two).

The digital process used does not have metal-insulator-metal capacitors, so routing
92
metal was used for capacitors. Metal layers 3,4,5 and 6 were used, giving a capacitance of 3
 40 aF=(m)2 = 120 aF=(m)2 . The bottom-plate capacitance is around 10 aF=(m)2 .
Capacitors were built from a unit cell of 15 fF. The diagonal elements Cii are 150 fF to
360 fF, and the oating capacitors Cij to realise the stopband zeroes are 15 fF and 45 fF.
Some rounding of tap values is necessary to avoid too small a unit cell. The frequency
response with rounded tap values was checked in Matlab. In general, lossless simulations
can tolerate a surprising amount of rounding - a Matlab Monte Carlo simulation with 5%
capacitance mismatch still had acceptable frequency response. This knowledge is helpful for
layout where a decision often has to be made between low parasitics and optimum
matching. Given the tolerance of mismatch, here the emphasis was on low parasitics so
common-centroid geometries were not used.

The lter is tuned by switching binary-weighted resistors. The resistors are un-silicided
polysilicon. The most signi cant bit (MSB) is a resistor with a single nger R =3.4 k
, and
proceeding towards the least signi cant bit (LSB) the resistors are R; 2R; 4R;    . A
practical lter would have ve or six bits of tuning. The parasitic capacitance of the LSB's
needs to be considered, but it is not prohibitive if the resistor width is kept below about
1 m. In this process the polysilicon matching is fairly good (around 1 % ; m), so
mismatch is not a signi cant problem. If matching was worse, the MSB could be
constructed from a series-parallel combination rather than a single nger. In the lter
fabricated, only three bits of tuning were included due to pin count restrictions for probing.

There is some subtlety in implementing switched tuning. Matching favours one resistor
nger or capacitor cell being an LSB rather than an MSB. With capacitors, this is practical.
Tuning the capacitors also sometimes allows the common-mode level of the switches to be
at ground, lowering the on-resistance. However, in this design the capacitors vary from
15 fF to 360 fF. The LSB's of a 15 fF capacitor are too small to tune. (Behbahani, Tan,
Karimi-Sanjaani, Roithmeier & Abidi (2000) avoid this problem by not tuning the small
capacitors which realise the stopband zeroes, with the consequence that the stopband zeroes
are at the wrong frequency at tuning extremes.) If resistors are tuned and it is desired that
one nger is an LSB rather than an MSB, then the switched resistors need to be in series
rather than parallel. However, switches in series need to be wider, and the parasitic

93
3.4k 6.8k 13.6k
7.2/0.18 3.6/0.18 1.8/0.18

Cij
14.4/0.18 7.2/0.18 3.6/0.18

j i j i

Cij

Figure 3.3: Filter taps. Left { unit resistor (element of F , G or H matrix). Right { element Cij of C
matrix.

capacitance is unacceptable. Sometimes there is merit in tuning both capacitors and


resistors, with the MSB's tuned on one and the LSB's tuned on the other (Behbahani
et al. 2000).

3.3.5 Replica oscillator


To tune the lter to the correct cuto frequency, the RC time constant must be measurable.
This is commonly done using a replica oscillator which uses identical resistor and capacitor
elements to the main lter (Nauta 1993) (Tsividis & Gopinathan 1994). Probe pad count
and schedule restrictions prevented a replica oscillator being fabricated, but the circuit in
g. 3.5 performed well in simulation.

Transistors are used in place of diodes to control the oscillation amplitude because
diodes would give too high an amplitude. The replica oscillator cuto frequency correlated
with the lter {3 dB frequency to 2% in Monte-Carlo simulations.

94
Unit resistor elements with switched tuning

Bus

Opamps

Bypass capacitors
Capacitors

Figure 3.4: Filter layout

9/0.18

20k

unit C unit C

unit R unit R
(tuned) (tuned)

Figure 3.5: Simulated replica oscillator

95
3.4 OPAMP DESIGN

3.4.1 Choice of opamps as the active element


Feedback circuits such as opamp stages have the advantage that the active element does not
have to be as linear (before feedback). This should allow greater signal swing for a given
distortion.

It is conventionally argued that the problem with opamps is that they do not have
sucient bandwidth (Toumazou, Lidgey & Haigh 1990, p. 347) (Nauta 1993, p. 7). More
cautious authors (Tsividis & Gopinathan 1994, p. 201) admit that this argument may not
be valid. The author does not know whether opamps are slower than transconductors, and
whether voltage-mode circuits are slower than current-mode circuits. The argument that
transconductors and current-mode circuits are faster because they avoid internal nodes with
long time constants seems spurious. Presumably it stems from the method of open-circuit
time constants (Gray & Searle 1969, p. 531). Unfortunately, for Miller-compensated and
feedforward opamps, this method seems invalid. For a feedforward opamp, the feedforward
path skips the long time constants. For a Miller-compensated opamp, the poles split and
hence pole-by-pole calculations are inappropriate.

The argument in (Tsividis & Gopinathan 1994, p. 201) that opamp-RC lters are slow
because CMOS opamps which drive resistors are slow is more plausible. However, the
ultimate limitations on opamp gain-bandwidth may turn out to depend more on load
capacitance rather than load resistance. The usable opamp bandwidth is some fraction of its
unity-gain bandwidth. (By the Bode gain-phase relationships, the fraction depends on the
phase lag.) At frequencies close to unity gain, the impedance of the capacitors in a lter is
signi cantly lower than the impedance of the resistors. (For the lter described here, at the
3 GHz unity-gain frequency the integrator capacitance of 250 fF is { j 210
. The resistive
load on each opamp is around 1 k
.) It is therefore the load capacitance rather than
resistance which limits the unity-gain frequency, and hence also the operating bandwidth.

96
3.4.2 Fundamental limitations on opamp bandwidth
What opamp bandwidth is possible? This problem can be split into two parts:

1. What opamp unity-gain frequency is possible?

2. At what fraction of unity-gain frequency can reasonable loop gain (30{40 dB) be
obtained?

Unity-gain frequency
Using a simple small-signal model of a transistor which consists of only gm , Cgs and Cds ,
the unity-gain frequency f0 of cascaded identical common-source stages is

f0 = 21 C g+m C = fT C C+gsC ' f2T :


gs ds gs ds
(Deep-submicron CMOS processes have a notably large Cds .) This is a fairly accurate
upper limit on opamp unity-gain frequency. A fundamental limit can be calculated using
the theory of Thornton (1957) and Glasser (1988). If perfect inductors and available, the
maximum unity-power-gain frequency is fmax . Without inductors, the techniques in
(Glasser 1988) can be used to calculate that oscillation and hence ampli cation is only
possible up to around fT =2 for a typical deep-submicron CMOS transistor. Capacitance
from pMOS current sources and routing reduce this a little further.

If the transistors could be biased for maximum fT (30{40 GHz for a 0.18 m process),
very impressive results would be obtained. In deep-submicron CMOS analog circuits, bias
for maximum fT is rarely possible, for two reasons:

 The Vgs (0.8 V) and Vds (> 0.5 V) required compromises signal swing or tolerance of
threshold, temperature and VDD variations.

 The current consumed is excessive. For optimum bandwidth, transistor capacitance


needs to dominate load capacitance. This requires wide transistors. A transistor with
100 m width operating at current density of 100 A=m consumes 10 mA.

97
In this circuit, the transistor current density was limited to 10 A=m. At this current
density, the fT is around 10 GHz. Therefore cascaded identical common-source stages have
a unity-gain frequency around 5 GHz. The opamp unity-gain frequency achieved was 3 GHz.

Bandwidth with reasonable loop gain


There is a bandwidth penalty in using feedback. With conventional one-pole gain rollo
({20 dB=decade), the bandwidth penalty is obviously severe { a factor of 100 in bandwidth
for 40 dB loop gain. For switched-capacitor circuits, one dominant pole is necessary for fast
settling. However, for continuous-time circuits it is usually bandwidth rather than settling
time which is important. Acceptable loop gain can be achieved over a greater bandwidths
by using a faster gain rollo , such as around {30 dB=decade (Bode 1945). Larger gain slope
of course reduces the phase margin, but 45 is usually quite sucient. An approximation to
this gain slope is two-pole compensation (Roberge 1975), where the gain falls o at a
two-pole rate at low frequencies and a one-pole rate closer to the unity-gain frequency. (This
is actually preferable to a constant {30 dB / decade slope as the phase margin is improved.)

More feedback than this is possible, such as by using a conditionally stable ampli er.
The results in section 2.10.1 are relevant here. The author guesses that 40 dB of loop gain is
possible at one fth of the unity gain frequency. The results obtained from the procedure in
section 2.10.1 depend heavily on the asymptotic rollo of the gain and the tolerable peaking
or out-of-band sensitivity. (For example, is an opamp which has a gain which falls to unity
at 10 GHz but stays at 0.9 up to 1000 GHz possible?)

Fig. 3.6 gives a summary of the situation regarding unity-gain bandwidth and
bandwidth for workable loop gain.

3.4.3 E ect of nite opamp gain-bandwidth


Given that opamp gain-bandwidth is limited, what opamp gain is actually required?

Consider a lter with transfer function H (s) built from integrators with gain k=s. If
integrator imperfections (notably nite opamp gain) result in an integrator transfer
function k=F (s), then the lter transfer function will be H (F (s)). This assumes that
98
Loop gain Can trade-off DC gain
vs. bandwidth

Estimated
ultimate limit
40dB This work with practical
sensitivity

Likely to
20dB be conditionally-stable

Conventional
Miller-comp. opamp Frequency
0dB
100MHz 300MHz 1GHz 3GHz 10GHz 30GHz

GBW loss with Lower Cload


Miller comp. Vgs, Vds pMOS C
ds

Opamp Opamp nMOS untuned stages


unity-gain unity-gain unity-gain freq.
freq. at freq. at around fT/2
practical bias max. bias

Figure 3.6: Achievable opamp gain vs. frequency.

99
imperfections in all integrators are the same. There will be some di erences because
di erent integrators have di erent loads, but much of the variation will be shared.

There does not seem to be a good reference to this simple analysis in the literature. For
LC lters, the equivalent assumption, namely uniform element losses, is considered in
(Bode 1945) and (Blostein 1967). The author is presumably not the rst person who has
had to re-invent this analysis, so he implores textbook writers to include useful
approximations like this rather than rigorous sensitivity analyses which o er less insight
and are superseded by Monte-Carlo simulations.

For an opamp-RC integrator, if the opamp has nite gain A(s) but no other
imperfections, it can be easily calculated that
F (s) = s + A(1s) (s + k): (3.5)
The opamp used here (and most others for that matter) has a phase shift 6 A(s) of around
;90 near the lter cuto . (s + k) is around +45 . Together, these result in the second
term in (3.5) having a phase shift around +135 , moving F (j!) to the left half plane. This
causes peaking as F (j!) is closer to the lter poles. The peaking is most pronounced
around the cuto frequency as the high-Q pole pair there is closest to the j! axis.

The amount of peaking is straight-forward to calculate. A reasonable approximation to


the opamp gain A(s) was chosen, and A(j!), F (j!) and H (F (j!)) were calculated using
Matlab. The result is that an opamp gain jA(j!)j of 32 dB at the cuto frequency results in
1 dB of peaking. A gain of 28 dB results in 2 dB of peaking. In practice, a few decibels more
gain is required because opamp input capacitance and routing parasitic capacitance on the
opamp inputs cut the e ective loop gain. In this lter the reduction in loop gain is around
3 dB. Therefore 35 dB opamp gain is required for 1 dB of peaking, and 31 dB for 2 dB of
peaking. The actual opamp gain achieved was 35 dB at 200 MHz and 30 dB at 350 MHz.

The tolerable deviations in the transfer function obviously depend on the application.
For a radio receiver, worst-case passband ripple of 2{3 dB is often speci ed. To allow for
process variation, nominal peaking should be kept to about 1 dB, ie. 35 dB opamp gain.
The opamp used achieves this at 200 MHz. At 350 MHz cuto frequency, Q-tuning would be
required. It is often said that 40 dB opamp gain is a good target, but with a well-rounded
100
transfer function like that used here slightly less suces.

Another way of doing the same analysis is to simulate an integrator to get F (j!)
(Nauta 1993). The integrator phase shift ( g. 3.7) is usually more signi cant than the
integrator gain as gain errors just cause slight tuning errors and will be tuned out by the
frequency tuning. (This may not be so true for disk-drive read channels where lter group
delay variations are a problem.) Acceptable phase error is often around 1{2 . This has the
advantage that opamp input capacitance is considered, although routing capacitance can
still cause unpleasant surprises.

A useful interpretation of this is that an integrator with phase lag less than 90 is like a
lossy inductor or capacitor in the LC ladder prototype. Dissipated energy results in
passband droop near the corner frequency. There is a quantitative theory of these e ects
(Blostein 1967) (Kishi & Nakazawa 1963) (Kishi & Kida 1967) (Livsic 1973) (Harrison &
Weste 2001) involving stored energy and group delay which makes similar predictions to
those from the dynamic range limitation. (The dynamic range limitation is superior in that
it applies to all transfer functions and better considers blocking dynamic range though.)
For the converse situation of integrator phase lag greater than 90 , the storage elements
generate rather than dissipate energy and the response peaks rather than droops. By
contrast, integrator gain errors preserve the passivity of the lter and cause less amplitude
deviation.

3.4.4 Gain stage principles


Close to optimal performance should be possible with simple common-source stages. There
is then the issue of how to connect multiple stages together.

In the past, this issue was complicated by the fact that common-source stages have
negative gain, and common-gate and source follower stages have positive gain. It was often
necessary to choose common-gate or source-follower stages simply because a positive gain
was required, not because there was any bandwidth or signal swing advantage. An example
of this is in two-stage single-ended feedback ampli ers.

101
Figure 3.7: Integrator phase shift { Monte Carlo simulation. 20 runs at each of {25C, 25 C and
125C.

102
With the rise of fully-di erential circuitry in the 1990's, this problem is less important.
The sense of the gain can be changed by swapping the positive and negative inputs (or
outputs). The problem is not entirely eliminated, because occasionally common-mode
latch-up or oscillation problems arise. (This is latch-up in the opamp not CMOS sense of
the word.)

(The complexity of transistor-level analog circuit design seems to be largely a


consequence of the unavailability of ideal transformers and ideal blocking capacitors. The
theory of Thornton (Thornton et al. 1966) (Thornton 1957) on fundamental limitations on
transistor circuits is very interesting in this context, and seems to be applicable to the
current-mode versus voltage-mode debate. The present era seems to be the Baroque period
of circuit design.)

Common-mode feedback
The diculty with fully-di erential circuits is that common-mode feedback is required
(Duque-Carrillo 1993). Building a good common-mode feedback circuit is harder than
might be expected. Switched-capacitor common-mode feedback is good, but it cannot be
used in a continuous-time lter. The following issues are signi cant:

 When the opamp is operating in a circuit in which the input common-mode level is
set by the output common-mode level, latch-up with both inputs and both outputs
stuck at a supply rail can occur. Common-mode feedback circuits that might be
expected to work often cannot recover from this condition. The circuit in (Banu,
Khoury & Tsividis 1988) is one which does avoid this problem.

 The common-mode feedback loop may need compensation. This is particularly an


issue in two-stage opamps. What bandwidth should the common-mode loop have?

 The gain of the common-mode feedback loop can be a ne trade-o between achieving
suciently precise control of the common-mode level, and not degrading the
di erential linearity. This is particularly a problem if a transistor circuit with limited
linearity is used to average the two outputs to derive the common-mode level.

103
The author has reached the conclusion that the best way to compensate a two-stage
opamp is for each stage to have its own common-mode feedback. The virtues of local
(single-stage) feedback are that it does not need compensation and is usually globally
stable. The virtue of global (multiple-stage) feedback is that the loop gain is higher so the
error is smaller. In common-mode feedback circuits, the common-mode level usually only
needs to be set to within 100 mV or so. This does not require the precision of multiple-stage
feedback, so the virtues of single-stage feedback are more important. Therefore two local
common-mode feedback loops should be used instead of one feedback loop around two
stages.

The nMOS-nMOS opamp


A two-stage opamp traditionally has a pMOS input stage and an nMOS output stage. The
two stages need to use opposite types of transistors to achieve good input common-mode
range and output swing (Cherry & Hooper 1968, p. 286). The pMOS / nMOS combination
is preferred over nMOS / pMOS as bandwidth of the output stage is more critical, and
pMOS transistors often have lower icker noise.

However, as transistor threshold voltage VT approaches VDD =2, an nMOS input stage
and an nMOS output stage becomes more viable. The maximum signal swing of cascaded
identical stages like this is 2VT peak-to-peak per side. Obviously, with a 15 V or 5 V power
supply this is sub-optimal. But with VDD =1.8 V and VT =0.5 V, the signal swing achieved is
over half of rail-to-rail. For high speed circuits, this is important because of the higher fT of
nMOS transistors. It also simpli es design somewhat, particularly when implementing
feedforward.

3.4.5 Compensation
A two-stage opamp is usually used for opamp-RC lters, as single-stage opamps do not
achieve sucient gain with resistive loading. The usual method of compensating a
two-stage opamp is Miller compensation. This is a good compensation technique { unity
gain frequencies of a signi cant fraction of fT are possible. It also has the virtue that the

104
VDD

36/0.36 108/0.36

0.9/18 20k

Out

In +
Bias 36/0.18 36/0.18 72/0.18
250uA

In -

9/0.18 36/0.18 36/0.18 72/0.18

Figure 3.8: Filter opamp

Miller capacitor provides additional feedback around the output stage. However, intuitively
Miller compensation can be expected to be sub-optimal for two reasons. Firstly, a one-pole
compensation with a gain slope of {20 dB=decade is not optimal loop-shaping. The loop
gain at moderate frequencies can be increased by a faster gain rollo , say {30 dB=decade
(Bode 1945). This can be approximately achieved by using two-pole compensation
(Roberge 1975). Secondly, the energy stored in the Miller capacitor might better be used by
an active device to generate signal power. Opamps which use the load capacitance for
compensation may have advantages.

The feedforward compensation used in the lter opamp does not have these drawbacks.
The gain declines at a two-pole rate (from the main signal path) at low frequencies, and a
one-pole rate (from the feed-forward path) at high frequencies. This is better shaping of the
sensitivity versus frequency, and it avoids compensation capacitors. How this is achieved is
easiest to see from the circuit.

105
3.4.6 Opamp circuit
As might be expected considering the discussion above, the opamp ( g. 3.8) is constructed
from nMOS di erential pairs with local common-mode feedback. The main low-frequency
signal path is through the two di erential pairs on the right. The di erential pair on the
left provides the high-frequency feedforward path which pulls the phase lag back to 90 at
high frequencies to provide acceptable phase margin.

The stages use simple local common-mode feedback circuits composed of pMOS current
source transistors with resistors or long transistors to derive the common-mode output.
Long, narrow (0.9/18) transistors are used for the input stage. These provide a higher
resistance and lower capacitance than resistors would. Resistors are used for the output
stage though as they are more linear.

For simplicity, transistors in all three di erential pairs run at the same current density,
10 A=m. The current density is limited by the need to provide sucient Vds across the
tail current sources at low temperature and slow (high VT ) process corner. For typical
transistors, the signal common-mode level is around 1.1 V, and the sources of the
di erential pairs sit at 0.4 V. In future designs, a more sophisticated CMFB which sets the
common-mode level at Vgs + Vsat rather than VDD ; Vgs could be used. This would reduce
the minimum power supply voltage from 2Vgs + Vsat to Vgs + 2Vsat , allowing higher
transistor current density for the same VDD and hence higher bandwidth.

3.5 MEASURED PERFORMANCE

3.5.1 Frequency response


The measured frequency response ( g. 3.10) has two main imperfections. One is the
peaking at high cuto frequencies due to nite opamp gain. This is in agreement with the
theory in section 3.4.3. The other e ect, which was not present in simulation, is some
passband droop. There are a number of imperfections which could cause this, such as
capacitor losses, routing resistance and lower than expected opamp DC gain. Routing

106
60
Gain (dB)
40 Phase (Deg.)

20

20

40

60

80

100

120

140

160
7 8 9 10
10 10 10 10
Frequency (Hz)

Figure 3.9: Opamp Bode plot. Load 300 fF in parallel with 1 k

resistance seems the most likely. It was not possible to include parasitic resistance in
extracted simulations, as it adds too many nodes to the circuit and the simulator crashes
when its matrix exceeds 4 Gbyte in size. (This problem is largely caused by a limitation of
the extraction program, Cadence Diva, which leads to a node being created at every via.)
The routes between the opamps and capacitors have an estimated resistance of 80
, which
is sucient to reduce the integrator phase lag by 2 and cause the observed passband droop.

Overall, the frequency response is suciently controlled for practical applications such as
radio receivers for cuto frequencies up to 200 MHz. A Monte-Carlo simulation at 180 MHz
nominal cuto frequency is show in g. 3.11. At higher cuto frequencies (350 MHz),
Q-tuning would be required to control the peaking due to nite opamp gain.

107
10

10

20
dB

30

40

50

60

70
2 3
10 10
Frequency (MHz)

Figure 3.10: Measured frequency response over the tuning range. Some of the stopbands are omitted
for clarity. Tuning 001 (40 MHz cuto ) to 111 (350 MHz cuto ).

108
Figure 3.11: Monte-Carlo simulation of the extracted circuit over expected process and mismatch
variations. 20 runs at each of {25C, 25C and 125C. Tuning at nominal cuto frequency of 180 MHz.

109
10
Prototype
Sim.
0
Sim.

