0% found this document useful (0 votes)
34 views195 pages

2018 Design of Power-Efficient Highly Digital Analog-To-Digital Converters Xing Eta

Uploaded by

Omar Abd Elrhman
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
34 views195 pages

2018 Design of Power-Efficient Highly Digital Analog-To-Digital Converters Xing Eta

Uploaded by

Omar Abd Elrhman
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 195

Xinpeng Xing Peng Zhu

Georges Gielen

Design of Power-Efficient
Highly Digital
Analog-to-Digital Converters
for Next-Generation Wireless
Communication Systems

123
Xinpeng Xing Georges Gielen
Graduate School at Shenzhen Departement Elektrotechniek,
Tsinghua University ESAT-MICAS
Shenzhen Katholieke Universiteit Leuven
China Leuven
Belgium
Peng Zhu
Zhongguancun Dongsheng Technology Park
Analog Devices, Inc.
Beijing
China

ISSN 1860-4862 ISSN 1860-4870 (electronic)


Signals and Communication Technology
ISBN 978-3-319-66564-1 ISBN 978-3-319-66565-8 (eBook)
https://doi.org/10.1007/978-3-319-66565-8
Library of Congress Control Number: 2017950276

© Springer International Publishing AG 2018

This Springer imprint is published by Springer Nature


The registered company is Springer International Publishing AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface

Mobile devices motivate the continuous development of the communication


industry, meaning low-power designs are in great demand. Analog-to-digital con-
verters (ADCs) build bridge between analog front-end and digital cores, and play a
more important role in emerging transceiver architectures. On the other hand, due to
CMOS technology scaling, the ADC design suffers from design issues such as
decreasing headroom voltage. Recently, one popular direction of ADC design is
shifting more functions from the voltage and the analog domains to the time and the
digital domains, by using a voltage-controlled oscillator (VCO). The implementa-
tion of circuits in the time domain immediately takes advantage again of the
technology scaling with reduced gate delay.
This book is a result of our research work in ESAT-MICAS, K.U. Leuven, in the
field of highly digital ADC design. Recently, highly digital ADC has drawn more
and more interest from both academy and industry, and the increasing number of
publication on this topic is an evidence. We believe this is a natural evolution when
IC designers are facing ongoing-scaling CMOS technologies. With this common
interest, however, we cannot find systematic book on this topic but only papers and
book chapters as far as we know. It has been our motivation to present our work in
highly digital ADC as a book. This collection is mainly based on our own five chip
designs in ESAT-MICAS, and also, some previous contributions by various
researchers to this field are also included.
This book focuses on the systematic design of power-efficient highly digital
ADC for future communication applications, with both architecture- and
circuit-level innovations. The first two designs are 40MHz-BW 12bit CT DR ADCs
implemented in 90nm CMOS, with quantizations done by highly digital
VCO-based quantizer. Various circuit-level techniques are applied for power
reduction, including the shaped switched capacitor digital-to-analog converter (SC
DAC), the look-up-table (LUT)-based digital calibration, and the current-sharing
feedforward-compensated OTA. The third design is a 40MHz-BW two-step
open-loop VCO-based ADC in 40nm CMOS. With hardware-economic structure
and mostly digital building blocks, an excellent FoM of 42fJ/step is obtained.
However, only first-order noise shaping is realized for the whole ADC, limiting its
SNR performance. To go further, in our fourth design, a nonlinearity cancellation
0–2 MASH DR ADC structure with innovative dual-input VCO-based quantizer is
adopted. The optimized systematic parameters and the highly digital circuit blocks
extend the FoM of the state-of-the-art high-bandwidth DR ADCs to 35fJ/step. With
second-order noise shaping, the ADC performance is improved, and one analog
front-end integrator is still needed, however. In the final design of this work, a
VCO-based integrator is proposed to replace the power-hungry analog integrator in
the traditional DR ADC topology, realizing second-order DR ADC without any
analog integrator. 74-dB SFDR and a FoM of 52fJ/step over a 40MHz bandwidth
have been achieved in the demonstration.
With the state-of-the-art power efficiencies, the presented highly digital ADCs
are very suitable for the applications of next-generation wireless communication
standards, including but not limited to 802.11n and LTE. Furthermore, the design
methodology and design innovations described in this book could also be applied to
ADCs for other applications.

Shenzhen, China Xinpeng Xing


Beijing, China Peng Zhu
Leuven, Belgium Georges Gielen
March 2017
Contents

1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Background and Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1.1 Communication Evolution . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1.2 CMOS Technology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.1.3 Wireless Receiver Architectures . . . . . . . . . . . . . . . . . . . . . 6
1.2 The Research Objective of the Book . . . . . . . . . . . . . . . . . . . . . . . 8
1.3 The Book Organization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2 A/D Converters and Applications. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.2 ADC Specifications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.2.1 ADC Speed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.2.2 ADC Accuracy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.2.3 ADC FoM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.3 ADC Architectures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.3.1 Flash ADC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.3.2 Two-Step ADC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.3.3 Pipelined ADC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.3.4 SAR ADC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.3.5 Delta-Sigma ADC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.3.6 ADC Architecture Summmary and Comparison . . . . . . . . . 26
2.4 Application of ADC in Communications . . . . . . . . . . . . . . . . . . . . 28
2.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3 Continuous-Time Delta-Sigma Modulators . . . . . . . . . . . . . . . . . . . . . 37
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.2 DSM Basics: Oversampling and Noise-Shaping . . . . . . . . . . . . . . . 38
3.3 DSM Structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.3.1 Discrete-Time and Continuous-Time DSMs . . . . . . . . . . . . 41
3.3.2 1st-Order and Higher-Order DSMs . . . . . . . . . . . . . . . . . . . 44
3.3.3 Single-Loop and MASH DSMs . . . . . . . . . . . . . . . . . . . . . 45
3.3.4 The D R-0 and 0-D R MASH Structures . . . . . . . . . . . . . . 46
3.3.5 Single-Bit and Multi-bit DSMs . . . . . . . . . . . . . . . . . . . . . . 48
3.3.6 Feedforward, Feedback and Hybrid DSMs . . . . . . . . . . . . . 50
3.3.7 Resonator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
3.3.8 Feedin Paths . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
3.4 CT DSM Nonidealities and Modeling . . . . . . . . . . . . . . . . . . . . . . 53
3.4.1 Loop Filter Nonidealities and Modeling . . . . . . . . . . . . . . . 54
3.4.2 DAC Nonidealities and Modelling . . . . . . . . . . . . . . . . . . . 59
3.4.3 Quantizer Nonidealities and Modeling . . . . . . . . . . . . . . . . 63
3.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
4 VCO-Based ADCs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
4.2 VCO-Based Quantizers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
4.2.1 Single-Phase Counting VCO-Based Quantizer . . . . . . . . . . 68
4.2.2 Multi-phase Counting VCO-Based Quantizer . . . . . . . . . . . 68
4.2.3 Frequency-Type VCO-Based Quantizer . . . . . . . . . . . . . . . 69
4.2.4 Phase-Type VCO-Based Quantizer . . . . . . . . . . . . . . . . . . . 72
4.3 Closed-Loop VCO-Based DSMs . . . . . . . . . . . . . . . . . . . . . . . . . . 72
4.3.1 DSM with Frequency-Type VCO-Based Quantizer . . . . . . . 73
4.3.2 DSM with Phase-Type VCO-Based Quantizer . . . . . . . . . . 74
4.3.3 DSM with Residual-Cancelling VCO-Based Quantizer . . . . 74
4.4 Open-Loop VCO-Based ADCs . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
4.4.1 VCO-Based ADC with Background Digital Calibration . . . 76
4.4.2 VCO-Based ADC with Counting and Foreground
Digital Calibration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
4.4.3 VCO-Based ADC with PWM Precoding . . . . . . . . . . . . . . . 78
4.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
5 CT DSM ADCs with VCO-Based Quantization . . . . . . . . . . . . . . . . . 83
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
5.2 A 40 MHz-BW 12-Bit CT DSM with Digital Calibration and
Shaped SC DAC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
5.2.1 Structure of the CT DSM . . . . . . . . . . . . . . . . . . . . . . . . . . 84
5.2.2 Delta-Sigma Modulator Building Blocks Design . . . . . . . . . 89
5.2.3 Measurement Setup and Experimental Results . . . . . . . . . . 95
5.3 A 40 MHz-BW 12-Bit CT DSM with Capacitive Local
Feedback and Current-Sharing OTA . . . . . . . . . . . . . . . . . . . . . . . . 97
5.3.1 System Design of the 40 MHz 12-Bit CT DSM . . . . . . . . . 99
5.3.2 Circuit Design of the DSM Building Blocks . . . . . . . . . . . . 101
5.3.3 Measurement Results and Discussions . . . . . . . . . . . . . . . . 104
5.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
6 Two-Step Open-Loop VCO-Based ADC . . . . . . . . . . . . . . . . . . . .... 109
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .... 109
6.2 Architecture Design of Two-Step Open-Loop VCO-Based
ADC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .... 110
6.2.1 A Two-Step Open-Loop VCO-Based ADC
Architecture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .... 110
6.2.2 Nonidealities of the Two-Step Open-Loop
VCO-Based ADC . . . . . . . . . . . . . . . . . . . . . . . . . . . . .... 113
6.3 Circuit Implementation of the Two-Step Open-Loop
VCO-Based ADC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
6.3.1 VCO-Based Quantizer Design . . . . . . . . . . . . . . . . . . . . . . . 117
6.3.2 DAC and Subtractor Design . . . . . . . . . . . . . . . . . . . . . . . . 119
6.4 Experimental Results and Discussions . . . . . . . . . . . . . . . . . . . . . . 122
6.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
7 VCO-Based 0-DR MASH ADC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
7.2 Architecture Analysis of the 0-DR MASH VCO-Based ADC . . . . 128
7.2.1 0-DR MASH VCO-Based ADC . . . . . . . . . . . . . . . . . . . . . 128
7.2.2 Nonlinearity-Cancellation Robustness Against PVT
Variations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .... 131
7.2.3 System Architecture of a 0–2 MASH VCO-Based
DR ADC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .... 132
7.2.4 Delay Matching Technique . . . . . . . . . . . . . . . . . . . . . .... 134
7.3 Circuit Implementation of the 0–2 MASH VCO-Based
DR ADC. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
7.3.1 Three-Input Adder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
7.3.2 VCO-Based Quantizer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
7.3.3 Integrator. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
7.3.4 DACs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
7.3.5 Interface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
7.4 Experimental Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
7.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
8 Fully-VCO-Based High-Order DR ADC . . . . . . . . . . . . . . . . . . . . . . . 153
8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
8.2 Integrators. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
8.2.1 Traditional Analog Integrator . . . . . . . . . . . . . . . . . . . . . . . 153
8.2.2 VCO-Based Integrator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
8.2.3 Fully-VCO-Based DR ADC Structure . . . . . . . . . . . . . . . . . 157
8.3 Design Example: A Fully-VCO-Based 0-2 MASH VCO-Based
DR ADC. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
8.3.1 System Architecture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
8.3.2 Circuit Implementation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160
8.3.3 Experimental Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
8.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
9 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
9.1 Summary and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
9.2 Suggestions for Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179
Abbreviations

2G Second Generation
3G Third Generation
AAF Anti-Aliasing Filtering
AC Alternating Current
A/D Analog-to-Digital
ADC Analog-to-Digital Converter
ADSL Asymmetric Digital Subscriber Line
AFE Analog Front-End
AGC Automatic Gain Control
APWM Asynchronous Pulse-Width Modulator
ASIC Application-Specific Integrated Circuit
BB BaseBand
BPF Band-Pass Filter
BW Bandwidth
CF Crest Factor
CM Common Mode
CMFB Common-Mode Feedback
CMOS Complementary Metal Oxide Semiconductor
CS Current-Steering or Current-Sharing
CT Continuous-Time
D/A Digital-to-Analog
DAC Digital-to-Analog Converter
DC Direct Current
DCM Duty Cycle Modulation
DEM Dynamic Element Matching
DFF Direct Feedforward
DNCF Digital Noise Cancellation Filter
DNL Differential Nonlinearity
DR Dynamic Range
DSM Delta-Sigma Modulator
DSP Digital Signal Processing
DT Discrete-Time
DVD Digital Versatile Disc
EDA Electronic Design Automation
ELD Excess Loop Delay
ENOB Effective Number of Bit
ERBW Effective Resolution Bandwidth
FET Field Effect Transistor
FF FeedForward or Flip-Flop
FIR Finite Impulse Response
FM Fading Margin
FoM Figure of Merit
FS Full Scale
GBW Gain bandwidth product
GP General Purpose
GSM Global System for Mobile Communications
HSDPA High Speed Downlink Packet Access
IC Integrated Circuits
ICO Current-Controlled Oscillator
IF Intermediate Frequency
IIT Impulse Invariant Transformation
IM2/3 2nd/3rd-order Intermodulation Distortion
INL Integrated Nonlinearity
IPTV Internet Protocol Television
I/Q In-phase/Quadrature
ISI Inter-Symbol Interference
ISSCC International Solid-State Circuits Conference
LFSR Linear Feedback Shift Register
LHP Left Half Plane
LMS Least Mean Square
LNA Low-Noise Amplifier
LO Local Oscillator
LPF Low-Pass Filter
LSB Least Significant Bit
LTE Long Term Evolution
LUT Look-Up Table
LVDS Low-Voltage Differential Signalling
MASH Multi-stAge noise SHaping
MDAC Multiplying DAC
MIM Metal-Insulator-Metal
MOS Metal-Oxide-Semiconductor
MSB Most Significant Bit
NAND Negated AND
NF Noise Figure
NFC Near Field Communication
NM Noise Margin
NMOS N-channel MOSTFET
NRZ Non Return-to-Zero
NTF Noise Transfer Function
OSR Oversampling Ratio
OTA Operational Transconductance Amplifier
PA Power Amplifier
PC Personal Computer
PCB Printed Circuit Board
PD Phase Detector
PLL Phase-Locked Loop
PMOS P-channel MOSFET
PM Phase Margin
PSD Power Spectral Density
PSRR Power Supply Rejection Ratio
PVT Pcocess Voltage Temperature
PWM Pulse-Width Modulator
RF Radio Frequency
RMS Root-Mean-Square
ROM Read-Only Memory
RSR Receivable Signal Range
RZ Return-to-Zero
SA Sense-Amplifier
SAFF Sense-Amplifier Flip-Flop
SAR Successive Approximation Register
SAW Surface Acoustic Wave
SC Switched-Capacitor
SFDR Spurious-Free Dynamic Range
S/H Sample and Hold
SiGe Silicon-Germanium
SNDR Signal-to-Noise-Distortion Ratio
SNR Signal-to-Noise Ratio
SoC System on Chip
SP Standard Performance
SQNR Signal-to-Quantization-Noise Ratio
STF Signal Transfer Function
TDMA Time Division Multiple Access
THD Total Harmonic Distortion
TSPC FF True-Single-Phase Clock Flip-Flop
TV Television
VCO Voltage-Controlled Oscillator
VDSL Very-High-Bit-Rate Digital Subscriber Line
V-F Voltage-Frequency
VGA Variable Gain Amplifier
VGLNA Variable Gain Low-Noise Amplifier
VTC Voltage-to-Time Converter
WLAN Wireless Local Area Network
XOR Exclusive OR
Figures

Fig. 1.1 Data rate comparison between different wireline internet


accesses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 2
Fig. 1.2 Evolution of wireless communications [3] . . . . . . . . . . . . . . . . .. 3
Fig. 1.3 Relative global shipments of four consumer electronics from
2012 to 2017 (estimated values for 2013 and 2017) [4] . . . . . .. 4
Fig. 1.4 CMOS technology scaling over the last 45 years and prediction
for the next 4 years . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 5
Fig. 1.5 Decreasing supply voltage and relatively constant threshold
voltage with CMOS scaling [7] . . . . . . . . . . . . . . . . . . . . . . . . .. 5
Fig. 2.1 Generalized architecture, operation and waveforms of an ADC,
both with Nyquist sampling and oversampling . . . . . . . . . . . . .. 17
Fig. 2.2 The linear model and input-output transfer function of the
quantizer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
Fig. 2.3 N-bit flash ADC architecture . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
Fig. 2.4 Block diagram of a (M þ N)-bit two-step ADC . . . . . . . . . . . . 22
Fig. 2.5 Simplified architecture of pipelined ADC . . . . . . . . . . . . . . . . . . 23
Fig. 2.6 Structure of SAR ADC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
Fig. 2.7 Block diagram of DSM ADC . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
Fig. 2.8 Accuracy-bandwidth tradeoff of ADC architectures . . . . . . . . . . . 27
Fig. 2.9 FoMs of different ADC architectures in ISSCC publications
[27] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 27
Fig. 2.10 A receiver is partitioned into three basic parts. . . . . . . . . . . . . .. 28
Fig. 2.11 Receiver AFE is a cascaded system . . . . . . . . . . . . . . . . . . . . .. 28
Fig. 2.12 Accuracy relationship between the ADC and the AFE in a
telecom receiver . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 32
Fig. 3.1 Structure of a whole CT DSM ADC. The corresponding
illustrative signals and their spectral plots illustrate the
structure’s operation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 39
Fig. 3.2 A generalized DSM structure (top) and its linear model
(bottom) for performance analysis . . . . . . . . . . . . . . . . . . . . . . .. 40
Fig. 3.3 Structure of a DT DSM (top) and its CT counterpart
(bottom). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 41
Fig. 3.4 Popular circuit implementations of DT (left) and CT (right)
integrators used in the DSM loop filter . . . . . . . . . . . . . . . . . . .. 42
Fig. 3.5 Different DAC current profiles used in CT DSMs: NRZ (left),
RZ (middle) and SC (right) . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 42
Fig. 3.6 Performance of Nth-order 3-bit DSM with different OSRs . . . .. 44
Fig. 3.7 The architecture of a discrete-time 2-2 MASH DSM,
the total noise shaping is 4th-order . . . . . . . . . . . . . . . . . . . . . . . 46
Fig. 3.8 Block diagram of a D R-0 MASH structure . . . . . . . . . . . . . . . . 47
Fig. 3.9 Block diagram of a modified DR-0 MASH structure . . . . . . . . . . 47
Fig. 3.10 Block diagram of a 0-DR MASH structure . . . . . . . . . . . . . . . . . 48
Fig. 3.11 Single-bit quantizer: the linear model, the input-output transfer
function and its highly nonlinear gain . . . . . . . . . . . . . . . . . . . .. 49
Fig. 3.12 The concept of DEM: 3-bit DACs without (left) and with
(right) DEM are illustrated for comparison . . . . . . . . . . . . . . . .. 50
Fig. 3.13 Different structures of generalized Nth-order DSMs:
feedforward (top), feedback (upper middle) and hybrid ones
(lower middle and bottom) . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 51
Fig. 3.14 Resonator in the loop filter for the generation of complex
NTF zeroes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 52
Fig. 3.15 Direct coupling paths from the DSM input to integrators
and quantizer for integrator output swing reduction . . . . . . . . .. 53
Fig. 3.16 An active-RC integrator with nonideal OTA and three input
voltages in a CT DSM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 54
Fig. 3.17 An active-RC integrator with nonideal OTA and input parasitic
capacitor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 57
Fig. 3.18 An accurate and convenient nonideal model of an
active-RC integrator in Matlab Simulink . . . . . . . . . . . . . . . . . .. 58
Fig. 3.19 Variable capacitor for RC time constant tuning in an
active-RC integrator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 59
Fig. 3.20 Clock-jitter-induced noise in single-bit DACs: RZ (left),
NRZ (middle) and SC (right) . . . . . . . . . . . . . . . . . . . . . . . . . .. 61
Fig. 3.21 Clock-jitter-induced noise in multi-bit DACs: RZ (left),
NRZ (middle) and SC (right) . . . . . . . . . . . . . . . . . . . . . . . . . .. 62
Fig. 3.22 Inter-symbol inteference (ISI) of a NRZ DAC (right);
the error is marked in darkgray . . . . . . . . . . . . . . . . . . . . . . . . .. 62
Fig. 3.23 The zero-order feedback path for ELD compensation in
feedback-type DSMs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 63
Fig. 3.24 The ELD compensation scheme without extra analog adder:
the zero-order feedback path is composed by a digital
differentiator and an analog integrator . . . . . . . . . . . . . . . . . . . .. 64
Fig. 4.1 Single-phase (left) and multi-phase (right) counting
quantizer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 69
Fig. 4.2 Structure, behavioral model and illustrative spectrum of the
frequency-type VCO-based quantizer . . . . . . . . . . . . . . . . . . . .. 70
Fig. 4.3 Structure of the phase-type VCO-based quantizer; the
additional digital frequency output is used for DSM ELD
compensation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 72
Fig. 4.4 Architecture of a 3rd-order CT DSM with frequency-type
VCO-based quantizer [4] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 73
Fig. 4.5 Topology of a 4th-order CT DSM with phase-type VCO-based
quantizer [8] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 74
Fig. 4.6 Structure of a 2nd-order CT DSM with residual-cancelling
VCO-based quantizer [10] . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 75
Fig. 4.7 Architecture of a 1st-order open-loop Delta-Sigma ADC
with VCO nonlinearity calibration (only one path
is shown here) [11] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 76
Fig. 4.8 Concept of the VCO nonlinearity background digital
calibration used in Fig. 4.7 [11] . . . . . . . . . . . . . . . . . . . . . . . .. 77
Fig. 4.9 Structure of an open-loop Delta-Sigma ADC with coarse and
fine quantizations and VCO nonlinearity calibration [13] . . . . .. 78
Fig. 4.10 Concept of the foreground digital calibration used in
Fig. 4.9 [13] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 78
Fig. 4.11 Architecture and illustrative signals of an open-loop
VCO-based ADC with PWM precoding [15] . . . . . . . . . . . . . .. 79
Fig. 4.12 Implementation of an asynchronous PWM with hysteresis
single-bit quantizer [16] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 79
Fig. 5.1 Structure of the 40 MHz 12-bit CT DSM . . . . . . . . . . . . . . . . .. 85
Fig. 5.2 CT DSM SNDR performance as a function of the scaled
OTA GBWs (normalized to 4, 2 and 2 GHz for loop filter
integrator one to three respectively) . . . . . . . . . . . . . . . . . . . . .. 86
Fig. 5.3 CT DSM SNDR performance as a function of the delay
for the three different feedback DACs . . . . . . . . . . . . . . . . . . . .. 87
Fig. 5.4 Block diagram of the digital calibration of the DAC. The
values in the look-up table are arbitrary values, for illustrative
purpose . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 89
Fig. 5.5 Block diagram of the 4-stage full-feedforward compensated
operational transconductance amplifier . . . . . . . . . . . . . . . . . . .. 90
Fig. 5.6 SNDR performance of the CT DSM as a function of the
relative position of the poles and zeros in the doublets . . . . . . .. 91
Fig. 5.7 AC simulation result of (internal) gain of the operational
tranconductance amplifier used in the first integrator . . . . . . . . .. 92
Fig. 5.8 Circuit implementation of the shaped SC DAC (right) and its
driving circuit (left) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 93
Fig. 5.9 The output current pulse of the shaped SC DAC, compared
to the traditional pulse shape, as used in our design . . . . . . . . .. 93
Fig. 5.10 Schematic of the current-steering DAC cells . . . . . . . . . . . . . . .. 94
Fig. 5.11 Die photo of the 40 MHz 12-bit CT DSM . . . . . . . . . . . . . . . .. 96
Fig. 5.12 Measurement setup for the 40 MHz 12-bit CT DSM . . . . . . . .. 96
Fig. 5.13 Measured spectra with (right) and without (left) digital
calibration for 938 kHz and 7.5 MHz input signals. . . . . . . . . .. 97
Fig. 5.14 SNR/SNDR versus input amplitude for 938 kHz input
signals (left) and 7.5 MHz input signals (right) . . . . . . . . . . . . .. 98
Fig. 5.15 Output spectrum for a 10:5 dB two-tone test with 10
and 10.5 MHz frequencies . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 98
Fig. 5.16 Block diagram of the 40 MHz 12-bit CT DSM. . . . . . . . . . . . .. 99
Fig. 5.17 Schematic of the current-sharing feedforward OTA. . . . . . . . . . . 103
Fig. 5.18 AC simulation results of the current-sharing feedforward
OTA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
Fig. 5.19 Die photo of the low-power 40 MHz 12-bit CT DSM . . . . . . . . 104
Fig. 5.20 Measured spectra for a 3 dBFS 938 kHz (left) and a 7.5
MHz (right) single-tone measurement . . . . . . . . . . . . . . . . . . . . . 105
Fig. 5.21 Measured SNR/SNDR versus input amplitude for a 938 kHz
input signal (left) and a 7.5 MHz input signal (right) . . . . . . . . . 105
Fig. 5.22 Spectrum of two-tone measurement (9:5 dBFS, 10 MHz
and 10.5 MHz input signals) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
Fig. 6.1 Proposed two-step open-loop VCO-based ADC architecture
and the VCO nonlinearity mitigation concept . . . . . . . . . . . . . . . 111
Fig. 6.2 ADC SNDR performance as a function of the normalized
gain of the coarse ADC, fine ADC and DAC . . . . . . . . . . . . . . . 114
Fig. 6.3 ADC SNDR performance as a function of the delays of the
coarse ADC and the DAC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
Fig. 6.4 ADC SNDR performance as a function of the standard
deviation of the current of the DAC cells. For comparison,
both cases with and without DEM are showed . . . . . . . . . . . . . . 116
Fig. 6.5 ADC SNDR performance as a function of the clock jitter:
around 15ps jitter can be tolerated for 10-bit accuracy . . . . . . . . 116
Fig. 6.6 Schematic of the 15-stage ring VCO in the quantizer
(single-ended); the delay cell circuit is shown in the blue
dashed box . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
Fig. 6.7 Large-signal simulation of a 5-bit frequency-type VCO-based
quantizer. Both the single-ended result (left) and the
pseudo-differential result (right) are presented . . . . . . . . . . . . . . . 118
Fig. 6.8 Small-signal simulation of a 5-bit frequency-type VCO-based
quantizer; with around 20 dBFS input, the 3rd-order
distortion is about 76 dB . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
Fig. 6.9 The simulated VCO phase noise; it is 113 dBc/Hz at an offset
frequency of 10 MHz . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
Fig. 6.10 Schematic of the current-steering DAC; the resistor connected
to the ADC input and the DAC output forms a passive
subtractor. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
Fig. 6.11 Noise simulation of the feedback current-steering DAC and the
passive subtractor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
Fig. 6.12 Transistor-level simulation result of the two-step open-loop
VCO-based ADC. Top left PSD plot of the coarse quantizer
output; top right PSD plot of the fine quantizer output; bottom
left PSD plot of the overall ADC output . . . . . . . . . . . . . . . . . . . 121
Fig. 6.13 Die photo of the two-step open-loop VCO-based ADC. . . . . . . . 122
Fig. 6.14 Measured results of the two-step 1st-order VCO-based DSM
with 1 dBFS 12 MHz sine input. Top left PSD plot of the
coarse quantizer output; top right PSD plot of the fine quantizer
output; bottom left PSD plot of the overall ADC output . . . . . . . 123
Fig. 6.15 The measured ADC SNR/SNDR as a function of the analog
input power . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
Fig. 7.1 VCO-based quantizer in a single-loop DR modulator . . . . . . . . . 128
Fig. 7.2 Block diagram of a 0-DR MASH VCO-based ADC . . . . . . . . . . 128
Fig. 7.3 The nonlinearity-cancellation principle in a 0-DR MASH
VCO-based ADC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
Fig. 7.4 Equivalent signal model of the proposed 0-DR MASH
VCO-based ADC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
Fig. 7.5 SNDR of the 0–1 MASH and 0–2 MASH VCO-based ADCs
as a function of the stage-gain mismatch . . . . . . . . . . . . . . . . . . . 132
Fig. 7.6 ADC SFDR performance as a function of the normalized
gain of the VCOs and DACs . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
Fig. 7.7 The presented 0–2 MASH VCO-based DR ADC . . . . . . . . . . . . 133
Fig. 7.8 SNDR performance as a function of the clock jitter . . . . . . . . . . 134
Fig. 7.9 Block diagram of a simplified CT loop to approximately
represent the 2nd stage in a MASH ADC . . . . . . . . . . . . . . . . . . 134
Fig. 7.10 Impulse invariance transformation of the 2nd stage in the
presented 0–2 MASH VCO-based ADC . . . . . . . . . . . . . . . . . . . 135
Fig. 7.11 Topology of the dual-input VCO-based fine quantizer
(single-ended) to realize the three-input adder . . . . . . . . . . . . . . . 137
Fig. 7.12 Schematic of the sense-amplifier-based flip-flop . . . . . . . . . . . . . 138
Fig. 7.13 Simulated phase noise performance of the VCO . . . . . . . . . . . . . 138
Fig. 7.14 Schematic of the integrator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
Fig. 7.15 Discretely tunable capacitor for the loop filter . . . . . . . . . . . . . . . 140
Fig. 7.16 Schematic of the operational amplifier in the integrator . . . . . . . 141
Fig. 7.17 Small-signal equivalent block diagram of the operational
amplifier . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
Fig. 7.18 ADC SNDR performance as a function of the standard
deviation of the current of the global and internal
DAC cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
Fig. 7.19 Block diagram of the Wallace tree encoder . . . . . . . . . . . . . . . . . 143
Fig. 7.20 Schematic of the mirror adder implemented in the
unary-to-binary encoder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
Fig. 7.21 Chip microphotograph of the 0–2 MASH DR ADC
prototype . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
Fig. 7.22 Measurement setup for the 0–2 MASH DR ADC prototype . . . . 146
Fig. 7.23 Measured PSD results of the 0–2 MASH DR ADC chip
prototype with an 8-MHz input. Top first stage; middle
second stage; bottom complete ADC output . . . . . . . . . . . . . . . . 147
Fig. 7.24 Measured PSD results of the 0–2 MASH DR ADC chip
prototype with a 4-MHz input . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
Fig. 7.25 Measured PSD results of the 0–2 MASH DR ADC chip
prototype with a 1-MHz input . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
Fig. 7.26 Measured SNDR versus the input amplitude for different input
frequencies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
Fig. 7.27 Output spectrum of the two-tone measurement
(at 10 and 11 MHz) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
Fig. 8.1 Traditional analog R-C and Gm -C integrator . . . . . . . . . . . . . . . . 154
Fig. 8.2 Simplified diagram of a R-C integrator, in which the opamp
has a finite DC gain and a finite GBW . . . . . . . . . . . . . . . . . . . . 154
Fig. 8.3 Equivalent model of an R-C integrator, in which the opamp
has a finite gain and GBW . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
Fig. 8.4 Frequency-domain transfer function of a VCO . . . . . . . . . . . . . . 156
Fig. 8.5 Block diagram of a N-stage VCO-based 1st-order integrator. . . . 156
Fig. 8.6 Block diagram of a single-loop DR ADC with a VCO-based
integrator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
Fig. 8.7 Block diagram of a 2nd-order VCO-based DR ADC
with a VCO-based integrator [2] . . . . . . . . . . . . . . . . . . . . . . . . . 158
Fig. 8.8 Block diagram of the proposed 0-DR MASH ADC with
VCO-based integrators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
Fig. 8.9 Block diagram of the presented fully-VCO-based 0-2 MASH
DR ADC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
Fig. 8.10 Clock timing of the presented fully-VCO-based ADC. . . . . . . . . 159
Fig. 8.11 Signal model of the presented fully-VCO-based 0-2 MASH
VCO-based ADC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
Fig. 8.12 Schematic of the pseudo-differential VCO-based integrator . . . . . 161
Fig. 8.13 Simulated phase noise performance of the VCO used in the
integrator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
Fig. 8.14 Schematic of the SA-based buffer in the VCO-based
integrator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
Fig. 8.15 Schematic of the two-state phase detector . . . . . . . . . . . . . . . . . . 163
Fig. 8.16 Transfer function of the two-state phase detector . . . . . . . . . . . . 163
Fig. 8.17 Schematic of the source-switched DAC cells in the
VCO-based integrator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164
Fig. 8.18 Chip microphotograph of the fully-VCO-based 0-2
MASH DR ADC prototype . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164
Fig. 8.19 Measured PSD results of the fully-VCO-based 0-2
MASH DR ADC chip prototype with an 8-MHz input . . . . . . . . 165
Fig. 8.20 Measured PSD results of the fully-VCO-based 0-2
MASH DR ADC chip prototype with a 4-MHz input . . . . . . . . . 165
Fig. 8.21 Measured PSD results of the fully-VCO-based 0-2
MASH DR ADC chip prototype with a 2-MHz input . . . . . . . . . 166
Fig. 8.22 Measured ADC SNDR versus input amplitude . . . . . . . . . . . . . . 166
Fig. 8.23 Output spectrum of the two-tone measurement
(at 10 and 11 MHz) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
Fig. 8.24 Digital correction setup of the fully-VCO-based 0-2 MASH
DR ADC chip prototype . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
Fig. 8.25 Digitally corrected PSD results of the fully-VCO-based 0-2
MASH DR ADC chip prototype . . . . . . . . . . . . . . . . . . . . . . . . . 168
Fig. 8.26 Comparison of the SNDR and the bandwidth of the presented
work with the state-of-the-art DSM ADC designs
(BW  8MHz) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
Fig. 8.27 Comparison of the FoM and the bandwidth of the presented
work with the state-of-the-art DSM ADC designs
(BW  8MHz) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
Fig. 9.1 Block diagram of the digital calibration for the two-step
open-loop VCO-based ADC . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
Fig. 9.2 Topology comparison between the traditional MASH ADC
and the presented VCO-based MASH ADC . . . . . . . . . . . . . . . . 176
Fig. 9.3 Improved 1–1 MASH structure for VCO-based ADC . . . . . . . . . 176
Fig. 9.4 Block diagram of a purely-time-domain DR ADC . . . . . . . . . . . 176
Tables

Table 2.1 Performance summary of different ADC architectures . . . . . . .. 26


Table 5.1 Measurement summary of the CT DSM ADC with digital
calibration and shaped SC DAC . . . . . . . . . . . . . . . . . . . . . . . . . 100
Table 5.2 Measurement summary of the CT DSM ADC with
capacitive local feedback . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
Table 6.1 Measurement summary of the two-step open-loop
VCO-based ADC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
Table 7.1 Measurement summary of the 0–2 MASH
VCO-based DR ADC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
Table 8.1 Measurement summary of the fully-VCO-based 0-2 MASH
DR ADC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
Chapter 1
Introduction

1.1 Background and Motivation

Integrated circuits (IC) is a multi-disciplinary and multi-level industry, which is


composed of product definition, system design, circuit design, electronic design
automation (EDA), chip fabrication, package, and test. As a connection between sys-
tem design and chip fabrication, circuit design needs to accommodate the needs of
both ends. For communications, both communication standards, the system architec-
ture and the widely-used complementary metal-oxide-field-semiconductor (CMOS)
technology have huge impacts on the circuit design.

1.1.1 Communication Evolution

1.1.1.1 High Date Rate and High Bandwidth

When we look back on the communication history, two facts are very evident and
accompany the whole communication evolution: higher data rates and more mobile
devices. Being the most important driving sources for developments in the commu-
nication and IC industries for the past decades, it is safe to say that these trends will
continue in the future years.
For both wireline and wireless communications, a higher data rate is the primary
characteristic to define a new generation communication standard. With a higher
data rate, communications can be improved in two aspects. Firstly, more information
can be transferred at the same time, so some new services are becoming reality. For
instance, video phone can be supported in 3G communications, while only audio and
messaging are available with 2G GSM systems. Secondly, for data with the same
size, less waiting time is needed for the transfer. For example, to download a 4.7 GB
2 1 Introduction

8 Dial up
10
ISDN
Cable Modem
ADSL
ADSL2
ADSL2+
SHDSL
7
10 VDSL
UpStream (bps)

VDSL2
VDSL2+

6
10

5
10

5 6 7 8
10 10 10 10

DownStream (bps)

Fig. 1.1 Data rate comparison between different wireline internet accesses

DVD from internet, it takes about 1.5 h with ADSL access versus only 6 min with a
VDSL2+ connection.
In Fig. 1.1, the typical downstream and upstream data rates of different wire-
line internet accesses are plotted. The data rate is increased by about 2000× from
the old Dial-up (1990s) to the latest VDSL2+ (2006). Furthermore, the bit rate of
optical communication can reach 10 Gbps [1]. Similar stories happen with wireless
communications, as shown in Fig. 1.2. For cellular systems, at least 5 generations
of standards have been defined, and the data rate has been upgraded by more than
10000×.
Finally, it is also a trend to combine several previously separated communication
services together. For example, ATT U-verse provides IPTV (several channels), IP
telephone and broadband internet services with a single cable [2]. This combination
simplifies the connection but calls for a larger data rate for the single connection.
With the same modulation, higher communication data rates require higher analog
bandwidths, setting new challenges for analog and mixed-signal circuit designs. It is
obvious that analog-to-digital converters (ADCs) with tens of MHz bandwidths are
in great demand to realize the next generation of communication systems, including
LTE, WLAN 802.11n and VDSL2.

1.1.1.2 Low Power for Mobile Devices

The ultimate goal of communication developments is that anybody can be connected


anytime and at anyplace. To achieve this objective, mobile devices are indispensable
for the whole communication system. Wireless communication not only expands its
1.1 Background and Motivation 3

Mobility 1995 2000 2005 2010


3GPP-
GPRS LTE+
GSM 3 G+
EDGE 3G
High CDMAone (Multimedia)

Speed 2G
(Digital)
UMTS
CDMA2000

802.16e

4 G Wireless
Medium Communication
Speed
1G
(Analog)

WIMAX

High rate
Low Speed/ 2.4 GHz 5 GHz WLAN
60 GHz
Stationary WLAN WLAN UWB
WPAN
Bluetooth

10K 100K 1M 10M 100M 1G Data Rate


(bps)

Fig. 1.2 Evolution of wireless communications [3]

data rate faster than cable communication, but it also costs less time to be a widely
accepted mature technique. Previously, with the invention of the GSM, more and
more telephones shifted from wireline to wireless communications. Nowadays with
the birth of the smartphone and tablet, more and more tasks are done with these
mobile devices, including email, internet browning and TV. In Fig. 1.3, the relative
global shipments of four consumer electronics (smartphones, tablets, desktop and
laptop PCs) are plotted from 2012 to 2017 (estimated values for 2013 and 2017) [4].
It is shown that the smartphone and tablet shipments have exceeded by far those of
PCs. Considering the large gap between the smartphone/tablet and cellphone (about
1.4 billion in 2010)/laptop sales, the growth in shipments of smartphones and tablets
will continue in future years.
Compared to cellphones and laptops, there are many more mobile connections
possible with smartphones and tablets, including LTE, WLAN, NFC and HSDPA.
While more connections and more functions integrated, also a longer operation time
is expected for smartphones and tablets. With limited battery volume, low power is
therefore a more stringent design criterion for chips used in smartphones and tablets.
In recent highly digital receiver design, the ADC is the performance bottleneck
and consumes a large part of the total power. Based on the above description and
deduction, low-power ADCs with tens of MHz bandwidths are key blocks for next-
generation smartphone and tablet developments.

1.1.2 CMOS Technology

Since its invention in 1967 [5], CMOS technology has been used to implement
most of the ICs, especially Systems on Chip (SoC). The primary reason for this
4 1 Introduction

Fig. 1.3 Relative global shipments of four consumer electronics from 2012 to 2017 (estimated
values for 2013 and 2017) [4]

fact is the low cost of the CMOS technology. In 2003, one could buy 10 million
CMOS transistors with only $1. Another cause of the wide application of CMOS
is the constant technology scaling, as shown in Fig. 1.4. In 1965, Moore Gordon
predicted the exponential growth of the number of transistors on one IC [6]. With
technology scaling, the minimum gate length of the CMOS transistor is reduced by
30% and the transistor area is halved every 18 months. As a result, the chip area, the
parasitic capacitance and the power consumption reduce correspondingly. Although
the investment cost of the scaled process line increases, the price of a function block
still decreases continuously due to this scaling. At the same time, with reduced
parasitic capacitance and increased oxide capacitance density, the CMOS transistor
can work at a higher frequency, increasing the circuit bandwidth and throughput.
While bringing lots of benefits for the IC industry, CMOS technology scaling
also introduces lots of design challenges into the chip design. For digital circuits,
the noise margin is becoming smaller and the circuit is more sensitive to noise and
interferences, increasing the circuit bit error rate. Furthermore, with thinner gate
oxide, the leakage current has increased, which has exceeded the dynamic power in
some advanced technologies.
In analog and mixed-signal circuit designs, the first design difficulty comes
from the voltage. To avoid transistor breakdown, the CMOS power supply voltage
decreases continuously. However, in order to prevent severe leakage problems and to
accommodate the doping and temperature variations, the transistor threshold voltage
scaling is much slower, as shown in Fig. 1.5. This means that the available signal head-
room voltage is becoming smaller and smaller. If the same SNDR is needed for an
analog circuit, more power consumption is needed to reduce the noise and distortion
levels. Another potential problem induced by the voltage scaling is the switch. With a
smaller supply voltage, the switch on-resistance becomes larger and its linearity be-
comes worse [8]. This introduces another challenge for example switched-capacitor
1.1 Background and Motivation 5

1
10
Existing
Estimated

0
10
Gate Length (µm)

−1
10

−2
10

1970 1975 1980 1985 1990 1995 2000 2005 2010 2015 2020
Year

Fig. 1.4 CMOS technology scaling over the last 45 years and prediction for the next 4 years

12

10

8
Vdd and Vth (V)

0 1 2 3 4
10 10 10 10
Gate Length (nm)

Fig. 1.5 Decreasing supply voltage and relatively constant threshold voltage with CMOS scaling [7]

circuit design. Although in some technologies more than one threshold voltage is
provided for different kinds of circuits, additional mask layer are required in such
fabrication, leading to a higher cost.
Secondly, the output impedance of a CMOS transistor in saturation is proportional
to its gate length. If the minimum gate length is adopted to make use of the technology
scaling, then an amplifier’s intrinsic gain is decreasing with gate length scaling:
6 1 Introduction

2Id VE L 2VE L
A DC = gm ro = · = (1.1)
Vgs − Vth Id Vgs − Vth

For example, in 65 nm CMOS, if the minimum gate length used, the intrinsic gain
of a single-stage amplifier is only 16.5 dB, and it can not exceed 28dB even when a
×16 minimum gate length is used [9]. Moreover, with smaller headroom voltages,
some techniques are excluded in circuit design to improve the amplifier gain, e.g.
cascode structures and gain boosting.
Passive components also suffer from the technology scaling. The overall stacking
and individual metal heights are decreasing, so both the parasitic resistance and the
vertical parasitic capacitance increase. These result in worse capacitor and inductor
quality factors [10]. All the mentioned challenges should be overcome when design-
ing low-power ADCs with tens of MHz bandwidths in advanced CMOS technologies,
by introducing innovations at both architecture and circuit level.

1.1.3 Wireless Receiver Architectures

In wireless communications, besides the functionality, also the power consumption,


the integrity (bulk) and the programmability are important design criteria when eval-
uating a transceiver. To achieve such goals, different transceiver architectures have
been proposed, including the heterodyne receiver [11], the homodyne receiver (direct
conversion receiver) [12], the low-IF receiver [13], the direct sampling receiver [14]
and the subsampling receiver [15]. The general trends for these receiver architectures
are that they use less off-chip RF components and realize more functions on chip. For
the on-chip circuits, less RF and analog blocks but more functions are implemented
digitally. The reasons underneath this fact are that on-chip and digital solutions show
their advantages in price, robustness and bulk. As a result, more design challenges
are shifted to the data converters, and ADCs with better specifications are required
to connect the receiver analog front-end (AFE) and the digital part.
In the heterodyne receiver, there are at least two mixers and the design of each
block (including the ADC) is relaxed. However, it is a complicated system and part of
filters (for image suppression) can not be implemented in CMOS technology. The al-
ternative off-chip solution leads to problems of huge bulk, parasitics, 50Ω impedance
matching and cost, making it not suitable for future mobile communications.
If the sampling and A/D conversion are moved from the baseband (BB) stage to
the intermediate frequency (IF) stage, a heterodyne receiver evolves into a digital IF
receiver. In a digital IF receiver, the AFE is simplified and the mismatch between
the IF I/Q paths is mitigated. However, the ADC needs to work at IF and not BB,
and the choice of the IF frequency is dictated by the ADC’s maximum speed. In the
digital IF receiver, since less filtering is performed preceding the ADC, a larger ADC
dynamic range (DR) is required to avoid overloading.
1.1 Background and Motivation 7

In the homodyne receiver, since the local oscillator (LO) frequency is the same
as the RF frequency, there is no image problem anymore, leading to a very simple
RF stage. Also only a low-performance ADC is required in the homodyne receiver.
However, the homodyne structure suffers from several nonidealities and generally
it is not suitable for high-performance receivers. Because of the parasitic coupling
between the two inputs of the mixer, a DC-offset is generated at the mixer output,
saturating the following stages. In some systems, the DC-offset is constant and can
be removed by digital calibration. However, in some systems (e.g. TDMA systems),
the DC-offset is different for different channels and bursts, which makes digital
calibration difficult. Another similar design challenge in the homodyne receiver is the
flicker noise of the active components, especially for some communication standards
with small bandwidth.
The low-IF receiver combines the advantages of both the heterodyne and the
homodyne receivers: full integration and good immunity to the DC-offset and flicker
noise. Because of these advantages, the low-IF receiver can reach a better perfor-
mance than the homodyne receiver. Since the selectivity is executed by the digital
filter in the low-IF receiver, an ADC with high DR is always required. Furthermore,
if a higher-accuracy ADC is applied, then the preceding polyphase bandpass filter
can be replaced by a simple 1st-order lowpass filter, resulting into a digital low-IF
receiver.
One design trend of modern receivers is to put the ADC closer to the antenna,
making the AFE simpler. If the sampling and A/D conversion are placed at the RF
stage, the direct sampling receiver [14] is developed, which has the simplest AFE. In
this architecture, after some amplification and filtering, the RF signal is sampled and
converted into a digital stream directly. There is no analog mixer and most functions
are shifted to the digital domain. In most direct sampling receivers, only a small
partition of the whole bandwidth is occupied by the desired signal, so a bandpass
ADC is preferred to save power consumption. The mostly digital characteristic makes
the direct sampling receiver very suitable for advanced CMOS technology scaling
and an excellent choice for high-reconfigurability receivers. However, the ADC and
the preceding variable gain amplifier (VGA) operating at RF are difficult to realizes.
Moreover, with little signal conditioning circuits, the DR requirement of the ADC
is much higher compared to those in other receivers. These two requirements make
the ADC design in existing CMOS technologies a huge challenge. Finally, with the
high input signal frequency and DR requirement, the jitter performance of the clock
generation circuit is very stringent. These ADC and PLL design challenges make the
direct sampling receiver almost impossible to realize for communication standards
with high performance in existing CMOS technologies.
In the subsampling receiver, the ADC speed requirement is relaxed compared to
the direct sampling receiver. However, the requirements of the AFE and the sample-
and-hold (S/H) remain the same. Moreover, if the input frequency is m× higher than
the sampling frequency, 2 m× more noise is folded into the Nyquist band during
sampling. At the output of the S/H, the phase noise is increased by m 2 ×, which
makes this architecture very sensitive to clock jitter.
8 1 Introduction

1.2 The Research Objective of the Book

The ADC is an indispensable block in most electronic systems, including but not
limited to consumer products, telecommunications, instrumentation, automotive and
military applications. Since digital signal processing (DSP) is becoming more eco-
nomic and more powerful, the AFE is becoming simpler and the ADC is playing a
more important role. At the same time, the ADC design needs to face the challenges
from new product definition standards and the CMOS technology scaling.
As one important high-accuracy ADC structure, the delta-sigma modulator (DSM)
ADC (i.e. oversampling ADC) is widely used in many applications such as audio
and sensing. In applications with 10–100 MHz bandwidths and 10–12 bit accuracy,
previously these A/D conversions were mostly done by pipeline ADCs. With their
good figure of merit (FoM) and intrinsic anti-aliasing filtering, the continuous-time
(CT) DSM ADC is considered potentially being a substitute of the pipeline ADC.
Thus it is meaningful to investigate the design challenges of low-power CT DSM
ADCs and enable them to be applicable in the next-generation products. From 1997
to 2011, the CMOS technology scaled down by 18× (from 0.8 µm to 45 nm), however
the DSM ADC bandwidth increased by 2604× (from 48 kHz to 125 MHz), and the
corresponding FoM is improved by 43×. In the future, the technology scaling is
believed not to be so fast as before and maybe 10 nm is the end. It is clear that
innovations at both architecture level and circuit level are required for the further
bandwidth enhancement and power reduction of DSM ADCs.
This book is about innovative highly-digital ADC design, targeting at high-
bandwidth (30–50 MHz) moderate-resolution (10–12 bit) ADCs for next-generation
communications, with as low as possible power consumption. For low-power design,
one popular solution is to keep the architecture simple and to avoid high-performance
analog blocks. To maintain the desired resolution, different necessary digital cali-
brations are added. To be compatible with deep nano CMOS technology and meet
the low-power requirement of mobile electronic systems, a popular trend of ADC
design is to shift more and more functions from the analog domain and the volt-
age domain to the digital domain and the time domain, by using voltage-controlled
oscillators (VCO) and pulse-width modulators (PWM) in the ADC. By using VCO-
based time-domain quantization, it is possible to reach 1st-order noise shaping with
an open-loop structure, solving the stability issue of traditional high-order DSMs
thoroughly. For the DSM ADCs based on VCOs, nonlinearity of the VCO voltage-
frequency transfer function is the ADC performance bottleneck and should be solved
primarily.
In this book, the design and silicon implementation of 5 oversampling ADCs are
discussed. Highly-digital low-power ADCs for next-generation wireless communi-
cations will be proven. The major contributions are summarized as below, and will
be presented with more detail from Chaps. 5 to 8.
1.2 The Research Objective of the Book 9

1. A two-step open-loop VCO-based ADC architecture is proposed. This architec-


ture realizes 1st-order noise shaping with only three performance-relaxed and
mostly digital blocks. The VCO nonlinearity is mitigated by distortion cancella-
tion and input swing reduction techniques. The DAC matching requirement is also
relaxed by the implicit dynamic element matching (DEM) of the frequency-type
VCO-based quantizer. The silicon implementation of the ADC is presented. The
excellent figure-of-merit (FoM) of the demonstrator in 40 nm CMOS technology
is 42fJ/Step.
2. A 2nd-order 0–2 multi-stage shaping (MASH) DSM ADC with fully-VCO-based
integrator and quantizers in 40 nm CMOS technology is proposed and the design
details are discussed, realizing a high-order noise shaping DSM ADC without any
analog amplifier. Measurement results with a state-of-the-art FoM (52fJ/Step) are
presented and discussed.
3. The systematic design of a 2nd-order 0–2 MASH DSM ADC over a 40-MHz
bandwidth in 40 nm CMOS technology is presented. A dual-input VCO-based
quantizer topology is proposed to implement a low-power multi-input adder and
integrator, with no penalty in terms of nonlinearity. The silicon implementation
of the ADC is presented. The measured result extends the FoM of the state-of-art
high-bandwidth DSM ADCs to 35 fJ/Step.
4. A current-shaped switched-capacitor (SC) DAC, a capacitive local feedback and
a current-sharing feedforward (FF) OTA are proposed to reduce analog power
consumption in closed-loop CT DSM ADCs. The systematic design of two 4th-
order 40 MHz-bandwidth 12-bit DSMs in 90 nm CMOS technology is presented,
measurement results are also given.