10 Measured

Measured
20
dB

30

40

50

60

2 3
10 10
Frequency (Hz)

Figure 3.12: Measured, simulated and prototype frequency responses (350 MHz responses are o set
by +10 dB)

110
3.5.2 Dynamic range
The maximum signal swing is 0.50 Vp-p di erential for {40 dB total harmonic distortion
(THD). This is with 150 MHz and 160 MHz input tones and 350 MHz cuto frequency. The
distortion does not vary much with input frequency, although it degrades a little if
intermodulation products fall near the cuto frequency, and improves a little at low
frequencies. The worst distortion products are at 460 MHz and 470 MHz (around {45 dBc);
the IM3's at 140 MHz and 170 MHz are around {50 dBc. Measured and simulated distortion
are similar. The table below summarises the performance.

Speci cation Performance


Power 14 mA @ 1.8 V
Gain 0 dB
Tuning range ({3 dB cuto ) 40 MHz { 350 MHz
Inband IP3 (150,160 MHz) {13 dBVRMS
Out-of-band IP3 (450,650 MHz) {10 dBVRMS
Max. input for {40 dB THD 0.50 Vp-p
p
Average passband noise 24 nV= Hz
Dynamic Range ({40 dB THD) 52 dB
Process digital CMOS 0.18 m
Area 0.5 mm2

A slightly higher signal swing could perhaps have been hoped for { as discussed
previously the voltage swing is limited to around 2VT peak-to-peak per side, ie. 2.0 Vp-p
di erential. This lter is limited more by output stage current than voltage swing; at
0.50 Vp-p the devices in the opamp output stages swing from 0.4 mA to 1.1 mA drain
current. The opamps clip from lack of output stage current when the output is around
1.0 Vp-p. There is also some contribution to the distortion from the switched tuning
transistors.
p
The average noise over the passband is 24 nV= Hz. The impedance levels were kept as
low as possible (a little too low in fact, because the opamps clip from lack of output stage

111
current) to minimise thermal noise. The resulting dynamic range is 52 dB (for {40 dB THD).

The table below compares these results to others published for lters with cuto
frequency > 100 MHz. \Swing" is di erential signal swing for {40 dB THD. \error " is the
maximum integrator phase error in Monte Carlo simulations, for a frequency of 200 MHz.

Reference Topology V DD Swing error Noise


p
This work opamp-RC 1.8 V 0.50 Vp-p 1:2 24 nV= Hz @ 25 mW
p
(Pavan et al. 2000) gm-C 3.3 V 0.38 Vp-p ? 14 nV= Hz @ 70 mW
(Gopinathan et al. 1999) opamp-gm-C 2.5 V 0.46 Vp-p 0:7 ?
(Rao et al. 1999) gm-C 3V 0.20 Vp-p 1:5 ?

It is dicult to make a fair comparison considering the di erent applications, but the
signal swing and dynamic range of the opamp-RC approach appears superior for a given
VDD . The noise in (Pavan et al. 2000) is identical to the lter presented here after scaling
for power consumption.

These results are all without Q-tuning. There are also circuits which use Q-tuning, such
as (Deng & Lee 2001) and (Nauta 1993). A signal swing of 0.3 Vp-p at 1 V VDD is achieved
in (Deng & Lee 2001) using a current-mode approach. However, the current ampli er in
(Deng & Lee 2001) still requires VDD > Vgs + 2Vsat . It is only by operating at lower current
densities and using Q-tuning to correct for the nite ampli er bandwidth (only 1.2 GHz)
that 1 V operation is possible. (There also seems to be no mention of threshold variations.)
Simple di erential pairs with switched-resistor source degeneration (to control gm ) may also
be attractive if Q-tuning can be used { an impressive dynamic range (albeit at lower
frequencies and higher VDD ) is achieved in (Behbahani et al. 2000). Nevertheless,
opamp-based high-feedback circuits seem the best possibility for achieving close to
rail-to-rail signal swings with low distortion.

3.6 ADDITIONAL SIMULATIONS


Some additional simulations were performed later to see what bandwidth and dynamic
range could be achieved by the opamp-RC topology. The lter capacitors were kept

112
Figure 3.13: Three-stage opamp. The opamp is constructed from identical di erential pairs (left).
The number on each stage (right) is the number of di erential pairs in parallel.

unchanged, and the unit resistors decreased to allow tuning up to 1 GHz. More opamp
power supply current was required to drive the lower impedances. Satisfactory opamp gain
at 1 GHz was achieved with a factor of six increase in power supply current { this is not too
bad considering that the impedances are one third of what they were. A remarkable opamp
unity-gain frequency, 10 GHz, was achieved ( g. 3.14).

The opamp ( g. 3.13) has three stages with two feedforward paths. This increases the
gain at low frequencies, with little expense in unity-gain bandwidth. The phase margin ( g.
3.14) is reduced to 20-30 . This may seem marginal, but the feedback path has around 20
lead, which increases the loop phase margin to 40 . The opamp phase margin also does not
vary much under process variations { it is closely controlled by the relative strengths of the
feedforward paths. Inadequate phase margin manifests itself as stopband peaking around
5{10 GHz; g. 3.15 shows this is acceptable.

It is interesting to note that even if the phase lag exceeded 180 at lower frequencies, this
opamp architecture will not necessarily be conditionally stable. This is because the main
path will tend to clip from insucient current before the feedforward path does. The
reduced gain from the main signal path and unchanged gain from the feedforward path will
reduce the phase lag. This o ers interesting possibilities for opamps using very fast gain

113
Figure 3.14: Three-stage opamp Bode plot. Monte-Carlo simulation over expected process and mis-
match variations. 20 runs at each of {25C, 25C and 125C. Load 400 fF in parallel with 400
.

114
roll-o to achieve high gain (> 60 dB) at gigahertz bandwidths.

The 10 A current source in each di erential pair's common-mode feedback ( g. 3.13,


left) increases the common-mode output level by 0.2 V, to around 1.0 V. This increases the
output swing. A xed current was used here, but a feedback loop or replica structure could
be used if a controlled output common-mode level was needed. A more sophisticated
approach could also optimise the signal swing over VDD and process variations.

A Monte-Carlo simulation of lter frequency response when tuned to 800 MHz is shown
in g. 3.15. This is about the practical limit to the achievable bandwidth for acceptable
frequency response without Q-tuning. The bandwidth achieved would actually be
signi cantly higher (around 1.5 GHz) if not for the parasitic capacitances of the tuning
switches. The tuning range was reduced to the minimum necessary to reliably cover process
variations, so as to minimise the e ect of switch parasitics. Zero-threshold transistors would
be very helpful here as acceptable on-resistance could be achieved with narrower transistors.
Switch width is a trade-o between capacitance and distortion from voltage across the
switch.

The simulated performance is as follows:

Speci cation Performance


Power 76 mA @ 1.8 V
Gain 0 dB
Tuning range ({3 dB cuto ) 500 MHz { 1 GHz
Inband IP3 (400,500 MHz) +8 dBVRMS
Out-of-band IP3 (1.4,2.4 GHz) +11 dBVRMS
Max. input for {40 dB THD 1.6 Vp-p
p
Average passband noise 18 nV= Hz
Dynamic Range ({40 dB THD) 60 dB
Process digital CMOS 0.18 m

A 10 dB higher signal swing and 20 dB higher IP3 than the fabricated 350 MHz lter has
been achieved. This is a consequence of higher opamp gain and more careful choice of

115
Figure 3.15: Monte-Carlo simulation over expected process and mismatch variations. 20 runs at each
of {25C, 25C and 125C. Tuning at nominal cuto frequency of 800 MHz.

opamp output stage current. The lter maintains low distortion (< {40 dB THD) until it
clips.

Hopefully this lter will be fabricated in the near future. Layout is likely to require more
care. The integrator capacitors will probably have to be placed next to their respective
opamps to prevent oscillation at 5{10 GHz. Routing capacitance will need to be minimised,
and routing resistance extracted or carefully estimated. Nevertheless, the operation of the
existing lter at 350 MHz suggests that the 800 MHz lter is practical. The parasitic
capacitances assumed for the schematic have been found to be conservative for past designs.

3.7 CONCLUSIONS
Opamp-RC lters can equal or better the performance of gm-C lters at the high
frequencies previously though to favour transconductor-based approaches. It seems almost
certain that they will be competitive at lower frequencies too. Signal swings of a large

116
fraction of rail-to-rail are possible by using the feedback of opamps to minimise distortion,
and switched polysilicon resistors to overcome the linearity limitations of transistor-based
resistors or transconductors.

Opamp-RC lters can operate at frequencies up to around fT =50. Two-pole feedforward


compensation is superior to Miller compensation for wide bandwidths. It can be
implemented using a simple and robust approach in which each di erential pair has its own
common-mode feedback. This approach has a certain logic to it in that the problems of
stabilising common-mode feedback loops around multiple stages are avoided.

The dynamic range limitation presented in the previous chapter can aid the choice of
lter transfer function. The insights are similar to those from pole Q considerations, but are
more rigorous and are applicable to consideration of dynamic range as well as sensitivity.
Plotting the kernel j[H (j!1 ) ; H (j!2 )]=j (!2 ; !1 )j of the Hankel operator is shown to be
particularly useful for comparing transfer functions H (s).

Despite its provably good dynamic range properties, the frequency-weighted


input-normal realisation is not well-suited to practical lter design as it has a dense tap
matrix and wide spread of tap coecients. State-space simulations of lossless systems such
as LC ladders seem better in practice and even in theory.

117
4 ANALYTIC LIMITATIONS ON SIGMA-DELTA
MODULATOR PERFORMANCE

4.1 INTRODUCTION
We now turn our attention from active lters to sigma-delta analog to digital converters. A
model using an additive noise proved useful in chapter 2 for analysing dynamic range, and
again it does here. In this case the additive noise is quantisation error, and the additive
noise analysis is the well-known noise transfer function analysis. Like lters, sigma-delta
modulators are interesting to study because they have a non-trivial relationship between
overall performance (say, signal to noise ratio) and the performance of the sub-blocks
(notably quantiser resolution).

In chapter 2 much of the attention was on deriving fundamental limitations on the


achievable performance, so as to guide design e ort. This will again be the case { the Bode
integral results in this chapter are in some sense analogous to the dynamic range limitation.
Sections 4.2 and 4.3 introduce sigma-delta modulators and the noise transfer function
(NTF) analysis respectively. Section 4.4 derives fundamental limitations on the noise
transfer function. There are two main results { one which applies to all sigma-delta
modulators, and another which considers excess delay in the modulator feedback loop. The
rst has been discovered by other authors but we elaborate on it here; the second appears
to be new. Section 4.5 considers links with other theory, including the work in chapter 2.
Section 4.6 shows how these fundamental limitations can guide design.

4.2 THE SIGMA-DELTA MODULATOR


Sigma-delta modulators \work" by using a feedback loop to achieve resolution higher than
that of their internal quantiser, at the expense of less signal bandwidth than the Nyquist
rate. A popular structure is that of 1A in g. 4.1, namely a loop lter L(z ) and a quantiser

118
1A 2A

L(z) L(z)

L(z)

e e
1B 2B

r y r f d y
L(z) L(z)

L(z)

Figure 4.1: Noise-shaping structures: 1A: Sigma-delta modulator. 1B: Linearised model of sigma-
delta modulator. 2A: Predictive coder form derived by loop transformation. 2B: Linearised model of
predictive coder. (The predictive coder form is used later in the chapter.)

in the forward path of a feedback loop. The feedback corrects for quantiser error in the
same way that a feedback ampli er corrects for output stage distortion. This requires jL(z )j
to be large in-band in the same way that an opamp needs high open-loop gain at
frequencies of interest.

In a sigma-delta ADC the loop lter is an analog lter, often switched-capacitor and
hence basically discrete-time. However, di erent structures are possible { for instance, the
sigma-delta ADC described in chapter 5 has the loop lter in the feedback path and does
not use a conventional analog loop lter. The quantiser block actually consists of a ash
ADC driving a DAC (digital to analog converter). However, these can usually be lumped
together for theoretical analysis. The modulators in g. 4.1 are single-loop modulators, but
cascades of feedback loops (multiple-stage analog signal shaping or MASH modulators) are
also commercially important. Multiple-loop modulators could be analysed using vector
signals and a matrix loop lter, but that is not done here.

The term sigma-delta modulator is used here. The \sigma" and \delta" refer to

119
summing / integrating and subtracting / di erentiating blocks in early modulators.
\Modulator" seems to date back to when the raw sigma-delta ADC output was considered
for transmission along telephone lines and the like. It is used to describe both sigma-delta
ADC's and DAC's. (Today's sigma-delta DAC's are not really sigma-delta de -modulators.)
Others use \delta-sigma" in place of \sigma-delta" as it better describes the signal ow in
early modulators. The author considers both terms largely meaningless with today's loop
lters, and wants to encourage a way of thinking in which the loop lter is a generic
transfer function rather than an integrator or integrators. \Sigma-delta modulator" seems
preferable to \oversampling ADC" as it excludes oversampling without feedback.

Despite sigma-delta modulators having a variety of di erent structures (see (Tewksbury


& Hallock 1978) for some of the diversity), a consistent analysis using the so-called noise
transfer function can be used. This can be done by adapting the noise transfer function
analysis to the structure used, or by a loop transformation to a standard structure such as
that of 1A in g. 4.1. The loop transformation approach is mostly used here so that the
fundamental limitations only need to be derived once. There is nothing unusual in such
loop transformations; for example control systems engineers transform practically every
problem they encounter into a standard feedback loop form.

4.2.1 Trade-o between resolution and oversampling ratio


As mentioned, sigma-delta ADC's trade o bandwidth for resolution. The nature of this
trade-o has di erent explanations (Candy & Temes 1992) (Norsworthy, Schreier &
Temes 1997). Many authors consider it a reshaping of the quantisation noise spectrum away
from frequencies of interest. In the terms of the lter analysis in chapter 2, we might say
that we want a low noise gain from quantiser to output in the same way that we want a low
noise gain G(s) from state to output in a lter. There is more interpretation in section 4.4.2.

The oversampling ratio is de ned as the Nyquist bandwidth (half the sample rate)
divided by the useful signal bandwidth. It is obviously important for design to know what
oversampling ratio (OSR) is required for what resolution. (\Resolution", \signal to noise
ratio (SNR)" and \dynamic range" here have similar meanings.)

120
The conventional approach to determining the oversampling ratio required for a given
resolution is to try a range of loop lters L(z ). The sensible place to start is with rst and
second order loop lters, and then increase the loop lter order. For low-pass modulators
with single-bit quantisers and discrete-time loop lters this research program seems
reasonably complete. Good loop lters are known for high-order single-loop and low-order
multiple loop (MASH) modulators (Schreier 1993) (Norsworthy et al. 1997). The results are
usually presented as graphs of signal to (quantisation) noise ratio (SNR) versus OSR and
loop lter order. Bandpass modulators, multi-bit quantisers and continuous-time loop lters
are the subject of ongoing research in this context.

4.2.2 A subjective comment regarding use of \fundamental limitations"


There is no doubting the practical merit of the SNR versus OSR curves. Where they seem
somewhat unsatisfactory is that they o er little insight into why a certain OSR is needed
for a certain SNR. It seems reasonable to try to improve a circuit by understanding what is
fundamentally limiting its performance, and what form the mathematical fundamental
limitations take. For instance, the work in the previous chapter stemmed from knowing
that the mathematical limitations on feedback ampli ers allowed far higher loop gains at
wide bandwidths than were currently being achieved. The author would not have
attempted to use opamps at such high frequencies if he did not know this. The derivations
of such limitations are also useful to understand, as they often give useful design hints as to
what factors lead to sub-optimal performance.

This approach of considering fundamental limitations seems to appeals strongly to some


designers and design schools and not at all to others. It seems to be popular with Philips
(Annema 1999) (Tellegen 1952) (Hurkx 1997) and particularly MIT researchers (Thornton
et al. 1966) (Glasser 1988) (Mason 1954) and former students (Lee 1998). This is
understandable, for two reasons. Firstly, the limitations usually make assumptions of
limited validity. In this chapter the main assumption is that a linearised analysis (noise
transfer functions and the like) is close enough, and that stability can be reasonably
predicted from out-of-band noise transfer function. This largely ignores a lot of good work
on nonlinear stability, such as (Baird & Fiez 1994), (Schreier, Goodson & Zhang 1997) and

121
(Farrell & Feely 1998).

Secondly, the form of the limitations is often unfamiliar. In this chapter they are
integrals over frequency. These obviously appeal more to people who analyse feedback
ampli ers using Bode plots and loop-shaping rather than root loci, poles and zeroes and
polynomials. The author has a preference for techniques based on frequency response or
impulse response rather than on realisations with a particular order and structure of tap
coecients. Mathematics went through a similar phase; the impression is that Newton
thought of a function and its polynomial or power series as sort of the same thing
(Stillwell 1989, p. 101) (Bourbaki 1994, p. 205). However, a more general concept of a
function makes it easier to consider brick-wall ideal responses. These tend to be
conceptually useful. It also avoids not nding the optimum because the wrong
parameterisation has been chosen. Of course, to build a sigma-delta modulator you
eventually have to choose a structure; the point is that there are advantages to doing this
later rather than sooner. The practice of lter designers is better { the approximation
(order, Butterworth vs. Chebyshev etc.) is separate from the realisation (integrators,
inductors and capacitors). Sigma-delta modulator loop lters are lters (albeit with unusual
sensitivity requirements), not some separate entity possessing a \life force" all their own.

4.3 NOISE TRANSFER FUNCTION ANALYSIS


In section 2.2 we calculated the dynamic range of state-space systems using an an additive
noise model. We do a similar thing for sigma-delta modulators in this section. The
sigma-delta modulator in g. 4.1 1A is modelled as 1B. This structure plays a similar role
to the one the state-space equations did in chapter 2. Quantisation noise is represented as
an additive noise input E (z ).

For the modulator structure 1B in g. 4.1, it is readily derived that

Y (z) = 1 +L(Lz()z) R(z) + 1 + 1L(z) E (z) (4.1)


= G(z )R(z ) + H (z )E (z ) (4.2)

where G(z ) is called the signal transfer function, and H (z ) is called the noise transfer

122
function (NTF). The NTF is more important for deriving fundamental limitations, as the
signal transfer function can be changed by an external pre- lter.

(Again we use the convention that upper-case quantities such as those in (4.2) are the
Laplace transforms of the lower-case time-domain quantities of the same letter, here as
labelled in g. 4.1 1B. The notation in this chapter is similar to that in (Norsworthy
et al. 1997), but with some control systems symbols. H (z ) is used as a generic transfer
function in other chapters; in this chapter it will be used for NTF, although the limitations
derived will sometimes be applicable to other quantities as well. L(z ) is the loop lter, but
it includes any gain or delay in the quantiser or elsewhere in the loop { this is simpler for
the long derivations which follow.)

At this point, no approximations have been made { the NTF and additive quantisation
noise analysis is perfectly general and accurate. However, we know neither the quantisation
noise at the output, the maximum signal input, nor whether the modulator will be stable.
The output quantisation noise is the most straightforward. It is H (z )E (z ), so we need the
noise added at the quantiser E (z ). Provided that the quantiser does not overload, the noise
power can be calculated fairly accurately, as the quantisation noise has a peak amplitude of
one half of an LSB (least signi cant bit of the quantiser). More problematic is what
spectrum the quantisation noise E (z ) has.

4.3.1 Quantisation noise spectrum


The spectrum of quantisation noise is extensively discussed in (Gray 1990). It is well known
that E (z ) is distinctly non-white in many practical modulators (Norsworthy et al. 1997).
The worst problem is usually that E (z ) contains tones. This reduced spurii-free dynamic
range (SFDR) is particularly a problem in audio applications, and radio systems susceptible
to narrowband interference. (The e ect of tones on radios varies a lot { a software radio or
mobile telephone base-station will obviously have a problem if a tone knocks out a channel;
on the other hand OFDM and CDMA systems depend more on high SNR than high
SFDR.) Tones are usually worst in modulators with single-bit quantisers, low loop- lter
order, no dither and little thermal noise, and for DC or other highly deterministic inputs.

123
Despite this, the linear NTF analysis is still useful even when tones are present. As the
noise output is H (z )E (z ) and the power of E (z ) is limited (provided that the modulator
does not overload), the amplitude of a large tone is still limited by the NTF H (z ) at the
tone frequency.

Fortunately, for the sigma-delta modulator presented in the next chapter the quantiser
noise spectrum appears close to white in-band. This is presumably because it has a
multi-bit quantiser and signi cant thermal noise.