1.3 The Book Organization

This book mainly covers the architecture design and circuit optimization of highly-
digital low-power DSM ADCs for next generation wireless communications. The
book is organized as follows:
Chapter 2 gives an introduction of popular ADC architectures, including flash,
two-step, pipelined, successive approximation register (SAR) and DSM ADCs. The
basic operation principle of each architecture is briefly presented; the performance
and speed limitations are also given. A brief comparison between different ADC
architectures is provided. Finally the application of ADCs in communication sys-
tems is discussed. The relationship between the ADC accuracy specification and the
receiver performance requirements is also given in detail.
Chapter 3 talks about CT DSM ADCs. First the linear analysis of a DSM is given
in short. Then different structural options for DSMs are discussed and compared,
including single-bit and multi-bit quantizations, 1st-order and higher-order noise
shapings, feedback and feedforward loop filter topologies, single-loop and MASH
structures, MASH-0 and 0-MASH topologies. The last part of this chapter focuses
10 1 Introduction

on building block nonidealities of CT DSM ADC. The nonideality modelling and


corresponding solutions are also presented.
Chapter 4 provides an overview of existing low-power VCO-based ADCs,
including closed-loop DSM ADCs with VCO-based quantizers and open-loop over-
sampling ADCs. First operation principles and characteristics of VCO-based quan-
tizations are given. Since the VCO nonlinearity is the bottleneck of highly-digital
ADC performance, the main content of this chapter is about how to solve the VCO
nonlinearity issue in VCO-based ADC design.
Chapter 5 describes the behavioral and circuit designs of two 40MHz-BW 12-bit
CT DSM ADCs with VCO-based quantizer in 90 nm CMOS technology. Also their
measurement results are presented. The first design uses a shaped SC DAC to relax
the OTA slew-rate requirement. A LUT-based calibration is also applied to eliminate
the multi-bit DAC distortion. Both system-level optimization (half-period delay) and
circuit-level improvements (capacitive local feedback and current-sharing FF OTA)
are applied in the second design to achieve lower power.
Chapter 6 proposes an open-loop two-step VCO-based ADC architecture. It re-
alizes 1st-order noise shaping with only three highly digital blocks, mitigating the
VCO nonlinearities and relaxing the DAC matching requirement. A 40 MHz-BW 10-
bit ADC implemented in 40 CMOS achieves 59.5 dB SNDR and shows an excellent
FoM of 42fJ/Step.
To address the VCO nonlinearity and circuit power challenges, design innova-
tions are implemented at architecture and circuit levels in Chap. 7. A nonlinearity-
cancellation technique in a 0–2 MASH VCO-based DSM ADC is proposed, a dual-
input VCO-based quantizer topology is also proposed to implement a low-power
multi-input adder and integrator, with no penalty in terms of nonlinearity. Fabricated
in a 40 nm CMOS process, a proof-of-concept prototype achieves 66.8-dB SNDR
with a 40 MHz bandwidth, consuming only 4.98 mW. This result extends the FoM
of state-of-the-art high-bandwidth DSM ADCs to 35 fJ/Step.
To realize a mostly-digital DSM ADC structure with high-order noise shaping,
a VCO is proposed to replace the power-hungry analog integrator in the traditional
DSM ADC topology. In Chap. 8, we report a 0–2 MASH DSM ADC with fully-VCO-
based integrator and quantizers, without any analog integrators. Taking advantage of
the highly-digital structure and relieving the VCO nonlinearity bottleneck, a 74-dB
SFDR and a FoM of 52 fJ/Step over a 40 MHz bandwidth are achieved.
Chapter 9 summarizes and concludes the research work on power-efficient highly-
digital ADC design. Some suggestions for future work are also proposed.

References

1. I.A. Young, E. Mohammed, J.T.S. Liao, A.M. Kern, S. Palermo, B.A. Block, M.R. Reshotko,
P.L.D. Chang, Optical I/O technology for tera-scale computing solid-state circuits. IEEE J.
45(1), 235–248 (2010)
2. ATT: ATT u-verse-digital TV, high speed internet and voice (2012), http://www.att.com/shop/
u-verse.html
References 11

3. P. Crombez, Full reconfigurable analog baseband circuits for multimode radios, KU Leuven
Thesis (2009)
4. S. McGlaun, Tablet shipments to outpace total PC shipments in Q4 2013 says IDC, https://
www.slashgear.com/tablet-shipments-to-outpace-total-pc-shipments-in-q4-2013-says-idc-
12297574/
5. F. Wanlass, Low stand-by power complementary field effect circuitry, US Patent 3356858
(1967)
6. M. Gordon, Cramming more components onto integrated circuits, Electronics Magazine (1965)
7. W. Zhao, Y. Cao, New generation of predictive technology model for sub-45 nm early design
exploration. IEEE Trans. Electron Devices 53(11), 2816–2823 (2006)
8. W. Sansen, Analog design essentials (Springer, 2008)
9. K. Cornelissens, Delta-Sigma A/D converter design in nanoscale CMOS, KU Leuven Thesis
(2010)
10. ITRS, International technology roadmap for semiconductors (2012), http://www.itrs.net/Links/
2012ITRS/Home2012.htm,2012
11. J.L. Hogan, Developments of the heterodyne receiver. Proc. Inst. Radio Eng. 3(3), 249–259
(1915)
12. A.A. Abidi, Direct-conversion radio transceivers for digital communications. IEEE J. Solid-
State Circ. 30(12), 1399–1410 (1995)
13. J. Crols, M.S.J. Steyaert, single-chip 900 MHz CMOS receiver front-end with a high perfor-
mance low-IF topology. IEEE J. Solid-State Circuits 30(12), 1483–1492 (1995)
14. T. Chalvatzis, T.O. Dickson, S.P. Voinigescu, 2-GHz direct-sampling delta-sigma tunable re-
ceiver with 40-GHz sampling clock and on-chip PLL, in 2007 IEEE Symposium on VLSI
Circuits (2007), pp. 54–55
15. B. Grayver, E. Daneshrad, VLSI implementation of a 100-µw multirate FSK receiver. IEEE J.
Solid-State Circuits, 36(11), 1821–1828 (2001)
Chapter 2
A/D Converters and Applications

2.1 Introduction

Considering that a signal can be continuous or discrete in time and amplitude, there
are four kinds of signals. Among these, the analog signal (continuous in time and
amplitude) and the digital signal (discrete in time and amplitude ) are widely used
in electronic circuits and systems. Because of the nature of the CMOS transistor, it
is a good switch component and a perfect choice for digital circuits (e.g. processor,
ASIC, etc.), which is suitable for complex computations. However, all signals in the
real world (e.g. sound, temperature, pressure, radio, etc.) are analog. As a result,
data converters, i.e. ADC and digital-to-analog converters (DAC), are indispensable
blocks for almost all electronic systems, behaving as bridges between the digital
processing cores and the analog world. Although in some recent communication
systems a power amplifier (PA) array is adopted to functionally combine a traditional
PA and a DAC [1], ADCs (or its simplest structure: comparators) cannot be replaced
or eliminated from systems.
In this chapter, first different specifications will be introduced as used to evaluate
an ADC. Also two FoMs are given to compare ADCs with different architectures and
specifications. For different applications and specifications, different ADC architec-
tures are needed to meet the requirements. In the next section, the basic operation
principles and the speed-performance limitations of different popular ADC architec-
tures are discussed. A short comparison between different ADCs is also presented.
Finally, we will give a brief introduction about the application of ADCs in communi-
cation systems, focusing on how to derive the block-level ADC accuracy requirement
from the system-level receiver performance.
14 2 A/D Converters and Applications

2.2 ADC Specifications

As basic blocks in electronic systems, ADCs are characterized by four parameters:


speed, accuracy, power consumption and power efficiency, three of which are inde-
pendent.

2.2.1 ADC Speed

The ADC speed defines how fast an ADC can convert its analog input signal into the
corresponding digital signal. We distinguish the following terms.

• Bandwidth (BW):
For a lowpass (or baseband) signal, the signal BW is its highest frequency; the
BW of a bandpass (or passband) signal is its highest frequency minus its lowest
frequency. The BW of an ADC is the maximum BW of the analog input beyond
which the frequency aliasing phenomenon happens.
• Effective resolution bandwidth (ERBW):
Normally the ADC accuracy is the highest for a low input signal frequency, and it
drops gradually with increasing input frequency. The ADC ERBW is the frequency
range where the ADC conversion accuracy is 3dB below its low-frequency value.
• Nyquist frequency (FN ):
The ADC FN is the minimum sampling rate required to prevent frequency aliasing.
It is twice the ADC BW.
• Sampling frequency (FS ):
FS is the frequency at which the ADC converts the analog signal into a digital
value. It is not directly related to the ADC bandwidth specifications. According to
the Nyquist theorem, the FS should be no smaller than the FN . In Nyquist ADCs,
the FS is equal to the FN ; while in oversampling ADCs, the FS is much larger than
the FN . In that case the oversampling ratio (OSR) is defined as the ratio of FS and
FN :
FS FS
OSR = = (2.1)
FN 2 · BW

2.2.2 ADC Accuracy

The ADC accuracy specifications measure how large the difference is between the
ADC analog input and the digital output. They are classified into static and dynamic
parameters, depending on the ADC analog input signal frequency. The ADC static
accuracy specifications are obtained in simulations or measurements for a direct-
current (DC) signal. They are always expressed relative to the least significant bit
(LSB).
2.2 ADC Specifications 15

• Gain error:
The gain error of an ADC describes how accurate the slope of a real transfer
function matches the slope of the ideal ADC transfer function.
• Offset:
Because of ADC nonidealities, when the input is zero, the ADC output is not
always zero. If a non-zero input voltage is needed to make the ADC output zero,
this non-zero input is called the ADC offset.
• Differential nonlinearity error (DNL):
Ideally, the difference between the analog voltages of two adjacent digital steps
should be 1 LSB, which is not the case in real circuits. The ADC DNL is the largest
difference between two adjacent analog voltage steps minus 1 LSB step.
• Integral nonlinearity error (INL):
The ADC INL is the maximum analog deviation of the actual transfer function
from the ideal ADC transfer function (after correcting for the offset and gain
errors).

When the ADC analog input is an alternating-current (AC) signal, its accuracy is
characterized by the dynamic specifications. All these specifications are calculated
in the frequency domain and are expressed in dB (except for ENOB).

• Signal-to-noise ratio (SNR):


The ADC SNR is the ratio between the signal power to the noise power integrated
overall the whole signal bandwidth.
• Signal-to-noise-distortion ratio (SNDR):
SNDR is the ratio between the signal power to the power of the integrated noise
and the harmonic distortion components within the signal bandwidth.
• Effective number of bits (ENOB)
A real ADC can be replaced by an ideal ADC with only quantization noise, whose
SNR is the same as the SNDR of the real ADC. The number of bits of the ideal
ADC is the ENOB of the evaluated ADC, it is a function of its SNDR:

SNDR − 1.76
ENOB = (2.2)
6.02
• Total harmonic distortion (THD):
The THD is the ratio of the ADC signal power to the root-mean-square (RMS)
power sum of all harmonic distortion components within the signal bandwidth.
• Dynamic range (DR):
The ADC DR is defined by the ratio between the ADC’s largest input signal power
and the smallest input signal power. The largest input signal is the input when its
SNDR drops by 3dB from the peak SNDR, and the minimum input signal is the
same as the ADC noise floor.
• Spurious-free dynamic range (SFDR):
The SFDR is the ratio between the fundamental signal power to the next largest
spurious component within the signal bandwidth, excluding the DC offset.
16 2 A/D Converters and Applications

• 2nd/3rd-order intermodulation distortion (IM2/IM3):


In a two-tone test, if the test signal frequencies are f and f + Δf , then the 2nd-order
intermodulation distortions are located at the frequencies Δf and 2f + Δf , while
the frequencies of the 3rd -order intermodulation distortions are f − Δf , f + 2Δf ,
3f + Δf and 3f + 2Δf . The ratio between the signal power (single-tone) to the
2nd- and 3rd-order intermodulation distortion powers are defined as IM2 and IM3
respectively.

Although several characteristics are described above to evaluate the ADC accu-
racy, in practical applications different ADC characteristics make sense and only
some ADC accuracy specifications are important. Generally speaking, DR is impor-
tant in sensor interfaces and instrumentation applications; THD is critical for audio
systems; DNL is essential for video processing; and SFDR is the key accuracy spec-
ification for multi-channel radio communications.

2.2.3 ADC FoM

Different ADCs have different architectures and a different range of specifications. To


make a comprehensive and fair comparison between these ADCs, some parameters
combine the ADC speed, its accuracy and power consumption into a so-called figure-
of-merit (FoM). In ADC design and applications, two FoMs [2, 3] are widely used:

P P
FoM1 = = (2.3)
2 · BW · 2 ENOB SNDR−1.76
2 · BW · 2 6.02
 BW 
FoM2 = SNDR + 10 · log10 (2.4)
P
where the P is the ADC power consumption. These two FoMs are expressed in
pJ/Conversion (or pJ/Step) and dB respectively.

2.3 ADC Architectures

Generally speaking, a full A/D conversion can be splitted into four steps: anti-aliasing
filtering, sampling and holding, quantization and encoding. In Fig. 2.1, a generalized
ADC architecture with the basic blocks is shown. In this figure the signal waveforms
and their corresponding spectrum plots for both Nyquist sampling and oversampling
are illustrated as well.

• Anti-aliasing filter (AAF):


According to the sampling theory, the sampled signal repeats its analog signal
spectrum every multiple of the sampling frequency. The function of the AAF is to
2.3 ADC Architectures 17

Ampl Ampl Ampl Ampl Ampl


Nyquist 10.5237 11 1011

Sampling 8.4723
8 1000
7.2896 7 0111

2.4184 2 0010

0 Time 0 Time 0 Time 0 Time 0 Time

Ampl Ampl Ampl Ampl Ampl


11 11
Over 10.5237
9.3176 9
10011 10100
8.4723
Sampling 7.2896
8
7
01100
4.8695 5
2.4184 2

0 Time 0 Time 0 Time 0 Time 0 Time

CLK Vref CLK

b0~b15 C0~C3

OSR
Analog AAF Band-limited S/H Sampled Quantizer Digital Digital
Signal Encoder
Analog Signal Signal Signal Code

Ampl Ampl Ampl Ampl Ampl

Nyquist
Sampling

-BW 0 BW freq -BW 0 BW freq -2FS -FS 0 BW FS 2FS freq -2FS -FS 0 BW FS 2FS freq -2FS -FS 0 BW FS 2FS freq
-BW -BW -BW

Ampl Ampl Ampl Ampl Ampl

Over
Sampling

-BW 0 BW freq -FN -BW 0 BW FN freq -FS -FN 0 BW FN FS freq -FS -FN 0 BW FN FS freq -FS -FN 0 BW FN FS freq
-BW -BW -BW

Fig. 2.1 Generalized architecture, operation and waveforms of an ADC, both with Nyquist sam-
pling and oversampling

suppress the out-of-band noise and interference to a low enough level to prevent
the desired signal from being polluted after being sampled. The different transfer
functions of the AAF required in Nyquist sampling and oversampling are shown
in Fig. 2.1.
• Sample-and-hold (S/H)
The S/H is used to sample the analog input voltage and keep it constant during
one conversion period, discretizing the signal in time. If the Nyquist theorem is
met, the ideal sampling is a reversible process and the original analog signal can
be extracted from the sampled signal by filtering. In a flash ADC architecture, the
conversion is so fast that no S/H is typically needed. In ADC architectures with
SC circuits (e.g. pipelined and discrete-time DSM ADCs), the S/H is combined
with other circuits and no separate S/H is needed.
• Quantizer
The sampled signal is now discrete in time, but still continuous in amplitude. The
quantizer uses a limited set of preset values to approximate the infinite-accuracy
input signal in amplitude, discretizing the sampled signal in amplitude. Due to the
18 2 A/D Converters and Applications

quantization, a quantization error is added to the signal, which is not reversible. The
quantizer block and its complexity vary largely among different ADC architectures,
as shown in the following paragraphs. In the time domain, the quantization error
is determined by the input voltage and the quantization levels (and therefore the
number of bits), as shown in Fig. 2.2. If the quantizer input changes randomly from
sample to sample, and it is not overloaded, the quantization error can be considered
distributes uniformly over the interval [− Δ2 , Δ2 ] and the total quantization error
power is:
 +∞  Δ
1 +2 2 Δ2
Eq2 = e2 pdfe de = e de = (2.5)
−∞ Δ − Δ2 12

In the frequency domain, this error distributes evenly over the range [− F2S , FS
2
] and
the power spectral density is:

1 Δ2
Sq2 (f ) = Eq2 · = (2.6)
FS 12 · FS

Without overloading, the quantizer maximum output signal power is:

1 2 1  (2B − 1)Δ 2 22B Δ2


P= Vo = ≈ (2.7)
2 2 2 8
where B is the number of quantization bits, so the peak SNR is:

22B Δ2
P 3 2B
SNRp = 2 = 8
Δ2
= · 2 = 1.76 + 6.02B[dB] (2.8)
Eq 12
2

By applying a larger FS , the power spectral density of the quantization error is


lower, and the quantization error integrated over the range [−BW , BW ] is reduced
by the OSR, as presented in Fig. 2.1:

Δ2 · 2BW Δ2
2
Eq,osr = Sq2 (f ) · 2BW = = (2.9)
12 · FS 12OSR

So the peak SNR is improved by OSR×:

P 3
SNRp,osr = 2
= · 22B OSR = 1.76 + 6.02B + 10log(OSR)[dB] (2.10)
Eq,osr 2

• Encoder
The output of the quantizer is always thermometer code, which is not efficient for
transmission. The encoder block transforms the quantizer output into the desired
encoding. For oversampling ADCs, decimation is also needed, filtering out the
noise beyond the signal bandwidth and reducing the data rate to FN .
2.3 ADC Architectures 19

Y(n)

111
110
101
X(n) Y(n) 100 k
Quantizer -Xmax 011 Xmax X(n)
010
001
000

Non-overload
E(n) E(n)
k 2 Xmax
X(n) Y(n) -Xmax X(n)
2

Fig. 2.2 The linear model and input-output transfer function of the quantizer

In this section, several popular ADC architectures will be introduced and dis-
cussed briefly, including Nyquist ADCs (flash, pipelined, two-step, SAR ADCs) and
oversampling ADCs (DSM ADC). It should be noticed that Nyquist ADCs are also
clocked in an oversampling way in most applications. There are two benefits from
this arrangement. Firstly, the design of the anti-aliasing filter (AAF) preceding the
ADC is relaxed. Furthermore, the ADC noise performance is improved by 3dB for
every doubling of the clock frequency.

2.3.1 Flash ADC

The architecture of a N-bit flash ADC is illustrated in Fig. 2.3. It is composed of a


resistor ladder, a pre-amplifier array, a comparator array and the digital processing
part. The resistor ladder is used to generate a string of reference voltages with identical
step size to bias the negative inputs of the pre-amplifier. The pre-amplifiers amplify
the difference between the analog input voltage and the reference voltages, before
driving the subsequent comparators. In some high-speed flash ADCs, more than one
stage of pre-amplifier is used to speed up the conversion. The comparators make
a logic decisions (0 and 1) based on the pre-amplifier outputs. In the digital part,
NAND operations are introduced to remove code bubbles and a ROM is used for any
thermometer-to-binary encoding.
The flash ADC sampling rate is limited by the speed of the pre-amplifier and
the settling time of the comparators. It is the ADC topology with the highest signal
bandwidth. A flash ADC with up to 40 GS/s sampling frequency in 0.12 µm SiGe
technology is shown in [4]; the SNDR is 18.6 dB and the power is about 3.8 W. For the
accuracy, flash ADCs highly depend on the transistor matching, which is determined
by technology parameters and the transistor area. As a result, its chip area, the power
consumption and the input capacitance are exponentially proportional to the number
20 2 A/D Converters and Applications

Fig. 2.3 N-bit flash ADC Analog


architecture Input
VREFP 2N-1 2N-1
R/2

R
Encoder

N-bit
R Digital Output

R/2
VREFN
Resistor Pre Comparator Digital
Ladder Amplifier Processing

of bits. If the resolution is increased by 1-bit, the number of pre-amplifiers and


comparators will double, and the area of each analog input transistor also needs to be
doubled to improve the matching. So the chip area and the power consumption will
be ×4. Moreover, the large input capacitance together with the output impedance of
the preceding stage would reduce the maximum flash ADC bandwidth. As a result,
flash ADC resolution is usually below 6 bit. So far, a 550 MHz-BW flash ADC with
up to 35.8 dB SNDR has been reported [5], which dissipates about 300 mW power.
In flash ADCs, due to the high throughput, a large portion of the power consumption
comes from the digital part.
To solve the issues mentioned above, some techniques have been proposed and
several new architectures have been invented. Analog interpolation is applied to
reduce the number of pre-amplifiers, and folding is used to reduce the number of
latch comparators. These two techniques can always be applied at the same time
to increase the ADC resolution and/or to save flash ADC power consumption. For
example, the SNR of a bipolar folding ADC reaches 75 dB for a 50 MHz signal BW
with 1.25 W power consumption [6]. Limited by the folding circuit, the maximum
sampling frequency of folding ADCs so far is about 2.2 GS/s, with 37.4 dB SNDR
[7]. The FoM of this folding ADC is only 205.7fJ/Step, much better than typical
value of flash ADCs (around 1pJ/Step).
Averaging is another important innovation for flash ADCs [8], which is used to
relax the transistor matching requirement, to reduce the chip area and to increase
the ADC bandwidth. The averaging technique is based on the fact that deterministic
2.3 ADC Architectures 21

signals are added in a linear way, while random signals (including mismatch and
noise) are summed in a root-mean-square (RMS) manner.

1 S +S +···+S
2 N
Saverage
SNRaverage = 20log = 20log  N
(2.11)
Eaverage E1 2 +E2 2 +···+EN 2
N2

Si NSi
= 20log Ei = 20log (2.12)
√ Ei
N

where Saverage and Si are signal with and without averaging; Eaverage and Ei are error
after and before√averaging. Here N comparator outputs are averaged, and the SNDR
is improved by N.

2.3.2 Two-Step ADC

As mentioned in subsection 2.3.1, the chip area and the power consumption of a
flash ADC will increase ×4 if its resolution is increased by 1 bit. One solution to
this fast-increasing cost is to execute the conversion in two steps instead of single
step, resulting into the two-step ADC architecture, as presented in Fig. 2.4. Due to
the two-step conversion, its latency is larger than that of a flash ADC, so an S/H is
needed to keep the analog input voltage constant during one conversion period. After
being sampled and held, it works as follows. The input voltage is first converted by a
M-bit coarse ADC; a M-bit DAC converts the digital bits back to the analog domain.
Then the residual error is generated by subtracting the DAC output from the whole
ADC input; the residual is amplified by 2M according to the resolution of the coarse
ADC and processed further by the N-bit fine ADC. The digital outputs of the coarse
and fine ADCs are the MSB and LSB sets of the whole (M + N)-bit ADC digital
output respectively. By doing the splitting, the hardware cost of a two-step ADC
(including the extra hardware) is much smaller than that of a flash ADC with the
same resolution:
2M + 2N << 2M+N (2.13)

The speed of a two-step ADC is determined by the delay of the critical path, which
includes the S/H, the coarse ADC, the DAC, the subtractor, the amplifier and the fine
ADC. Until now, a 1 GS/s two-step ADC with 33.8 dB SNDR has been presented,
with about 55 mW power consumption [9]. Because of the preceding amplification,
the fine-ADC accuracy requirement is relaxed. However, the accuracy requirements
of the S/H, the coarse ADC, the DAC, the subtractor and the amplifier are the same
as the whole ADC specification, which limits the two-step ADC resolution to about
10 bit. Also the matching between the ranges of the coarse and fine ADCs is critical
to the overall ADC performance. One solution is assigning an “over-range” to the
22 2 A/D Converters and Applications

Analog
Input
Coarse Fine
S/H DAC X2M ADC
ADC
M-bit N-bit

Digital Logic

(M+N)-bit
Digital Output

Fig. 2.4 Block diagram of a (M + N)-bit two-step ADC

fine ADC [10]. By addressing these bottlenecks, a 58.7 dB SNDR two-step ADC
with 20 MS/s and 55 mW has been reported in [11].

2.3.3 Pipelined ADC

In a two-step ADC, if an additional S/H is inserted between the amplifier and the
fine ADC and the number of stages is extended, then the pipelined ADC structure is
generated, as shown in Fig. 2.5 (analog input buffer is not included). The analog part
of a pipelined ADC is N cascaded conversion stages; all stages except the last stage
(usually a flash ADC) are multiplying DACs (MDACs), as illustrated in the dashed
box. Each MDAC first samples the input voltage, quantizes the sampled voltage,
and then extracts the quantization error by using a DAC and a subtractor. Finally
the residual quantization error is amplified and forwarded to the succeeding MDAC
stage. All operations of the MDAC stage can be merged and executed by some SC
circuit with one sampling phase and one amplifying phase. Because of the cascade
structure, the accuracy requirement of each stage can be scaled, and full accuracy
is only required in the first stage. Although the pipelined ADC can process a new
sample in each clock cycle, the latency of a complete digital output is N clock periods,
which limits the use of pipelined ADCs in control loops.
The simplest pipelined ADC architecture converts a single bit per stage. However,
the gain error and offset of the MDAC stage influence the ADC performance directly.
One solution to this problem is to use a 1.5-bit structure to compensate for the range
mismatch [12]. Generally speaking, the bandwidth and performance of a pipelined
ADC are mainly determined by the input buffer and the front-end MDAC OTAs,
and these building blocks consume most of the power of the whole ADC. For a
single-channel pipelined ADC, the maximum sampling frequency can reach 1 GS/s,
with 69dB SNDR and 1.2 W power consumption [13]; with time interleaving, up to
20 GS/s has been achieved [14], with a SNDR of 29.5 dB and a power of 10 W.
To reach high resolutions (>12bit), digital calibration is always needed. A 130 mW
pipelined ADC with 84.9 dB SNDR for 2.5 MHz bandwidth has been presented [15].
To ease the ADC design and to optimize the power consumption, simple amplifiers
or even comparators are used in the MDAC design [16], and any resulting error can
2.3 ADC Architectures 23

Vin_i Vout_i
S/H X2Mi

ADC DAC
Mi-Bit Mi-Bit

Analog Stage N
Stage 1 Stage 2 Stage i MN-Bit
Input M1-Bit M2-Bit Mi-Bit Flash

Digital Delay Cells and Calibration

(M1+M2+ ...Mi ...+MN)-Bit


Digital Output

Fig. 2.5 Simplified architecture of pipelined ADC

be solved by digital calibration. Since there is no OTA used during the sampling
phase, it can be shared between the sampling phase and the amplifying phase (of
another MDAC stage), reducing the ADC power dissipation and chip area [17].

2.3.4 SAR ADC

The architecture of an N-bit SAR ADC is presented in Fig. 2.6. It consists of a


S/H, a comparator, an N-bit DAC, a register and some control logic. The internal
clock frequency (for the comparator, the DAC and the register) is N times the ADC
frequency. At the beginning of each ADC clock period, the register bits are reset to
zero, and the analog input voltage is sampled and held constant during the whole
period. In the first internal clock cycle, the MSB of the register bits is preset to 1 and
the DAC output is updated. Then the sampled input voltage is compared to the DAC
output and the register MSB is kept to 1 or set back to 0, depending on the comparison
result. In the next internal cycle, the MSB-1 bit is preset to 1 and the updated DAC
output is again compared to the sampled input voltage. The comparison result decides
the MSB-1 bit value. After N “preset-update-compare-determine” cycles in N internal
clock periods, the LSB is determined and the complete N-bit conversion is done.
Because only one comparator and one DAC are analog in a SAR ADC, it can reach
a very high power efficiency: a design with a FoM as low as 0.85fJ/Step have been
reported [18]. In SAR ADCs, the performance directly depends on the linearity of
the DAC, whose area always dominates the chip area. The nonidealities of the com-
24 2 A/D Converters and Applications

Fig. 2.6 Structure of SAR Analog


ADC Input S/H
SAR &
Comp Control Logic

N-bit
Digital Output
DAC
(N-bit)

parator, e.g. the offset voltage and the kick-back noise are also critical to the overall
performance. So far, an 12-bit ENOB SAR with 35 MS/s and 54.5 mW power dissi-
pation has benn published [19]. The SAR ADC’s maximum frequency is determined
by its resolution and the total delay of the comparator and the DAC. With a single
channel, SAR ADCs can reach 1.2 GS/s maximum sampling frequency with about
39.3 dB SNDR and 3.1 mW power [20]. With 80 channels time interleaving, the clock
frequency impressively reaches 56 GS/s, with 5.7bit ENOB and 2 W power in 65 nm
CMOS [21]. However, because of the additional power from the clock generation
circuit and the multiplexing, the ADC FoM is 0.69pJ/Step, much larger compared
to single-channel SAR ADCs. By using time interleaving and oversampling, the
resolution of a SAR ADC can reach 14 bit [22].

2.3.5 Delta-Sigma ADC

Until now all ADC architectures introduced above in this section are Nyquist ADCs.
In theory their sampling rate can be as low as the Nyquist frequency. Contrary to
Nyquist ADCs, the sampling frequency of a Delta-Sigma (DSM) ADC is much higher
than its Nyquist frequency. When doubling the sampling rate of a Nyquist ADC, the
SNR is improved by 3 dB, so the accuracy-speed exchange efficiency is still too low.
One solution to this problem is the noise shaping technique, which is always done by
negative feedback. The combination of oversampling and noise shaping results in the
DSM ADC, which is shown in Fig. 2.7. The DSM core is a negative feedback loop
with both analog and digital signals. The loop filter compares the analog input and the
feedback DAC’s output and amplifies/integrates the difference. The filter output goes
into the quantizer and the digital output stream is obtained. Because of the global
negative feedback and the high gain of the loop filter within the signal bandwidth,
the M-bit digital stream tracks the analog input on average. The quantization noise
is not only oversampled, but also partly pushed from the band of interest to the band
beyond. In the digital decimation circuit, the out-of-band quantization noise of the
digital stream is suppressed, and the data rate is then reduced to the Nyquist rate by
a sinc filter.
2.3 ADC Architectures 25

DSM Core Digital Process


M-bit
Analog Digital Stream
N-bit
Input Digital Output
Quantizer
Loop Filter Decimator

DAC

Fig. 2.7 Block diagram of DSM ADC

In a DSM ADC, the performance is determined by three parameters: the number


of quantization bits B, the oversampling ratio OSR and the noise shaping order N [3]:
 π 2N 
SNR = 6.02B + 1.76 + 10(2N + 1)log(OSR) − 10log (2.14)
2N + 1

Generally, high-accuracy DSM ADCs are implemented by discrete-time (DT) DSMs,


which are widely used in high-quality audio systems. The performance of DT DSMs
is limited by the settling accuracy of the integrators in the loop filter. So far, a
116 dB SNDR 1 kHz-BW ADC has been presented in [23], with 12.7 mW power. The
highest BW of a DT DSM ADC is 20 MHz, with 64dB SNDR and 27.9 mW power
consumption [24]. With relaxed settling requirements on the loop filter, continuous-
time (CT) DSM ADCs have extended the bandwidth to tens of MHz (with lower
accuracy) and found application in communications. In such ADCs, the bandwidth
is limited by the speed of the loop filter and the multi-bit quantizer. For example, a
465 MHz signal bandwidth CT DSM ADC has been reported, with 64.7 dB SNDR
and 930 mW power consumption [25]. With 24 kHz and 280 µW power, a CT DSM
ADC has reached a highest SNDR of 98.5 dB [26].
Compared to Nyquist ADCs, one advantage of DSM ADCs is the relaxed (e.g.
simple RC filter) or even eliminated anti-aliasing filter, especially for CT DSM ADCs.
DSM ADCs need one extra digital decimation filter, however, instead of an analog
filter; it is robust and easy to design, furthermore its area and power dissipation can
be reduced with CMOS technology scaling. In DSM ADCs, different choices of
each block will lead to different structures and they all show benefits and limitations.
Since the digital part can be synthesized by a digital design flow using standard
libraries from the foundry, it is not covered in this work. The book is about the
design consideration of the DSM core (closed-loop ones and open-loop ones), which
will be presented in the following chapters.
26 2 A/D Converters and Applications

2.3.6 ADC Architecture Summmary and Comparison

As a summary of this subsection, the performance limitations, the time needed for
one complete conversion and the hardware cost of the five discussed ADC topologies
are listed in Table 2.1. Here N stands for the ADC resolution and T for the conversion
period of a basic comparator. In Fig. 2.8, the achievable ENOB and bandwidth of each
ADC in today’s CMOS technology are shown. It is observed that for different ADC
architectures and for each architecture itself, there are tradeoffs between bandwidth
and accuracy, and the choice among these ADC architectures highly depends on the
application and its requirement. With CMOS technology scaling and circuit design
innovations, both the bandwidth and the ENOB range of each ADC type will become
larger and there will be more overlap between different ADC topologies. In that case,
the power efficiency, chip area, design complexity and robustness will be the selection
criteria.
To give a general impression about power efficiencies of different ADC architec-
tures, a ADC FoM plot based on ISSCC publications [27] is presented in Fig. 2.9.
Here several phenomena and trends can be observed.

• The FoMs of SAR, pipelined and DSM (both DT and CT) ADCs improve by
around 1.5× every year, but the FoM improvements of flash and two-step ADCs
are much slower.
• Besides flash ADC, all other ADC architectures can reach a FoM smaller than
100fJ/Step, which is a consequence of technology advancements and circuit
design innovations. Only pipelined and SAR ADCs can achieve a FoM better
than 10fJ/Step, and these two topologies and their time-interleaving counterparts
are popular recently. Among all topologies, SAR ADC has the lowest FoM, and
it is the only choice for a FoM lower than 1fJ/Step.

Table 2.1 Performance summary of different ADC architectures


ADC architecture Accuracy limitation Clock cycles Hardware cost
Flash Transistor intrinsic T ∝ 2N
matching, kick-back
noise
N+2
Two-step coarse and fine stages 2T ∝2 2

range matching
N
Pipelined (M stages) MDAC gain and offset M · T ∝ M · 2M
SAR Feedback DAC N ·T ∝N
matching, comparator
offset and kick-back
noise
DSM (2nd-order, Integrator settling 20.4N+1 · T −
1-bit)
2.3 ADC Architectures 27

ADC
ENOB 24
22
20
18
DT DSM
16 Intrinsic Matching Limitation
14 Pipelined
12 CT DSM
10
SAR
8 Folding
6
Two-Step
4 Flash

10 100 1K 10K 100K 1M 10M 100M 1G 10G


ADC Bandwidth (Hz)

Fig. 2.8 Accuracy-bandwidth tradeoff of ADC architectures

Fig. 2.9 FoMs of different ADC architectures in ISSCC publications [27]

• Both pipelined and CT DSM ADCs are qualified for communication applications
with less than 100 MHz bandwidth, and they have almost the same FoM. However,
considering the large power consumed by the AAF preceding a pipelined ADC,
CT DSM ADCs will replace more and more pipelined ADCs in communication
terminal applications in the future.
28 2 A/D Converters and Applications

2.4 Application of ADC in Communications

Every complete receiver can be divided into three parts: the analog front end (AFE),
the ADC and the DSP, as shown in Fig. 2.10. All circuits preceding the ADC are
grouped into the AFE part, consisting out of the RF circuits, IF circuits, baseband
(BB) circuits and mixers. Different types of AFE have been presented and discussed
in Chap. 1, and a typical receiver AFE is shown in 2.11. The function of the AFE is
to perform frequency downconversion, and the necessary filtering and amplification.
It precondition the antenna output into a stable signal which can be digitized by the
ADC. The DSP is used to do demodulation to recover the information data. The
ADC in between converts the signal from the analog domain to the digital domain
and behaves as the unidirectional signal bridge between the AFE and the DSP. Today
it is the performance bottleneck in more and more receiver designs [28].
In receiver system-level design, the receiver topology determines the optimized
ADC architecture. Two lowpass ADCs with quadrature clocks are suitable in a zero-
IF receiver. On the other hand, in a low-IF receiver, a quadrature ADC is preferred to
avoid the DC offset problem. In heterodyne receivers, a lowpass ADC is enough for
a BB conversion while a bandpass ADC is optimum in power for a IF/RF conversion.
For the ADC bandwidth requirement, it is mainly determined by the communication
channel bandwidth, the number of channels converted by the ADC and the amount
of oversampling. The accuracy demanded from an ADC is however based on more
factors and the derivation is much more complex, as will be covered in this section.

AFE ADC DSP

Fig. 2.10 A receiver is partitioned into three basic parts

LO
Preselection Channel Selection
LNA Filter VGA AAF
Filter

G1 G2 G3 G4 G5 G6
NF1 NF2 NF3 NF4 NF5 NF6

GAFE
NFAFE

Fig. 2.11 Receiver AFE is a cascaded system


2.4 Application of ADC in Communications 29

First, we treat the AFE as a whole entity and calculate its gain (G AFE ) and noise
figure (NFAFE ), using the classic Friis formulas for cascaded stages:

G AFE (dB) = G 1 + G 2 + · · · + G 5 + G 6 (2.15)


NFi
Fi = 10 10 (2.16)

F2 − 1 F3 − 1 F6 − 1
FAFE = F1 + + + ··· + (2.17)
G1 G1G2 G1G2G3G4G5

NFAFE = 10log(FAFE ) (2.18)

where G i , Fi and NFi are the gain, noise factor and noise figure of the ith stages
respectively.
In general, the ADC accuracy must be high enough so that it can handle both the
smallest and the largest AFE outputs in the receiver. Also the A/D conversion may
not cause too much accuracy degradation in the signal chain. The adopted analysis
aims to find the highest power level (Pmax ) and the lowest power level (Pmin ) at the
ADC input, and the required ADC accuracy is the difference between them:

SNDR(dB) = Pmax − Pmin (2.19)

The highest power level (Pmax ) is always found when the receiver receives the
highest signal power as determined by several factors listed below [29].

• The received maximum signal power (Rmax )


In communication standards, to avoid too much inteference to other communi-
cation systems, the power received by the antenna should not exceed the defined
maximum power (Rmax ).
• The AFE minimum gain (G AFE_min )
To accommodate to the high input power from the free space, the AFE gain is set
to its lowest value G AFE_min to avoid saturating the ADC.
• The crest factor (CF)
In ADC evaluation, the SNDR is obtained with a peak signal amplitude, while in
many communication applications the RMS value of the carrier waveform does
not reach this peak value. This carrier peak-to-rms ratio is also called the crest
factor (CF). The CF value depends on the modulation. For example, the CF of 64
QAM is 3.7 dB [28].
• DC offset (OFST )
In some receiver architectures, because of VCO leakage and mixer self-mixing,
there is a DC offset (OFST ) with the AFE output signal, which is larger than the
desired channel amplitude. To make sure that the receiver works properly, the DC
offset should not saturate the ADC.
30 2 A/D Converters and Applications

• Blocker (BLKR)
After filtering and suppression, there may still be some residual blocker signal
(BLKR) left, which is nearby the desired channel in spectrum. This blocker should
also not affect the A/D conversion of the signal channel.
• Fading Margin (FM)
Because of the complex propagation environment, some margin is needed for the
channel signal to allow some fading, which is called the fading margin (FM). A
typical FM for a GSM system is 3 dB.

If we take all these factors into consideration, then the ADC’s highest power level
(Pmax ) can be expressed as:

Pmax (dB) =Rmax + G AFE_min + CF


+ max{OFST , BLKR} + FM (2.20)

The ADC lowest power level is decided by the ADC’s input-referred noise, which
can be calculated based on the following terms.

• The receiver sensitivity (Rss )


In the AFE input, the received channel noise Nin is proportional to the channel
bandwidth BWRF :
Nin = kTBWRF (2.21)

where k is the Boltzman constant and T is the absolute temperature. There is a


minimum receiver input signal power needed to provide an adequate accuracy
for the receiver to do the demodulation correctly, which is called the receiver
sensitivity (Rss ). Like the received maximum power Rmax , the required receiver
sensitivity Rss is also defined in the communication standard.
• The required SNR (SNRreq )
The difference between the receiver sensitivity and the total channel noise (sum of
the received channel noise and the AFE thermal noise) is defined as the required
SNR SNRreq . The SNRreq value is a function of the signal modulation and the
required bit error rate (BERreq ). With a larger SNRreq , the channel capacity will
increase and the BER can be reduced. For example, in 16QAM, a BERreq of 1e −
9 results in a SNRreq of 35dB. With this relationship, the noise figure of the AFE
is decided:
NFAFE_max = Rss − kTBWRF − SNRreq (2.22)

• The AFE maximum gain (G AFE_max )


With the minimum input signal, the AFE gain is set to its maximum value
(G AFE_max ) to avoid the signal from being polluted by ADC noise.
2.4 Application of ADC in Communications 31

• The ADC noise margin (NMADC )


For the ADC, the noise of the analog input is the received channel noise amplified
by the maximum AFE gain, plus the AFE noise figure:

Nch = kTBRF + G AFE_max + NFAFE_max


= Rss + G AFE_max − SNRreq (2.23)

The ratio between the ADC input noise (channel noise) and the ADC noise (quanti-
zation noise, thermal noise and clock jitter noise) is called the ADC noise margin,
this margin must be large enough so that the ADC noise does not degrade the
receiver noise figure too much (normally less than 1 dB). A typical value of the
ADC noise margin in wideband communication systems is 17 dB.

NADC = Nch − NMADC (2.24)

ΔNF = NFAFE_ADC_max − NFAFE_max


Nch NADC
= 10log[10 10 + 10 10 ] − NCH − NFAFE_max
kTBWRF +G AFE_max +NFAFE_max
= 10log[10 10

kTBWRF +G AFE_max +NFAFE_max −NMADC


+ 10 10 ]
− kTBWRF − G AFE_max − NFAFE_max < 1dB (2.25)

Considering all the above mentioned factors, the ADC lowest power level is
given by:

Pmin (dB) = Rss + G AFE_max − SNRreq − NMADC (2.26)

With the ADC maximum and minimum input power levels, the ADC’s required
SNDR can be calculated with the following formula. The complete ADC SNDR
deduction procedure is graphically illustrated in Fig. 2.12.

SNDR(dB) = Rmax + G AFE_min + CF


+ max{OFST , BLKR} + FM
− Rss − G AFE_max
+ SNRreq + NMADC (2.27)

Here we can define the difference between the received maximum power Rmax and
the receiver sensitivity Rss as the receivable signal range (RSR), and the difference
between the AFE maximum and minimum gains as the VGA gain range (GRV GA ).
32 2 A/D Converters and Applications

Rmax+FM+GAFE_min+CF+max{OFST, BLKR} ADC highest level

DC offset &
residual blocker
Rmax+FM+GAFE_min+CF
Crest factor
Rmax+FM+GAFE_min

Rmax+FM Rmax+GAFE_min

fading margin AFE minimum gain ADC


SNDR
Rmax
Rss+GAFE_max
Required SNR kTB+GAFE_max
Receivable +NFAFE_max
AFE Noise figure
signal range
kTB+GAFE_max

ADC noise margin


AFE
Rss maximum gain
ADC lowest level

kTB

AFE ADC

Fig. 2.12 Accuracy relationship between the ADC and the AFE in a telecom receiver

RSR = Rmax − Rss (2.28)


GRV GA = G AFE_max − G AFE_min (2.29)
SNDR(dB) =RSR − GRV GA + CF
+ max{OFST , BLKR} + FM
+ SNRreq + NMADC
− 10log10(BWADC /BWRF ) (2.30)

It can be seen that if the communication channel is oversampled by the ADC, the
ADC SNDR can be relaxed. To summarize, all factors influencing the required ADC
SNDR in a wireless receiver are as follows:
• the receivable signal range (RSR)
• the VGA gain range (GRV GA )
• the crest factor (CF)
• the DC offset (OFST )
• the blocker (BLKR)
2.4 Application of ADC in Communications 33

• the fading Margin (FM)


• the required SNR (SNRreq )
• the ADC noise margin (NMADC )
• the oversampling ratio (OSR)

In certain receiver architectures, some factors mentioned above do not exist, which
can be set to 0dB during the calculation. With formula 2.30, some tradeoffs between
the AFE and ADC design are shown clearly. For example, with a larger VGA gain
range and a better suppression of blockers in the AFE, a smaller ADC SNDR is
needed and the ADC power consumption can be reduced. On the other hand, with a
larger ADC SNDR, the VGA and filter can be relaxed or even eliminated.

2.5 Conclusions

In this chapter, the ADC and its application in communication system have been
discussed briefly. First, several specifications have been presented to evaluate the
ADC in speed and accuracy. Two FoMs have also been introduced to compare differ-
ent ADCs. In practical systems, the relevant accuracy specifications of interest are
different.
For different specifications and applications, different ADC architectures are
needed, and have been described. The flash ADC is the topology achieving the
highest sampling frequency and with the lowest resolution. The two-step ADC splits
the A/D conversion into two steps and can achieve a higher resolution, but its speed is
reduced compared to the flash ADC. The SAR ADC is based on a binary search algo-
rithm, and is known as the topology with the best power efficiency. The pipelined
ADC is the unique architecture which can achieve high sampling frequency and
high resolution simultaneously, and it is the most frequently used ADC structure in
communications. Normally either high-performance OTA or digital calibration are
required in a pipelined ADC. The DSM ADC is based on oversampling and noise
shaping techniques. Because of the intrinsic AAF, more and more applications shift
from Nyquist ADCs to DSM ADCs, especially CT DSMs.
In a receiver, the design of the ADC and the AFE depends on each other. The ADC
must be able to handle the AFE output signal without causing too much accuracy
loss. The full procedure of how to determine the ADC accuracy requirement has been
illustrated, and all effecting factors have been listed. With the tradeoffs expressed
by the resulting equation, a system-level optimization between the AFE, the ADC
and even the signal modulation can be done. The next chapter will discuss about the
design of closed-loop DSM ADCs.
34 2 A/D Converters and Applications

References

1. M. Park, M.H. Perrott, R.B. Staszewski, A time-domain resolution improvement of an RF-DAC.


IEEE Trans. Circuits Syst. II Express Briefs 57(7), 517–521 (2010). July
2. R.H. Walden, Analog-to-digital converter survey and analysis. IEEE J. Sel. Areas Commun.
17(4), 539–550 (1999). Apr
3. R. Schreier, G.C. Temes, Understanding Delta-Sigma Data Converters (Wiley, New York,
2005)
4. W. Cheng, W. Ali, Moon-Jung Choi, K. Liu, T. Tat, D. Devendorf,L. Linder, R. Stevens,
A 3b 40GS/s ADC-DAC in 0.12μm SiGe, in 2004 IEEE InternationalSolid-State Circuits
Conference, Digest of Technical Papers (ISSCC) (2004), pp. 262–263
5. G. Geelen, 6b 1.1GSample/s CMOS A/D converter, in 2001 IEEE International Solid-State
Circuits Conference, Digest of Technical Papers (ISSCC) (2001), pp. 128–129
6. C. Moreland, M. Elliott, F. Murden, J. Young, M. Hensley, R. Stop, A 14b 100MSample/s
3-stage A/D converter, in 2000 IEEE International Solid-State Circuits Conference, Digest of
Technical Papers (ISSCC) (2000), pp. 34–35
7. M. Miyahara, I. Mano, M. Nakayama, K. Okada, A. Matsuzawa, 22.6 A 2.2GS, s 7b 27.4
mW time-based folding-flash ADC with resistively averaged voltage-to-time amplifiers, in
IEEE International Solid-State Circuits Conference Digest of Technical Papers (ISSCC). San
Francisco, CA (2014), pp. 388–389
8. K. Kattmann, J. Barrow, A technique for reducing differential non-linearity errors in flash A/D
converters, in Solid-State Circuits Conference, Digest of Technical Papers (1991), pp. 170–171
9. P. M. Figueiredo, P. Cardoso, A. Lopes, C. Fachada, N. Hamanishi,K. Tanabe, J. Vital, A 90 nm
CMOS 1.2 V 6b 1GS/s two-step subranging ADC, in Solid-State Circuits Conference, ISSCC
2006. Digest of Technical Papers. IEEE International (2006), pp. 2320–2329
10. S.H. Lewis, P.R. Gray, A pipelined 5-Msample/s 9-bit analog-to-digital converter. IEEE J.
Solid-State Circ. 22(6), 954–961 (1987). June
11. B. Brandt, J. Lutsky, A 75-mW 10-b 20-MSample/s CMOS subranging ADC with 59-dB
SNDR, in 1999 IEEE International Solid-State Circuits Conference, Digest of Technical Papers
(ISSCC) (1999), pp. 322-323
12. X. Wang, P.J. Hurst, S.H. Lewis, A 12-bit 20-MS/s pipelined ADC with nested digital back-
ground calibration, in Proceedings of IEEE Custom Integrated Circuits Conference (2003), pp.
409–412
13. M.A. Ahmed, Ali, H. Dinc, P. Bhoraskar, C. Dillon, S. Puckett, B. Gray, C. Speir, J. Lanford,
J. Brunsilius, P.R. Derounian, B. Jeffries, U. Mehta, M. McShea, R. Stop, A 14 bit 1 GS/s
RF sampling pipelined ADC with background calibration. IEEE J. Solid-State Circ. 49(12),
2857–2867 (2014)
14. K. Poulton, R. Neff, B. Setterberg, B. Wuppermann, T. Kopley, R. Jewett, J. Pernillo, C. Tan,
A. Montijo, A 20 GS/s 8 b ADC with a 1 MB memory in 0.18 μm CMOS, in 2003 IEEE
International, Solid-State Circuits Conference, Digest of Technical Papers, ISSCC (2003), pp.
318–319
15. Sung-Ung Kwak, Bang-Sup Song, K. Bacrania, A 15b 5MSample/s low-spurious CMOS ADC,
in 1997 IEEE International Solid-State Circuits Conference, Digest of Technical Papers. 43rd
ISSCC (1997), pp. 146–147
16. L. Brooks, Hae-Seung Lee, A 12b, 50 MS/s. Fully differential zero-crossing based pipelined
ADC. IEEE J. Solid-State Circ. 44(12), 3329–3343 (2009)
17. N. Sasidhar, Youn-Jae Kook, S. Takeuchi, K. Hamashita, K. Takasuka, P.K. Hanumolu, Un-
Ku Moon, A low power pipelined ADC using capacitor and opamp sharing technique with
a scheme to cancel the effect of signal dependent kickback. IEEE J. Solid-State Circ. 44(9)
2392–2401 (2009)
18. H.Y. Tai, Y.S. Hu, H.W. Chen, H.S. Chen, 11.2 A 0.85fJ, conversion-step 10b 200kS, s subrang-
ing SAR ADC in 40 nm CMOS, IEEE International Solid-State Circuits Conference Digest of
Technical Papers (ISSCC). San Francisco, CA (2014), pp. 196–197
References 35