4.3.2 Stability and maximum input signal


As with the state-space lter theory, maximum input signal must be considered as well as
noise. Sigma-delta modulators have the rather peculiar feature that they can overload from
their own quantisation noise. Moreover, they are often conditionally stable. Therefore, the
maximum input signal problem cannot be decoupled from the noise problem like it can be
for lters. A number of theories of sigma-delta modulator stability have been developed
(Baird & Fiez 1994) (Hein & Zakhor 1993) (Schreier et al. 1997). Two are introduced here.
The rst, from (Schreier & Yang 1992) and (Kenney & Carley 1993), uses a 1-norm in the
time domain and guarantees stability. The second uses an 1-norm in the frequency
domain. It does not guarantee stability, but gives a reliable and less conservative indication
of it. It also ts very well with the integral limitations derived later in the chapter. The
best expositions of it are (Norsworthy et al. 1997) and (Schreier 1993), but it was widely
used before these.

jjhjj1 stability theory


The quantiser input is (R(z ) ; E (z ))G(z ) = (R(z ) ; E (z ))(1 ; H (z )). If we can guarantee
that this does not overload the quantiser, the modulator will be stable. For a quantiser with
n reconstruction levels with the usual equally-spaced decision and reconstruction levels, this
can be shown (Kenney & Carley 1993) (Schreier & Yang 1992) to occur if

jjhjj1 < n + 1 ; (n ; 1)jjrjj1 (4.3)

124
where jjrjj1 is the maximum input and

jjhjj1 = 1 + jh1 j + jh2 j + jh3 j +   


given the NTF impulse response

H (z) = 1 + z;1h1 + z;2 h2 + z;3h3 +    : (4.4)

Unfortunately, the sucient condition for stability (4.3) is rarely satis ed by practical
modulators with one-bit (n = 2) quantisers. A maximum input of jjrjj1 = 0:5 is typical,
requiring jjhjj1 < 2:5. However, modulators with jjhjj1 = 4:5 are usually stable (Schreier &
Yang 1992), and o er much wider bandwidth. Interestingly, quite competitive bandwidth
can be achieved with guaranteed stability { g. 4.2 shows an NTF with jjhjj1 = 2:5, but
optimisation results in a strange FIR NTF with most of the taps equal to zero which would
be near-impossible to realise in practice (except perhaps for sigma-delta DAC's where it
could be realised digitally). Using a more conventional NTF with jjhjj1 = 2:5 results in a
bandwidth of only one-tenth of that achieved with jjhjj1 = 4:5.

jjH jj1 stability theory


When making practical stability predictions, another norm of the NTF is often more
convenient and accurate than the 1-norm. This is the peak NTF in the frequency domain,
jjH (ej )jj1. For high-order single-loop modulators with one-bit quantisers, it is empirically
known that an jjH (ej )jj1 of 1.5 or 1.6 gives a good practical modulator (Schreier 1993)
(Norsworthy et al. 1997) with a maximum input of around one half of the quantiser range.
This frequency-domain condition will turn out to be very convenient when we consider
integral limitations later in this chapter. Of course, as this design guideline is empirical the
actual modulator being considered should be simulated to check the stability.

For multi-bit quantisers, the 1-norm approach guaranteeing stability is practical (Kenney
& Carley 1993). Indeed, it is used in the following chapter. However, it is still conservative.
Multi-bit quantisers are stable with NTF's which have much higher jjH (ej )jj1 than those
for one-bit quantisers. Many researchers in the eld appear to be unaware of this fact or its
importance. The literature rarely mentions it { (Ju & Vallancourt 1992) is the only study

125
10

10

20
NTF (dB)

30

40

50

60 Guaranteed stability
Conventional (OOB NTF = 1.6)

70
0 0.5 1 1.5 2 2.5 3
Frequency (rad)

Figure 4.2: An NTF optimised with guaranteed stability, H (z ) = 1 ; 1:1972z ;1 + 0:2321z ;8 ;


0:0523z ;26 + 0:0189z ;37, and a more conventional good 5th order NTF with passband zeroes spread
and an out-of-band gain of 1.6.

126
the author knows. A modulator with a 16-level (four-bit) quantiser and an out-of-band gain
of jjH (ej )jj1 = 5 is stable with an input of 0.8 of the quantiser range (Ju &
Vallancourt 1992), whereas (4.3) only guarantees its stability with inputs below 0.4 of the
quantiser range. A modulator with jjH (ej )jj1 = 5 achieves four times the signal bandwidth
of one with jjH (ej )jj1 = 1:5. The integral limitations derived shortly will show why.

Despite the fact that the jjH (ej )jj1 criterion is fairly reliable and appears close to
optimum for many practical cases, nding a better nonlinear stability criterion appears to
be the \holy grail" of many researchers. This work is important, but more because it is
likely to advance the art of nonlinear stability theory than because it will signi cantly
improve practical sigma-delta modulators. The author's opinion is that the existing state of
the art is fairly close to fundamental limits. This will be discussed further later in the
chapter.

Future developments are also likely to make a rigorous nonlinear stability analysis
including both quantisation noise and maximum input less practical. Multi-bit modulators
seem to be gaining in popularity for both ADC's and DAC's (with dynamic element
matching). The linearised model works better for multi-bit modulators as the quantisation
noise is whiter. (Checking for conditional stability with a Liapunov model is likely to
remain important { the point is that the noise and stability questions become separate for
many-level quantisers.) As speed and power consumption continue to be optimised,
quantisers and integrators will become less ideal and hence require a more complicated
mathematical model.

4.3.3 Applicability of the NTF analysis to the modulator investigated


The following chapter describes a sigma-delta ADC with a four-bit (n = 16) quantiser. The
linear analysis is particularly suitable for such a modulator. Indeed, even if a reliable and
tight nonlinear stability test was found, it would probably be of little bene t. This is for the
following reasons:

 The four-bit quantiser is likely to result in a quantisation noise spectrum which is


closer to white than that of a one-bit quantiser.
127
 The modulator has such a wide bandwidth that thermal noise is probably large
enough to have a dithering e ect.

 The comparators in the quantiser have signi cant o set. An ideal quantiser cannot be
assumed. Not much is lost by assuming that the additive quantisation noise is random.

 The modulator is absolutely stable and will recover from overloads which may result
from an occasional bad combination of quantisation errors over successive cycles.

 The modulator is largely intended for radio applications such as wireless LAN's. Such
an input signal is unlikely to have a tight jjrjj1 . Overloads will occur from time to
time. They are not catastrophic to the system.

All of these considerations favour a simple analysis based on average signal and
quantisation noise powers. The NTF and signal transfer function (which is roughly 0 dB
in-band) can be used to calculate an average power at the quantiser. Assuming a
normally-distributed signal, an overload probability can be calculated. The NTF and input
signal level can be chosen to make this acceptable. In the following chapter a 1-norm
stability criterion was used, but a power (2-norm) criterion should work as well or better.

4.4 INTEGRAL LIMITATIONS

4.4.1 Links with Bode's sensitivity integral


The NTF has the same expression 1=(1 + L(z )) as the quantity called sensitivity in control
theory. This is useful, because Bode (1945) derived a limitation on the integral of the log
magnitude of sensitivity over frequency. This is the way the author originally derived the
limitations which form the bulk of this chapter. Others have had the same idea
independently (Nawrocki et al. 1999) and earlier (Gerzon & Craven 1989), although they
seem not to have developed the links with linear prediction and the limitations regarding
excess loop delay which are presented later in this chapter. H (z ) denotes both NTF and
sensitivity in what follows. The equivalent continuous-time quantity will be
H (s) = 1=(1 + L(s)).
128
Bode's integral is
Z1
log jH (j!)jd!  0: (4.5)
;1
That is, the sensitivity of a control system or feedback ampli er cannot be small at all
frequencies. Distortion reduction can only be achieved at the expense of distortion increase
at other frequencies. For a feedback ampli er, this practically manifests itself as distortion
increase around the loop unity-gain frequency.

Bode's result is for continuous time. The discrete-time equivalent is in (Sung &
Hara 1988) and (Mohtadi 1990). It is
Z
log jH (ej )jd  0: (4.6)
;

The interpretation is the same { the sensitivity or NTF cannot be small at all
frequencies. A small in-band NTF (eg. {100 dB in a 16 bit sigma-delta ADC) can only be
o set by an NTF above 0 dB out of band (eg. jH (ej )j = 1:6, over a much wider
bandwidth).

The situation where (4.6) is equal to zero is desirable, so it is worth explaining when this
condition is met. It is met when the sensitivity is minimum-phase, ie. has no zeroes outside
the unit circle. For (4.5), the integral is satis ed with equality when the sensitivity has no
zeroes in the right-half plane; again this means that it is minimum-phase. (The integrals
could be negative if the sensitivity had poles outside the unit circle or in the right
half-plane, but this would mean that the system was closed-loop unstable and hence
useless.) NTF zeroes outside the unit circle correspond to loop lter poles outside the unit
circle, as H (z ) = 1=(1 + L(z )). That is, the loop lter is open-loop unstable.
Open-loop-unstable loop lters have been proposed as a way of making a modulator chaotic
so that it has less tonality (Schreier 1994). (This technique is viable despite a necessary
degradation in NTF at some frequency, as (4.6) is only a little greater than zero if the loop
lter poles are just outside the unit circle.)

129
10
Outofband
gain

10
NTF (dB)

20
These areas are equal

30

40

50
0 0.5 1 1.5 2 2.5 3
Frequency (rad)

Figure 4.3: The Bode integral states that the two shaded areas on this plot are equal. (The NTF
shown is a 5th order inverse-Chebyshev with out-of-band gain around 1.5 (+3.5 dB). A fairly moderate
in-band NTF ({40 dB) is chosen to make the diagram clearer.)

4.4.2 Interpretation
Eq. (4.6) has a good graphical description, namely that the areas above and below the
x-axis on a plot of NTF in decibels versus frequency on a linear scale are the same. Fig. 4.3
provides a good example. (A chaotic modulator would have slightly larger area above the
x-axis.)

This result is central to the author's understanding of noise shaping. It is important to


explain clearly why it is signi cant:

 It is possible to realise any NTF magnitude response which obeys the integral. The
only limitation on an NTF is that it must be causal and have rst sample of one (4.4).
We shall see shortly why the rst sample is always one and why it implies (4.6). It is
a basic fact of lter theory that any magnitude response jH (ej! )j is realisable by a
causal, stable transfer function. The Bode gain-phase relationships (Bode 1945) imply
that there is a certain associated minimum phase shift, but the phase of an NTF is
130
basically irrelevant. What matters is that any NTF magnitude response satisfying
(4.6) can be realised. (This is meant with an engineering level of rigour; brick-wall
lters can only be approximated arbitrarily closely, and presumably mathematicians
have more pathological counter-examples.)
 The integral makes it clear that what OSR you need for a given in-band NTF comes
down to two factors: how high the out-of-band NTF can be (whilst keeping acceptable
stability), and how much bandwidth is \wasted" in the transition between the desired
band (where the NTF is low) and the out-of-band region where the NTF is high
(jH (ej! )j = 1:6 or whatever). Fig. 4.3 should make this clearer. A sigma-delta loop
lter with high order and spread passband zeroes results in a modulator with less
OSR for a given resolution solely because it wastes less transition bandwidth.
 The conventional explanation (Candy & Temes 1992, p. 2{3) of noise-shaping in
terms of quantisation noise power on a linear scale is reasonable and correct, but gives
no indication that it is impossible to have, say, no quantisation noise at all over half of
the bandwidth. This is the strength of the Bode integral; the y-axis is a logarithmic
scale on which a {100 dB NTF is a lot di erent from a {50 dB NTF, and an NTF is
realisable if it satis es the integral.

Overall, the point is that the result is useful for design, not just of theoretical interest.

Provisos and possible objections


Of course, the limitations of the linear NTF analysis should be kept in mind. For instance,
whilst spreading the NTF zeroes should increase the useful bandwidth, sometimes it results
in tones at frequencies between the zeroes and there is little advantage (Thoss, Li &
Tenhunen 2000) (Vleugels 2000). Also, with a one-bit quantiser the loop lters cL(z ) (for a
positive constant c) and L(z ) result in identical modulators. The single comparator in the
quantiser does not measure magnitude. However, the NTF's 1=(1 + cL(z )) and 1=(1 + L(z ))
are di erent. The author's opinion with respect to these and similar limitations is as follows:

 Design improvement often proceeds along the lines of Popper's model of science
(Popper 1959): come up with a hypothesis regarding design any way you can, and
131
then test the predictions it makes. The NTF analysis is good for making predictions.
For example, the next chapter is based on the predictions that OSR can be decreased
by increasing the out-of-band gain and using a multiple-bit quantiser to provide
stability, and the prediction that an FIR loop lter will work as well as a conventional
loop lter provided that it gives an appropriate NTF. Testing predictions about
sigma-delta modulator performance is usually the easy part { just run a behavioral
simulation with realistic non-idealities for 100 Monte-Carlo runs of 10000 cycles. It is
not so important that the NTF analysis is sometimes of limited accuracy; what
matters is that it points the designer in the correct direction about 80% of the time.
 The linearised theory is at least applicable to multiple-bit quantisers, which have a
well-de ned quantiser gain and less problems with tones.

There is another legitimate objection, with respect to treating loop lters L(z ) as just a
lter transfer function to be approximated and realised. This is that the sensitivity and
dynamic range requirements are quite di erent to those for a conventional lter. A lter
like the anti-alias lter in the previous chapter might typically have 1{2 dB tolerable
passband error. Stopband errors of 10{20 dB are quite acceptable provided that the
stopband stays below the minimum stopband attenuation. Stopband phase shift is
irrelevant. By contrast, a sigma-delta modulator loop lter can have large passband gain
error provided that the gain stays suciently high. However, in the \stopband" of L(z )
where the NTF is high, 0.5 dB gain errors or 20 phase errors are quite signi cant. These
di erent sensitivity requirements are presumably why cascades of integrators are used. As
an aside, the author suspects that ADC's with continuous-time loop lters often achieve
considerably less bandwidth than is possible because L(z ) is not suciently well controlled
in the \stopband". Dynamic range requirements are also peculiar { in-band distortion is
reduced by feedback, and out-of-band distortion is mostly removed by the digital post- lter.

The main counter to this objection is that it still seems excessive to throw away ninety
years of lter realisation theory. The realisation step of lter design has to be adapted to
the di erent sensitivity requirements. However the approximation step, performed on the
NTF H (z ) rather than the loop lter L(z ), is still classical. The NTF can be speci ed with
rectangular acceptable / unacceptable regions on the magnitude versus frequency plot in
132
10

10
NTF (dB)

20

30

40

50
0 0.5 1 1.5 2 2.5 3
Frequency (rad)

Figure 4.4: Choosing the NTF is just a lter design problem. Note that the excluded region (shaded)
is on one side only { the Bode integral provides the other constraint. A moderate in-band NTF
speci cation ({40 dB) is again used for clarity.

the same way as any other lter transfer function { see g. 4.4.

Practical NTF calculation


It is worth describing how an arbitrary NTF and loop lter are calculated. This is also
discussed in (Norsworthy et al. 1997). A lter prototype A(z ) with the desired NTF shape
(but probably a gain error, eg. out-of-band gain = 1) is found. For example, a high-pass
inverse-Chebyshev A(z ) will result in a conventional low-pass modulator with stopband
zeroes. (Experience shows that an inverse-Chebyshev prototype is usually a good choice; an
Elliptic prototype wastes less transition bandwidth but usually requires lower out-of-band
NTF for acceptable stability.) The NTF has rst sample of one, which is satis ed by
choosing H (z ) = A(z )=A(1). The loop lter L(z ) is calculated from H (z ) = 1=(1 + L(z )).

133
In practice this is done by equating numerator and denominator polynomials:

H (z) = 1 + 1L(z)
L(z) = 1 ;HH (z )
(z )
NL (z) = 1 ; DNHH ((zz)) = DH (z) ; NH (z)
DL (z) NH (z) NH (z)
DH (z)

The Matlab commands are shown below. Here the lter prototype is a 5th order
inverse-Chebyshev with {100 dB stopband. The \0.016" is the useful bandwidth (OSR =
1/0.016 = 63) and sets the out-of-band gain. It is adjusted to get jjH (ej )jj1 = 1:6 or
whatever is desired for stability. The resulting NTF is about +4 dB (1.6) out-of-band and
{96 dB ({100 dB + 4 dB) over the useful bandwidth.

% Prototype

[numa,dena]=cheby2(5,100,0.016,'high')

% Scale to get h0 = 1 in NTF H(z) = h0 + h1 z^-1 + h2 z^-2 + ...

numh = numa / numa(1)

denh = dena

% Check out-of-band NTF visually

freqz(numh,denh)

% Loop filter L(z)

numl = denh - numh

denl = numh

The main point of including this is to show that it is straightforward to convert an


arbitrary lter prototype to an NTF and loop lter transfer function.

Bitstream information capacity interpretation


A single-loop sigma-delta modulator with a single-bit quantiser which provides 16 bit
resolution in-band (98 dB SNR) must have an OSR of at least 16. Otherwise the output
bitstream would not have sucient information capacity to represent the input signal. This

134
sounds banal, but thinking about it can give an indication of whether existing modulators
can be signi cantly improved by yet to be discovered techniques.

The OSR of the 16 bit example above is 63. Therefore, about 1=4 of the bitstream
information capacity represents the input signal. What does the other 3=4 represent?

Some bits represent the input signal in the transition band. (In many applications a
pre- lter will make this close to zero, but unless the modulator has a peculiar structure
which overloads for very small out-of-band inputs, it still consumes bits to represent it.)
This can be estimated by integrating the SNR in bits versus frequency. This is logarithmic
in magnitude and linear in frequency like the Bode integral (4.6) at high SNR. Other
authors (Gerzon & Craven 1989) (Nawrocki et al. 1999) have therefore given the Bode
integral an information capacity interpretation. This seems reasonable, although it should
be noted that the two expressions behave di erently at low SNR. Integrating over the
region from  = 0:016 (the passband edge, NTF = {96dB) to  = 0:08 (where the NTF is
around 0 dB) indicates that around 30% of the output bitstream capacity represents the
transition band.

This accounts for 25% (in-band) + 30% (transition) = 55% of the bitstream capacity.
Presumably the other 45% is related to the stopband where the NTF is close to 1.6.
However, it is dicult to calculate exactly how much information is present, as small
di erences between the NTF and SNR lead to large di erences in the calculated
information content.

The \wasted" 30% transition bandwidth can reduced by using a higher-order loop lter
(at the expense of complexity and increased power consumption). How about the 45%
suspected to be representing the stopband? If a sigma-delta modulator was found which did
not \waste" this bandwidth, and the SNR agreed with the NTF, the modulator would have
an out-of-band NTF of a little over two. (Increased out-of-band NTF corresponds to
increased useful bandwidth.) This is a useful link between the bitstream information
capacity and the Bode integral { information capacity seems to guarantee that an NTF
which is too large out-of-band cannot result in a stable modulator. Practical modulators
with high-order loop lters are only stable with out-of-band NTF's up to about 1.75. This

135
limit (rather than a little over two) is probably the consequence of the raggedness of the
mapping from the previous state and input to the current state. The stability boundaries
(with respect to maximum input, state or parameters) of existing high-order NTF's tend to
be complicated fractal shapes (Schreier et al. 1997) (Farrell & Feely 1998) (Schreier &
Yang 1992); presumably if a loop lter could be found which more neatly mapped the
previous state and input to the current state, the maximum out-of-band NTF for stability
would increase.

To sum up this discussion, the bitstream information capacity limitation can either be
considered to be manifested mathematically by the Bode integral (as Nawrocki et al. (1999)
and Gerzon & Craven (1989) tend to do), or to impose a necessary limitation on the
out-of-band NTF for stability. The author tends to think in the second way, perhaps out of
familiarity with the mathematics behind the Bode integral, or perhaps because of a stronger
preference for the linearised analysis from experience largely with multi-bit modulators
where it has worked well.

4.4.3 Derivation of integrals


It is helpful to understand where the Bode integral comes from. Sometimes examining the
derivation of a fundamental limitation o ers possibilities by showing a premise which can be
negated by a di erent architecture or circuit. In this case it is more because it leads on to
an additional restriction involving excess loop delay.

The loop lter is

L(z) = l1 z;1 + l2 z;2 + l3z;3 +   


It must be strictly proper (no l0 term) or else there would be a delayless path around the
feedback loop. (The notation here is that the quantiser is instantaneous, and there is no
separate z ;1 or gain factor elsewhere in the loop. This avoids having to carry a complicated
loop gain expression through the derivation.) The loop lter L(z ) is related to the NTF
H (z) by

H (z) = 1 + 1L(z) :

136
Rearranging and equating coecients of z ;n ,

(1 + L(z ))H (z ) = 1
(1 + l1 z ;1 + l2 z ;2 +    )(h0 + h1 z ;1 + h2 z ;2 +    ) = 1
h0 + (l1 h0 + h1 )z;1 + (l2 h0 + l1h1 + h2 )z;2 +    = 1

h0 = 1 (4.7)
l1 h0 + h1 = 0 ) h1 = ;l1: (4.8)

Eq. (4.7) has been mentioned; this is where it comes from. Eq. (4.8) will be used when
we consider an excess loop delay.

These coecient relationships can be converted to integral limitations. There are a


number of ways to proceed. Bode used contour integration, but a series expansion is used
here. This follows Kolmogorov's work on linear prediction (Shiryayev 1992, vol. 2 p.
272{280).

Let log H (z ) have the series expansion

log H (z ) = B (z ) = b0 + b1 z ;1 + b2 z ;2 +    ;

so

H (z) = eB(z) = eb0 eb1 z;1 eb2 z;2    :


Using the power series for the exponential function,
2 ;2 2 ;4
H (z) = eb0 (1 + b1z;1 + b1z2 +    )(1 + b2z;2 + b2z2 +    )   
2
= eb0 + b eb0 z ;1 + (b + b1 )eb0 z ;2 +   
1 2 2
= h0 + h1 z ;1 + h2 z ;2 +   

Equating coecients,

h0 = eb0 ) b0 = log h0
h1 = b1 eb0 ) b1 = hh1 :
0

137
Around the unit circle z = ej ,

log H (ej ) = b0 + b1 e;j +   


= log h0 + hh1 e;j +   
0
Now

log jH (ej )j2 = log H (e;j )H (ej )


= log H (e;j ) + log H (ej )
=    + hh1 e;j + 2 log h0 + hh1 ej +   
0 0
Equating the Fourier series coecients,
Z
log jH (ej )j2 d = 4 log h0 (4.9)
;
Z
ej log jH (ej )j2 d = 2h
h0
1 (4.10)
;
Z
e;j log jH (ej )j2 d = 2h1
h0 : (4.11)
;
Adding the last two gives
Z
cos  log jH (ej )j2 d = 2h1
h : (4.12)
;
0

These relationships look too good to be true for arbitrary H (z ), and indeed this is the case.
They rely on the series for B (z ) converging on the unit circle z = ej . This will occur if
B (z) = log H (z) is analytic for jz;1 j < 1. (Remember that a power series in a complex
variable converges out to the radius where it rst hits a pole. The power series here are in
z;1, not z.) For B (z) to be analytic in jz;1 j < 1, H (z) must have no poles or zeroes in
jz;1 j < 1 (outside the unit circle). The condition for the integrals above to hold is therefore
that H (z ) has no poles or zeroes outside the unit circle.