19. M. Krmer, E. Janssen, K. Doris, B. Murmann, 14b 35MS, S SAR ADC achieving 75dB SNDR
and 99dB SFDR with loop-embedded input buffer in 40nm CMOS, in IEEE International
Solid-State Circuits Conference – (ISSCC) Digest of Technical Papers. San Francisco, CA
(2015), pp. 1–3
20. L. Kull1, T. Toifl, M. Schmatz, P. A. Francese, C. Menolfi, M. Braendli, M. Kossel, Thomas
Morf, T. M. Andersen, Y. Leblebici, A 3.1 mW 8b 1.2 GS/s single-channel asynchronous SAR
ADC with alternate comparators for enhanced speed in 32 nm digital SOI CMOS, 2013 IEEE
International Solid-State Circuits Conference Digest of Technical Papers, San Francisco (2013)
21. I. Dedic, 56Gs/s ADC: Enabling 100GbE, 2010 Conference on Optical Fiber Communication
(OFC/NFOEC), collocated National Fiber Optic Engineers Conference, San Diego, CA, (2010),
pp. 1–3
22. M. Hesener, T. Eichler, A. Hanneberg, D. Herbison, F. Kuttner, and H. Wenske, A 14b 40MS/s
redundant SAR ADC with 480MHz clock in 0.13μm CMOS, in IEEE International Solid-State
Circuits Conference, ISSCC 2007, Digest of Technical Papers, pp. 248–249 (2007)
23. M. Steiner, N. Greer, A 22.3b 1 kHz 12.7 mW switched-capacitor Δ modulator with stacked
split-steering amplifiers, IEEE International Solid-State Circuits Conference (ISSCC). San
Francisco, CA (2016), pp. 284–286
24. P. Malla, H. Lakdawala, K. Kornegay, and K. Soumyanath, A 28mW spectrum-sensing recon-
figurable 20MHz 72dB-SNR 70dB-SNDR DT ADC for 802.11n/WiMAX receivers, in Solid-
State Circuits Conference, 2008. ISSCC 2008. Digest of Technical Papers. IEEE International,
pp. 496-631, Feb. 2008
25. Y. Dong, J. Zhao, W. Yang, T. Caldwell, H. Shibata, R. Schreier, Q. Meng, J. Silva, D. Paterson,
J. Gealow, 15.5 A 930 mW 69 dB-DR 465 MHz-BW CT 1–2 MASH ADC in 28 nm CMOS,
IEEE International Solid-State Circuits Conference (ISSCC). San Francisco, CA (2016), pp.
278–279
26. S. Billa, A. Sukumaran, S. Pavan, A 280 μW 24 kHz-BW 98.5 dB-SNDR chopped single-bit
CT ΔM achieving <10 Hz 1/f noise corner without chopping artifacts, in Solid-State Circuits
Conference, Digest of Technical Papers, ISSCC, pp. 276–277 (2016)
27. B. Murmann, ADC Performance Survey 1997–2016 (2017), http://www.stanford.edu/
murmann/adcsurvey.html
28. J. Mikael Gustavsson, Jacob Wikner, and Nianxiong Tan (Springer, CMOS data converters for
communications, 2000)
29. Q. Gu, RF system design of transceivers for wireless communications (Springer, 2005)
Chapter 3
Continuous-Time Delta-Sigma Modulators

3.1 Introduction

DSM ADCs trade off speed for accuracy by using oversampling and noise shaping
techniques. Compared to discrete-time (DT) DSMs, usually continuous-time (CT)
DSMs are applied for higher-bandwidth and lower-accuracy A/D conversions. Three
system-level parameters determine the ideal DSM ADC performance: the number
of quantization bits (B), the oversampling ratio (OSR) and the noise shaping order
(N). For the quantizer, both single-bit and multi-bit have their advantages and disad-
vantages. Generally the noise shaping order should not be higher than 5th (lowpass
case) in a single-loop structure for modulator stability. The multi-stage noise shaping
(MASH) topology increases the shaping order and maintains the stability simulta-
neously. However, a noise leakage issue comes from the analog/digital mismatch. In
single-loop DSMs, feedforward, feedback and hybrid structures have been developed
to accommodate different design requirements. In circuit design, both the active-RC
and the Gm-C integrators can be used for loop filter realization. Current, voltage
and charge (switched-capacitor) type DACs provide different choices for the feed-
back DACs. DSMs with different combinations of these parameters have different
characteristics and trade off in different aspects (noise, distortion, power, bandwidth,
clock jitter, etc.). It is important to select the most suitable structure based on the
exact specification requirements and condition limitations. For each block in a CT
DSM, different nonidealities have different mechanisms and effects on the modulator
performance, such as the amplifier’s limited gain and GBW, DAC mismatch, clock
jitter, quantizer delay and so on. The effect of the nonideality also depends on the
modulator structure. To quickly predict the modulator performance and to determine
the design requirements for each building block, all important nonidealities need to
be modelled correctly in behavioral-level simulations in Matlab.
This chapter covers the design of CT DSM ADCs. Firstly, the DSM basic operation
principle is presented. Then, comparisons between different DSM topologies are
given to help choose the optimized structure. Finally, the mechanisms and modelling
38 3 Continuous-Time Delta-Sigma Modulators

of important building block nonidealities in CT DSM are presented; also potential


solutions for these nonidealities are discussed. Based on these contents, a complete
CT DSM design flow can be summarized.

3.2 DSM Basics: Oversampling and Noise-Shaping

Being different from all other Nyquist ADC architectures, a DSM ADC has a sam-
pling frequency much higher than its Nyquist frequency, so it is also named as over-
sampling ADC. A whole DSM ADC structure is presented in Fig. 3.1; it is composed
of an AAF filter, a DSM core and a decimation filter.
The AAF filter is applied for out-of-band noise and distortion suppressing to
avoid aliasing phenomena during sampling. Since in DSM ADCs the AAF filter
transition band is large, the filter design is relaxed. In CT DSM ADCs, with the
intrinsic AAF of the CT loop filter, only a simple RC filter is required as explicit
AAF, saving considerable power consumption. The DSM core is the interface circuit
which converts the analog signal into digital bits at the sampling frequency and
moves part of the in-band quantization noise beyond the signal bandwidth. In the
DSM core output, the digital bit rate is much higher than the Nyquist frequency and
the out-of-band noise is high. A digital decimation filter (always implemented as a
sinc filter) is needed after the A/D conversion to filter out the out-of-band noise and
reduce the data rate back to the Nyquist frequency.
As mentioned previously, doubling the ADC sampling frequency will improve the
SNR by 3dB. To improve the speed-accuracy tradeoff, the noise-shaping technique is
introduced and combined together with the oversampling technique, resulting into a
DSM [1]. A generalized DSM is illustrated in the top of Fig. 3.2, which is composed
of a loop filter, a quantizer and a feedback DAC. Here the loop filter is a linear block
and the feedback DAC can be considered as a path with unit gain. To analyse the
DSM in the frequency domain, the quantizer can also be modelled as a linear block
as long as its input is random and it is not overloaded [2]. In this model, if we assume
that the loop filter is a N-stage cascaded integrator and the gain of the quantizer k is
unity, then the DSM output Y can be expressed as a function of the input X and the
quantization noise E:
H(s)k 1
Y= X+ E = STF(s) · X + NTF(s) · E (3.1)
1 + H(s)k 1 + H(s)k

And the signal transfer function (STF) and the noise transfer function (NTF) are
respectively:

fsN sN sN
STF(s) = ≈ 1; NTF(s) = N ≈ N (3.2)
sN + fsN s + fsN fs
3.2 DSM Basics: Oversampling and Noise-Shaping 39

Ampl
Ampl

0 Time -FS -FS/2 -BW BW FS/2 FS freq

Ampl Ampl

AAF
0 -FS 0 BW
Time -FS/2 -BW FS/2 FS freq

D Ampl
Ampl A

CT B
DSM
0 -FS 0 BW
Time -FS/2 -BW FS/2 FS freq

Ampl
Ampl

N
0 -FS 0 BW
Time -FS/2 -BW FS/2 FS freq

OSR Ampl
Ampl

Deci. N
Filter
0 -FS 0 BW
Time -FS/2 -BW FS/2 FS freq

Fig. 3.1 Structure of a whole CT DSM ADC. The corresponding illustrative signals and their
spectral plots illustrate the structure’s operation

With a much larger sampling frequency fs than the in-band signal frequency, the
STF is close to 1. If the quantizer contains B bits, and the quantization level is Δ,
then the output signal power is:

1 2 1 (2B − 1)Δ 2 22B Δ2


P= Vo = ( ) ≈ (3.3)
2 2 2 8
Considering the oversampling and noise-shaping effects, the integrated output
quantization noise can be calculated as:
40 3 Continuous-Time Delta-Sigma Modulators

Fig. 3.2 A generalized


X Y
DSM structure (top) and its
Quantizer
linear model (bottom) for Loop Filter
performance analysis

DAC

k E
X H(s) Y

 BW
2
Eq,dsm = |NTF(f )|2 Sq2 (f ) · df
−BW
  
BW j2π f 2N Δ2
=  
 f  12f df
−BW s s
π 2N 1 Δ2
= (3.4)
2N + 1 OSR2N+1 12

So the peak SNR is:


 
P 3π OSR 2N+1
SNRp,dsm = 2 = 2 (2N + 1)
2B
Eq,dsm 2 π
 
2N + 1
= 6.02B + 1.76 + 10log + 10(2N + 1)logOSR [dB] (3.5)
π 2N

It is clearly shown in the Eq. 3.5 that the performance of a DSM depends on the
number of quantization bits B, the noise-shaping order N and the oversampling ratio
OSR. In theory, if the number of quantization bits is increased by 1 bit, then the
DSM peak SNR is improved by 6dB. Considering the complexity of the digital part,
generally the number of quantization bits is not larger than 5bit. With a doubled OSR,
the DSM SNR is improved by 3(2N + 1)dB. However, the sampling frequency is
limited by the maximum speed of each block. For the noise-shaping order N, 5th or
higher values would impair the modulator’s stability, so MASH structures are pro-
posed. With all these limitations and design considerations, different combinations
can be generated, resulting in different DSM structures, which will be presented and
discussed in the next section.
3.3 DSM Structures 41

Fig. 3.3 Structure of a DT Q


Loop Filter
DSM (top) and its CT UD(z)
counterpart (bottom) X(s) HD(Z) Y(z)
XD(z) ED(z) Quantizer

DT DSM
ZD(z)
DAC

Loop Filter Q
UC(s) UC(z)
X(s) HC(S) Y(z)
EC(s) Quantizer

ZC(s)
CT DSM
DAC

3.3 DSM Structures

3.3.1 Discrete-Time and Continuous-Time DSMs

DSMs are classified into two basic types: discrete-time (DT) DSMs and continuous-
time (CT) DSMs. Their basic structures are shown in Fig. 3.3. Here two differences
can easily be observed.
• Different sampling positions: in the DT DSM, the sampling is preceding the whole
DSM feedback loop. While in the CT counterpart, the sampling is behind the loop
filter, which is inside the feedback loop.
• The loop filter and DAC implementations are different. Loop filters in DT DSMs
are based on switched-capacitor (SC) integrators, and CT DSM filters are con-
structed by active-RC or Gm-C integrators. Commonly used DT and CT integra-
tors are presented in Fig. 3.4. DACs in DT DSMs are also SC circuits, while there
are more choices for CT DSM DACs, including current-steering (NRZ and RZ)
DACs, SC DACs and switched-resistor DACs, as shown in Fig. 3.5.
Due to their different structures, DT and CT DSMs have different design chal-
lenges, bandwidths and performances [3]. The most critical block in a DT DSM
is the front-end S/H, as its nonidealities are not shaped by any filter. As the hold
capacitor value is determined by the kT /C noise, a small value and a high linearity
are required from the switch on-resistance. Since bipolar technology is not a good
choice for such a switch design, DT DSMs became popular only after the introduc-
tion of CMOS technology. With CMOS technology scaling, the transmission gate
on-resistance becoming larger and larger, it is more and more difficult to design
high-accuracy DT DSMs with simple switches. To tackle this challenge, both boot-
strapping switches [4] and switched-OpAmp [5] techniques have been proposed. In a
42 3 Continuous-Time Delta-Sigma Modulators

Cint Cint

Cs Ø2
Vin Rin

Ø1 Vin
Vout Vout
Ø2 Ø1

Cint Active-RC
Non-inverting
Cs Ø2 Vout
Vin Vin

Ø2
Vout gm Cint
Ø1 Ø1

Inverting Gm-C

Discrete-time integrator Continuous-time integrator

Fig. 3.4 Popular circuit implementations of DT (left) and CT (right) integrators used in the DSM
loop filter
Current

Current
Current

NRZ DAC RZ DAC SC DAC

nT (n+0.5)T (n+1)T nT (n+0.5)T (n+1)T nT (n+0.5)T (n+1)T

Fig. 3.5 Different DAC current profiles used in CT DSMs: NRZ (left), RZ (middle) and SC (right)

CT DSM, being shaped by the preceding analog loop filter, the quantizer switch non-
idealities almost do not contribute to the modulator performance degradation. For a
DAC cell, its input has only two states (1 and 0), and there is no switch on-resistance
nonlinearity problem.
In DT DSMs, the achievable SNR depends on the SC circuit settling accuracy,
which requires OTAs with larger SR and GBW than CT ones. Generally speaking, an
OTA GBW of at least 5× the sampling frequency is needed in DT DSMs, while the
ratio in CT DSMs is only 1 to 3×. With the larger ratio between the OTA GBW and
the sampling frequency, the analog power consumption in DT DSMs is higher and the
maximum allowed sampling frequency is lower. With existing CMOS technology,
DT DSMs are suitable for bandwidths <10MHz and accuracies >12bit designs.
In a DT DSM, the feedback loop can tolerate up to one period delay. As a result,
any delay in the CT DSM loop will affect its stability and changes the NTF. To
accommodate the excess loop delay (ELD) in CT DSMs, half-period and full-period
compensations are widely adopted and an extra zero-order feedback path is added
to maintain the stability and to keep the NTF the same. Half-period compensation
tolerates a smaller ELD, but with a larger delay budget for the DAC and loop filter,
relaxing the speed requirement for these blocks.
3.3 DSM Structures 43

Once the SC circuit settling in the DT DSM is sufficient, it is very robust to clock
jitter. This is different in CT DSMs, where the DAC is more sensitive to clock jitter,
especially when a RZ DAC or a single-bit DAC is applied. To address the clock
jitter noise issue, several solutions can be adopted in CT DSM design, for example
multi-bit qantization together with NRZ DAC [6], SC DAC [7] and FIR single-bit
DAC [8] techniques. With such performance limitations, CT DSMs are mostly used
for high-bandwidth (>10MHz) moderate-accuracy (10-12bit) applications.
Another disadvantage of CT DSMs compared to their DT counterparts is their
large variation in coefficients. In SC integrators, the SC part behaves as an equivalent
resistor, and the integrator gain is determined by the ratio between the integration
capacitor and the feedback capacitor (if the OTA is ideal):

1
Req = (3.6)
fs Cs

Cs z−0.5
H(z) = ± (3.7)
Cint 1 − z−1

With good capacitor matching in CMOS technology, this ratio can reach a very
high accuracy. Differently, the CT integrator transfer function is based on the absolute
values of the resistor transconductor and capacitor:

1
H(s) = − (3.8)
sRin Cint
gm
H(s) = − (3.9)
sCint

In CMOS design, the absolute accuracies of single component value are not under
good control: the variation can be as high as ±20%. So in a CT DSM design, a
capacitor bank [9] is always required to adjust the integrator gain.
One merit of DSMs compared to Nyquist ADCs is the relaxed preceding AAF.
In DT DSMs, because of the front-end sampling and the DT loop filter, the AAF is
relaxed by the oversampling. In CT DSMs, since the quantizer is behind the analog
loop filter, the AAF is not only relaxed by the oversampling, but also by the intrinsic
filtering of the loop filter. If a feedback loop filter topology is selected, usually the
separate AAF can be eliminated.
Although there are several differences between the DT and CT DSMs, in the CT
DSM design flow, it is usually preferred that first a DT NTF is designed, which is
then converted to the desired CT NTF by using the impulse invariant transformation
(IIT) [10].
44 3 Continuous-Time Delta-Sigma Modulators

3.3.2 1st-Order and Higher-Order DSMs

In the 1st-order DSM, the loop filter is only an integrator, which introduces a 90° degree
phase shift. Considering the 180° degree phase shift caused by the negative feedback,
the phase margin (PM) of a 1st-order DSM is 90° degree, which means that it is uncon-
ditionally stable if it is not overloaded. However, in the 1st-order DSM, the quantizer
input is not random and its quantization error can not be treated as white noise. The
consequence is that the 1st-order DSM is susceptible to limit cycles, which means that
its output contains some pattern noise if its analog input is a DC signal. One proof of
the limit cycles is that the SNR of the 1st-order DSM with DC input highly depends on
the input amplitude. To solve this limit cycle problem, a white noise used for dithering
can be injected into the DSM loop [11]. Another drawback of a 1st-order DSM is the
much lower performance compared with higher-order DSMs. In Fig. 3.6, the SNR of
a Nth-order (N = 1 · · · 5) 3-bit DSM with different OSRs is plotted using Eq. (3.5). It
can be concluded that a much larger OSR is needed for a 1st-order DSM to achieve
the same performance as higher-order modulators.
In some high-performance DSM design, the required OSR for 1st-order DSM
may be too high to be implemented in the technology. One solution to the too high
sampling frequency is to adopt a higher-order DSM. In a 2nd-order DSM, there is
more phase shift in the loop filter and it is only conditionally stable. If the gain of the
first integrator is larger than 1, there are both stable and unstable limit cycles, resulting
into a unstable modulator. To design a stable 2nd-order DSM, the first integrator gain
should be less than 1 and the modulator can not be overloaded. 3rd-order and higher-
order DSMs are even more prone to instability, including DC instability and dynamic
instability. According to Lee’s empirical rule [12], a single-bit DSM is stable if its
maximum out-of-band gain Hinf is not larger than 2. However, this criterion is proved
not necessary nor sufficient. To design a stable higher-order DSM, several solutions
are helpful.

Fig. 3.6 Performance of 250


Nth-order 3-bit DSM with
different OSRs N=5
200 B=3
N=4

150
SNR (dB)

N=3

N=2
100

N=1
50

0 1 2
10 10
OSR
3.3 DSM Structures 45

1. Use the proven design tool to synthesize the NTF and loop filter, for example
the Delta-Sigma toolbox [1] based on Matlab, which is developed by Richard
Schreier.
2. Scale the integrator gain and the input signal swing, which will however degrade
the DSM performance.
3. A RC filter preceding a high-order DSM can effectively reduce the dynamic
instability problem.
4. Use a clamp circuit to limit the integrator voltage swing to the normal stable
operating range.
5. Reset the integrator capacitor once the integrator output is beyond the stable range
or there are long strings of consecutive 1’s and 0’s at the output.
6. Use multi-bit quantization and/or MASH structures, which will however intro-
duce other design issues, as presented later.

3.3.3 Single-Loop and MASH DSMs

As discussed in the last subsection, higher-order DSMs suffer from stability issues
and consequently have a smaller overload level. For DSMs with 3rd- or higher-order
noise shapings, to keep the same noise-shaping order and simultaneously avoid the
potential instability problem, the multi-stage noise-shaping (MASH) topology [13]
has been proposed. A DT 2-2 MASH DSM is shown in Fig. 3.7. In this structure, there
are two stages and each stage is a single-loop 2nd -order modulator. The quantization
noise of the first stage E1 is obtained by a subtractor and used as the input of the
second stage. The outputs of both stages are respectively:

Y1 (z) = X(z)z−2 + E1 (z)(1 − z−1 )2 (3.10)

Y2 (z) = −E1 (z)z−2 + E2 (z)(1 − z−1 )2 (3.11)

The digital outputs of the two stages are processed further by the digital noise
cancellation filter (DNCF) and its output is the output of the entire MASH DSM:
 
Y (z) = Y1 + Y2 = Y1 z−2 + Y2 (1 − z−1 )2
= X(z)z−4 + E1 (z)(1 − z−1 )2 z−2
− E1 (z)z−2 (1 − z−1 )2 + E2 (z)(1 − z−1 )4
= X(z)z−4 + E2 (z)(1 − z−1 )4 (3.12)

It is shown in this equation that if the DNCF is properly designed, the quantization
noise of the first-stage DSM E1 is totally cancelled and the quantization noise of
the second stage E2 is 4th -order noise-shaped. In a real implementation, the noise
cancellation depends on the matching between the NTF of the first-stage DSM and
46 3 Continuous-Time Delta-Sigma Modulators

E1
-1 -1 Z Y1
Z Z
X Z-2
1-Z-1 1-Z-1 Y1' Y
2

DAC

-E1
E2 Y2'

Z-1 Z-1 Y2
(1-Z-1)2
1-Z-1 1-Z-1

2
Digital noise
cancellation filter
DAC

Fig. 3.7 The architecture of a discrete-time 2-2 MASH DSM, the total noise shaping is 4th-order

the DNCF (1−z−1 )2 . If the matching is not perfect, part of the first-stage quantization
noise E1 , which is only 2nd -order noise-shaped, appears at the DSM output. This is
called MASH DSM noise leakage. In a DT modulator realization, although both the
NTF and the DNCF are implemented in DT, lots of circuit nonidealities affect the
NTF, including the OTA limited slew-rate, the GBW, etc. So the requirements for
the building blocks in the MASH structure are more rigorous than in a single-loop
one.
In CT MASH DSMs, the NTF is analog while the DNCF remains digital. These
different natures make the function matching very difficult. Furthermore, compared
to the DT NTF, the CT NTF is more sensitive to circuit nonidealities, which will be
covered in the next section. As a result, it is much more difficult to achieve a good
function matching and noise cancellation. This is the reason why most of the MASH
DSMs are realized in DT. To relax the matching requirement and to ease the block
design, usually digital calibration is applied in CT MASH structures [14] to tune the
path gain in the DNCF. In CT MASH DSMs, all stages not only contribute to the
total noise shaping, but also provide AAF for the input signal [14]. Since all noise
and nonlinearity of the second stage is shaped by the first stage, the design of the
following stages can be relaxed greatly to reduce the overall power consumption and
area.

3.3.4 The Δ-0 and 0-Δ MASH Structures

Specially, an open-loop quantizer without preceding filters, or so-called a zero-order


quantizer, can be adopted in the stages of the MASH modulator. With a zero-order
stage, a two-stage MASH Δ ADC can be implemented with a Δ-0 or a 0-Δ
structure. This subsection will discuss about the performance of these structures.
3.3 DSM Structures 47

Fig. 3.8 Block diagram of a


Δ-0 MASH structure

Fig. 3.9 Block diagram of a


modified Δ-0 MASH
structure

As shown in Fig. 3.8, if the zero-order quantizer is used as the second stage, the
cascade structure is a Δ-0 MASH, also known as the Leslie-Singh architecture
[15]. Here, H1 = z−k is a delay implemented to compensate the latency of the 2nd
stage, and H2 = NTF1 . Then the whole modulator output becomes:

V = H1 · STF1 · U − H2 · NTF2 · E2 = z−k · (STF1 · U − NTF1 · E2 ) (3.13)

Compared to a single-loop modulator, E2 replaces E1 appearing at the output.


Since the 2nd stage is an open-loop structure, it has a relaxed latency requirement, and
thus can be realized with a low-complexity multi-bit pipeline structure [1], making
E2 much smaller than E1 . Therefore, the peak SQNR performance can be enhanced.
However, the matching requirement H2 = NTF1 is still difficult to be satisfied.
A modified Δ-0 MASH structure is shown in Fig. 3.9 to eliminate the substractor
preceeding the 2nd stage. Keeping H1 = z−k and choosing H2 = NTF1 /(NTF1 − 1),
the final output becomes:

z−k · STF1 z−k · NTF1


V = ·U + · E2 (3.14)
1 − NTF1 1 − NTF1

In the signal band, it is safe to assume |NTF|  1, and therefore the result is
close to Eq. (3.13). However, the drawback of this structure is that the input of the
2nd stage may contain the input U. To aviod that the 2nd stage needs to handle a
48 3 Continuous-Time Delta-Sigma Modulators

Fig. 3.10 Block diagram of


a 0-Δ MASH structure

large input signal, a feedforward topology should be implemented for the integrators
in the Δ loop of the first stage.
Another structure is the 0-Δ MASH, as presented in Fig. 3.10. A crucial insight
from the above discussion is that, if the first stage is a zero-order open-loop quantizer
with an approximately unity NTF1 , resulting in the constraint H2 = NTF1 = 1, then
the matching requirement can be relaxed. Although the 2nd stage has the capability
of processing signals above the full scale (FS) of the 1st-stage quantizer [16], the
input signal is typically limited within the FS range. Therefore, the 2nd stage only
processes the quantization noise, and power-efficient opamps with improved linearity
can be implemented here. The final output of the 0-Δ MASH becomes:

V = H1 · U + (H1 − STF2 ) · E1 − NTF2 · E2 (3.15)

However, different from Eq. (3.12), the quantization noise E2 is only shaped by NTF2 ,
which is the same as is the case in a single-loop structure. Unfortunately, the first
stage consumes power, but cannot offer additional noise shaping.

3.3.5 Single-Bit and Multi-bit DSMs

In single-bit DSMs, the quantizer is only one comparator. With its small input par-
asitic capacitor, the comparator can achieve a very high speed. This explains why
most of the high-sampling-frequency DSMs [17, 18] choose the single-bit topology.
In single-bit DSMs, compared to the analog part, the power consumption and chip
area of the digital part is only a small fraction. On the other hand, the comparator
output only depends on the polarity of its input, but not on the input signal amplitude,
as shown in the right part of Fig. 3.11. This means that the gain of a single-bit quan-
tizer is highly nonlinear, as illustrated in the left part of Fig. 3.11. Because of this
nonlinearity, the single-bit DSM is more prone to instability than multi-bit DSMs
and has a lower overload level.
Correspondingly, the DAC of a single-bit DSM also is the simplest case, as it has
only two output levels. With only two levels, no matter what values they have, they
3.3 DSM Structures 49

Y(n)
E(n)
k 2
X(n) Y(n) -Xmax
Xmax X(n)
2
k
Non-overload
E(n)

2
Xmax
-Xmax X(n)
X(n)
2

Fig. 3.11 Single-bit quantizer: the linear model, the input-output transfer function and its highly
nonlinear gain

can be connected by a straight line. In other words, a single-bit DAC is intrinsically


linear, which is its most important advantage. The intrinsic linearity makes single-bit
DSMs very suitable for high-linearity designs, e.g. digital audio. However, if a NRZ
CS DAC is applied, a single-bit DSM is more sensitive to clock jitter than multi-
bit DSMs, as the clock-jitter-induced noise is proportional to the DAC output step
variation. One solution to the single-bit DSM clock jitter issue is to insert a digital
FIR multi-tap filter between the quantizer and the DAC, in such a way that the DAC
behaves as a multi-bit DAC [8]. Another challenge of the single-bit DSM is that there
is more quantization noise generated than in multi-bit DSMs, and the amplitude of
the loop filter residual signal is larger, so a larger voltage swing and gain are needed
in single-bit DSMs.
If a multi-bit quantization is applied in a DSM design, different advantages and
disadvantages are obtained. First, with more quantization steps, less quantization
error is introduced to the feedback loop. Every extra bit of quantization results in
DSM SNR improvement of 6dB. Using a multi-bit topology, the feedback DAC
output tracks the analog input signal more accurately. This leads to a smaller residual
signal in the loop filter, relaxing the filter amplifier design. Another positive aspect of
multi-bit DSMs is that the quantizer gain is well defined and so the modulator stability
is better under control. Similar to a flash ADC design, the multi-bit quantizer can
not operate as fast as a simple comparator: the input parasitic capacitor together with
the output impedance of the proceeding stage introduces another pole and limits the
sampling frequency. The exponentially increasing hardware of a multi-bit quantizer
also costs considerable power and chip area.
For the DAC of a multi-bit DSM, especially the front-end DAC, its performance
tradeoff is very meaningful for the whole DSM design, as its nonidealities determine
the DSM SNDR and SFDR directly. First of all, a multi-bit NRZ DAC is more robust
to clock jitter than a single-bit one. In an implementation, because of the component’s
systematic and random variations, the values of different cells in a multi-bit DAC are
50 3 Continuous-Time Delta-Sigma Modulators

Fig. 3.12 The concept of

Input Code
1
DEM: 3-bit DACs without 3
(left) and with (right) DEM 6
are illustrated for comparison 4
5
2

DAC cell DAC cell

different, which is called multi-bit DAC mismatch. One serious result of the DAC
mismatch is the appearance of harmonic distortion at the DSM digital output. To
solve the multi-bit DAC mismatch and to improve the DSM linearity, three kinds of
techniques can be applied:

1. Design the DAC current source array with sufficient area and a proper layout (e.g.
the centroid, double centroid, triple centroid) to reduce the static mismatch. The
driving circuit for each DAC cell should also be designed carefully to eliminate
any dynamic mismatch. A multi-bit DSM with 12-bit intrinsic DAC matching
accuracy has been reported [9] in this way.
2. Use a dynamic element matching (DEM) technique [19], the concept of which
is shown in Fig. 3.12. In a DAC without DEM, the first DAC cell is used most
frequently, and the probabilities of cells being under used decrease with their
number. If a DEM block is placed preceding the DAC, then all cells are used
almost equally. By doing so, the DAC mismatch is 1st -order shaped. However, a
high-bit DEM generates a large delay, which may destroy the modulator stability,
thus it is not a good choice for high-bit DSMs with large sampling frequencies.
The high-bit DEM block will also consume considerable digital power and area.
3. Calibrate the DAC mismatch via a digital look-up table (LUT) [20]. In such a
calibration, in the extraction phase first the DAC cell mismatch information is
extracted and stored in a LUT. Then in the use phase, based on the quantizer
output, a calibrated value is withdrawn from the table as the final DSM output.
The benefit of digital LUT-based calibration is that it is out of the feedback
loop and its delay does not affect the DSM stability. The drawback of digital
calibration is that it can not track different variations, including temperature and
aging variations.

3.3.6 Feedforward, Feedback and Hybrid DSMs

Once the DSM NTF is synthesized, the loop filter transfer function can be deduced.
Several kinds of topologies are available for the filter realization and they all shown
their merits and drawbacks. In Fig. 3.13, generalized Nth-order DSMs with four
3.3 DSM Structures 51

KFF1

KFF2

A_IN KFF(n-1) D_OUT


∫ ∫ ∫ ∫
KFF(n)
Feedforward
DAC

A_IN D_OUT
∫ ∫ ∫ ∫

KFB(n) KFB(n-1) KFB(n-2) KFB2 KFB1 Feedback

DAC

KFF2

A_IN KFF(n-1) D_OUT


∫ ∫ ∫ ∫
KFF(n)
KFB(1) Feedforward-feedback
DAC

KFF2

KFF3
A_IN D_OUT
∫ ∫ ∫ ∫
KFF(n)
Feedforward-feedback KFB(1)
DAC

Fig. 3.13 Different structures of generalized Nth-order DSMs: feedforward (top), feedback (upper
middle) and hybrid ones (lower middle and bottom)

different kinds of filter topologies (one feedforward, one feedback and two hybrid
ones) are illustrated. They have the same NTF but different STFs.
The first architecture is the feedforward DSM [21]. Here, there is only one (front-
end) feedback DAC. The outputs of all integrators are summed with different weights
and fed to the quantizer. The benefit of the feedforward DSM is that only noise goes
through the loop filter, so the output swings of the integrator OTAs are relaxed.
However, one extra adder is needed preceding the quantizer. An active adder con-
sumes extra power and a passive one can only provide attenuation. For the STF of a
feedforward DSM, its AAF has only a 1st-order roll-off, which limits its application
potential.
52 3 Continuous-Time Delta-Sigma Modulators

Another widely used topology is the feedback one [22], where in total N DACs are
needed to feed the output signal with different gains back to the loop filter, which is its
biggest shortcoming. The output of the ith integrator minus the (i+1)th DAC output
results in the residual signal for the (i+1)th integrator. One important advantage of
the feedback-type DSM is that no extra adder is needed. Moreover, a feedback DSM
has a smaller STF in-band ripple than a feedforward DSM. Last but not least, the
AAF characteristic of the feedback DSM STF has a Nth-order roll-off, which can
greatly relax or even eliminate the preceding AAF.
To make use of the merits of both the feedforward and feedback topologies, hybrid
DSM structures have been proposed. The first hybrid architecture is similar to the
feedforward one, except that the 1st-order feedback loop is composed of the last
integrator, but not the first integrator. To do that, one more DAC is needed to feed
back the signal into the last integrator. In such topology, the extra adder is still needed,
but with one input less.
To eliminate the extra adder in the hybrid structure, one popular solution is to
combine the last integrator and the adder together. In such a case, the 2nd-order
feedback loop is not constructed by the first two integrators, but by the first and the
last integrators. In both hybrid topologies, the STF AAFs have a 2nd-order roll-off.
With different STF AAF orders and different DAC numbers, several different hybrid
topologies can be applied in design, depending on the exact design specifications
and considerations.

3.3.7 Resonator

With all the kinds of loop filter topologies as seen in the above subsection, all poles
of the loop filter are at the DC frequency, and so are the NTF zeroes. In 3rd- and
higher-order DSMs, if some of NTF zeroes are moved from the DC frequency to
higher non-zero frequencies, the integrated in-band quantization error can be several
dB lower. This SQNR improvement due to a NTF zero distribution optimization is
more obvious in high-bandwidth DSMs. In the circuit implementation, such complex
non-zero NTF zero pair is generated by a resonator in the loop filter, as presented
in Fig. 3.14. The resonator is a negative feedback loop across two integrators, with a
180° phase shift from the negative feedback and a 90° phase shift from each integrator.
The resonator in Fig. 3.14 introduces two NTF zeroes at the frequency of:

Fig. 3.14 Resonator in the k1fs k2fs


loop filter for the generation S S
of complex NTF zeroes
kb
3.3 DSM Structures 53

KFI1 KFI2 KFI(n-2) KFI(n-1) KFI(n)


D_OUT
∫ ∫ ∫ ∫
A_IN
KFB1 KFB2 KFB3 KFB(n-1) KFB(n)
Feedback

DAC

Fig. 3.15 Direct coupling paths from the DSM input to integrators and quantizer for integrator
output swing reduction

kb k1 k2 fs
f = (3.16)

3.3.8 Feedin Paths

As shown in Sect. 3.3.6, DSMs especially with a feedback loop filter face the design
challenge of having large OTA output swings. To resolve the stringent OTA swing
requirement and to reach high linearity, direct coupling paths from the modulator
analog input to the inputs of integrators and the quantizer are added in a feedback-type
DSM, as shown in Fig. 3.15. If the gains of these coupling paths are designed prop-
erly, the direct-coupled signals cancel the DAC outputs and only noise is processed
by the integrators, relaxing the OTA output swings largely. These direct-coupling
paths can also be applied in feedforward-type DSMs [21]. However, the integrator
output swing reduction and linearity improvement is not so obvious. In a real circuit
implementation, not all direct-coupling paths are necessary. For integrators closer
to the quantizer, their nonlinearities are suppressed by the preceding stages, so the
corresponding direct coupling-paths can be removed.
These direct-coupling paths do not change the modulator NTF, but they modify
the STF in two aspects. First, with these direct-coupling paths, the in-band ripple of
the STF becomes worse. Furthermore, the STF AAF characteristic degrades in a CT
DSM, especially in a feedback-type DSM. With such paths, the AAF no longer has a
Nth-order roll-off, but a kth-order roll-off (0≤k≤(N-1)), depending on the injection
positions of the paths.

3.4 CT DSM Nonidealities and Modeling

In CMOS circuit implementations, no building block of a CT DSM can be perfect.


Different circuit nonidealities have different mechanisms and different effects on
the modulator performance [23]. The nonideality effect also highly depends on the
54 3 Continuous-Time Delta-Sigma Modulators

location. Generally speaking, the front-end nonidealities are more critical than the
back-end ones, as the latter ones are suppressed by the preceding gain stage.

3.4.1 Loop Filter Nonidealities and Modeling

The loop filter is important for a DSM design, since the modulator’s noise shaping
capability is determined by the loop filter transfer function. Inside the loop filter, the
first stage is the most critical one, as its nonidealities are not suppressed by any gain,
determining the DSM performance directly. Both active-RC and Gm-C integrators
can be used in the loop filter design, and the active-RC one is typically applied in
high-accuracy (>10-bit) DSMs.

3.4.1.1 Finite OTA Gain

In a CT DSM design, without stringent settling requirement, the OTA requirement is


relaxed compared to a DT DSM. However, amplifier nonidealities still introduce lots
of errors, degrading the modulator performance or even destroying its stability. For
signal summation and resonator generation, an active-RC integrator in a CT DSM
can have as many as three inputs, as shown in Fig. 3.16. If the OTA is ideal, the
transfer function from the ith input to the output of the active-RC integrator is:
1 ki fs
H(s)i = − =− (3.17)
sRin(i) Cint s

Its pole is located at DC, thus it has an infinite gain at DC. First assume that the OTA
gain is ADC at any frequency:

Fig. 3.16 An active-RC Cint


integrator with nonideal OTA Rin1
and three input voltages in a Vin1
CT DSM Rin2
Vin2
Rin3 A(s) Vout
Vin3
3.4 CT DSM Nonidealities and Modeling 55

A(s) = ADC (3.18)

Then the integrator transfer function can be derived as:

1 ADC
H(s)i = − 
Rin(i) Cint s (ADC + 1) + 3i=1 1
Rin(i) Cint
ki fs
=   3 (3.19)
i=1 ki fs
s 1+ 1
ADC
+ ADC


Hence the integrator pole is shifted from DC to 3i=1 ki fs /(1 + ADC ) and its gain
(especially at low frequency) is reduced. Since the loop filter in-band gain determines
the suppression for the quantization error and the other circuit nonidealities, the
limited OTA gain results in a noise leakage phenomenon. If a certain integrator
gain is required at the signal bandwidth, it can be concluded that the OTA gain
requirement relies on the OSR and the integrator coefficient. A larger OSR always
calls for a larger amplifier gain. The relationship between the required amplifier gain
and the integrator coefficients is however more complex and can be investigated by
behavioural simulations. A rule of thumb for the required OTA gain in a single-loop
DSM is that it should not be less than the OSR. Another effect of the OTA finite gain
is that the signal swing of the OTA input nodes is √not zero anymore.
√ For a differential
OTA input pair, the linear input range is around [- 2(Vgs − Vth ), 2(Vgs − Vth )], here
the Vgs and Vth are the transistor gate-source and threshold voltages respectively. The
too large input voltage swing together with the nonlinear OTA input stage introduce
harmonic distortion in the CT DSM output PSD.

3.4.1.2 Finite OTA GBW

Next, we consider the OTA in an active-RC integrator as a 1st-order block, with not
only a finite gain, but also a finite bandwidth:
ADC 1
A(s) = = (3.20)
1 + ωsp 1
ADC
+ GBW
s

Applying this expression to the equation of the integrator, we obtain the nonideal
transfer function:
56 3 Continuous-Time Delta-Sigma Modulators

H(s)i =
1 1
−  3  3
Rin(i) Cint 1 1 1 1 s + A1 1
GBW s + 1 + ADC + GBW
2
i=1 Rin(i) Cint DC i=1 Rin(i) Cint
1 1
=−  
Rin(i) Cint 1 s2 + 1 + 1 3 1
GBW GBW i=1 Rin(i) Cint s
3 1
i=1 Rin(i) Cint
1− 
1 GBW + 3i=1 1
Rin(i) Cint
=−
sRin(i) Cint 1 + s
GBW + 3i=1 1
Rin(i) Cint

ki fs 1 − G E
=− (3.21)
s 1 + ωsE

where G E and ωE stand for the gain error and the phase error respectively:
3
i=1 ki fs
GE =  (3.22)
GBW + 3i=1 ki fs

3
ωE = GBW + ki fs (3.23)
i=1

From the above derivation, the finite OTA GBW introduces two extra terms to
the active-RC integrator transfer function: a gain error and an extra pole. The gain
error reduces the loop filter suppression capability for quantization noise and other
nonidealities. The more severe consequence is that the extra pole generates extra
phase shift and time delay for the DAC signal (high-frequency signal) when it passes
through the integrator, changing the modulator NTF and degrading the modulator
stability. It is clear that a large OTA GBW is helpful for building a high-performance
active-RC integrator, at the cost of extra power consumption. If the OTA GBW is
fixed, for the integrator coefficient value, there is a trade-off between the gain and
phase errors. In our designs later on, the integrator gain is decided by parameter
explorations in behavioural simulations.

3.4.1.3 Parasitic Capacitors of the OTA and the Integration Capacitor

Another nonideality of an active-RC integrator is the input parasitic capacitance,


including the OTA input parasitic capacitance and the parasitic capacitance of the
integration capacitor (positive end). To investigate its effect on the integrator per-
formance, a single-input active-RC integrator with input parasitic capacitor Cin is
presented in Fig. 3.17. Here the OTA transfer function is the same as in the above
analysis. Now the transfer function of the nonideal active-RC integrator is:
3.4 CT DSM Nonidealities and Modeling 57

Fig. 3.17 An active-RC Cint


integrator with nonideal OTA
and input parasitic capacitor
Rin
Vin
A(s) Vout
Cin

kfs GBW
H(s) = −   (3.24)
s GBW + ω + s Cin + 1 + kfs
p Cint s

As expected, the OTA input parasitic capacitance Cin degrades the integrator
performance further, and the extra degradation depends on the OTA GBW and the
ratio Cin /Cint . A rule of thumb is to keep the input parasitic capacitance less than
one third of the integration capacitor. The output parasitic capacitance, including
the OTA output parasitic capacitance and the parasitic capacitance of the integration
capacitor (negative end) does not change the integrator transfer function. However,
extra power consumption is needed in the OTA to drive these capacitances.

3.4.1.4 Limited OTA Output Swing

The final important OTA nonideality is its limited output swing. If the integrator
output swing is larger than the OTA output swing, the integrator output is clamped
and distortions are generated. This nonideality is more severe when nanoscale CMOS
technology is used in the design, as the available headroom voltage is much smaller.
It is essential to note that the effects of these mentioned OTA nonidealities depend on
the system-level parameters, i.e. the OSR, the loop filter architecture and the number
of quantization bits.
In behavioral-level modeling, one method is to deduce the transfer function of the
whole integrator with a nonideal OTA, and symbolically express the error induced
by the OTA nonidealities. Another solution is to use the model from Fig. 3.18. Here
the complex nonideal integrator transfer function is modeled and more OTA non-
idealities can be added, resulting in a higher-accuracy model. The OTA input stage
nonlinearity and the output limited swing are represented by a hyperbolic tangent
function and a signal clamp, respectively. The limited gain, GBW and doublet (comes
from feedforward frequency compensation) are represented by the linear filter stages
in Matlab Simulink. In this model, the effect of the input parasitic capacitor can also
be considered conveniently, by calculating the transfer function of the two voltage
dividers with one extra component.
58 3 Continuous-Time Delta-Sigma Modulators

Rz Cint

Rin
Vin
Vout

R
RC= + C
R C

Linear Amplifier
Vin RzCintS+1 ADC S/(2*pi*zero)+1 1 Vout
tanh(x)
(Rin+Rz)CintS+1 S/(2*pi*BW)+1 S/(2*pi*pole)+1 1+ RC

RzCintS+1
(Rin+Rz)CintS+1

Fig. 3.18 An accurate and convenient nonideal model of an active-RC integrator in Matlab Simulink

3.4.1.5 RC Time Constant Variation

Besides the nonideal OTA, the input resistor and the integration capacitor also effect
the CT DSM performance, as the gain of an ideal active-RC integrator is determined
by their absolute values. In CMOS technology, the variation on passive component
absolute value can be as high as ±30%. As a result, the active-RC integrator gain
deviation can be as large as [−41%, +104%]. With a negative gain variation, the
suppression of the quantization error and other circuit nonidealities will not be so
effective and the CT DSM performance is degraded. On the other hand, when the
integrator gain varies positively, the shaping of noise and other nonidealities is better,
increasing the modulator SNDR. However, once the integrator gain exceeds some
threshold level, the CT DSM becomes unstable. In behavioral-level simulations, a
gain error stage can be added to model the RC time constant variation (see Fig. 3.18):

1 1 1
H  (s) = − ≈− (3.25)
S(R + ΔR)(C + ΔC) SRC 1 + ΔR/R + ΔC/C

To design a CT DSM with stable performance, its RC time constant needs to be


tuned, which is done by a variable capacitor. In the variable capacitor in Fig. 3.19, a
capacitor Co with fixed value is connected in parallel with a 3-bit binary-controlled
capacitor bank. The normal, minimum and maximum capacitor values are Co +
3Cu , Co and Co + 7Cu respectively. Thus the reachable tuning range and step are
[− Co3C u
, 4Cu ] and Co +3C
+3Cu Co +3Cu
Cu
u
respectively. To reduce the nonlinearity, both the zero-
cancellation resistor Ro and the switches are connected to the integrator vitual ground,
whose signal swing is smaller than that of the integrator output.
3.4 CT DSM Nonidealities and Modeling 59

Fig. 3.19 Variable capacitor R0 C0


for RC time constant tuning
in an active-RC integrator
D1 CU

PLUS D2 2*CU MINUS

D4 4*CU

D*

3.4.2 DAC Nonidealities and Modelling

In a CT DSM, the front-end DAC directly influences the modulator performance.


Thus the DAC nonidealities must be modelled and addressed in the different design
phases, including clock-jitter-induced noise, inter-symbol interference (ISI) and
multi-bit DAC mismatch.

3.4.2.1 Clock-Jitter-Induced Noise

Clock jitter is the manifestation in the time domain of phase noise. It can be classified
into two types: the pulse-delay jitter and the pulse-width jitter. The effect of the pulse-
delay jitter on the CT DSM performance can be neglected, as it can be considered as
a variable loop delay and such delay can be tolerated by the modulator. On the other
hand, the pulse-width jitter stochastically changes the amount of charge transferred
from the feedback DAC to the loop filter in every clock period. This random variation
of the amount of charge increases the noise contribution to the CT DSM output. The
CT DSM performance degradation due to the pulse-width jitter highly depends on
the DAC current waveform and the number of quantization bits.
In Fig. 3.20, the effect of clock jitter on a single-bit RZ, NRZ and SC DAC is
shown respectively. In a single-bit RZ DAC, if the variance of the clock jitter is σs2 ,
the variance of the feedback charge error can be expressed as:
60 3 Continuous-Time Delta-Sigma Modulators

σRZ
2
= 2σs2 IRZ
2
(3.26)

where IRZ is the RZ DAC output current, and the factor 2 originates from the fact
that both the rising and the falling edges are modulated by clock jitter. Assuming
that the DAC duty cycle is 50%, the amount of charge transferred by the RZ DAC in
one clock period is:
I2 T 2
2
QRZ = RZ s (3.27)
4
In a CT DSM, the clock-jitter-induced noise is oversampled, and so the SNR due
to clock jitter is:
Q2 OSR OSR
SNRRZ = RZ 2 = 2 2 (3.28)
σRZ 8fs σs

For a single-bit NRZ, there are three differences with the RZ DAC case. First,
because the duty cycle of the NRZ DAC is 100%, the NRZ DAC current INRZ is
half the RZ DAC one IRZ . For the charge transferred in one period, only one edge is
influenced by clock jitter. Finally, clock jitter only introduces a charge error when
there is a transition in the DAC output current. These three factors improve the NRZ
DAC clock jitter performance by about 4.6dB, compared to a RZ DAC.
For a single-bit CT DSM with RZ or NRZ DAC, the clock jitter requirement is
stringent. For example, in a single-bit CT DSM with 20MHz-BW, 20× OSR and
12-bit accuracy, only 0.4 and 0.67ps clock jitters can be tolerated respectively if a
RZ and a NRZ DAC are used.
To relax the clock jitter requirement in single-bit CT DSMs, the switch-capacitor
(SC) DAC is proposed. In a SC DAC, half a clock period is assigned for capacitor
charging, and half a period is used for discharging to the loop filter, with exponentially
decaying output current:
−t
Io e τ nT < t < (n + 0.5)T
ISC (t) = (3.29)
0 (n + 0.5)T < t < (n + 1)T

Here Io is the initial value of the discharge current and τ is the discharge time
constant:
Vref
Io = (3.30)
R
τ = RC (3.31)

The total charge transferred by the SC DAC in one clock period is:
 −0.5T

QSC = Vref C 1 − e RC (3.32)

At the end of the discharge phase, the DAC current value and the charge error
variance are respectively:
3.4 CT DSM Nonidealities and Modeling 61

IDAC IDAC IDAC

2T 3T 4T 2T 3T 4T 2T 3T 4T
T t T t T t

Fig. 3.20 Clock-jitter-induced noise in single-bit DACs: RZ (left), NRZ (middle) and SC (right)

−0.5T Vref −0.5T


Itail = Io e τ = e RC (3.33)
R
 2
Vref −0.5T
σSC
2
= σs2 e RC (3.34)
R

So the DSM SNR caused by clock jitter is:


 −0.5T
2
2
QSC OSR (RC) 1 − e 2
RC

SNRSC = =  −0.5T 2 OSR (3.35)


σSC
2 σs2
e RC

If the RC time constant of the SC DAC is 0.2T , then the clock jitter tolerance of
a DSM with the above specifications is about 2.5ps.
In CT DSMs with multi-bit quantization, the clock jitter performances of the RZ
and SC DACs are similar to the single-bit cases, as shown in the left and right plots
in Fig. 3.21. Since their currents return to zero every period, the ratio between the
delivered charge in one clock period and the charge error variance is the same for
the single-bit and multi-bit cases. The only difference is that in the multi-bit case the
modulator overload level is higher and thus other nonidealities are relatively lower
compared to the peak signal and the corresponding clock-jitter-induced error.
Fortunately, if a multi-bit NRZ DAC is applied, the clock jitter performance can
be improved by (2B − 1)2 × compared to the single-bit case, where B is the number
of quantization bits. The reason is that in a NRZ DAC, the current doesn’t return
to zero every period and the clock-jitter-induced noise is proportional to the current
transition step (middle in Fig. 3.21), which is inversely proportional to the number
of quantization steps. As a result, the combination of multi-bit quantization together
with a NRZ DAC is an effective solution to clock jitter, at the expense of inter-symbol
interference.
62 3 Continuous-Time Delta-Sigma Modulators

IDAC IDAC IDAC

2T 3T 4T 2T 3T 4T 2T 3T 4T
T t T t T t

Fig. 3.21 Clock-jitter-induced noise in multi-bit DACs: RZ (left), NRZ (middle) and SC (right)

IDAC IDAC

nt (n+1)t (n+2)t (n+3)t t nt (n+1)t (n+2)t (n+3)t t

Fig. 3.22 Inter-symbol inteference (ISI) of a NRZ DAC (right); the error is marked in darkgray

3.4.2.2 Inter-symbol Interference

In CT DSMs, another nonideality of NRZ DACs is the inter-symbol interference


(ISI), which is caused by the output current’s unequal rising and falling edges. With
unequal edges, the amount of charge transferred by one DAC cell not only depends
on the number of codes “1”, but also on the data sequence. In Fig. 3.22, assuming
that the DAC current rise time is larger than the fall time, on the left the DAC input
sequence is “101” and on the right it is “111”. The ratio between the transferred
charge amounts in these cases is not exactly 2/3; the error is marked in darkgray on
the right. This sequence-dependency causes distortion tones in the CT DSM output
spectrum and degrades the modulator SNDR. Furthermore, this degradation can not
be shaped by DEM. One solution to the ISI is to adopt RZ and SC DACs, since their
output current returns to zero every period, hence their transferred charge in one
period is not related to the data sequence, eliminating the ISI issue.