Let us apply these results to an H (z ) which is a noise transfer function. If the


sigma-delta modulator is closed-loop stable, as any practical modulator would be, H (z ) has
no poles outside the unit circle. For H (z ) to have no zeroes outside the unit circle, it has
been shown previously that the loop lter L(z ) must be stable. (Recall the discussion about
138
chaotic modulators.) If these conditions hold, substituting (4.7) and (4.8) into (4.9) and
(4.12) and using log jH (ej )j2 = 2 log jH (ej )j results in
Z
log jH (ej )j d = 0 (4.13)
;
Z
cos  log jH (ej )j d = ;l1 : (4.14)
;

Eq. (4.13) is the Bode integral (4.6).

4.4.4 E ect of excess delay


Eq. (4.14) is a new integral. It will be interpreted in this subsection. Recall that l1 is the
rst sample of the impulse response of the loop lter L(z ) = l1 z ;1 + l2 z ;2 +    . Also, in
the notation used any delay around the modulator loop is included in the loop lter. The
quantiser is assumed to be instantaneous; if it has delay then that is included in L(z ).

At high clock rates, it would often be convenient to allow more than one cycle of latency
in the loop lter and quantiser combination. The probability of comparator metastability is
greatly reduced by passing the comparator output through a few latches. In
continuous-time modulators, DAC pulses can be made cleaner. Latency was a signi cant
design limitation for the sigma-delta ADC presented in the next chapter.

If the total latency around the modulator loop is two samples rather than one sample,
then the loop lter (including any quantiser delay) is L(z ) = l2 z ;2 + l3 z ;3 +    , ie. l1 = 0.
Therefore (4.14) becomes
Z
cos  log jH (ej )j d = 0: (4.15)
;

The question is then whether this integral is satis ed by good conventional NTF's, or at
least something reasonably close. For low-pass modulators, it usually is not. The log of the
NTF is very negative at low frequencies. cos  is close to one there. The log of the NTF is
positive at frequencies closer to Nyquist, where cos  is negative. Both of these result in a
negative integrand in (4.15), so it will not integrate to zero. Conventional low-pass NTF's

139
like the one in g. 4.3 therefore cannot be implemented with an excess sample of latency in
the loop.

For bandpass modulators with a centre frequency of fs=4, the situation is better. (fs is
the clock frequency. fs=4 is  = =2.) If the NTF is symmetric around fs=4, (4.15) will be
satis ed. The loop lter in this case is of the form L(z ) = l2 z ;2 + l4 z ;4 + l6 z ;6 +    . Some
asymmetric NTF's are also possible using nonzero l3 etc.

Are there low-pass modulator NTF's with excess loop delay (no l1 term) and close to
optimum performance? To satisfy both (4.15) and (4.13), a little thought shows that the
NTF needs to be more positive at frequencies just above the passband, and less positive or
negative at high frequencies, than if just (4.13) is satis ed. This uneven out-of-band gain
will push up jjH (ej )jj1 for a given useful bandwidth, almost certainly degrading stability.
Alternatively, for the same stability the useful bandwidth will decrease.

There does not appear to be any NTF which maintains jjH (ej )jj1 and satis es (4.13)
and (4.15) without \wasting" about half of the bandwidth. The best way to satisfy (4.15) is
for the NTF to be equal to jjH (ej )jj1 for all out-of-band frequencies except for a deep
notch right at the Nyquist frequency ( = ). This deep notch must have nearly the same
area as that of the useful bandwidth at low frequency. Then to satisfy (4.13) the useful
bandwidth must be halved compared to the case with no excess delay.

Continuous-time loop lters


The desire to use excess loop delay is particularly strong for modulators with
continuous-time loop lters and very high clock rates. It is therefore worth explaining how
these limitations relate to continuous-time loop lters.

An equivalent discrete-time loop gain L(z ) can be calculated by evaluating the samples
of the DAC pulse ltered by the continuous-time loop lter fed back to the quantiser at
successive sampling instants (Cherry & Snelgrove 1999b). If the overall latency around the
modulator loop is more than one clock period, then the rst sample l1 is necessarily zero.
The NTF will therefore be limited by (4.15). This has the same implications as with
discrete-time loop lters, namely that either stability or bandwidth su ers.

140
4.4.5 Excess delay with unstable loop lters
Eq. (4.14) and (4.13) assume that the loop lter is stable. There is therefore a possibility
that an unstable loop lter (chaotic modulator) might help. Recall that an unstable loop
lter leads to a non-minimum-phase NTF.

To analyse this case, the NTF is factorised into a minimum-phase part Hmp (z ) and an
allpass part Hap (z ):

H (z) = Hmp(z)Hap (z):


This approach is standard { mathematicians refer to the minimum-phase part as \outer"
and the allpass part as \inner". The minimum-phase part satis es (4.13) and (4.14). The
allpass part can be factorised into a product of allpass sections
Y
Hap(z) = Hap(i)(z) (4.16)
i
where
;1
Hap(i) (z) = kiz ;;11 : (4.17)
ki ; z
Hap(z) is called a Blaschke product in mathematics. If the ki are complex, they occur in
conjugate pairs. The allpass terms can be integrated using contour integration, giving
Z
log jHap(i) (ej )j d =  log jki j (4.18)
;
Z 2
cos  log jHap(i) (ej )j d = (jki2jk ; 1) : (4.19)
i
;
Eq. (4.18) is zero or positive. This is where the inequality in (4.6) comes from. The integral
contributions from the allpass factors (4.17) add to (4.13) and (4.14), as taking log jH (z )j
converts the products (4.16) and (4.17) into sums:

141
Y
H (z) = Hmp(z) Hap(i)(z) (4.20)
i X
log jH (z )j = log jHmp (z )j + log jHap(i) (z )j (4.21)
i
Z X
log jH (ej )j d =  log jkij (4.22)
; i
Z X (jki j2 ; 1)
cos  log jH (ej )j d = ;l1 + 2ki : (4.23)
; i

A positive (4.22) is detrimental as the NTF has to be larger out of band for the same
in-band noise reduction. However, this can allow (4.23) to be negative despite the excess
loop delay (l1 = 0). This allows a low-pass modulator with excess loop delay to have lower
peak out-of-band NTF (better stability) or lower in-band quantisation noise. The question
is therefore whether the bene ts from the (possibly) negative (4.23) outweigh the
detrimental e ects of the positive (4.22).

For ki close to {1, (4.22) increases by the same amount amount that (4.23) decreases.
The net e ect of this is that no more \noise reduction area" is available to improve in-band
NTF or stability. Positive or complex ki only make matters worse. However, for large
negative real ki (4.23) decreases faster than (4.22) increases. An unstable loop lter could
therefore be bene cial. Most bene t is derived from a single allpass section with large
negative k1 rather than multiple allpass terms. (This bene t is only to recover the losses
caused by excess loop delay; without excess loop delay an unstable loop lter is always
detrimental as far as NTF is concerned.) To use a large negative k1 , the modulator must be
stable with sizable out-of-band NTF, as (4.22) is signi cantly positive. The situation looks
more hopeful for multiple-bit quantisers which are stable with greater out-of-band NTF.

Sample non-minimum-phase NTF


The easiest way to experiment with such NTF's is to take an existing minimum-phase NTF
Hmp(z), and multiply it by an allpass factor. (In practice, this may be slightly sub-optimal
as it raises the degree of the NTF by one without narrowing the NTF transition band.) A
raw allpass factor will not leave the h0 = 1 coecient of the NTF unchanged, so it is
142
necessary to multiply the allpass factor by ;k1 :
 k z;1 ; 1 
H (z) = ;k1 1 ;1 Hmp(z)
k1 ; z
;1
= 1 ; k11 z ;1 Hmp (z ):
1 ; k1 z
This multiplication by ;k1 is how the degradation due due to the excess loop delay
manifests itself. The NTF is multiplied by jk1 j at all frequencies, satisfying (4.22). This is
primarily a problem out-of-band as it degrades stability. To compensate, a Hmp with
smaller jjHmp (ej )jj1 must be chosen. The key question is how much smaller jjHmp (ej )jj1
is for l1 = 0.

This was tried for two cases; jjH (ej )jj1 = 1:6 as might be used for a modulator with
single-bit quantiser, and jjH (ej )jj1 = 5 as might be used with a four-bit quantiser. A fth
order low-pass modulator with spaced zeroes and {96 dB passband was used. The results
were:

jjH(ej )jj1 1.6 5


OSR with no excess delay 63 17
jjHmp(ej )jj1 1.26 2.37
k1 {1.26 {2.15
OSR with excess delay 125 33

The loss of bandwidth is very close to 50% in both cases. This is also the approximate
bandwidth loss without unstable zeroes. Unless a chaotic modulator was desired to decrease
tones, an unstable loop lter is of no bene t.

4.5 CONNECTIONS WITH OTHER THEORY


The Bode and excess delay integrals may appear to be an obscure corner of circuit theory.
However, they have received the attention of a surprising number of notable authors. This
section presents two links with other theory. The rst is with linear prediction. The author
has found it to be conceptually useful, particularly when considering excess loop delay. The
second is a discussion of how the theory in this chapter relates to other system theory, such
143
as the state-space theory in chapter 2. In this case, the links are tantalising but rather hard
to pin down.

4.5.1 Linear predictor interpretation


There is another interpretation of the integral limitations which particularly explains why
excess delay is detrimental. This is to consider the modulator as a predictive coder. This
approach is also useful because linear prediction has been studied by leading
mathematicians such as Kolmogorov, Wiener and Szego, and is therefore well understood.

Older papers such as (Candy 1985) and (Tewksbury & Hallock 1978) consider predictive
coding. Such works tend to distinguish between predictive coding and noise shaping
modulators, but Tewksbury & Hallock (1978) clearly make the point discussed below that
the two are interchangeable by using a pre-emphasis (input) lter. The links are roughly
that di erent modulator structures have similar noise transfer functions but di erent signal
transfer functions. Predictive coding is presumably mentioned less today because the
sigma-delta modulator structure has been found to be superior in practice (it is less
sensitive to loop- lter gain errors and the like).

The easiest way to link modern sigma-delta modulator structures to predictive coding is
with a loop transformation (2A, g. 4.1). The loop ltering function is split between an
input lter and a feedback lter. The input lter shapes the signal input so that it is much
larger in-band. The feedback loop then has to predict this weighted input to keep the input
to the quantiser small. The overall resolution of the converter is the quantiser resolution
multiplied by how accurately the signal is predicted. This is because the prediction is
subtracted from the input, leaving only the residual to quantise. (There is another form of
the predictive coding modulator in which the prediction is made from the reconstructed
signal, not the error signal. This can also be obtained by loop transformation.)

Note that this loop transformation leads to diculties with unstable loop lters. The
case of loop lters with poles on the unit circle (including at DC) would make
mathematicians nervous, but engineers would handle it by saying that the integrators or
resonators have a very high but nite gain so the poles are just inside the unit circle. This

144
case was of practical importance for delta modulators, and was handled by including a
small integrator \leak" to make the input lter stable (Wooley 2000).

Theoretical limitations on how accurately a coloured noise input can be predicted are
well known (Grenander & Szego 1958) (Scharf 1991). The usual assumptions that the
coloured noise is stationary and stochastic are made. The optimum one-step-ahead
predictor of a coloured noise F (z ) with power spectral density jF (ej )j2 has a prediction
error (Grenander & Szego 1958, p. 181)
0 Z 1
E (jxi ; xij2 ) = exp @ 1 log jF (ej )j2 dA
2 (4.24)
;

where xi is the prediction of xi from previous inputs xi;1 , xi;2 ,    and E () is expectation.
The optimum predictor is the one which whitens the residual error (Bode & Shannon 1950).

The logarithmic integral suggests links with (4.13), and indeed the two are closely
related. For the linearised model 2B of the transformed modulator 2A in g. 4.1, if the
input is R(z ) then the signal being predicted is F (z ) = L(z )R(z ). From 2B in g. 4.1, the
prediction error is
D(z) = 1 +L(Lz()z) R(z):
The optimum predictor whitens the prediction error, D(z ) = c for a constant c. This results
in
L(z) R(z) = c
1 + L(z )
R(z) = c 1 +L(Lz()z)
F (z) = L(z)R(z)
= c (1 + L(z )):
Eq. (4.24) then becomes
0 Z 1
c2 = exp @ 1 log jc (1 + L(ej ))j2 dA
2
;
1 Z 
2 log c = 2 log c + 2 log j1 + L(ej )j2 d
;
Z
0 = 21 log jH (ej )j2 d
;
145
using the de nition of the NTF, H (z ) = 1=(1 + L(z )).

This is an unnecessarily complicated way of deriving (4.13). The derivation of the NTF
integrals in this chapter is actually an intermediate \spectral factorisation" step in the
proof of (4.24) (Scharf 1991). The linear prediction interpretation is most useful
conceptually rather than for calculation.

The Bode integral (4.13) is thus related to one-step-ahead prediction. With a sample of
excess delay in the modulator loop, the problem becomes two-step-ahead prediction. The
multi-step prediction error was calculated by Kolmogorov (Shiryayev 1992, vol. 2, p. 274)
(and likely also by Wold or Wiener).

The main bene t in relating excess delay to two-step-ahead prediction is conceptual. It


becomes a lot more reasonable to expect that excess delay cannot be fully compensated for.
The case where one sample of excess delay is not detrimental, namely fs=4 bandpass
modulators, is instructive. A sample of excess delay is not detrimental because with fs =4
centre frequency, the samples go I; Q; ;I; ;Q;    and the in-phase signal I is no use in
predicting the following quadrature sample Q. By contrast, for low-pass inputs the previous
sample is usually the most important for predicting the current sample. Predictive coders
will therefore perform more poorly (require larger quantisation steps and hence have lower
SNR) without it. Degradation of the performance of sigma-delta modulators then follows
using the loop transformation described above.

Three-step-ahead prediction could be considered, but the mathematics becomes


complicated and it is rare to need so much latency. It is also possible to apply the the
results for prediction of continuous-time processes (Grenander & Szego 1958, p. 196) to
sigma-delta modulators with continuous-time loop lters. However, the results do not seem
to have a form particularly conducive to understanding or useful for design.

4.5.2 Links with other system theory


The work in this chapter shares many features with the state-space theory in chapter 2. In
particular:

146
 Quantisation noise is analysed as an additive noise in the same way that thermal noise
(and perhaps distortion) is in the state-space theory.

 The noise transfer function H (z) in this chapter plays the same role as the noise gain
G(s) in the state-space theory.
 Fundamental limitations in both theories follow from requiring the transfer function
to be causal. Recall that the Hankel operator was a map from past inputs to future
outputs, and that the Bode integral followed from the causality of the noise transfer
function.

It is peculiar then that the two theories have such a di erent mathematical form. Perhaps
we are just using di erent notation for the same thing?

The author has tried to reconcile the two theories but has made little progress. They are
not identical { most notably, the quantisation noise is only added at one point, whereas the
state-space theory had a noise source adding to each state. This may be why the Bode
integral is a weaker logarithmic restriction on the dynamic range. A future theory which
could deal with noise added to only some of the states would surely be useful.

It should be noted that others have almost certainly also tried to nd links between
Bode integrals and state-space realisation. The Bode integral and the frequency-weighted
balanced realisation are generally considered to be amongst the most important results in
robust control theory. One of the nest recent textbooks, (Zhou 1996), has the two theories
in successive chapters. No link is made between them.

Logarithmic integrals also occur in broadband impedance matching theory (Bode 1945)
(Fano 1950). Interestingly, Helton (1981) has applied Hankel operators to this theory.
Helton connects the theories by stating that the logarithmic integral uses one basis for the
Hilbert spaces involved, and the Hankel operator theory uses another. The logarithmic
integral basis is notable for giving limitations in a nite number of dimensions, but is
dicult to use for automatic computations.

147
4.6 DESIGN IMPLICATIONS OF THE INTEGRAL LIMITATIONS
For standard low-pass modulators with one-bit quantisers, most of the design limitations
are well enough known that the theory developed here does not predict any new avenues to
explore. The most useful feature of the integral limitations is perhaps the negative result
that resolution for a given OSR can only be improved by increasing the out-of-band noise
gain, or by sharpening the loop lter response.

It is worth mentioning one other point which may have occurred to the reader. It is
known that spreading the NTF zeroes widens the bandwidth. The NTF zeroes are still on
the unit circle. Aren't the deep notches in the NTF a waste of noise reduction area {
wouldn't it be better for the NTF to be constant across the passband? This is the case, but
a atter in-band NTF would require a higher-order loop lter for a given transition
bandwidth so there is little advantage.

4.6.1 Multiple-loop modulators


The theory so far concentrates on single-loop modulators. However, the best sigma-delta
modulators are often cascades of low-order loops (the MASH architecture) (Norsworthy
et al. 1997). For multiple-loop modulators, performance is usually limited by a combination
of the NTF of the rst loop and the degree of matching between loops ( g. 4.5). By
applying the integral limitations to the rst loop, useful conclusions can still be drawn.

Suppose that the rst loop of a multiple-loop modulator has {36 dB or less quantisation
noise in-band. The second and subsequent stages cancel this noise. If the second stage DAC
has a 0.1% ({60 dB) gain error due to capacitor mismatch and the like, the lowest
quantisation noise which is possible is around {36 dB { 60 dB = {96 dB. To satisfy the Bode
integral (4.6) with an out-of-band gain of jjH (ej )jj1 = 1:6 (+4 dB), the OSR must be
 
1 + 4 dB 1 ; 1
 
;36 dB OSR OSR  0 dB
OSR  10

The more general result is that the SNR achieved by a multiple-loop modulator is limited
to that achieved by a single-loop modulator in accordance with the Bode integral, plus the
148
20

20

40
dB

60

80

100

120
0 0.5 1 1.5 2 2.5 3
Frequency (rad)

Figure 4.5: NTF's for the rst loop of a multiple-loop (MASH) modulator. The fundamental limit
to performance is usually the NTF (solid curves), of the rst loop, multiplied by the mismatch error
(here 0.1% or {60 dB) between the rst and second loop (dotted curves). Note that the second-order
loop- lter is not much worse than the fth order loop- lter.

149
degree of matching.

This analysis predicts that the best performance would be to use a rst loop with
high-order loop lter and spaced NTF zeroes to achieve the {36 dB NTF over as wide a
band as possible. Practical design deviates from this for two reasons. Firstly, only a
second-order loop lter is usually used in the rst loop as it is absolutely stable and simpler.
It is also stable with higher out-of-band gain that jjH (ej )jj1 = 1:6 (Schreier & Yang 1992),
which at low OSR almost compensates for the larger wasted transition bandwidth. (The
transition band is a lot narrower if only {36 dB NTF is required rather than {96 dB NTF.)
Secondly, spaced NTF zeroes seem to cause larger tones (Thoss et al. 2000) (Vleugels 2000).
It may be that this would be less of a problem if a higher-order loop lter was used.

4.6.2 Multiple-bit quantisers


For multiple-bit quantisers, the theory seems more useful. The main prediction is that large
out-of-band NTF's are the way to get high bandwidth at very low OSR (4{10). (Ju &
Vallancourt 1992) have shown that jjH (ej )jj1 = 5 (+14 dB) is stable with a four-bit
quantiser. The stability is for inputs up to 0.8 of the maximum DAC output so presumably
it is robust. To get 80 dB SNR with a four-bit quantiser, an NTF around {60 dB is required.
A fth-order loop lter with spaced zeroes and +14 dB out-of-band gain can an achieve this
at OSR=8.

Conventional wisdom (Norsworthy et al. 1997, p. 244) is that the SNR bene t of
multi-bit quantisers comes from the 6 dB / bit reduction in raw quantisation noise. It is
widely known that a multi-bit quantiser gives a more stable modulator, but this is seen only
as a incidental bonus. At most, an NTF of the form H (z ) = (1 ; z ;1 )n with n greater than
two might be used.

The integral limitations make quite a di erent prediction. To maximise bandwidth,


out-of-band NTF must be aggressively maximised, constrained of course by stability for a
reasonable maximum input. The NTF zeroes should be spread across the passband.
Fortunately, multiple-bit modulators usually have few problems with tones and their
stability correlates well with the linear NTF-based model.

150
4.6.3 Compensation for excess loop delay
The restrictions resulting from excess loop delay have been calculated in detail in previous
sections. The aim here is to discuss the practical e ects, particularly whether a performance
degradation is inevitable or whether it can be compensated for by changing the loop lter.
The theoretical results will also be checked against simulations in the literature. As the
sections on excess loop delay were detailed and mathematical, a reminder of the results is in
order:

 A modulator with one sample of excess delay around the modulator loop (l1 = 0 in
L(z) = l1z;1 + l2z;2 + l3 z;3    ) and no loop lter poles outside the unit circle obeys
(4.13) and (4.15). To satisfy both of these, the NTF of a low-pass modulator is
inevitably distorted in such a way that either the peak out-of-band NTF increases or
the useful bandwidth decreases. The optimum NTF with excess loop delay is one
which has a large wide notch near the Nyquist frequency to satisfy (4.15) with
minimum loss of \noise reduction area" in (4.13). This still halves the useful
bandwidth.

 A modulator with or without loop lter poles outside the unit circle obeys (4.22) and
(4.23). Loop lter poles outside the unit circle give a chaotic modulator
(Schreier 1994). The limitation resulting from (4.22) and (4.23) is dicult to assess
exactly, but sample NTF calculations show that there appears to be negligible bene t
in using an unstable loop lter.

 For continuous-time loop lters, there should be no loss of bandwidth provided that
some feedback gets around the loop in less than one clock period.