3.4.2.3 Multi-bit DAC Mismatch

Single-bit DACs, since there are only two output levels, are intrinsically linear. How-
ever in a real implementation of a multi-bit DSM, because of the systematic and
random variations of the component values, the transferred amounts of charge of
different feedback DAC cells are not identical. One result of the multi-bit DAC mis-
match is the harmonic distortion in the DSM output. Together with the distortion,
more noise is introduced into the signal bandwidth during the DAC sampling. The
3.4 CT DSM Nonidealities and Modeling 63

FS

A_IN D_OUT
∫ ∫ ∫

KFB(n) KFB(n-1) KFB2 KFB1 KFB0


FS

Latch
DAC

Fig. 3.23 The zero-order feedback path for ELD compensation in feedback-type DSMs

solutions for the multi-bit DAC mismatch have been briefly addressed in the last
section.

3.4.3 Quantizer Nonidealities and Modeling

3.4.3.1 Quantizer Delay

In a CT DSM, most nonidealities of the quantizer are not so critical as those of the
loop filter and feedback DAC, because they are suppressed by the preceding gain
stage. However, the quantizer generates most of the excess loop delay (ELD), and
the ELD will change the modulator NTF or even make the feedback loop unstable.
Generally speaking, higher-order CT DSMs are more sensitive to ELD than lower-
order ones. In quantizer design, the delay is determined by its time constant τ , and
a smaller delay always corresponds to a larger power consumption.
To absorb the quantizer delay, usually one or half a clock period is assigned for
the quantizer settling, which is executed by inserting a resampling FF between the
quantizer and the feedback DAC. At the same time, to keep the NTF the same, a
zero-order feedback path is added, as shown in Fig. 3.23. If the modulator is of the
feedback type and the quantizer is not current mode, one extra adder is needed.
To remove the extra analog adder, the second hybrid topology in Fig. 3.13 can be
adopted, and a digital differentiator together with an extra DAC are used to feed
back the digitally differentiated derivative signal to the last integrator, as shown in
Fig. 3.24. By doing this, the digital differentiator and the analog integrator form the
zero-order feedback path for the ELD compensation.

3.4.3.2 Multi-bit Quantizer Input Parasitic Capacitor

In high-bandwidth high-accuracy CT DSMs, multi-bit quantization is indispensable


and the sampling frequency is very high. The exponentially-increasing number of
64 3 Continuous-Time Delta-Sigma Modulators

KFF2
FS
KFF3
A_IN D_OUT
∫ ∫ ∫ ∫
KFF(n)

DAC

DAC
DAC

FS

1-Z-1

Latch
Fig. 3.24 The ELD compensation scheme without extra analog adder: the zero-order feedback
path is composed by a digital differentiator and an analog integrator

comparators results in a considerable quantizer input parasitic capacitance. This


parasitic capacitance together with the output impedance of the preceding stage
generates an extra pole. With such a pole, the order of the CT DSM is increased by
1 and extra phase shift (time delay) is introduced.
To avoid performance degradation or instability due this extra pole, the output
impedance of the preceding stage should be designed to be low enough to push the
extra pole much higher than the sampling frequency. Another solution is to use the
quantizer parasitic capacitance as part of a passive capacitive adder.

3.5 Conclusions

This chapter has provided the basics of CT DSMs, including the operation principle,
the structural options, the modelling and the potential solution for building block
nonidealities. By applying oversampling and noise shaping techniques, the DSM
ADC is a high-accuracy ADC architecture without high-accuracy components. For
the DSM ADC architecture, there are different options with different design aspects:
discrete-time or continuous-time; single-bit or multi-bit quantization; 1st-order or
higher-order noise shaping; single-loop or MASH structure; feedforward, feedback
or hybrid structure. The combinations of these choices result in numerous candidate
implementations. They have to be investigated and compared to each other to find
the optimized structure for the required specifications in any given application. The
source and the modeling of the important CT DSM building block nonidealities
has also been given in this chapter. With these the modulator performance can be
predicted and the specifications for the different building blocks can be determined
at behavioral level, shortening the time to market and cutting the design cost.
3.5 Conclusions 65

Based on these design considerations, two CT DSM ADC designs will be pre-
sented in Chap. 5. In the next chapter, the recently developed time-domain quantiza-
tion and VCO-based A/D conversion will be discussed, as a substitute for traditional
DSM ADCs.

References

1. R. Schreier, G.C. Temes, Understanding Delta-Sigma Data Converters (Wiley, New York,
2005)
2. S.R. Norsworthy, R. Schreier, G.C. Temes, Delta-Sigma Data Converters: Theory, Design, and
Simulation (IEEE Press, IEEE Circuit and Systems Society, 1997)
3. J.A. Cherry, W.M. Snelgrove, Continuous-time Delta-Sigma Modulators for High Speed A/D
Conversion: Theory (Kluwer Academic Publishers, Practice and Fundamental Performance
Limits, 1999)
4. M. Dessouky, A. Kaiser, Very low-voltage digital-audio modulator with 88-dB dynamic range
using local switch bootstrapping. IEEE J. Solid-State Circuits 36(3), 349–355 (2001). Mar
5. J. Crols, M. Steyaert, Switched-opamp: an approach to realize full CMOS switched-capacitor
circuits at very low power supply voltages. IEEE J. Solid-State Circuits 29(8), 936–942 (1994).
Aug
6. S. Yan, E. Sanchez-Sinencio, A continuous-time sigma-delta modulator with 88-dB dynamic
range and 1.1-MHz signal bandwidth. IEEE J. Solid-State Circuits 39(1), 75–86 (2004)
7. M. Ortmanns, F. Gerfers, Y. Manoli, A continuous-time modulator with reduced sensitivity to
clock jitter through SCR feedback. IEEE Trans. Circuits Syst. I: Regul. Pap. 52(5), 875–884
(2005). May
8. P. Shettigar, S. Pavan, Design techniques for wideband single-bit continuous-time modulators
with FIR feedback DACs. IEEE J. Solid-State Circuits 47(12), 2865–2879 (2012). Dec
9. G. Mitteregger, C. Ebner, S. Mechnig, T. Blon, C. Holuigue, E. Romani, A 20-mW 640-MHz
CMOS continuous-time ADC with 20-MHz signal bandwidth, 80-dB dynamic range and 12-bit
ENOB. IEEE J. Solid-State Circuits 41(12), 2641–2649 (2006). Dec
10. F.M. Gardner, A transformation for digital simulation of analog filters. IEEE Trans. Commun.
34(7), 676–680 (1986). July
11. S.R. Norsworthy, Effective dithering of Sigma-Delta modulators, in Proceedings International
Symposium on Circuits and Systems, 1992, pp. 1304–1307
12. W.L. Lee, A novel higher-order interpolative modulator topology for high resolution oversam-
pling A/D converters, M.S. thesis, MIT, Cambridge, MA, 1987
13. M. Rebeschini, N.R. van Bavel, P. Rakers, R. Greene, J. Caldwell, J.R. Haug, A 16-b 160-
kHz CMOS A/D converter using sigma-delta modulation. IEEE J. Solid-State Circuits 25(2),
431–440 (1990). Apr
14. L.J. Breems, R. Rutten, G. Wetzker, A cascaded continuous-time Δ modulator with 67-dB
dynamic range in 10-MHz bandwidth. IEEE J. Solid-State Circuits 39(12), 2152–2160 (2004).
Dec
15. T. Leslie, B. Singh, An improved sigma-delta modulator architecture, in Proceedings of the
IEEE International Symposium on Circuits and Systems (ISCAS), pp. 372–375, May 1990
16. A. Gharbiya, D.A. Johns, A 12-bit 3.125 MHz bandwidth 0-3 MASH Delta-Sigma modulator.
IEEE J. Solid-State Circuits 44(7), 2010–2018 (2009)
17. V. Srinivasan, V. Wang, P. Satarzadeh, B. Haroun, M. Corsi, A 20mW 61dB SNDR (60MHz
BW) 1b 3rd-order continuous-time Delta-Sigma modulator clocked at 6GHz in 45nm CMOS,
in 2012 IEEE International on Solid-State Circuits Conference Digest of Technical Papers
(ISSCC), pp. 158–160, Feb 2012
66 3 Continuous-Time Delta-Sigma Modulators

18. P. Shettigar, S. Pavan, A 15mW 3.6GS/s CT-ADC with 36MHz bandwidth and 83dB DR
in 90nm CMOS, in 2012 IEEE International on Solid-State Circuits Conference Digest of
Technical Papers (ISSCC), pp. 156–158, Feb 2012
19. Y. Geerts, M.S.J. Steyaert, W. Sansen, A high-performance multibit CMOS ADC. IEEE J.
Solid-State Circuits 35(12), 1829–1840 (2000). Dec
20. J.G. Kauffman, P. Witte, J. Becker, M. Ortmanns, An 8.5 mW continuous-time modulator with
25 MHz bandwidth using digital background DAC linearization to achieve 63.5 dB SNDR and
81 dB SFDR. IEEE J. Solid-State Circuits 46(12), 2869–2881 (2011)
21. J. Silva, U. Moon, J. Steensgaard, G.C. Temes, Wideband low-distortion delta-sigma ADC
topology. Electron. Lett. 37(12), 737–738 (2001). June
22. L. Yao, M. Steyaert, W. Sansen, A 1-V, 1-MS/s, 88-dB sigma-Delta modulator in 0.13-μm
digital CMOS technology, in 2005 Symposium on VLSI Circuits, 2005. Digest of Technical
Papers, June 2005, pp. 180–183
23. F. Gerfers, M. Ortmanns, Continuous-Time Sigma-Delta A/D Conversion: Fundamentals, Per-
formance Limits and Robust Implementations (Springer, Berlin Heidelberg, 2006)
Chapter 4
VCO-Based ADCs

4.1 Introduction

With CMOS technology scaling, the analog and mixed-signal circuits face more and
more design challenges and suffer a lot in accuracy. At the same time, digital circuits
benefit from technology scaling in terms of improved timing accuracy and reduced
power consumption. As a result, more and more quantization and A/D conversions
are transferred from the analog and the voltage domains to the digital and the time
domains, by using ring VCOs and pulse-width modulation (PWM). Four different
kinds of VCO-based quantizer have been developed for power-efficient conversion,
with a highly-digital structure and intrinsic 1st-order noise shaping. In a VCO-based
counting quantizer, the VCO maximum frequency is higher than half of the sampling
frequency and a counter is required. By reducing the VCO maximum frequency, the
counter can be replaced by a FF. In the frequency-type VCO-based quantizer, there is
an implicit DEM on its digital output. However, the VCO voltage-to-frequency (V-F)
nonlinearity is the performance bottleneck. For the phase-type quantizer, the VCO
nonlinearity is not a limitation, but there is no implicit DEM and the nonlinearity of
the multi-bit feedback DAC restricts the ADC SNDR. In closed-loop DSMs with a
frequency-type VCO-based quantizer, the modulator linearity can be improved by
analog loop filter suppression and input swing reduction. If the VCO-based quantizer
is applied in an open-loop ADC, there is only 1st-order noise shaping, and the quan-
tizer nonlinearity can be solved by digital LUT-based calibration (both foreground
and background) or by PWM precoding.
This chapter provides an overview and a discussion about the design of recently
developed VCO-based ADCs. First, the structures and the circuits of four VCO-
based quantizers are presented; their merits and shortcomings are also explored and
compared. Then three closed-loop DSM ADC structures with VCO-based quantiz-
ers are shown, focusing on how to solve the frequency-type quantizer nonlinearity.
Finally, three open-loop VCO-based ADC architectures are also introduced, together
with their linearity enhancement techniques: digital LUT-based calibration and PWM
precoding.
68 4 VCO-Based ADCs

4.2 VCO-Based Quantizers

In SoC design, the analog and mixed-signal circuits are expected to follow the CMOS
technology scaling, which is mainly optimized for digital circuits. With process
scaling, the transistor gate length and parasitic capacitance become smaller, while
the oxide capacitance density increases. As a result, the operating frequency of CMOS
transistor increases. However, analog and mixed-signal circuits face several design
challenges, which have been introduced in Chap. 1 and need to be addressed to
accommodate the developments in SoC design. On the other hand, the gate delay
and the power consumption of digital circuits are continuously decreasing, as the
timing accuracy in the time domain is improved. Based on these facts, more and
more ADC designs are executed in the time and the digital domains.
Both PWM and VCO are common blocks used for the transformation from the
voltage domain to the time domain. They have the same characteristic as a compara-
tor: the input amplitude is continuous while the output amplitude is discrete. which
implies that with proper arrangement, they can be used as time-domain quantizer to
replace the voltage-domain quantizer (comparator and flash ADC). The major dif-
ference is that their information is stored in the edge positions, but not in the voltage
levels. Until now several VCO-based quantizer structures have been proposed, some
of them using a counter and register to process the VCO output phase, including a
single-phase and multi-phase quantizer. In the frequency- and phase-type VCO-based
quantizers, a register and XOR gate are used for the information extraction.

4.2.1 Single-Phase Counting VCO-Based Quantizer

The first VCO-based quantizer [1] has been proposed in 1977 for power applications.
It comprises a single-phase VCO, a counter and a sampling register, as shown on
the left in Fig. 4.1. The oscillator frequency is controlled continuously by the analog
input voltage. With a rising edge of the reference clock, the counter is reset and
begins to accumulate the number of VCO output state transitions. After one clock
period, the counting result is sampled by the register and one full process is finished.
The quantizer digital output DOU T is proportional to the VCO frequency, and thus
to the analog input voltage AIN . It is clear that the VCO maximum frequency is
exponentially proportional to the required number of bits, making this approach not
practical.

4.2.2 Multi-phase Counting VCO-Based Quantizer

To improve the quantization accuracy and to reduce the VCO maximum frequency,
a multi-phase counting VCO-based quantizer [2] has been developed, as illustrated
on the right in Fig. 4.1. A multi-phase VCO, which is usually implemented by a ring
4.2 VCO-Based Quantizers 69

AIN
AIN

Reset
Reset Counter
Counter

Ref
Register
Ref
DOUT Register

DOUT
Fig. 4.1 Single-phase (left) and multi-phase (right) counting quantizer

VCO, is modulated by the analog input voltage. Each output phase of the VCO is
used to trigger one counter. Once the counter is reset by the clock edge, the numbers
of state transitions for all VCO phase outputs are counted and all counter outputs
are summed together. With the next clock edge, the summation output is sampled
and all counters are reset. With more than one phase output, the VCO maximum
frequency of this structure is reduced compared to that of the single-phase counting
VCO-based quantizer. However, the design of the counters and the adder are more
complex. In both the multi-phase and the single-phase counting quantizers, there is
an inherent 1st-order noise shaping for the quantization noise [3].

4.2.3 Frequency-Type VCO-Based Quantizer

If the number of VCO output transitions is no more than 1 in one clock period, the
counter for the multi-phase counting quantizer can be replaced by a flip flop (FF),
simplifying the design of the digital part. The structure and model  of a widely
 used
VCO-based quantizer (B-bit) [4, 5] is shown in Fig. 4.2. First a 2B − 1 -stage ring
VCO converts the input voltage into
 phase information, which is illustrated by the
rising/falling edges of the 2B − 1 VCO phase outputs. Then each VCO phase is
fed to a slice of frequency detector. A sense-amplifier flip flop (SA FF) is used to
synchronize each VCO output, introducing a quantization error in the phase domain.
Finally a digital differentiator (comprised of a TSPC FF and an XOR gate) processes
the sampled phase and outputs the digital frequency information. The digital output
actually is the number of VCO delay cells, that change their state once (from “1” to
“0” or from “0” to “1”) in one clock period. In this quantizer, since there is no register
assigned for the transition counting, the number of state transitions of each phase
70 4 VCO-Based ADCs

Ring VCO SA FF First_Order


Difference XOR
1
z
1
z
TSPC FF
CLK

Voltage-to Frequency
-Frequency -to-Phase
2 1
S TS 1-Z-1
KVCO
|H(f)| |H(f)| |H(f)|

Fin Freq Freq Fin Freq

Fig. 4.2 Structure, behavioral model and illustrative spectrum of the frequency-type VCO-based
quantizer

output may not be larger than 1, or else the digital frequency output is reset and the
desired information is ruined. With this limitation, the VCO maximum frequency can
not exceed half of the quantizer sampling frequency. A proper design is to set the VCO
maximum frequency to half the sampling frequency, and the free-running frequency
to one quarter of the sampling frequency. Since the VCO minimum frequency is
much smaller compared to the maximum frequency, the VCO gain is:
fmax − fmin fs
KV CO = ≈ (4.1)
Ain 2Ain

Compared to a flash ADC, the VCO-based quantizer has two extra advantages,
which make it very suitable for DSM design. First, with the VCO integration and
digital differentiation, the transfer function from the signal input to the output is
unity, while the transfer function from the quantization error source to the output
is a differentiator. Because of the VCO integration 2πKsV CO , the quantization noise is
inherently 1st-order noise-shaped. With this 1st-order noise shaping and sufficient
OSR, even a single frequency-type VCO-based quantizer can reach moderate noise
performance. Several examples will be presented later.
In a ring VCO, the state transition of a delay cell corresponds to a “1” at its
frequency detector output. Since the delay cells change their state one by one, the
frequency detectors output “1” consecutively. In other words, a DEM is intrinsically
implemented with this frequency-type VCO-based quantizer, without any extra tim-
ing penalty and power consumption. If this quantizer is applied in a closed-loop DSM,
then the implicit DEM relaxes the front-end DAC matching requirement greatly.
4.2 VCO-Based Quantizers 71

On the other hand, the relationship between the input voltage and the VCO fre-
quency is not linear. In behavioral simulations, a polynomial can be used to model
the V-F nonlinearity:

f = fo + kvco Ain + k2 A2in + k3 A3in + k4 A4in (4.2)

Such nonlinearity is not suppressed by any filtering and generates harmonic dis-
tortion in the quantizer output. Although the even-order distortion can be cancelled
by a pseudo-differential topology, the 3rd-order distortion still exists and limits the
quantizer linearity to about 5-bit. How to solve this nonlinearity is the key challenge
to apply this quantizer in DSM designs.
In the VCO-based quantizer design, the VCO phase noise must also be considered,
especially if it is used in open-loop ADCs, where there is no preceding analog loop
filter suppression. Let’s assume that the VCO phase noise is L at the offset frequency
foff , and the input signal amplitude and frequency are Ain and fin respectively, then
the peak signal-to-phase-noise-ratio of the quantizer can be expressed as [6]:
 
(KV CO Ain )2
SPNR ≈ 10log 2
(4.3)
16Lfoff fin

If the clock of the VCO-based quantizer is not clean enough, it introduces an error
in the quantization through two mechanisms: the sampling uncertainty triggered by
pulse delay jitter and the integration error caused by pulse width jitter. The sampling
error in the VCO-based quantizer is the same as in conventional ADCs:

SJNRd = −20log(2π fin σd ) (4.4)

Here σd2 is the pulse delay jitter variance and fin is the input signal frequency. If the
variance of the pulse width jitter is σw2 , then the quantizer performance is:
⎛ ⎞
2 fin
2 sinc
⎜T fs ⎟
SJNRw = 10log ⎝ s2 2⎠
(4.5)
σw ffr
1 + 2 KV CO Ain

whereffr is the VCO free-running frequency. In addition, the VCO-based quantizer


has a 2B − 1 × smaller probability of being metastable than a flash ADC. If the
delay cells are implemented by minimum gate length transistors, then the offset volt-
age can reach tens of mV. Fortunately, the quantization error due to the offset is also
1st-order shaped. Finally, the VCO-based quantizer is also statistically monotonic.
72 4 VCO-Based ADCs

Fig. 4.3 Structure of the INPUT


phase-type VCO-based
quantizer; the additional
digital frequency output is 1
used for DSM ELD z
compensation First Order
Difference

CLOCK

4.2.4 Phase-Type VCO-Based Quantizer

As mentioned in the last subsection, the VCO nonlinearity is the performance bot-
tleneck in the frequency-type VCO-based quantizer. To avoid this disadvantage, a
phase-type VCO-based quantizer [7, 8] has been proposed, as shown in Fig. 4.3. In
this quantizer, the VCO phase and not the frequency is adopted as the digital output.
To remove the common-mode information in the digital phase output, an xor oper-
ation between the VCO outputs and quadrature pulses with the VCO free-running
frequency is taken after the SA FF sampling. Now the transfer function from the
signal input to the output is an integration, while the function from the quantization
error source to the output is unity. As a result, the intrinsic 1st-order quantization
noise shaping is the same as that of the frequency-type VCO-based quantizer.
Because the signal transfer function is an integration and not unity, the phase-type
VCO-based quantizer cannot be used separately as an open-loop ADC; it can only be
applied in a closed-loop system. In a closed-loop DSM with a phase-type VCO-based
quantizer, because of the large gain provided by the VCO integration, the VCO input
voltage swing is reduced greatly, and so is its nonlinearity. Behavioral simulations
show that the harmonic distortion of this quantizer can be as low as −90 dBc [8],
which is enough for most DSM designs.
Another drawback of the phase-type VCO-based quantizer is that there is no
implicit DEM on the quantizer digital outputs. Consequently, the DAC matching is
a potential design challenge and needs to be solved by others methods.
In CT DSM design, usually a zero-order feedback path is added to compensate the
ELD. To implement a zero-order feedback path, frequency detectors are also applied
in the phase-type quantizer to provide the digitized frequency. For this secondary
output, the intrinsic DEM still remains. Because of the reduced VCO input voltage
swing, the digital frequency output is also a low-swing signal.