 Interpreting the sigma-delta modulator as a predictive coder, it seems reasonable that


performance should necessarily be degraded by excess delay, as the modulator has to
predict its input further into the future. For loop lter poles outside the unit circle,
this interpretation is not really valid as the process being predicted is unstable.

 Bandpass modulators with fs=4 centre frequency can achieve the same performance
with a sample of excess delay.

151
Considering these theoretical results, it does not seem possible to fully compensate for
excess delay in low-pass modulators. It is reasonable to expect that a loop lter optimised
for use with excess delay will perform better than one optimised for no excess delay.
However, the theory shows that the best sigma-delta modulator with an excess delay of one
sample will still only have about half the useful bandwidth of the best one without excess
delay.

Reading papers such as (Cherry & Snelgrove 1999b) and (Yahia, Benabes &
Kielbasa 2001), it is easy to come away with the impression that excess delay can be
compensated for without bandwidth reduction. Which is wrong, the theory here or the
literature? Closely examining the results and discussion in (Cherry & Snelgrove 1999b), this
author cannot see anything which contradicts the theory presented here. Cherry &
Snelgrove (1999b) present simulations which show the following:

 Excess delay can be fully compensated for in bandpass modulators with fs=4 centre
frequency.

 Low-pass modulators can be stable with excess delay, but it is only achieved at the
expense of bandwidth or resolution by using lower jjH (ej )jj1 .

 A return-to-zero DAC works well with less than 1/2 cycle of delay. As Cherry &
Snelgrove (1999b) point out, this allows an l1 to be realised.

All of these results agree with the analysis here. It is only the conclusion \excess delay
can be rendered e ectively a non-problem in high-speed CT M's" (Cherry &
Snelgrove 1999b, p. 388) which is contentious. Certainly a loop lter which has been
optimised for excess delay (by making the NTF droop at high frequencies as discussed
above) is better than one optimised without loop delay, but there is still an inherent loss of
about 50% of the bandwidth for a given resolution. To compensate for excess delay, (Cherry
& Snelgrove 1999b) recommend using an NTF with jjH (ej )jj1 = 1:3 before the excess
delay (of one sample) is added. Matlab NTF calculations show that this requires OSR=112
for {96 dB NTF with a fth order modulator with spaced zeroes, as compared to OSR=63
with no excess delay and jjH (ej )jj1 = 1:6.

152
There could conceivably be nonlinear e ects which degrade the NTF but preserve the
SNR. This is not the case { the simulations in (Cherry & Snelgrove 1999b, g. 14) show a
clear di erence in SNR between an NTF with jjH (ej )jj1 = 1:6 and no excess delay, and an
NTF of jjH (ej )jj1 = 1:3 with excess delay. For a 5th order modulator with spaced zeroes
and an OSR of 64, the dynamic range is 19.2 bits with jjH (ej )jj1 = 1:6 and no excess
delay, and 15 bits with jjH (ej )jj1 = 1:3 (before excess delay) and one sample of excess
delay. These results are for continuous-time modulators, but given the NTF equivalence
there should be little di erence for discrete-time modulators.

Similarly, (Yahia et al. 2001, p. 344) claims that \Theoretically, high loop delay can be
achieved and non-ideal DAC can be used without e ect on the performance of the
modulator, if some special feedback schematics are used". However, in the body of their
paper they acknowledge that a feedback term necessary for compensating for quantiser
delay (their \location #1") degrades the NTF and produces a resolution loss.

To end this discussion on a more positive note, it should be acknowledged that di erent
modulator architectures are still useful in combating excess loop delay. The theory shows
that for continuous-time low-pass modulators, some form of feedback must get around the
loop in one sample to realise the l1 in the loop lter L(z ) = l1 z ;1 + l2 z ;2 +    . There are
good and bad ways of doing this. The best that the author has seen is to have an extra
DAC which feeds directly to the quantiser input rather than through the integrators (Luh,
Choma & Draper 2000). This is good because the output of that DAC only needs to be
ready after one sample of delay, not after just one half of a sample of delay as would be
required with a return-to-zero DAC feeding an integrator. The other DAC's feeding the
integrators can have longer latency. A similar \express" feedback path is used in the ADC
in the next chapter.

The other positive outcome which is clari ed by the theory is that there is no reason not
to operate fs =4 bandpass modulators with a sample of excess delay.

153
4.7 CONCLUSIONS
The noise transfer function (NTF) is constrained by the Bode integral (4.6). This integral
limitation is clearly expressed as a conservation of area on a plot of NTF in decibels versus
frequency on a linear scale. The limitation also has interpretations in terms of information
capacity of the output bitstream, and in terms of predictive coding. The following
conclusions follow from the integral limitation:

 The oversampling ratio (OSR) for a given resolution depends on two factors: how
large the out-of-band NTF can be made whilst retaining acceptable stability, and how
much \noise reduction area" is wasted in transition regions of intermediate NTF. The
traditional curves relating resolution to OSR and loop lter order are not mysterious;
they follow simply from these factors.

 Modulators with multiple-bit quantisers necessarily need to be operated with high


out-of-band NTF's to achieve the widest bandwidth for a given resolution. The
advantage that they are stable with higher out-of-band NTF is as important or more
important than the 6dB reduction in raw quantisation noise with each extra quantiser
bit.

Another integral limitation (4.15) shows that a sample of excess delay around the
modulator loop roughly halves the useful bandwidth achievable for low-pass modulators.
\Achievable" includes attempts to compensate for excess loop delay. The result is derived
in discrete-time, but it applies to the important case of modulators with continuous-time
loop lters if no DAC pulse returns to the quantiser in time for the next sample. Bandpass
modulators with fs=4 centre frequency need not lose bandwidth. Simulations in the
literature are consistent with this.

154
5 A 520MHZ SIGMA-DELTA ADC WITH A FIR
DAC LOOP FILTER

5.1 INTRODUCTION
This chapter describes a high-speed, low-resolution sigma-delta ADC which was fabricated
in a 0.25 m process. The aim was to build a converter with bandwidth beyond that
practical with switched-capacitor circuits. The ADC achieves 60 MHz signal bandwidth,
albeit at only 40 dB signal to noise and distortion ratio. The ADC uses a novel
nite-impulse-response (FIR) digital to analog converter (DAC) structure in place of a
conventional loop lter.

This section introduces the FIR DAC and the reasons for using it. Section 5.2 discusses
choice of noise transfer function. The theory in the previous chapter suggested using a
multi-bit quantiser and high out-of-band NTF; the converter presented uses a four-bit
quantiser and +17 dB peak out-of-band NTF. Section 5.3 describes the architecture at the
block level, and section 5.4 describes it at the circuit level. Section 5.5 describes the testing
procedure and test printed circuit board. This section is included as testing
high-performance ADC's is non-trivial. The measured results are given in section 5.6. The
nal parts of the chapter describe possible modi cations to the architecture, and compare it
with other approaches such as pipelined converters.

5.1.1 Argument for FIR DAC loop lters


The previous chapter shows that a sigma-delta modulator can achieve high resolution at
low oversampling ratio (OSR) if the noise transfer function (NTF) is high out-of-band. A
multi-bit quantiser is needed for stability with high out-of-band NTF. (Intuitively, a
multi-bit quantiser has less quantisation noise, so it takes a higher out-of-band gain for the
quantisation noise to overload it.) Multi-bit quantisers also o er increased resolution even

155
before the out-of-band gain is increased, simply because the quantisation steps are smaller.

The de nition of NTF is identical to that of sensitivity. A high NTF implies that a small
gain or phase error can make the modulator unstable. To realise the advantages of high
out-of-band NTF, the loop lter L(z ) must be accurately realised.

At clock frequencies below 200 MHz, switched-capacitor circuits allow an accurate loop
lter to be realised. This approach has been heavily studied (Norsworthy et al. 1997)
(Candy & Temes 1992), so the aim here was to try to operate at clock frequencies above
those possible with today's switched-capacitor circuits. These frequencies are currently the
domain of sigma-delta modulators with continuous-time loop lters.

Almost all modulators built with continuous-time loop lters have used one-bit
quantisers and hence out-of-band NTF around 1.4{1.7. An argument for one-bit quantisers
is that they are inherently linear, so the converter resolution is not limited by transistor
matching. However, such converters rarely achieve resolutions beyond the matching level of
MOS transistors (often 10-12 bits), so there is no need for one-bit quantisers. If higher
resolutions are needed, there is also an opportunity to use dynamic element matching.
One-bit quantisers also use less power, but in CMOS the transistors are small enough that
3-4 bit quantisers are quite practical.

A more serious reason why sigma-delta modulators with continuous-time loop lters
cannot be operated with out-of-band NTF's of +15 dB or so is that the continuous-time loop
lters simply are not accurate enough. A gain controlled to 0.5 dB and a phase controlled to
10 is required. It seems unlikely that structures such as Q-enhanced LC resonators (Gao &
Snelgrove 1998) or integrators based on simple di erential pairs could achieve this level of
control. The opamp-RC lter technology in chapter 3 might be suitable. This lter work
was done after the ADC described here was fabricated, so it was not an option at the time.
A di erent modulator structure, using a FIR DAC, was used. The relative merits of FIR
DAC and continuous-time loop lters are discussed later in the chapter.

156
Loop transformation

1. Conventional sigma-delta modulator 2. Transformed "delta" modulator

L(z) L(z)

L(z)

1A. Continuous-time loop filter 2A. FIR DAC architecture


L(s) L(s) T/H

FIR
DAC
DAC

2B. Architecture with digital filter


L(s) T/H

DSP
DAC
L(z)

Figure 5.1: Wideband sigma-delta ADC architectures

157
D Q D Q D Q D Q

DAC DAC DAC DAC

Figure 5.2: FIR DAC

5.1.2 FIR DAC loop lter


The sigma-delta modulator presented in this chapter realises a nite-impulse-response
(FIR) transfer function in the DAC. The DAC current fed back is a weighted sum of the
last 12 quantiser outputs. The quantiser has four bits, or 16 levels. The DAC's are unary
(thermometer-coded), so in total there are 16  12 = 192 switched current sources in the
FIR DAC.

How this can replace the loop lter in a conventional sigma-delta ADC can be seen from
a loop transformation ( g. 5.1). In a conventional sigma-delta ADC, the loop lter lters
both the input signal and the fed-back signal. In this modulator, these two functions are
handled by separate lters { an analog pre- lter and the FIR DAC. The requirements on
the analog pre- lter are similar to those of an anti-alias lter for a Nyquist-rate converter.
For simplicity, it was omitted from the test chip { an opamp-RC lter like that in chapter 3
would easily meet the requirements. It is important to point out that there is a big
advantage in making this split { the analog pre- lter is no longer in the feedback loop, so
NTF and stability do not depend on its accuracy. Its phase response can be arbitrary, and
its gain only needs to be at to within a few decibels in the passband and to be suciently
low out-of-band. The FIR DAC does need an accurate transfer function, but DAC's and
digital delays are naturally more accurate.

Another way of looking at this loop transformation is that the converter is a delta
modulator (albeit with unusual NTF) rather than a sigma-delta modulator (Wooley 2000).

158
5.2 CHOICE OF NOISE TRANSFER FUNCTION
The NTF is constrained by the loop lter being of the form

L(z) = l1 z;1 + l2z;2 +    + lnz;n:


The NTF was numerically optimised in Matlab using the Nelder-Mead simplex algorithm
optimiser fmins. A variety of FIR lengths n were tried before settling on n = 12. Higher
OSR's require higher n. A one-norm constraint on the noise gain (Kenney & Carley 1993)
(Schreier & Yang 1992) was used, as this o ers a guarantee of stability. This was discussed
in chapter 4. The in-band NTF was minimised subject to this, for an OSR of around ve.
(A range of OSR's were tried; OSR=4{6 was most likely to exceed the state-of-the-art in
CMOS ADC's as >100 Ms=s is dicult for pipelined converters.)

High-order single-loop sigma-delta modulators are usually conditionally stable, and


hence require a reset. In early simulations this modulator was no exception. Implementing
a reset turned out to be surprisingly dicult. During reset the loop gain has to be reduced
by over 30 dB so that the loop gain is below 0 dB in-band. Reset switch capacitance
increased the error ampli er settling time. Error ampli er o set voltage and the state of the
FIR DAC meant that recovery from reset was somewhat uncertain. It was therefore a
pleasant surprise when one choice of FIR taps revealed an absolutely-stable modulator in
circuit simulation. What had happened was that the maximum phase lag in this simulation
was slightly less, around ;170 (except for frequencies where the loop gain was less than
0 dB) as compared to ;200 in others.

An additional constraint was therefore added to the NTF optimisation to make sure that
the phase of L(z ) could be no worse than ;160 at frequencies where the gain exceeded
0 dB. All resulting modulators were absolutely stable, and the performance degradation
from the additional constraint was small. (The condition (Bode 1945) that an ampli er is
absolutely stable if the phase lag does not exceed 180 is an approximation { see
(Vidal 1969). Counter-examples led to the development of rigorous criteria such as the
circle and Popov tests for continuous-time, and the Jury-Lee test for discrete-time
(Vidal 1969). Nevertheless, Matlab simulations show that the < 180 phase lag criterion is
fairly accurate for the sigma-delta modulators examined.)
159
20

10

0
NTF (dB)

10

20

30

40
0 0.5 1 1.5 2 2.5 3
Angular Frequency (rad)

Figure 5.3: Noise Transfer Function of the described ADC

5.3 ARCHITECTURE DETAILS


The architecture is basically that of 2A in g. 5.1. A more detailed stage-by-stage diagram
is g. 5.6.

5.3.1 Error ampli er


In-band, the error signal which is fed to the quantiser is about {30 dB relative to the input
and FIR DAC signals. This is like any feedback loop - the signal fed back closely matches
the input signal. Now the four-bit quantiser operates with signal levels around 1 V, so the
error signal must be ampli ed before being quantised. (It is conceivable that the input and
DAC currents, each around 3 mA, could be subtracted, and the di erence across a resistor
of around 10 k
fed straight to the quantiser. However, speed and DAC accuracy would
greatly su er from using such a high load impedance.)

The error ampli er gain used was around 25 dB. The error ampli er consists of cascaded
di erential pairs. As only the error signal is being processed, the error ampli er does not

160
40

30
L(z) Gain (dB)

20

10

10
0 0.5 1 1.5 2 2.5 3
Angular Frequency (rad)

0
L(z) Phase (Degrees)

45

90

135

180

225
0 0.5 1 1.5 2 2.5 3
Angular Frequency (rad)

Figure 5.4: Loop gain of the described ADC. Note that the phase lag does not exceed 180 when the
gain is > 0 dB, resulting in an absolutely-stable modulator.

161
6

0
0 2 4 6 8 10 12

Figure 5.5: Impulse response of the FIR DAC, L(z )

need to operate with the full precision of the ADC. Around 30 dB of dynamic range is
sucient. Therefore, feedback ampli ers are not required. This is vital considering that the
error ampli er has to settle in less than one cycle at 520 MHz, ie. 2 ns.

Replica reference ladder


The disadvantage with simple di erential pairs is that they do not have accurately known
gain. This problem was overcome by feeding the reference voltages for the quantiser
through replica di erential pairs ( g. 5.6), giving a closely-controlled error ampli er /
quantiser overall gain.

Track and hold


The input signal is continuous-time, and the signal fed back from the FIR DAC is
discrete-time. At some point the input signal must be sampled. This could be done by the
quantiser, but the error signal would have a very high dv=dt by that point. This would
make both error ampli er and quantiser design very dicult.

A track and hold at the input could also be used. But it is better to sample the error

162
signal rather than the input signal, as this reduces the required track and hold dynamic
range by the in-band loop gain (30 dB). This is a considerable advantage of this
architecture over Nyquist ADC's. The track and hold dynamic range required is only
around 30 dB, so simple series CMOS switches and hold capacitors can be used.

The track and hold was eventually placed in the middle of the error ampli er. Some gain
(around 8 dB here) before the track and hold decreases its thermal noise contribution. Too
much makes the dv=dt too high. Thermal noise is not a limiting factor for this converter,
but in the future this advantage over Nyquist ADC's may prove to be of considerable
importance.

5.3.2 Latency and rst-tap problems


The greatest design diculty is that an error signal must be fed back with a latency of only
one cycle. This is not a particular drawback of the architecture used here; it is inherent to
all conventional single-loop low-pass sigma-delta modulators. We saw in chapter 4 that
omitting a l1 z ;1 term from the loop gain L(z ) = l1 z ;1 + l2 z ;2 +    roughly halves
achievable bandwidth.

Getting a signal through an error ampli er, track and hold, quantiser, DAC and
subtractor in 2 ns (1/520 MHz) is genuinely dicult. The comparators in the quantiser, and
the error ampli er each basically take half a cycle (1 ns) to settle, and that does not leave
time for the track and hold. It was therefore necessary to feed the l1 z ;1 FIR tap back after
the track and hold. This increases the error signal which includes only the
l2 z;2 +    + l12 z;12 terms, requiring a little more linearity in the early stages of the error
ampli er. It also leads to gain mismatches between the l1 z ;1 term and the other ones. The
DC gain errors can be overcome using replica structures, but the dynamic mismatch is a
problem.

Overall, the speed of the converter is limited by the error ampli er settling time,
particularly for the l1 z ;1 FIR tap. At 520 MHz the error ampli er only settles to about 85%
of its nal value for the rst tap, and 90% for the others. The resulting loop gain is therefore
L(z) = 0:85l1 z;1 + 0:9l2 z;2 + 0:9l3 z;3 +    + 0:9l12 z;12 :
163
This needs to be taken into account when choosing the tap coecients l1 -l12 . Typical
process variation should not be enough to alter these numbers enough to cause instability
(about 10% di erence is needed).

5.4 CIRCUIT AND LAYOUT DETAILS


The circuits used for the error ampli er, quantiser, shift register and DAC are conventional.
All signals are fully di erential to maximise power supply rejection and common-mode
rejection. It would be practically impossible to use single-ended signals as the package has
signi cant inductance and digital power supply noise is present. A stage-by-stage diagram
of the ADC is shown in g. 5.6, and a die micrograph is shown in g. 5.7.

5.4.1 Error ampli er


The stages in the error ampli er are nMOS di erential pairs with pMOS diode-connected
transistors as loads ( g. 5.6). By making the pMOS devices narrower than the nMOS
devices, a gain per stage of around 8 dB is achieved. Diode-connected pMOS transistors
give slightly less distortion than resistors as loads, because the nonlinearity in the nMOS
di erential pair and pMOS loads tends to cancel. The only exception is the nal stage which
drives the comparators, which is of the same form as those used in the lter in chapter 3.
The resistors in its common-mode feedback are replicas of the reference ladder ( g. 5.6). It
also has resistive source-degeneration to decrease distortion. Each signal (input, main
feedback and rst-tap feedback) passes through identical cascades of di erential pairs. The
track and hold is after the rst error ampli er stage. It uses pMOS switches with dummy
pMOS devices to cancel charge injection (these may not be necessary). The hold capacitors
are the gate capacitance of the second stage { this is sub-optimal with respect to linearity,
but the dynamic range required is only 30 dB or so and speed is critical.

164
Error Amplifier Comparator FIR DAC

Reference

the error ampli er.


Iref Ladder

Source degen.
Comp Latch Latch Latch Latch

Vin T/H

Cascoded Source degen.

DAC DAC DAC


First tap
Common-gate fed back
after T/H
15 more comparators
and FIR DAC's

Common-gate

165
9 more DAC's

48/0.24 120/0.24
Transistors and resistors vary from stage to stage.
These sizes are typical for the error amplifier;
250 comparators are around 1/10 width.
120/0.24

120/0.24 Typical bias 1mA / differential pair for error amplifier,


100uA / differential pair for comparators.

Figure 5.6: ADC block diagram and typical gain stages. Note the identical parallel signal paths in
Bypass caps.

FIR DAC

Error
amplifier

Clock
spine

Comparators
Clock buffer
Bypass
caps.

Shift register for


output storage

Figure 5.7: ADC die micrograph

166
Reset

Vdd

In+ In-

Out+ Out-

Figure 5.8: Comparator and latch (blocks labelled \Comp" and \Latch" in g. 5.6). (Note that
the comparator is preceded by di erential pairs, not shown.) Comparator transistors are all 3/0.24,
except for 1.44/0.24 in the inverters.

5.4.2 Quantiser
The quantiser is a four-bit ash ADC. As the state of the art is six bits at above 1 Gs=s in a
few hundred milliwatts (Choi & Abidi 2001), a quantiser which achieves four bits at
520 Ms=s in 30 mW is not particularly dicult. The comparator consists of a pair of
di erential pairs (to compare a di erential reference to a di erential input), followed by a
di erential pair as second-stage preamp, a comparator of the same topology as that in
(Kim, Kusayanagi & Abidi 1997) and a di erential conventional CMOS latch { see g. 5.8.
Comparators with di erential-pair input were preferred to those using switches as they
avoid injecting charge onto the reference ladder, and the o set voltage of a di erential pair
is acceptable for a four bit quantiser.

5.4.3 Shift register and DAC


The shift register uses a latch built in the conventional static CMOS manner from four
CMOS switches and four CMOS inverters ( g. 5.8). It has di erential input and di erential
output so that the DAC is driven with a symmetrical signal. Switch transistors are
0.72/0.24 and inverter transistors are 1.44/0.24 (nMOS = pMOS). A single-phase clock is
distributed and the complement generated locally to avoid skew between the clock and its

167
complement. A large clock spine runs along the middle of the converter ( g. 5.7). It is
18 m wide, 1.2 mm long, and consists of metal four and ve in parallel. The clock bu er
drives both the converter and a 128 stage shift register which allows output data to be
stored and read out later at a much lower rate. The clock bu er transistors are 3840/0.24
pMOS and 1920/0.24 nMOS.

The clock input from o -chip is di erential. This greatly reduces problems with clock
jitter from ground bounce. A peculiar circuit based on the author's work with quadrature
generators (Harrison & Weste 1999) was used, although a more conventional ECL to CMOS
converter would have worked equally well. There is also a synchronous clock enable which
allows the clock to be latched whilst changing over to a lower frequency to clock out the 128
stage test shift register data.

The DAC cell is conventional, with 1.44/0.24 nMOS switches and an uncascoded 3/3
nMOS current source transistor. The DAC outputs operate into an nMOS common-gate
bu er, so a cascoded current source is not required. (This common-gate bu er is needed to
provide a high output resistance at the nodes where the input signal and FIR DAC
feedback are summed into a resistor. This resistor matches the reference current input, so
any transistor output conductance would cause a gain error.)

The DAC cells operate at currents from 0.5 A to 30 A, depending on which tap ( g.
5.5) they are 1/16 of. The 12 reference biases are generated by a small feedback ampli er
which provides the same bias to the DAC cell as it feeds back to the reference input pin.
More will be said later about this circuit as it caused problems during testing.

5.5 TEST CIRCUIT BOARD AND TESTING


Testing of an ADC with such a high clock rate is non-trivial. The clock rate was too high
for any available logic analyser to store the digital output. Achieving full dynamic range
required careful board layout and consideration during integrated circuit design. Useful
information on test PCB layout for high-speed ADC's is provided in ADC manufacturer
data sheets, such as (Analog Devices 2000). The test printed circuit board (PCB) is shown

168
12 tap coefficients Signal
balun
Bias
currents

Clock
balun

Run /
store in
shift register Switch
debounce DAC Chokes to separate
output analog, digital Vdd

Figure 5.9: Test PCB { top

in g. 5.9 and g. 5.10.

5.5.1 Data and clock inputs


The data and clock inputs are both di erential. Baluns consisting of a twisted pair around
a ferrite toroid were used to generate the di erential signal from a 50
BNC input. A small
RC low-pass lter with a cuto around 50 MHz was placed right next to the signal input
pins, to minimise high-frequency noise on the inputs. This was needed because with no
input pre- lter on chip, the ADC was sensitive to overload from even small out-of-band
signals. The clock power was fairly large so as to minimise jitter due to digital noise and
ground bounce. About +5 dBm was the minimum for reasonable operation; +20 dBm would
destroy the clock inputs. There were a few bad clock frequencies where the minimum clock
input for correct operation was larger and the in-band noise oor was higher, most notably
from 550{600 MHz. This suggests some resonances in the clock or power lines.

169
16 resistor DAC
Bypass
caps.

Resistors
to stop
tap bias
resonances

Input
attenuator
+filter

Figure 5.10: Test PCB { bottom

5.5.2 Digital output


It was originally intended that the output data would be stored in the 128 bit test shift
register and clocked out later. This worked, but a 128 point FFT has little frequency
resolution. (Only 128 samples were available due to chip area restrictions on the shift
register. The 16 levels were stored as thermometer code rather than four bit binary so that
bubbles in the quantiser could be seen and so that ground bounce would be linear with
output.) It was also possible to use the 16 digital outputs as primitive unary
(thermometer-coded) DAC using 16 o -chip 4.7 k
chip resistors and a 51
termination.
This seemed to give the same SNDR, so in later testing it was used instead. The primitive
DAC output was fed directly into a spectrum analyser. For future converters, an on-chip
DAC seems natural { it makes testing fast and simple, and as ADC performance lags DAC
performance it is not dicult to design the DAC.

170
5.5.3 Power rails
The chip had analog VDD , digital VDD , and a shared ground. The logic behind sharing
analog and digital ground was that they are connected by less than 1
to each other via
the substrate anyway, and so may as well share pins. This resulted from a failure to
understand that digital VSS is usually kept separate from the substrate. The author
presumed that keeping digital VSS separate would cause latch-up problems. This
misunderstanding was not revealed by reading several papers on substrate noise, so the
author points it out to textbook and expository article writers. Nevertheless, the shared
digital ground was not a serious problem as it allowed more pins (four, limited by chip
perimeter) to be used for ground. The use of a thermometer-decoded digital output also
helped as it meant that ground bounce was linear with the input signal. This means that it
should not cause nonlinearity if coupled back into the analog circuitry.

The power rails were bypassed with o -chip 1 F surface-mount ceramic capacitors as
close as possible to the IC package. A low-pro le IC socket was used to minimise
inductance. The circuit is not too sensitive to decoupling as it has about 1 nF of on-chip
decoupling capacitance. Package inductance was included in simulation. It was originally
feared that the on-chip decoupling capacitance would resonate with the package inductance
to such a degree that o -chip series resistors of about 0.5
in the power rails would be
required. However, simulation of the nMOS / pMOS decoupling capacitors used showed
that their Q was only about two and there would not be a problem. The analog and digital
VDD were fed from the same LM317 regulator (set to 2.5 V) via RF chokes and ceramic and
tantalum chip decoupling capacitors.

5.5.4 Reference currents


The quantiser resistor ladder reference current and error ampli er and comparator bias
currents were supplied by LM334 current sources. The quantiser reference is a di erential
input so it did not have ground bounce diculties. The bias inputs have fairly low input
resistance and high input capacitance and did not cause problems. (On-chip bypassing using
gate capacitance should have been used however.) There is also a gate bias voltage used by

171
the common-gate stages in the DAC bu er and input ampli er. It had on-chip bypassing as
it had oscillated in simulations with package inductance. It caused no further problems.

The pins which caused the most problems were the reference currents which control the
FIR tap coecients. As mentioned earlier, these have small feedback ampli ers in them.
These have a bandwidth of around 100 MHz. The reference currents were set by series
trimpots, with a xed resistance in series with the trimpot used to measure the reference
current. The PCB traces to the trimpots tended to resonate. There are 12 of them, so the
probability of one resonating at the clock frequency was high. The feedback ampli ers run
at di erent currents, which leads to them having di erent gain-bandwidths. They are
slightly underdamped, heightening the resonances excited by ground-bounce energy. The
problem was severe enough that the rst PCB did not work at clock frequencies above
100 MHz, and had good and bad clock frequencies. The resonances do not cause
nonlinearity, but the change the tap coecients to such a degree that the sigma-delta
modulator oscillates.

The resonances in the FIR reference current lines were overcome on the second PCB by
including 1 k
series chip resistors close to the IC package. Nevertheless, they may still
have had an e ect { the degradation of the SNDR as the clock rate was increased from
520 MHz to 760 MHz (the maximum for correct operation) was uneven. (At frequencies
above 520 MHz, the taps, particularly the rst, also had to be retuned.)

5.6 MEASURED PERFORMANCE


The best performance that might be expected is as follows. An ideal four-bit quantiser has
a signal to noise ratio of 26 dB. The NTF averages about {25 dB over the passband at an
OSR of 5. There is a little additional noise reduction from oversampling due to the noise
being spread over a wider bandwidth. At an OSR of 5, this is 10 log10 5 = 7 dB. To avoid
overloading the quantiser from a combination of input signal and quantisation noise, the
maximum input is around {6 dB. Therefore the overall SNR is 26 dB + 25 dB + 7 dB { 6 dB
= 52 dB.

172
The measured SNDR (signal to noise and distortion ratio) was 40 dB across a 60 MHz
bandwidth (an OSR of 4.3) when clocked at 520 MHz. It was not signi cantly better at
higher OSR. The di erence between this and ideal is largely due to two factors:

1. The comparators in the quantiser have signi cant o set. (Comparator transistor size
is a trade-o between o set and capacitive load on the error ampli er.) It is estimated
that this reduces the SNDR of the quantiser from 26 dB to about 20 dB.

2. Error ampli er distortion reduces the input level for optimum SNDR by about 6 dB.
Separate feedback for the rst FIR tap worsens the problem as the error signal is
larger before the rst FIR tap feedback is subtracted. This problem could have been
picked up in simulation, but it was dicult to spot as simulations had to run for a day
to get even a rough FFT.

Measured performance is shown in g. 5.11, g. 5.12 and the table below. A clock
frequency of 520 MHz was generally used as the clock jitter tended to vary with frequency
and 520 MHz was the highest frequency before it hit a bad patch around 550{600 MHz,
above which the converter was running too fast for the error ampli er to settle properly.
(The problem was presumed to be clock jitter because the observed e ect was an in-band
noise oor which depended on clock amplitude.) An SNDR of 30 dB could be obtained with
760 MHz clock and some retuning of taps. The noise oor does not vary with input signal
(except at very high input amplitudes). In-band tones do not seem to be a problem { the
combination of a multi-bit quantiser and thermal noise presumably breaks them up.
Intermodulation distortion (and hence SNDR) does not vary much with signal frequency
below 50 MHz. Above 60 MHz, the maximum input falls o quickly with increasing input
frequency. (Remember that the required pre- lter was not implemented.)

173
Figure 5.11: Two-tone measurement (38 MHz, 40 MHz) { 10 MHz=div: horizontal, 10 dB=div: vertical

Speci cation Performance


Signal bandwidth 60 MHz
Peak SNDR 40 dB
SFDR (at peak SNDR) 49 dB
Clock freq. 520 MHz
OSR 4.3
Process 0.25 m digital CMOS
Area 1.5 mm2
Power DAC 10 mW
Error amp. 100 mW
Comparators 30 mW
Clock bu er 40 mW
Shift register 40 mW
Total 220 mW (88 mA @ 2.5 V)

5.7 ALTERNATIVE ARCHITECTURES


The FIR DAC sigma-delta ADC fabricated is a replacement for conventional Nyquist-rate
ADC's. However, sigma-delta ADC's may be more attractive with alternative system
architectures. For radio receivers, other authors have suggested digitising the intermediate
174
0

10

20
dB

30

40

50

60
0 50 100 150 200 250
Frequency (MHz)

Figure 5.12: Single-tone measurement taken using the 128 tap on-chip shift register. The 128 point
FFT does not have enough frequency resolution for a meaningful two-tone test.

frequency (IF) signal with a bandpass sigma-delta ADC (Pellon 1998) (Gao &
Snelgrove 1998). This avoids baseband I/Q mismatch and icker noise problems, although
sensitivity to ADC clock jitter is greater. It is more attractive for oversampling converters
than Nyquist-rate converters as the clock rate need not be increased { only the passband
where the NTF is low needs to be moved. Section 5.7.1 considers how to achieve this with
the FIR DAC architecture.

There is also the idea of a \software radio" in which almost all signal processing is digital
(and preferably in software). The FIR DAC sigma-delta ADC is interesting in this context
as the loop lter is more digital in nature. Section 5.7.2 considers a variant in which the
loop lter is a digital lter rather than a FIR DAC.

In light of the performance achieved by the active lter in chapter 3, it is also worth
examining the possibilities with continuous-time loop lters. This is discussed in section
5.7.3

175
5.7.1 Bandpass option
Only minor changes would be required to change from a low-pass to a band-pass NTF. The
natural centre frequency is fs=4 = 130 MHz, as a symmetrical response around that
frequency allows half of the FIR DAC taps to be omitted. Most signi cantly, the
troublesome rst tap is not required; the loop lter is of the form

L(z) = l2 z;2 + l4 z;4 + l6z;6 +   

This would allow the sample rate to be increased somewhat, or the converter made more
resistant to process variation at 520 Ms=s. The only down-side is that the dv=dt in the error
ampli er due to the continuous-time input signal being compared with discrete-time
feedback would increase. This should not be a limiting factor however as the track and hold
only requires around 30 dB dynamic range; at worst the track and hold may need to be
moved closer to the input of the error ampli er. With a little more work on the existing
imperfections (o set averaging in the quantiser and reduced error ampli er distortion by
removing the rst tap problem) to increase the SNDR by about 10 dB, and an NTF which
gave 10 dB more in-band noise reduction in return for an OSR of six rather than four, the
converter could clock at 800 Ms=s and digitise a bandwidth of 70 MHz centred at 200 MHz
with 60 dB SNDR. This IF digitisation would be quite attractive for software radios and
wireless LAN's.

5.7.2 Digital lter


In the ADC designed, the FIR DAC is used to shape the NTF. There is another option,
namely digital ltering. That is, rather than a quantiser and a FIR DAC in the feedback
loop, there would be the quantiser, some digital signal processing, and a conventional DAC
(2B, g. 5.1). As with the FIR DAC architecture, an input pre- lter before the feedback
would be needed.

This is a promising architecture, as it follows the trend towards increasing use of digital
signal processing (DSP) in place of analog signal processing which is expected as CMOS
feature size is reduced. Using the linear predictor interpretation in chapter 4, the analog

176
pre- lter provides an input signal which is predictable. The quantiser, DSP and DAC then
predict the next sample of the ltered input signal. Subtracting this prediction allows the
quantiser input to be kept small (say, {30 dB) so that it can be quantised accurately with a
low-resolution (say, four-bit) quantiser. A software radio built using this approach would be
an interesting proposition.

For the converter designed, the aim was to achieve a clock rate and bandwidth beyond
that possible with switched-capacitor sigma-delta ADC's. This made the use of DSP less
attractive. The latency limitations derived in chapter 4 limit pipelining. An FIR or IIR
(in nite impulse response) digital lter could be built at 520 MHz { witness what
microprocessor arithmetic units can do { but it would be a major project in itself. About
the only approach which seemed simple and practical enough to work at 520 MHz was if a
simple ROM-based state machine could be used.

A ROM-based state machine seemed possible, so it was investigated further. (In


practice, a static RAM would probably have been used in place of a ROM to allow
programmability.) The four bits (or so) from the quantiser and n bits of state would
address the ROM, from which it would calculate a DAC output of around eight bits, and n
bits for the next state. The ROM would therefore require (n + 8)2(n+4) bits of storage. The
question was whether n would be small enough for this to be practical.

ROM size calculation


The ROM size can be calculated using the state-space dynamic range theory in chapter 2,
as the problem is that of realising a lter, L(z ), with sucient dynamic range. The
discrete-time version of the theory in which a delay z ;1 replaces an integrator is used. This
has not been presented, but it is practically identical except that z replaces s and quantities
such as Gramians are integrated over the unit circle z = ej rather than the j! axis
(Thiele 1986). The main e ort before using this theory is to calculate the input signal to
the lter (z ), and what weight (z ) to give output noise.

Assume that the modulator uses the same NTF and four-bit quantiser as the FIR DAC
realisation. For calculation purposes, assume the quantiser and DAC have unity gain, and

177
e d

r u y
L(z) L(z)

DAC

Figure 5.13: Sigma-delta ADC with digital loop lter. Left: modulator structure. Right: Linearised
model: e is quantisation noise and d is lter nite-wordlength rounding noise.

there is no gain before the quantiser. (Such gains will not a ect the dynamic range
required.) Fig. 5.13 shows the modulator structure.

Using the linear NTF analysis, the input to the loop lter is
U (z) = 1 + 1L(z) [R(z) + E (z) ; D(z)]
Assume the DAC full-scale output is 1. The maximum input will then be around 0:5;
assume that R(z ) is 0.25 RMS. The power spectral density is therefore 5  0:252 = 0:31,
assuming an oversampling ratio of 5. (All spectral densities here are with respect to 2 rad,
as we want a unity power white noise to give a weight of one.)

The NTF 1=(1 + L(z )) averages around {26 dB (0.05) in-band, so the peak error
R(z)=(1 + L(z)) will be of the order of 0:5  0:05 = 0:025. The quantiser needs to cope with
this and some ampli ed out-of-band quantisation noise, so the quantiser range will be
around 0:05. The quantiser step is therefore 0:1=16 = 0:00625. Hence the quantisation
noise E (z ) is 0.0031 peak, or 0.0018 RMS. This corresponds to a power spectral density of
0:00182 = 3:2  10;6 . We therefore choose the input weight as
j 2
(e ) = j1 +0L:31 (e )j
j 2 ; j  j  
5
; 6
= j13:+2  10
L(ej )j2 ; jj > 5


Now consider the lter rounding noise. The output is


Y (z) = 1 + 1L(z) D(z) + 1 +L(Lz()z) (R(z) + E (z)):
Assume the allowable in-band error contribution from lter rounding D(z )=(1 + L(z )) is
0.001 ({60 dB). Out-of-band, D(z )=(1 + L(z )) must be small enough that it does not
178
overload the quantiser. Assume that 0.01 ({40 dB) is acceptable { remember the maximum
quantiser input is 0:05. If we take the output weight as
j 2 6 
(e ) = j1 + 10
L(e )j
j 2 ; j  j  5
4 
= j1 + 10
L(e )j
j 2 ; j  j > 5
then the maximum weighted output noise (z )D(z ) allowed is unity.

We next calculate the frequency-weighted Hankel singular values using these weights and
the desired transfer function L(z ). For convenience, we use the same loop lter transfer
function as that of the FIR DAC architecture. We only have the magnitudes of the weights
not their transfer functions, but we can still calculate the Gramians according to their
de nitions (2.10) and (2.11). (The only change for discrete-time is to integrate around the
unit circle rather than along the j! axis.) This requires the gain to states F (z ) and the
noise gain G(z ) and hence a state-space realisation, which we obtain simply from the FIR
structure. We then use the property that the Hankel singular values i are the same for all
realisations.

This results in

i2 = i(K  W ) = 2740; 230; 5:2; 0:091;    :


The other eight i2 are smaller than 0.02. We recall from section 2.9.3 that for the
input-normal realisation, with unit noise e added to each state, the frequency-weighted
output noise power from the ith state is i2 . We also recall that the frequency-weighted
input-normal realisation has a property of minimising the number of bits of storage required
in a digital lter (Mullis & Roberts 1976).

We said above that the maximum weighted output noise allowed is unity. We therefore
need to keep the rounding noise power contribution to each state below 1=i2 . (Actually it
needs to be a little lower, as the noise powers from di erent states add.) For the last nine
states, i is less than unity. This means that even if we round those states to zero, the
output noise will be tolerable. The rounding noise power decreases by a factor of four for
each bit of precision used to store a state. For the rst three states, roughly 6, 4 and 2 bits
of precision respectively will give suciently low rounding error. Thus the total information
179
storage in the state is 12 bits. The ROM must take 12 bits of state and 4 bits of input, and
from these calculate 12 bits of state and around 8 bits of output. The total ROM size is
therefore (12 + 8)  2(12+4) = 1:3  106 .

This memory size is impractical for an embedded memory, although not hugely so.
SRAM bit area in 0.18 m logic CMOS is around 5 (m)2 , and ROM is presumably around
1 (m)2 , so the area needed is 6.5 (mm)2 in SRAM or 1.3 (mm)2 in ROM. Of course, this
would change greatly if margin for overload required a few more bits, or if a better transfer
function allowed a few less bits. A large ROM may also be unattractively slow.

5.7.3 Continuous-time loop lter versus FIR DAC


As mentioned earlier in the chapter, the FIR DAC was used because the author did not
believe that a continuous-time loop lter would be accurate enough at hundreds of
megahertz. Given the 800 MHz lter simulation results at the end of chapter 3, using a
continuous-time loop lter looks more promising.

The FIR DAC approach has advantages and disadvantages compared to a


continuous-time loop lter. The advantages are that the FIR DAC transfer function is well
controlled, and that the FIR DAC output is only used once it has settled, so DAC glitches
are not signi cant like they are when using a DAC in continuous-time. This also reduces
sensitivity to clock jitter. The disadvantages are that the FIR DAC modulator requires a
discrete-time ampli er which settles in around 1ns, and that a continuous-time input is
compared with discrete-time feedback, leading to a high dv=dt for the track and hold. A
continuous-time loop lter also allows slightly better shaping of the NTF { the FIR DAC,
like all short FIR lters, has a broad transition band. A combination of a FIR DAC and
conventional loop lter may be attractive { this approach has been tried in
switched-capacitor technology by Okamoto, Maruyama & Yukawa (1993).

180
5.8 SIGMA-DELTA VERSUS NYQUIST-RATE CONVERTERS AT
20-200MHZ BANDWIDTH

5.8.1 Performance comparison


A signal to noise and distortion ratio of 40 dB over 60 MHz bandwidth does not sound
particularly impressive. However, it was equal to the state of the art in CMOS when this
work was published, and is still close. (Bipolar ADC's achieve better performance, but o er
less potential for system-on-a-chip integration.) At these bandwidths, the competitors are
Nyquist-rate converters. These require an anti-alias lter. An anti-alias lter similar to that
in chapter 3 with the passband edge at 60 MHz, has a stopband edge around 100 MHz. A
160 Ms=s Nyquist-rate converter would therefore be required to avoid aliasing. The
pre- lter required by the FIR DAC sigma-delta architecture allows the full 60 MHz
bandwidth to be used { the pre- lter would roll o above 60 MHz and needs to be {30 dB or
so by 90 MHz to avoid overloading the modulator.

The following are the best CMOS ADC's the author could nd for sample rates
100 Ms=s. Note how the state of the art is without exception recent work.

Reference Topology Speed SNDR SFDR Power Process


(Harrison & Weste 2000) FIR DAC  \160 Ms=s" 40 dB 49 dB 220 mW 0.25 m
(Choe et al. 2001) folding 100 Ms=s 35 dB 45 dB 165 mW 0.5 m
(Park et al. 2001) pipeline 100 Ms=s 57 dB 64 dB 180 mW 0.18 m
(Wang & Razavi 2000) subranging 150 Ms=s 40 dB 43 dB 395 mW 0.6 m
(Nagaraj et al. 2000) ash 200 Ms=s 41 dB 45 dB 143 mW 0.25 m
(Sumanen et al. 2001) 4  pipeline 200 Ms=s 46 dB 55 dB 280 mW 0.5 m
(Choi & Abidi 2001) ash 1025 Ms=s 36 dB 45 dB 500 mW 0.35 m

These SFDR's and SNDR's are for similar input frequencies. The SNDR and SFDR for
the ADC described is for 38 MHz / 40 MHz two-tone input, with energy above 60 MHz
ignored. (A sigma-delta will have lots of out-of-band noise and possibly harmonics, but
these are removed by digital ltering.) An 8 MHz single-tone input ( g. 5.12), putting the

181
harmonics in-band, gave similar noise and distortion. Data from the other papers for
40{53 MHz input is used, except (Choe et al. 2001) which uses 12 MHz input, and
(Sumanen et al. 2001) uses 71.3 MHz input (although the converter is not signi cantly
better at 40 MHz).

The recent work by Sumanen et al. (2001) is the rst to clearly outperform the ADC
described here. Sumanen et al. (2001) use a 200 Ms=s track and hold and four 50 Ms=s
parallel pipelines. This is quite elegant as track and holds do not require as precise a gain
as pipeline stages and hence can use lower gain, faster opamps. Bootstrapped switches and
digital correction are also used.

Unless it was an application which particularly favoured a four-bit output at 520 Ms=s
and had a readily available clock at that frequency, the sigma-delta approach would
probably not be preferable. Software radios or wireless LAN's come to mind as an
application where the higher clock rate might be available as a local oscillator, and might
reduce spurii problems. The four-bit output might be convenient if pin count limitations
were otherwise going to require a separate high-speed serial I/O. Note that Sumanen et al.
(2001) require an additional 125 mW for their digital I/O, which is not included in the
280 mW above. The digital ltering required by sigma-delta ADC's would probably also be
already present for radio applications.

The more relevant question is therefore whether the FIR DAC architecture has
advantages which will allow future implementations to outperform Nyquist-rate converters.

5.8.2 Limitations of sigma-delta ADC's and Nyquist-rate ADC's


The idea that sigma-delta modulators with continuous-time loop lters should be attractive
as ADC's at bandwidths of tens of megahertz has been around for a number of years.
Digitising at intermediate frequency (IF) with a bandpass modulator, rather than at
baseband with a Nyquist-rate converter, is particularly mentioned (Gao & Snelgrove 1998).
Unfortunately, continuous-time modulators have never really lived up to expectations, with
degradations attributed to loop lter dynamic range limitations and poor signal scaling,
clock jitter (Cherry, Snelgrove & Schvan 1997) and metastability. Should such modulators

182
be written o , or is just a little more attention to detail needed?

Limitations of pipelined converters


It seems reasonable to look at sigma-delta modulators only if Nyquist-rate converters will
hit fundamental performance limits which can be circumvented by sigma-delta modulators.
Pipelined converters seem the most attractive architecture for achieving bandwidths of
50{100 MHz at 60{80 dB SNDR, so it is worth checking what limits they face. Folding and
ash converters seem to have more fundamental limitations relating to transistor mismatch
and complexity which make them less likely to jump from the 8 bit level to the 12 bit level,
although they will presumably still be competitive at higher sample rates.

The fundamental limitations of pipelined converters appear to be that the sampling rate
is limited by opamp settling time, and that the SNDR for a given power consumption is
limited by thermal noise. ADC's using bipolar transistors o er some possibility to get high
dynamic range without feedback, but with deep submicron CMOS and its inferior transistor
gain, feedback around ampli er transistors seems inevitable.

Existing switched-capacitor opamps, even in deep submicron CMOS, achieve


gain-bandwidths around 1 GHz (Mehr & Singer 2000) (Park et al. 2001) and settling times
around 4 ns. Speed might be expected to scale with transistor fT , but it tends to fall short
because it is harder to operate at maximum fT at lower VDD , and the stage gain of
minimum-length transistors is falling.

The author has tried minimising the settling time of variants of the opamps in chapter 3.
The results were similar to those achieved with existing techniques such as
Miller-compensated opamps. (Gain-boosted cascodes (Bult & Geelen 1990) tend to be very
hard to implement at low VDD .) If power consumption is not an issue, settling times of
1{2 ns are possible. Once power consumption is restricted and process variations are
included, fast settling with the capacitive loads needed for low thermal noise is dicult.
Fast settling is more dicult than wide bandwidth because optimum phase margin is higher
(60 ; 70 ), and because doublets must be pushed to high frequencies (Bult & Geelen 1990).
Stages in a pipelined converter also run at a gain of two or more, not unity-gain.

183
Thermal noise advantages of sigma-delta modulators
The main advantage the author sees for sigma-delta ADC's with continuous-time loop
lters is that they should be able to achieve lower thermal noise. Thermal noise formulae
are in (Goes, Vital & Franca 1998) for pipelined ADC's, (Dias, Palmisano, O'Leary &
Maloberti 1992) for switched-capacitor sigma-delta ADC's, and (Dias, Palmisano &
Maloberti 1992) for continuous-time (and mixed continuous-time / discrete-time)
sigma-delta ADC's.

Kelly, Yang, Mehr, Sayuk & Singer (2001) state that the rst stage of a practical
pipelined converter needs 4 pF capacitors for 70 dB SNR. This is with a multi-bit rst stage
so that the noise of later stages is less signi cant, and 2 Vp-p di erential signal swing (using
3 V VDD ). The opamp for the stage with 4 pF capacitors consumed 100 mW to get fast
enough settling for 75 Ms=s. With lower power supply voltages and the single-bit-per-stage
pipelines needed for maximum sample rate it only gets worse.

The thermal noise of a sigma-delta modulator with continuous-time loop lter is largely
determined by the rst integrator (Dias, Palmisano & Maloberti 1992). For 50 MHz signal
bandwidth and 4 pF integrator capacitors, the loop lter bandwidth is likely to be around
100 MHz, resulting in input resistors around 400
. If the input signal is 1 Vp-p di erential,
the input current is 2.5 mAp-p, requiring a DAC tail current around 5 mA. Such a DAC is
likely to have transistor gm of around 5 mS per side. The total thermal-noise-producing
conductance at the input is 2.5 mS from the input resistor, 5 mS from the DAC (the
theoretical value is 32 gm , with extra in practice from hot carriers), and perhaps another
5 mS from the opamp and other resistors in the lter. This 12.5 mS contributes 100 nA
RMS noise over 50 MHz bandwidth, or 140 nA RMS counting both sides. With respect to
the 2.5 mAp-p input signal, this is an SNR of 76 dB. The opamp output current depends on
how well the DAC output matches the input signal and on second-stage loading, but 5 mA
output-stage tail current (assuming a similar architecture to that in chapter 3) seems
reasonable. The total opamp power is likely to be around 10 mA or 20 mW. This clearly
looks more promising than the pipelined converter.

FIR DAC architectures are also potentially quieter than pipelined architectures, as the

184
error signal can be ampli ed before sampling. The signal is also only sampled once and
does not accumulate multiple kT=C noises like that in a pipelined converter.

Limitations of sigma-delta modulators


Having observed that pipelined converters face fundamental thermal noise and settling time
limits, and that sigma-delta modulators o er advantages with respect to thermal noise, we
should check that sigma-delta modulators do not have other worse fundamental limitations
of their own. Clock jitter, metastability and loop lter dynamic range have been raised as
signi cant issues (Cherry & Snelgrove 1999a) (Tao, Toth & Khoury 1999) (Gao &
Snelgrove 1998). If a multi-bit DAC is to be used, DAC resolution is another signi cant
consideration.

Loop lter dynamic range is calculated in the example above. We saw in section 2.4 that
the noise of an opamp-RC lter is mostly that of its resistors (and DAC in a sigma-delta
ADC). Bandpass modulators could pose greater diculties, because of the dynamic range
reduction experienced in accordance with the dynamic range limitation of section 2.6. Note
that anti-alias lters are potentially a bigger dynamic range problem than sigma-delta
modulator loop lters, as the dynamic range required of loop lters is reduced by being in
the feedback loop.

Metastability can be reduced to acceptable levels by using one or more latches after the
comparator. The fundamental limitations with respect to excess loop delay in chapter 4 are
relevant here. However, the fT of deep submicron CMOS is very high, so 1{2 Gs=s seems
achievable with negligible metastability.

Clock jitter is a particular problem if continuous-time loop lters are used. (A


switched-capacitor or FIR DAC sigma-delta ADC has no disadvantage compared to a
Nyquist-rate ADC { they all sample the input signal. The FIR DAC architecture samples
the error rather than the input, but as the DAC pulse being subtracted has settled there is
no di erence.) With a continuous-time loop lter, the clock is used in two places, the
quantiser and the DAC. The quantiser clock is non-critical, as the quantiser resolution is
low. The DAC clock is critical.

185
6

2
Digital output

6
0 10 20 30 40 50 60 70 80 90 100
Time (samples)

Figure 5.14: Typical time-domain output of the FIR DAC sigma-delta ADC. (Data collected at
520 MHz, then clocked out at 1 kHz from the on-chip shift register.) The high-frequency noise is due
to the high out-of-band NTF.

Clock jitter is analysed in Cherry & Snelgrove (1999a) and Tao et al. (1999). Cherry &
Snelgrove (1999a) conclude that it need not constrain performance if a conventional on-chip
LC oscillator is used. This is for one-bit quantisers, so we should check that multi-bit
quantisers and high out-of-band NTF's do not cause problems.

Clock jitter can be analysed using the NTF analysis. The gain from the DAC to the
output is equal to the signal transfer function. In-band (where the jitter noise matters), this
is close to unity. The magnitude of the jitter noise input basically depends on the average
change in DAC output each clock cycle (Cherry & Snelgrove 1999a). For a one-bit
quantiser, this is large as the output is either +1 or ;1, and alternates rapidly between the
two. For a low-pass modulator with multi-bit quantiser and low out-of-band gain, the DAC
change in output is usually only a small fraction of full scale, so clock jitter is less of a
problem. Using a multi-bit quantiser with high out-of-band gain, the change in output is a
signi cant fraction of full scale { see g. 5.14. As this is somewhat better than the one-bit
case, which Cherry & Snelgrove (1999a) do not consider a problem, we can conclude that

186
clock jitter will not fundamentally limit performance. Certainly care needs to be taken to
avoid digital data disturbing the clock however.

The other major challenge with such a sigma-delta modulator is to build a 1{2 Gs=s
DAC with sucient SNDR. Decoding-induced delay variations are not a problem as the
natural approach is to put the thermometer outputs from the quantiser straight into the
DAC. (The FIR DAC converter also used this approach.) Signal-dependent glitches are
more of a problem, although return-to-zero DAC pulses could be used. Van den Bosch,
Borremans, Steyaert & Sansen (2001) describe a DAC with 70 dB SFDR at 1 Gs=s for
100 MHz input without return-to-zero, so thermal-noise-limited performance is not too far
away. Only having 16 or so current cells and having the virtual ground of an opamp input
to work into should help.

5.9 CONCLUSION
The FIR DAC architecture is competitive with Nyquist-rate ADC's at wide bandwidths,
but needs further development to be of real interest to industry. The future of it and other
wide bandwidth sigma-delta modulators depends a lot on the extent to which pipelined
converters improve in the future. The FIR DAC architecture and other sigma-delta ADC's
seem to have an advantage with respect to thermal noise.

The combination of a multi-bit quantiser and high out-of-band NTF to achieve high
resolution at low OSR seems useful. It is dicult to decide whether a continuous-time or
FIR DAC loop lter is superior for clock rates beyond the capacity of switched-capacitor
circuits. The FIR DAC architecture is a little inelegant in that it subtracts a discrete-time
signal from a continuous-time signal. With low-pass modulators loop latency problems also
limit speed; fortunately with fs=4 bandpass modulators this is far less of an issue as an
extra cycle is available. The FIR DAC architecture avoids problems with clock jitter and
DAC nonlinear glitch energy su ered by modulators with continuous-time loop lters.

The variant in which the loop lter is fully digital has conceptual elegance. The use of a
DAC to subtract linearly predicted interferers before the signal is quantised seems a good

187
prospect for a practical software radio. At present, the information storage is prohibitive for
a ROM-based state machine, but this may change with future digital CMOS scaling.

188
6 CONCLUSION AND SUGGESTIONS FOR
FUTURE WORK

6.1 CONCLUSION
Let us return to two questions posed in the introduction. Why do high-Q lters composed
of integrators tend to have poor dynamic range? High-Q lters have poor dynamic range
because the transfer function H (s) maps closely spaced points s1 and s2 in the s-plane to
far away points H (s1 ) and H (s2 ) in the H (s)-plane. This is an ill-conditioned operation,
which requires taking the di erence of nearly equal quantities. The mathematical
description of this is the dynamic range limitation
H (s1) ; H (s2) = G(s )F (s ):
s2 ; s1 1 2

We saw that high dynamic range required making the gain to states F (s) and the noise
gain G(s) small.

The dynamic range limitation is a particularly useful result because when considered
over all frequencies !1 and !2 , the limitation is achievable, namely by the
frequency-weighted input-normal realisation or frequency-weighted balanced realisation.
Unfortunately, this realisation is dicult to implement because of the dense coecient
matrices (A; B; C; D). However, we saw in chapter 3 that practical realisations from
simulations of LC ladders had very similar performance. This means that the dynamic
range limitation is still useful as an accurate guide to the choice of transfer function.
Low-ripple Elliptic transfer functions work well for lters.

A second question asked how sigma-delta modulators traded bandwidth for dynamic
range. The existing explanation that the feedback loop corrects for errors is reasonable.
Another explanation is that the modulator predicts the next sample, and hence only needs
to quantise a small error in the prediction rather than the complete input. This reduces the
number of bits required in the quantiser. This linear predictor interpretation seems

189
particularly applicable to the sigma-delta ADC in chapter 5, where the input signal is
pre- ltered to make it predictable.

The question then arises as to how much feedback is possible, or how accurately the
input signal can be predicted. We attempted to answer this in section 2.10.1 for ampli ers,
using the dynamic range limitation, but the results were numerical rather than analytic.
Optimising the amount of opamp loop gain was also an important part of the lter design
in chapter 3. This was the major innovation which allowed the lter to operate at 350 MHz.
Again, the approach was empirical. For the sigma-delta modulators in chapter 4, we have a
stronger result however, namely the Bode integral. This states that the noise transfer
function in decibels integrates to zero over frequency. As the noise transfer function is
constrained out of band by stability, we can accurately calculate how much quantisation
noise reduction is possible.

The Bode integral states that the oversampling ratio of a sigma-delta modulator can
only be reduced by increasing the out-of-band gain or sharpening the loop lter rollo . The
fact that increased out-of-band gain is possible with a multi-bit quantiser was the key to
achieving a wide bandwidth from the ADC in chapter 5. This bene t of multi-bit quantisers
has been known previously to some degree, but it becomes a lot clearer with the Bode
integral that high out-of-band gain is vital for maximum bandwidth.

The relevance of the Bode integral has been identi ed by previous authors, but a second
integral concerning excess loop delay is new. It makes the prediction that a sample of excess
loop delay is necessarily detrimental (to the extent of halving the bandwidth) for low-pass
modulators. This harks back to interpretation of a sigma-modulator predicting its future
input. If the recent past is unavailable, the predictions are less accurate.

To nish, a few words on the e ect of scaling on analog CMOS. The author's belief is
that the greater use of feedback possible with faster transistors will more than counter the
falling swing as a percentage of supply voltage down to at least VDD = 1 V. Low-threshold
transistors may be required, but microprocessors use these too. Today's 10 MHz gm-C lter
will be successively replaced by the class-A opamp-RC lter and the class-B opamp-RC
lter. With more loop gain, output stages can be driven closer to cuto and saturation for

190
the same distortion. Cascodes are no longer needed. Loop gain is not going to fall because
the transistor stage gain falls; the feedforward techniques in chapter 3 allow three and four
stage opamps to operate at wide bandwidth. Pipelined ADC's will have the bandwidth of
today's ash ADC's, and sigma-delta ADC's will have the bandwidth of today's pipelined
ADC's. The techniques in this work are adaptable to at very least the 0.13 m process
generation. It should be emphasised that this is the author's opinion; this work is about the
state-space and sigma-delta modulator dynamic range theories, not scaling.

6.2 SUGGESTIONS FOR FUTURE WORK

6.2.1 Circuit design


Feedback ampli ers
The improvement of opamp bandwidth from 1 GHz to 10 GHz achieved in chapter 3
suggests that opamps should be used more at high frequencies. Active lters and
intermediate-frequency ampli ers are obvious applications, but what could be more
interesting is the use of feedback ampli ers at RF in the commercially important 900 MHz {
2.4 GHz bands. Feedback ampli ers potentially allow higher IP3 and tightly-controlled gain.
Feedback ampli ers tend to become practical at frequencies around fT =100, although this
can be pushed to fT =30 or so (as in the lter at the end of chapter 3). Transistor fT is
entering this range with the 0.18{0.13 m generations.

Feedback ampli ers at RF face peculiar design challenges. The input impedance should
be resistive (eg. 50
) rather than an open circuit. Inductive sources and loads present
stability challenges. It may be possible to deal with these using a combination of shunt and
series feedback, and Zobel networks to resistively-terminate the feedback ampli er
out-of-band.

Sigma-delta modulators
As mentioned in chapter 5, a sigma-delta modulator with continuous-time loop lter using
the opamp-RC technology in chapter 3 appears promising. The opamp-RC technology gives

191
a controlled transfer function at up to 1 GHz, which should allow high out-of-band NTF
(with a multi-bit quantiser) to enhance performance. An fs=4 bandpass version of the FIR
DAC architecture was also highlighted as worth investigating, as it avoids latency problems.

6.2.2 Circuit theory


The frequency-domain state-space realisation theory in chapter 2 seems fertile. There is
work to be done in converting the time-domain state-space realisation results of Kalman
and coworkers into the frequency-domain. The theory also may be helpful for deriving the
long sought-after (Anderson & Moore 1989) (Zhou 1996) frequency-domain error bound for
frequency-weighted balanced truncation.

Linear Fractional Transformations


A generalisation of the state-space equations is the linear fractional transformation

u = Ax + Br
y = C x + Dr
x = S u + e:
Here r is the system input, and y is the system output. In a state-space lter, S would be a
diagonal matrix of integrators, and A, B , C and D would be frequency-independent. These
equations result in a gain from input r to input to the integrators u of F , and a gain from
added noise e to output y of G, where

F = (I ; AS );1 B
G = C (I ; SA);1 :
The overall transfer function from r to y is

H = CS (I ; AS );1 B + D = C (I ; SA);1 SB + D
An equivalent of the dynamic range limitation can be derived. If A, B , C and D are
frequency independent, and S = S (j!),

H (j!1 ) ; H (j!2) = G(j!1 )[S (j!1 ) ; S (j!2 )]F (j!2 ):


192
If this could somehow be given a singular value decomposition in analogy to the work in
chapter 2, we would have a synthesis which allows a transfer function to be realised from an
arbitrary resonator S (j!). There may also be possibilities for circuit-level synthesis; linear
fractional transformations are the basis of the Youla (1961) cascade synthesis variant of
Darlington synthesis. Linear fractional transformations are also widely used in control
theory (Zhou 1996).

193
BIBLIOGRAPHY

Abidi, A. A. (1992), `Noise in active resonators and the available dynamic range', IEEE
Trans. Circuits Syst. I 39(4), 296{299.

Adamyan, V. M., Arov, D. Z. & Krein, M. G. (1971), `Analytic properties of Schmidt pairs
for a Hankel operator and the generalized Schur-Takagi problem', Math. USSR Sbornik
15(1), 31{73.
Analog Devices (2000), `AD6644 (14-Bit 40 MSPS / 65 MSPS A/D converter) data sheet'.
Rev. 0.
Anderson, B. D. O. & Moore, J. B. (1989), Optimal Control: Linear Quadratic Methods,
Prentice-Hall, Englewood Cli s, NJ.
Anderson, B. D. O. & Vongpanitlerd, S. (1973), Network Analysis and Synthesis: A Modern
Systems Theory Approach, 1st edn, Prentice-Hall, Englewood Cli s, NJ.

Annema, A.-J. (1999), `Analog circuit performance and process scaling', IEEE Trans.
Circuits Syst. II 46(6), 711{725.

Antsaklis, P. J. (2000), `At the gates of the millenium: Are we in control?', IEEE Contr.
Syst. Mag. 20(1), 50{55.

Baird, R. T. & Fiez, T. S. (1994), `Stability analysis of high-order delta-sigma modulation


for ADC's', IEEE Trans. Circuits Syst. II 41(1), 59{62.
Banu, M., Khoury, J. M. & Tsividis, Y. (1988), `Fully di erential operational ampli ers
with accurate output balancing', IEEE J. Solid-State Circuits 23(6), 1410{1414.
Behbahani, F., Tan, W., Karimi-Sanjaani, A., Roithmeier, A. & Abidi, A. A. (2000), `A
broad-band tunable CMOS channel-select lter for a low IF wireless receiver', IEEE J.
Solid-State Circuits 35(4), 476{489.

Blostein, M. L. (1967), `Sensitivity analysis of parasitic e ects in resistance-terminated LC


two-ports', IEEE Trans. Circuit Theory 14(1), 21{25.
194
Bode, H. W. (1945), Network Analysis and Feedback Ampli er Design, Van Nostrand, New
York.

Bode, H. W. & Shannon, C. E. (1950), `A simpli ed derivation of linear least square


smoothing and prediction theory', Proc. IRE 38(4), 417{425.

Bourbaki, N. (1994), Elements of the History of Mathematics, Springer-Verlag, Berlin.

Bult, K. (2000), Analog design in deep sub-micron CMOS, in `European Solid-State


Circuits Conf.', number 305.

Bult, K. & Geelen, G. J. G. M. (1990), `A fast-settling CMOS opamp for SC circuits with
90-dB DC gain', IEEE J. Solid-State Circuits 25(6), 1379{1384.

Buss, D. (1999), Device issues in the integration of analog / RF functions in deep submicron
digital CMOS, in `IEEE Int. Electron Devices Meeting Tech. Digest', p. 424.

Candy, J. C. (1985), `A use of double integration in sigma delta modulation', IEEE Trans.
Commun. 33(3), 249{258.

Candy, J. C. & Temes, G. C. (1992), Oversampling Delta-Sigma Data Converters, IEEE,


New York.

Castello, R. & Gray, P. R. (1985), `Performance limitations in switched-capacitor lters',


IEEE Trans. Circuits Syst. 32(9), 865{876.

Cherry, E. M. & Hooper, D. E. (1968), Amplifying Devices and Low-Pass Ampli er Design,
Wiley, New York.

Cherry, J. A. & Snelgrove, W. M. (1999a), `Clock jitter and quantizer metastability in


continuous-time delta-sigma modulators', IEEE Trans. Circuits Syst. II
46(6), 661{676.
Cherry, J. A. & Snelgrove, W. M. (1999b), `Excess loop delay in continuous-time
delta-sigma modulators', IEEE Trans. Circuits Syst. II 46(4), 376{389.

Cherry, J. A., Snelgrove, W. M. & Schvan, P. (1997), `Signal-dependent timing jitter in


continuous-time  modulators', Electronics Lett. 33(13), 1118{1119.

195
Choe, M.-J., Song, B.-S. & Bacrania, K. (2001), `A 8-b 100-MSample/s CMOS pipelined
folding ADC', IEEE J. Solid-State Circuits 36(2), 184{194.
Choi, M. & Abidi, A. A. (2001), A 6b 1.3Gsample/s A/D converter in 0.35m CMOS, in
`Int. Solid-State Circuits Conf. Digest Tech. Papers', pp. 126{127.
Deng, Y. & Lee, E. K. F. (2001), Design of a 1V 250MHz current-mode lter in
conventional CMOS process, in `IEEE Int. Symp. Circuits Syst.', Vol. 1, pp. 236{239.
Dias, V. F., Palmisano, G. & Maloberti, F. (1992), `Noise in mixed continuous-time
switched-capacitor sigma-delta modulators', IEE. Proc. Part G 139(6), 680{684.
Dias, V. F., Palmisano, G., O'Leary, P. & Maloberti, F. (1992), `Fundamental limitations of
switched-capacitor sigma-delta modulators', IEE. Proc. Part G 139(1), 27{32.
Doetsch, G. (1974), Introduction to the Theory and Application of the Laplace
Transformation, Springer { Verlag, Berlin.

Dunford, N. & Schwartz, J. T. (1964), Linear Operators { Part I: General Theory, 1st edn,
Wiley Interscience, New York.
Duque-Carrillo, J. F. (1993), `Control of the common-mode component in CMOS
continuous-time fully di erential signal processing', Analog Integrated Circuits and
Signal Processing 4, 131{140.

Enns, D. F. (1984), Model reduction with balanced realizations: an error bound and a
frequency weighted generalization, in `IEEE Conf. Decision Contr.', Vol. 1,
pp. 127{132.
Fano, R. M. (1950), `Theoretical limitations on the broadband matching of arbitrary
impedances', J. Franklin Inst. 249, 57{154.
Farrell, R. & Feely, O. (1998), `Bounding the integrator outputs of second-order sigma-delta
modulators', IEEE Trans. Circuits Syst. II 45(6), 691{702.
Fettweis, A. (1986), `Wave digital lters: theory and practice', Proc. IEEE 74(2), 270{327.
Gao, W. & Snelgrove, W. M. (1998), `A 950-MHz IF second-order integrated LC bandpass
delta-sigma modulator', IEEE J. Solid-State Circuits 33(5), 723{732.
196
Gerzon, M. & Craven, P. G. (1989), `Optimal noise shaping and dither of digital signals',
87th Convention of the Audio Engineering Society. Preprint No. 2822.

Glasser, L. A. (1988), `Frequency limitations in circuits composed of linear devices', IEEE


Trans. Circuits Syst. 35(10), 1299{1307.

Glover, K. D. (1984), `All optimal Hankel-norm approximation of linear multivariable


systems and their L1-error bounds', Int. J. Control 39(6), 1115{1193.

Goes, J., Vital, J. C. & Franca, J. E. (1998), `Systematic design for optimzation of
high-speed self-calibrated pipelined A/D converters', IEEE Trans. Circuits Syst. II
45(12), 1513{1526.
Golub, G. H. & Van Loan, C. F. (1996), Matrix Computations, 3rd edn, John Hopkins
University Press, Baltimore.

Gopinathan, V., Tarsia, M. & Choi, D. (1999), `Design considerations and implementation
of a programmable high-frequency continuous-time lter and variable-gain ampli er in
submicrometer CMOS', IEEE J. Solid-State Circuits 34(12), 1698{1707.

Gray, P. E. & Searle, C. L. (1969), Electronic Principles: Physics, Models, and Circuits, 1st
edn, Wiley, New York.

Gray, R. M. (1990), `Quantization noise spectra', IEEE Trans. Inform. Theory


36(6), 1220{1244.
Greer, N. P. J., Henderson, R. K., Ping, L. & Sewell, J. I. (1991), Matrix methods for the
design of transconductor ladder lters, in `IEEE Int. Symp. Circuits Syst.', Vol. 3,
pp. 1355{1358.

Grenander, U. & Szego, G. (1958), Toeplitz Forms and Their Applications, 1st edn,
University of California Press, Berkeley.

Groenewold, G. (1991), `The design of high dynamic range continuous-time integratable


bandpass lters', IEEE Trans. Circuits Syst. 38(8), 838{852.

Groenewold, G. (1992), `Optimal dynamic range integrators', IEEE Trans. Circuits Syst. I
39(8), 614{627.
197
Harrison, J. & Weste, N. (1999), A wideband quadrature LO generator in digital CMOS, in
`Proc. Custom Integrated Circuits Conf.', pp. 657{659.

Harrison, J. & Weste, N. (2000), A multi-bit sigma-delta ADC with an FIR DAC loop
lter, in `Proc. IEEJ Int. Analog VLSI Workshop', pp. 26{31.

Harrison, J. & Weste, N. (2001), Energy storage and Gramians of ladder lter realisations,
in `IEEE Int. Symp. Circuits Syst.', Vol. 1, pp. 29{32.

Harrison, J. & Weste, N. (2002), A limitation on active lter dynamic range, in `IEEE Int.
Symp. Circuits Syst.'.

Hein, S. & Zakhor, A. (1993), `On the stability of sigma delta modulators', IEEE Trans.
Signal Processing 41(7), 2322{2348.

Helton, J. W. (1981), `Broadbanding: Gain equalization directly from data', IEEE Trans.
Circuits Syst. 28(12), 1125{1137.

Helton, J. W. & Merino, O. (1998), Classical Control Using H 1 Methods, SIAM,


Philadelphia, PA.

Ho, B. L. & Kalman, R. E. (1966), `E ective construction of linear state-variable models


from input / output functions', Regelungstechnik 14(12), 545{548.

Hurkx, G. A. M. (1997), `The relevance of ft and fmax for the speed of a bipolar CE
ampli er stage', IEEE Trans. Electron Devices 44(5), 775{781.

Hwang, S. Y. (1977), `Minimum uncorrelated unit noise in state-space digital ltering',


IEEE Trans. Acoust., Speech, Signal Processing 25(4), 273{281.

IEEE Std. 802.11a (1999), `Wireless LAN medium access control (MAC) and physical layer
(PHY) speci cations: High-speed physical layer in the 5GHz band'.

Johns, D. A. & Sedra, A. S. (1987), `State-space simulation of LC ladder lters', IEEE


Trans. Circuits Syst. 34(8), 986{988.

Johns, D. A., Snelgrove, W. M. & Sedra, A. S. (1989), `Orthonormal ladder lters', IEEE
Trans. Circuits Syst. 36(3), 337{343.

198
Ju, P. & Vallancourt, D. G. (1992), `Quantisation noise reduction in multibit oversampling
 ;  A/D convertors', Electronics Lett. 28(12), 1162{1164.

Justesen, J. (1978), `Finite state predictors for Gaussian sequences', Information and
Control 39, 35{45.

Kelly, D., Yang, W., Mehr, I., Sayuk, M. & Singer, L. (2001), A 3V 340mW 14b 75MSPS
CMOS ADC with 85dB SFDR at Nyquist, in `Int. Solid-State Circuits Conf. Digest
Tech. Papers', pp. 134{135.

Kenney, J. G. & Carley, L. R. (1993), `Design of multibit noise-shaping data converters',


Analog Integrated Circuits and Signal Processing 3, 259{272.

Kim, H. K. (1992), `The e ects of lower Q values on the lters having equal ripples in
passband', IEEE Trans. Circuits Syst. I 39(4), 305{309.

Kim, K. Y., Kusayanagi, N. & Abidi, A. A. (1997), `A 10-b 100-MS/s CMOS A/D
converter', IEEE J. Solid-State Circuits 32(3), 302{311.

Kishi, G. & Kida, T. (1967), `Energy theory of sensitivity in LCR networks', IEEE Trans.
Circuit Theory 14(4), 380{387.

Kishi, G. & Nakazawa, K. (1963), `Relations between reactive energy and group delay in
lumped-constant networks', IEEE Trans. Circuit Theory 10(1), 67{71.

Kung, S. (1978), A new identi cation and model reduction algorithm via singular value
decompositions, in `Proc. 12th Asilomar conf. on circuits, systems and computers',
pp. 705{714.

Kuo, B. C. (1995), Automatic Control Systems, 7th edn, Prentice-Hall, Upper Saddle River,
NJ.

Lee, T. H. (1998), The Design of CMOS Radio-Frequency Integrated Circuits, 1st edn,
Cambridge University Press, Cambridge.

Leeson, D. B. (1966), `A simple model of feedback oscillator noise spectrum', Proc. IEEE
54(2), 329{330.

199
Li, G., Anderson, B. D. O., Gevers, M. & Perkins, J. E. (1992), `Optimal FWL design of
state-spece digital systems with weighted sensitivity minimization and sparseness
consideration', IEEE Trans. Circuits Syst. I 39(5), 365{377.

Livsic, M. S. (1973), Operators, Oscillations, Waves (Open Systems) (Translations of


Mathematical Monographs v. 34), American Mathematical Society, Providence, Rhode
Island.

Luh, L., Choma, J. & Draper, J. (2000), Performance optimization for high-order
continuous-time  modulators with extra loop delay, in `IEEE Int. Symp. Circuits
Syst.', Vol. 5, pp. 669{672.

Lutz, W. J. & Hakimi, S. L. (1988), `Design of multi-input multi-output systems with


minimum sensitivity', IEEE Trans. Circuits Syst. 35(9), 1114{1121.

Martin, S., Archer, V. D., Boulin, D. M., Frei, M. R., Ng, K. K. & Yan, R. H. (1997),
`Device noise in silicon RF technologies', Bell Labs Tech. J. 2(3), 30{45.

Mason, S. J. (1954), `Power gain in feedback ampli er', IEEE Trans. Circuit Theory
1(2), 20{25.
Mehr, I. & Singer, L. (2000), `A 55-mW, 10-bit, 40Msample/s Nyquist-rate CMOS ADC',
IEEE J. Solid-State Circuits 35(3), 318{325.

Mohtadi, C. (1990), `Bode's integral theorem for discrete-time systems', IEE Proc. Part D
137(2), 57{66.
Moreira, J. P. & Silva, M. M. (2001), `Limits to the dynamic range of low-power
continuous-time integrators', IEEE Trans. Circuits Syst. I 48(7), 805{817.

Moreira, J. P. & Verhoeven, C. J. M. (1998), Fundamental limits to the dynamic range of


integrated continuous-time integrators, in `IEEE Int. Symp. Circuits Syst.', Vol. 1,
pp. 147{150.

Mullis, C. T. & Roberts, R. A. (1976), `Synthesis of minimum roundo noise xed point
digital lters', IEEE Trans. Circuits Syst. 23(9), 551{562.

200
Nagaraj, K., Martin, D. A., Wolfe, M., Chattopadhyay, R., Pavan, S., Cancio, J. &
Viswanathan, T. R. (2000), `A dual-mode 700-Msamples/s 6-bit 200-Msamples/s 7-bit
A/D converter in a 0.25-m digital CMOS process', IEEE J. Solid-State Circuits
35(12), 1760{1768.
Nauta, B. (1993), Analog CMOS Filters for Very High Frequencies, Kluwer, Boston.

Nawrocki, R., Goldberg, J. M. & Sandler, M. B. (1999), `Information-theoretic constraints


on noise shaping networks with minimum noise power gain', Int. J. Circ. Theor. Appl.
27, 437{441.
Norsworthy, S. R., Schreier, R. & Temes, G. C. (1997), Delta-Sigma Data Converters, 1st
edn, IEEE Press, New York.

Ogata, K. (1997), Modern Control Engineering, 3rd edn, Prentice-Hall, Upper Saddle River,
NJ.

Oizumi, J. & Kimura, M. (1957), `Design of conditionally stable feedback systems', IRE
Trans. Circuit Theory 4(3), 157{166.

Okamoto, T., Maruyama, Y. & Yukawa, A. (1993), `A stable high-order delta-sigma


modulator with an FIR spectrum distributor', IEEE J. Solid-State Circuits
28(7), 730{735.
Orchard, H. J., Temes, G. C. & Cataltepe, T. (1985), `Sensitivity formulas for terminated
lossless two-ports', IEEE Trans. Circuits Syst. 32(5), 459{466.

Park, Y.-I., Karthikeyan, S., Tsay, F. & Bartolome, E. (2001), A 10b 100Msample/s CMOS
pipelined ADC with 1.8V power supply, in `Int. Solid-State Circuits Conf. Digest Tech.
Papers', pp. 130{131.

Pavan, S., Tsividis, Y. P. & Nagaraj, K. (2000), `Widely programmable high-frequency


continuous-time lters in digital CMOS technology', IEEE J. Solid-State Circuits
35(4), 503{511.
Pecseli, H. L. (2000), Fluctuations in Physical Systems, 1st edn, Cambridge University
Press, Cambridge.

201
Pelgrom, M. J. M., Tuinhout, H. P. & Vertregt, M. (1998), Transistor matching in analog
CMOS applications, in `IEEE Int. Electron Devices Meeting Tech. Digest', pp. 915{918.
Pellon, L. E. (1998), RF-to-digital receivers employing bandpass multibit  ADC
architectures, in `GaAs IC Symp.', pp. 11{14.
Popper, K. R. (1959), The Logic of Scienti c Discovery, Hutchinson, London.
Premoli, A. (1973), `The MUCROMAF polynomials: An approach to the maximally at
approximation of RC active lters with low sensitivity', IEEE Trans. Circuits Syst.
20(1), 77{80.
Rabiner, L. R. & Gold, B. (1975), Theory and Application of Digital Signal Processing, 1st
edn, Prentice-Hall, Englewood Cli s, NJ.
Rao, N., Balan, V. & Contreras, R. (1999), `A 3-V, 10{100-MHz continuous-time
seventh-order 0.05 equiripple linear phase lter', IEEE J. Solid-State Circuits
34(11), 1676{1682.
Razavi, B. (1998), RF Microelectronics, 1st edn, Prentice Hall, New Jersey.
Roberge, J. K. (1975), Operational Ampli ers: Theory and Practice, Wiley, New York.
Roberts, R. A. & Mullis, C. T. (1987), Digital Signal Processing, 1st edn, Addison-Wesley,
Reading, MA.
Scharf, L. L. (1991), Statistical Signal Processing, 1st edn, Addison-Wesley, Reading, MA.
Schreier, R. (1993), `An empirical study of high-order single-bit delta-sigma modulators',
IEEE Trans. Circuits Syst. II 40(8), 461{466.
Schreier, R. (1994), `On the use of chaos to reduce idle-channel tones in delta-sigma
modulators', IEEE Trans. Circuits Syst. I 41(8), 539{547.
Schreier, R., Goodson, M. V. & Zhang, B. (1997), `An algorithm for computing positively
invariant sets for delta-sigma modulators', IEEE Trans. Circuits Syst. I 44(1), 38{44.
Schreier, R. & Yang, Y. (1992), Stability tests for single-bit sigma-delta modulators with
second-order FIR noise transfer functions, in `IEEE Int. Symp. Circuits Syst.',
pp. 1316{1319.
202
Semiconductor Industry Association (1999), International Technology Roadmap for
Semiconductors: 1999 edition, SEMATECH.

Shiryayev, A. N. (1992), Selected Works of A. N. Kolmogorov, Kluwer, Dordrecht.

Skellern, D. J., Lee, L. H. C., McDermott, T., Weste, N., Dalton, J., Graham, J., Wong,
T. F., Myles, A. F., Percival, T. M. P. & Ryan, P. J. (1997), `A high-speed wireless
LAN', IEEE Micro 17(1), 40{47.

Smith, W. E. (1965), `Energy and dispersion in electromagnetic systems', Aust. J. Phys.


18(4), 287{295.
Snelgrove, W. M. & Sedra, A. S. (1986), `Synthesis and analysis of state-space active lters
using intermediate transfer functions', IEEE Trans. Circuits Syst. 33(3), 287{301.

Spang, H. A. & Schultheiss, P. M. (1962), `Reduction of quantizing noise by use of


feedback', IRE Trans. Commun. Syst. 10(4), 373{380.

Stillwell, J. (1989), Mathematics and Its History, Springer-Verlag, New York.

Sumanen, L., Waltari, M. & Halonen, K. A. I. (2001), `A 10-bit 200-Ms/s CMOS parallel
pipeline A/D converter', IEEE J. Solid-State Circuits 36(7), 1048{1055.

Sung, H.-K. & Hara, S. (1988), `Properties of sensitivity and complementary sensitivity
functions in single-input single-output digital control systems', Int. J. Control
48(6), 2429{2439.
Tao, H., Toth, L. & Khoury, J. M. (1999), `Analysis of timing jitter in bandpass sigma-delta
modulators', IEEE Trans. Circuits Syst. II 46(8), 991{1001.

Tellegen, B. D. H. (1952), `A general network theorem, with applications', Philips Res. Rep.
7(4), 259{269.
Temes, G. C. & Orchard, H. J. (1973), `First-order sensitivity and worst case analysis of
doubly terminated reactance two-ports', IEEE Trans. Circuit Theory 20(5), 567{571.

Tewksbury, S. K. & Hallock, R. W. (1978), `Oversampled, linear predictive and


noise-shaping coders of order n>1', IEEE Trans. Circuits Syst. 25(7), 436{447.

203
Thiele, L. (1986), `On the sensitivity of linear state-space systems', IEEE Trans. Circuits
Syst. 33(5), 502{510.

Thompson, D. W. (1917), On Growth and Form, 1st edn, Cambridge University Press,
Cambridge.

Thornton, R. D. (1957), `Active RC networks', IEEE Trans. Circuit Theory 4(3), 78{89.

Thornton, R. D., DeWitt, D., Chenette, E. R. & Gray, P. E. (1966), Characteristics and
Limitations of Transistors, Wiley, New York.

Thoss, S., Li, B. & Tenhunen, H. (2000), Loop coecient optimization for cascaded
sigma-delta modulators { a case study, in `IEEJ International Analog VLSI Workshop
(June 2-3, 2000, Stockholm, Sweden)', pp. 51{56.

Toumazou, C., Lidgey, F. J. & Haigh, D. G. (1990), Analogue IC design: the current-mode
approach, IEE, London.

Tsividis, Y. P. & Gopinathan, V. (1994), Continuous-time lters, in `Design of


Analog-Digital VLSI Circuits for Telecommunications and Signal Processing', Prentice
Hall, Engelwood Cli s, New Jersey.

Van den Bosch, A., Borremans, M. A. F., Steyaert, M. S. J. & Sansen, W. (2001), `A 10-bit
1GSample/s Nyquist current-steering CMOS D/A converter', IEEE J. Solid-State
Circuits 36(3), 315{324.

Veendrick, H. (2001). personal communication.

Vidal, P. (1969), Non-linear Sampled-data Systems, 1st edn, Gordon and Breach Science
Publishers.

Vittoz, E. A. (1993), Very low power circuit design: fundamentals and limits, in `IEEE Int.
Symp. Circuits Syst.', Vol. 2, pp. 1439{1442.

Vittoz, E. A. (1994), Low-power design: ways to approach the limits, in `Int. Solid-State
Circuits Conf. Digest Tech. Papers', pp. 14{18.

Vleugels, K. (2000). personal communication.

204
Wang, Y.-T. & Razavi, B. (2000), `An 8-Bit 150-MHz CMOS A/D converter', IEEE J.
Solid-State Circuits 35(3), 308{317.

Wong, S. & Salama, C. A. T. (1983), `Impact of scaling on MOS analog performance', IEEE
J. Solid-State Circuits 18(1), 106{114.

Wooley, B. (2000). personal communication.

Yahia, A., Benabes, P. & Kielbasa, R. (2001), Bandpass delta-sigma modulators synthesis
with high loop delay, in `IEEE Int. Symp. Circuits Syst.', Vol. 1, pp. 344{347.

Youla, D. C. (1961), `A new theory of cascade synthesis', IRE Trans. Circuit Theory
8(3), 244{260.
Zames, G. (1981), `Feedback and optimal sensitivity: Model reference transformations,
multiplicative seminorms, and approximate inverses', IEEE Trans. Automat. Contr.
26(2), 301{320.
Zeiger, H. P. & McEwen, A. J. (1974), `Approximate linear realizations of given dimension
via Ho's algorithm', IEEE Trans. Automat. Contr. 19(2), 153.

Zhou, K. (1996), Robust and optimal control, Prentice-Hall, Upper Saddle River, NJ.

Zverev, A. I. (1967), Handbook of Filter Synthesis, Wiley, New York.

205

You might also like