4.3 Closed-Loop VCO-Based DSMs

In this section, three different published closed-loop DSMs with VCO-based quan-
tizers are presented, and the design solutions for the VCO nonlinearity are discussed.
4.3 Closed-Loop VCO-Based DSMs 73

VCO_Based_Frequency_Quantizer

D_OUT
A_IN
1 First_Order
KFB2 z Difference

RZ NRZ CLK
DAC DAC

Implicited
DEM

CLK

Fig. 4.4 Architecture of a 3rd-order CT DSM with frequency-type VCO-based quantizer [4]

4.3.1 DSM with Frequency-Type VCO-Based Quantizer

A closed-loop DSM [4, 5] with frequency-type VCO-based quantizer is shown in


Fig. 4.4. It is a 3rd-order feedback structure with 5-bit time-domain quantization.
1st-order noise shaping comes from the VCO-based quantizer, and the other 2nd-
order shaping is done by the analog loop filter. The nonlinearity of the VCO is
suppressed by the preceding loop filter. In the 2nd-order loop filter, the front-end
stage is implemented by a passive RC filter and the second stage by an active-RC
integrator. Since the passive low-pass filter doesn’t provide any gain, the quantization
noise and VCO nonlinearity are suppressed by 2nd and 1st-order filters respectively.
Furthermore, the resistor of the RC filter contributes to the DSM noise power and
degrades its SNR. However, the passive front-end filters out the high-frequency signal
component from the RZ DAC, and makes the input signal of the active-RC integrator
smooth. This is beneficial for the linearity of the active integrator.
A RZ DAC is chosen as the outermost DAC based on two design considerations.
First, a RZ DAC is robust to ISI. Secondly, by using a RZ DAC, more ELD can be
tolerated by the modulator. As a result the CT DSM is still marginally stable, even
if the inner compensation DAC is disabled. On the other hand, compared to a NRZ
counterpart, the RZ DAC has double the output current and requires a larger OTA
slew rate. As mentioned before, this disadvantage can be mitigated by the front-
end passive filtering. Besides, the RZ DAC is much more sensitive to clock jitter,
which means that a very clean clock generator is required for measurements and in
application. With the implicit DEM of the frequency-type VCO-based quantizer, the
RZ DAC matching is not the performance bottleneck.
74 4 VCO-Based ADCs

KFF1
KFF2 VCO_Based_Phase_Quantizer

Kint1 Kint2 Kint3


A_IN KFF3 D_OUT
1
z
First_Order
Difference

KLFB CLK

LATCH
KFB2 HRZ Implicited
D Q
DAC DEM
KDFB
CLK_B CLK
LATCH
NRZ D Q
Explicited
DAC DEM

CLK CLK

Fig. 4.5 Topology of a 4th-order CT DSM with phase-type VCO-based quantizer [8]

4.3.2 DSM with Phase-Type VCO-Based Quantizer

In the above design, even with filter suppression, the VCO nonlinearity is still con-
siderable and the modulator SFDR is only about 75 dB. To achieve better linearity
performance, the phase-type VCO-based quantizer is applied to a 4th-order 4-bit CT
DSM with hybrid structure [7, 8] as illustrated in Fig. 4.5.
1st-order noise shaping arises from the VCO integration, and the other 3rd-order
noise shaping is from the active-RC loop filter. Because of the VCO integration, the
VCO input swing is so small that its V-F characteristic can be considered to be very
linear. Hence the VCO nonlinearity never limits the DSM performance.
Without quantizer implicit DEM, the front-end DAC linearity is a primary design
challenge, which is solved by an explicit DEM. Since only half a clock period is
assigned for the DEM timing, high-speed dynamic logic is used to implement the
DEM, with a large power consumption penalty. In the outermost NRZ DAC design,
a degeneration resistor is use to reduced the DAC output noise by a factor of (gm R)2 ,
where gm is the transconductance of the current source transistor in the DAC cell,
and R is the degeneration resistor. For the inner NRZ DAC, since its nonidealities are
shaped by the 3rd-order analog filter, its noise and matching requirements are relaxed
to save chip area. For the RZ DAC, the low input swing and implicit DEM reduce
its area further. To save power consumption, a passive resistive adder is adopted to
sum the integrator voltages and DAC currents.

4.3.3 DSM with Residual-Cancelling VCO-Based Quantizer

Based on the frequency-type VCO-based quantizer, the residual-cancelling concept


has been proposed to solve the VCO nonlinearity. In Fig. 4.6, a 2nd-order CT DSM
with a residual-cancelling VCO-based quantizer [9, 10] is presented. The residual-
cancelling VCO-based quantizer (in the dashed box) comprises a flash ADC, a sub-
tractor, a DAC and a frequency-type VCO-based quantizer. First, the quantizer analog
4.3 Closed-Loop VCO-Based DSMs 75

Residual-cancelling VCO-based quantizer

Z-1+Z-2

Flash NRZ
ADC DAC
VCO_Based_Frequency_Quantizer

Kint1
A_IN D_OUT
1
z
First_Order
CLK Difference

NRZ NRZ
DAC DAC
Implicited
DEM

CLK

Explicited
DEM

CLK

Fig. 4.6 Structure of a 2nd-order CT DSM with residual-cancelling VCO-based quantizer [10]

input is converted into digital code by the flash ADC. Then, by using a DAC and a
subtractor, the residual error of the flash conversion is generated, which is processed
further by the frequency-type VCO-based quantizer. The outputs of the flash ADC
and the VCO-based quantizer are summed to create the whole quantizer digital out-
put. From the flash ADC output to the overall quantizer output, there are two paths:
the digital path and the path composed by the DAC, the subtractor and the VCO-based
quantizer. If the gains of these two paths are equal, then the flash ADC noise and
distortion are cancelled. With the low-swing residual error as input, the frequency-
type VCO-based quantizer is very linear, not limiting the DSM performance. In the
frequency-type VCO-based quantizer, the quantization is in the time domain and not
in the voltage domain. So even with a low voltage swing input, all the quantization
levels are used and the quantization error is small.
In the design of Fig. 4.6, 1st-order noise shaping is provided by the VCO-based
quantizer and another 1st-order shaping comes from the active-RC integrator. The
analog loop filter also suppresses all nonidealities of the residual-cancelling VCO-
based quantizer. For the multi-bit DAC feedbacking the flash ADC output, its digital
input has full swing and no intrinsic DEM, so an explicit DEM is required to ease
its matching. The matching requirement of the DAC connected to the VCO-based
quantizer is very loose because of two reasons: first this DAC’s input swing is much
smaller than the full swing; secondly there is implicit DEM on its digital input.
76 4 VCO-Based ADCs

4.4 Open-Loop VCO-Based ADCs

In this section, three published open-loop 1st-order VCO-based ADCs are presented,
with different solutions to address the VCO nonlinearity. Without global feedback,
these ADCs have no stability issue like closed-loop DSMs.

4.4.1 VCO-Based ADC with Background Digital Calibration

In Fig. 4.7, the architecture of a 1st-order open-loop Delta-Sigma ADC [11, 12] is
presented, it is a two-path structure, but only one path of the ADC is shown here.
In each path, to cancel 2nd-order harmonic distortion, a pseudo-differential ADC
topology is utilized. First the fully differential transconductance converts the input
voltage into current, which is fed to the pseudo-differential current-controlled ring
oscillator (ICO). Each ICO phase output is sampled and processed by the digital
differentiator to obtain the digital frequency information. Although the V-to-I con-
version can be very linear, because of the nonlinearity of the capacitor of the ICO
delay cell, the I-to-F conversion is still nonlinear. This nonlinearity determines the
VCO-based ADC performance directly. It is removed by the nonlinearity correction
block, which is explained in the next paragraph. To remove the idle tone in this 1st-
order Delta-Sigma ADC, a dithering signal is generated by the linear feedback shift
register (LFSR) and added to the ICO inputs. The polarities of the dithering signal
added to the two paths are opposite, and by adding the two path outputs, the dither
signal is cancelled at the output of the whole ADC.
To eliminate the VCO distortion, a background digital nonlinearity correction
block is designed. The correction coefficients are as shown in Fig. 4.8. To measure the
ICO 2nd- and 3rd-order distortion coefficients, a replica single-ended path is added
and a calibration sequence t1 [n] + t2 [n] + t3 [n] is used as its input. Here ti [n] (i = 1,
2, 3) is a 2-level, independent, zero mean, pseudo sequence generated by another
low-speed LFSR. In one calibration cycle, the replica path output is correlated with
three 2-level sequences respectively: t1 [n], t1 [n] + t2 [n] and t1 [n] + t2 [n] + t3 [n], and

Phase Nonlinearity
decoder 1-Z-1
Correction

V/I From D_OUT


Calibration Phase Nonlinearity
1-Z-1
decoder Correction

ICO
From Calibration
Dither RZ
DAC
CLK

Fig. 4.7 Architecture of a 1st-order open-loop Delta-Sigma ADC with VCO nonlinearity calibra-
tion (only one path is shown here) [11]
4.4 Open-Loop VCO-Based ADCs 77

t1[n]+t2[2]+t3[n] RZ
C[n]
DAC

CLK/64

Phase 1-Z-1
decoder

V/I
32-bit
sum & dump
t1[n]
32-bit
sum & dump
t1[n]t2[2]
Low-rate 32-bit
coefficient sum & dump
calculation
t1[n]t2[2]t3[n]
32-bit
sum & dump

CLK/228 CLK

Fig. 4.8 Concept of the VCO nonlinearity background digital calibration used in Fig. 4.7 [11]

228 correlation outputs are summed together to calculate the distortion coefficients.
With such digital calibration, the VCO-based ADC SFDR is improved from 50 dB
to better than 80 dB.

4.4.2 VCO-Based ADC with Counting and Foreground


Digital Calibration

An alternative open-loop VCO-based ADC with foreground digital foreground cal-


ibration [13] is shown in Fig. 4.9. In this design, the VCO maximum frequency is
much higher than the sampling clock. To avoid that the VCO output phase transi-
tion information is ruined by the next round of transitions, one delay cell output is
connected to an asynchronous counter. The bit-length of the counter depends on the
ratio of the VCO maximum frequency and the sampling frequency. The counter out-
put is sampled and handled by a multi-bit digital differentiator as the coarse output
of the VCO-based ADC. The fine ADC output is the result of frequency detectors
processing all the VCO output phases.
With such a coarse-fine ADC architecture, the power consumption can be reduced
based on the tradeoff between different design parameters. The global optimization is
based on combinations for different values of the OSR and the number of quantization
bits B. In the local optimization, different numbers of quantization bits comes from
the asynchronous counter and from the frequency detectors. With a M-bit counter,
the VCO maximum frequency can be as high as (2M − 1)× the sampling frequency.
78 4 VCO-Based ADCs

Asynchronous
Counter
1
z
CLK

Thermometer
to binary
1
z
CLK

Fig. 4.9 Structure of an open-loop Delta-Sigma ADC with coarse and fine quantizations and VCO
nonlinearity calibration [13]

Distortion
AIN estimation
VCO-based
DC sweep ADC
DOUT
DOUT_Cal
Look-up
table

Fig. 4.10 Concept of the foreground digital calibration used in Fig. 4.9 [13]

 
Since (B − M)-bit quantization
 is from the frequency detectors, a 2B−M − 1 -stage
VCO and 2B−M − 1 frequency detectors are required. If the maximum speed of
each block is known and the power estimation models are developed for the different
blocks, the parameter combination with the lowest power consumption can be found.
Although a pseudo-differential architecture is applied, the 3rd-order and the resid-
ual 2nd-order harmonic distortions limit the VCO-based ADC performance to about
5 bit. To remove these distortions, a LUT-based digital foreground calibration is
developed; the concept is illustrated in Fig. 4.10. In the calibration mode, 13 DC lev-
els are applied to the ADC input and the digital outputs are processed by a distortion
estimation algorithm. The algorithm fits a straight line to the measured nonlinear
V-F transfer function, with a least squares method. The error between the line and
the measured function is calculated, interpolated and stored in a LUT. In the normal
operation mode, the VCO-based ADC output is corrected by the coefficients stored
in the LUT.

4.4.3 VCO-Based ADC with PWM Precoding

In the above two open-loop VCO-based ADCs, the VCO nonlinearity is eliminated
by digital calibration. Another method used for VCO-based ADC linearity improve-
ment is PWM precoding [14, 15]. The block diagram of an open-loop VCO-based
4.4 Open-Loop VCO-Based ADCs 79

Voltage-to
-Frequency
Asynchronous 2 1
PWM S TS 1-Z-1
KVCO Frequency
-to-Phase
|H(f)| |H(f)| |H(f)| |H(f)|

Fin Freq Fc Freq Fc Freq


Fin Freq Fin

Fig. 4.11 Architecture and illustrative signals of an open-loop VCO-based ADC with PWM
precoding [15]

Fig. 4.12 Implementation of


an asynchronous PWM with
hysteresis single-bit
quantizer [16] ∫

ADC with PWM precoding is presented in Fig. 4.11. The ADC is composed of an
asynchronous PWM modulator and a frequency-type VCO-based quantizer. First the
analog input signal is modulated into a PWM signal. Because of the 2-level charac-
teristic of the PWM signal, the VCO only oscillates at two frequencies, and the VCO
V-F transfer function is intrinsically linear.
One implementation of the asynchronous PWM modulator is shown in Fig. 4.12. It
is similar to a 1st-order single-bit DSM, except that there is no clock in the modulator
and hysteresis is intentionally added to the comparator. Since there is no sampling in
the PWM modulator, there is no quantization noise. However, it introduces intrinsic
3rd-order distortion and some high-frequency signal components around the oscil-
lation frequency. The 3rd-order distortion amplitude depends on the input signal
frequency. The worst case is that the 3rd-order harmonic is located at the signal
bandwidth, and the HD3 can then be estimated as [16]:

π2 M2 1
HD3 = 2
(4.6)
216 M2 OCR2
1− 2

Here M is the modulation depth (the ratio between the input signal amplitude and
the reference value of the feedback path) and OCR is the overcycling ratio (the ratio
between the self-oscillation frequency fc and the input signal bandwidth).
80 4 VCO-Based ADCs

4.5 Conclusions

To be compatible with SoC design in nanoscale CMOS technologies, highly-digital


time-domain ADCs gain more and more attention and several novel ADC architec-
tures have been proposed. In this chapter, recently developed time-domain quantiz-
ers and A/D converters have been presented and discussed. Based on the number
of VCO output phase transitions, two kinds of VCO-based counting quantizer have
been developed. If the VCO maximum frequency is less than half the sampling fre-
quency, the counter can be replaced by a simple FF and two types of VCO-based
quantizer are proposed. With such highly-digital VCO-based quantizers, not only
sampling and quantization are done, but also 1st-order noise shaping and intrin-
sic DEM (for the frequency-type VCO-based quantizer) are provided. The signal
transfer function of the phase-type VCO-based quantizer is an integrator, which can
only used in closed-loop DSM designs. For frequency-type VCO-based quantizers,
the VCO nonlinearity is the major design challenge when designing moderate- or
high-accuracy ADCs. One solution is to apply this quantizer in closed-loop DSMs
and to suppress the VCO nonlinearity by the analog loop filter. Another solution
is to use the VCO-based quantizer for residual signal conversion, resulting into the
residual-cancelling ADC structure.
To further digitalize the ADC design, three open-loop ADCs with VCO-based
quantization have been introduced. Without global feedback, these ADCs are very
stable and their quantization noise is intrinsically 1st-order shaped. In these ADC
designs, if there is a counter assigned for one VCO phase output, then less VCO
stages are required and the VCO frequency can be set higher. One popular solution
to the VCO nonlinearity is the digital LUT-based calibration, using either a random
sequence or a DC voltage as calibration input. Another architecture is using a PWM-
modulated signal and not an analog input signal to drive the frequency-type VCO-
based quantizer; the PWM two-points characteristic eliminates the VCO nonlinearity
issue. However, the PWM will introduce inherent 3rd-order harmonic distortion.
In Chaps. 6, 7 and 8, we will introduce three different VCO-based ADCs (both
closed-loop ones and open-loop one) with excellent power efficiencies that mitigate
the VCO nonlinearity.

References

1. V.B. Boros, A digital proportional integral, and derivative feedback controller for power con-
ditioning equipment, in IEEE Power Electronics Specialists Conference (1977), pp. 135–141
2. J.P. Hurrell, D.C. Pridmore-Brown, A.H. Silver, Analog-to-digital conversion with unlatched
SQUID’s. IEEE Trans. Electron Devices 27(10), 1887–1896 (1980)
3. M. Hovin, A. Olsen, T.S. Lande, C. Toumazou, Delta-sigma modulators using frequency-
modulated intermediate values. IEEE J. Solid-State Circuits 32(1), 13–22 (1997)
4. M.Z. Straayer, M.H. Perrott, A 12-bits, 10-MHz bandwidth, continuous-time Delta-Sigma
ADC with a 5-bits, 950-MS/s VCO-based quantizer. IEEE J. Solid-State Circuits 43(4), 805–
814 (2008)
References 81

5. M.Z. Straayer, M.H. Perrott, A 10-bit 20-MHz 38-mW 950-MHz CT ΔΣ ADC with a 5-bit
noise-shaping VCO-based quantizer and DEM circuit in 0.13-µm CMOS, in VLSI Symposium
Dig. (2007), pp. 246–247
6. J. Kim, T.K. Jang, Y.G. Yoon, S.H. Cho, Analysis and design of voltage-controlled oscillator
based analog-to-digital converter. IEEE Trans. Circuits Syst. I Regul. Pap. 57(1), 18–30 (2010)
7. M. Park, M.H. Perrott, A 0.13 µm CMOS 78 dB SNDR 87 mW 20 MHz BW CT ΔΣ ADC
with VCO-based integrator and quantizer, in 2009 IEEE International Solid-State Circuits
Conference - Digest of Technical Papers, ISSCC (2009), pp. 170–171
8. M. Park, M.H. Perrott, A 78 dB SNDR 87 mW 20 MHz bandwidth continuous-time ΔΣ ADC
with VCO-based integrator and quantizer implemented in 0.13 µm CMOS. IEEE J. Solid-State
Circuits 44(12), 3344–3358 (2009)
9. K. Reddy, Rao S., R. Inti, B. Young, A. Elshazly, M. Talegaonkar, P.K. Hanumolu, A 16 mW
78 dB-SNDR 10 MHz-BW CT-DSM ADC using residue-cancelling VCO-based quantizer, in
2012 IEEE International Solid-State Circuits Conference Digest of Technical Papers (ISSCC)
(2012), pp. 152–154
10. K. Reddy, Rao S., R. Inti, B. Young, A. Elshazly, M. Talegaonkar, P.K. Hanumolu, A 16-mW
78-dB SNDR 10-MHz BW CT ΔΣ ADC using residue-cancelling VCO-based quantizer. IEEE
J. Solid-State Circuits, 47(12), 2916–2927 (2012)
11. G. Taylor, I. Galton, A mostly-digital variable-rate continuous-time ADC ΔΣ modulator, in
2010 IEEE International Solid-State Circuits Conference Digest of Technical Papers (ISSCC)
(2010), pp. 298–299
12. G. Taylor, I. Galton, A mostly-digital variable-rate continuous-time Delta-Sigma modulator
ADC. IEEE J. Solid-State Circuits 45(12), 2634–2646 (2010)
13. J. Daniels, W. Dehaene, M. Steyaert, A. Wiesbauer, A 0.02 mm2 65 nm CMOS 30 MHz BW
all-digital differential VCO-based ADC with 64 dB SNDR, in 2010 IEEE Symposium on VLSI
Circuits (VLSIC) (2010), pp. 155–156
14. L. Hernandez, S. Paton, E. Prefasi, VCO-based sigma-delta modulator with PWM precoding.
Electron. Lett. 47(10), 588–589 (2011)
15. S. Rao, B. Young, A. Elshazly, W. Yin, N. Sasidhar, P.K.Hanumolu, A 71 dB SFDR open-
loop VCO-based ADC using 2-level PWM modulation, in 2011 Symposium on VLSI Circuits
(VLSIC) (2011), pp. 270–271
16. P. Gao, X. Xing, J. Craninckx, G. Gielen, Design of an intrinsically-linear double-VCO-based
ADC with 2nd-order noise shaping, in 2012 Design, Automation and Test in Europe Conference
and Exhibition (DATE) (2012), pp. 1215–1220
Chapter 5
CT DSM ADCs with VCO-Based
Quantization

5.1 Introduction

Ever higher data rates in both wireless and wireline communications lead to an
ongoing increase in signal bandwidths. In receiver chains, the variable gain ampli-
fiers (VGAs) and the automatic gain control loops (AGCs) can be relaxed or even be
removed if the dynamic range (DR) of the following analog-digital converter (ADC)
is high enough, simplifying the analog front-end [1]. For better integration, the ADC
is placed closer and closer to the antenna. Low-IF conversion with a digital channel
selection filter and second-stage mixer is considered to be more cost-efficient than
the traditional heterodyne receiver architecture for multi-channel applications [2].
This simplifies the RF front-end and improves the system flexibility, but calls for
ADCs with larger bandwidth and higher accuracy. Therefore power-efficient ADCs
with tens of MHz of bandwidth and 12–14 bit resolution are in great demand for
future communication applications. Delta-Sigma modulators (DSMs) are an excel-
lent candidate for this. For the higher bandwidth, continuous-time (CT) DSM ADCs
are preferred, because of their high power efficiency [3], the inherent anti-aliasing
characteristic [4] and the possibility of re-configurability [5].
This chapter presents the full designs and evaluation results of two 40 MHz-BW
12-bit CT DSM ADCs with phase-type VCO-based quantizer. In our first design
[6], the CT DSM ADC is comprised of feedback DACs, a 3rd-order analog filter
and a 5-bit phase-type VCO-based phase quantizer. The proposed shaped SC DAC
has a reduced peak current while consuming no extra hardware or power. As a
result, about 30% operational transconductance amplifier (OTA) power consumption
is saved due to the relaxed slewing requirement. The front-end DAC nonlinearity is
cancelled via a LUT-based digital calibration. In this way, a power-efficient 40 MHz-
bandwidth modulator is obtained with 12-bit resolution and over 80 dB spurious-free
dynamic range (SFDR) in a 90 nm CMOS technology. The second CT DSM is an
improved design based on the first one, targeting a lower power consumption and no
digital calibration [7]. Here a half-period delay but not a one-period delay scheme
84 5 CT DSM ADCs with VCO-Based Quantization

is applied to compensate the modulator ELD. A capacitive local feedback without


noise contribution is proposed for the NTF zero optimization. The FF OTA used in
the loop filter is implemented by a current-sharing structure, leading to a considerable
analog power reduction. With these techniques, about 70 dB SNR is reached with
only 45 mW power, resulting in a FoM of 0.26 pJ/Step.

5.2 A 40 MHz-BW 12-Bit CT DSM with Digital Calibration


and Shaped SC DAC

5.2.1 Structure of the CT DSM

5.2.1.1 System Parameters

In high-bandwidth DSMs, the signal-to-noise ratio (SNR) should be limited by the


thermal noise of the analog front-end circuit [8], and the contribution of the quanti-
zation noise should be small. To incorporate some design margin, we chose to design
for about 90 dB signal-to-quanzation noise ratio (SQNR) to achieve a 12-bit thermal
noise limited design. In CT DSM design, the maximum OTA GBW and the mini-
mum digital gate delay determine the maximum sampling frequency, and hence the
maximum available OSR. Compared to lower-order DSMs, single-loop CT DSMs
of 5th and higher order become prone to instability. Alternatively, a MASH structure
could reduce the noise shaping order in each loop and improve the stability, but the
mismatch between the CT integrator and the digital filter imposes another design
difficulty, usually requiring digital calibration [9]. Trading off all considerations,
an architecture with OSR of 12, 4th-order noise shaping and 5-bit quantization is
adopted in this design for its robust performance and relaxed requirements on the
building blocks.

5.2.1.2 Modulator Topology and Block Selection

The block diagram of the CT DSM is shown in Fig. 5.1: 1st-order noise shaping
and 5-bit quantization are executed by a phase-type VCO-based quantizer [10]. The
remaining 3rd-order noise shaping is done by the analog inverse Chebyshev filter
preceding the quantizer. A non-return-to-zero (NRZ) DAC and a SC DAC delayed
by one sample period are used to form the different feedback loops. To solve the
modulator stability problem caused by the delayed DAC signals, an extra return-to-
zero (RZ) DAC is added to form a zero-order feedback loop.
The VCO-based phase quantizer consists of a 31-stage ring VCO and 31 slices
of the phase and frequency detection circuit, which is comprised of basic digital
gates (flip-flop, XOR gate). The VCO converts the analog input control voltage into
phase, which is indicated by the occurrence in time of the voltage transition edge.
5.2 A 40 MHz-BW 12-Bit CT DSM with Digital Calibration and Shaped SC DAC 85

DFF path

VCO_Based_Phase_Quantizer
Reference Pulse

Dout Digital Dcal


Ain Calibration

1
z
31 slices
CLK

LATCH
NRZ RZ
DAC DAC D Q
SC
DAC
CLK_B
LATCH
D Q

CLK

Fig. 5.1 Structure of the 40 MHz 12-bit CT DSM

The quantization error is generated when the VCO outputs are synchronized by the
sampling flip-flops; the XOR gate together with the delay cell is used to differentiate
the digital phase to obtain the digitized VCO frequency. The function of the second
XOR gate with reference pulses is to remove the DC information of the digital
phase, which corresponds to the VCO common-mode control voltage (0.6 V) and
free-running frequency (one quarter of the sampling frequency). Because of the high
gain of the VCO inherent integration, the input swing of the VCO is small, and as a
result the effect of the nonlinearity of the voltage-to-frequency/phase relationship is
greatly reduced as well. Simulation results show that the harmonic distortion of the
quantizer is around −90 dBc [10], plus the attenuation provided by the loop filter.
Therefore the distortion of the VCO-based quantizer is not the bottleneck for the
modulator performance.
The 3rd-order analog loop filter is a feedforward structure, for its reduced integra-
tor output swings and lower requirements for the transconductors [11]. Although the
direct feedforward (DFF) path in the analog filter leads to a larger signal-transfer-
function (STF) out-of-band peaking and reduced anti-aliasing filtering at higher fre-
quencies (a deterioration of about 2 dB for both), it can reduce the integrator output
swing further. This relaxes the linearity requirements for the OTA. Because of the
inherent integrating operation in the VCO, at least a 1st-order inherent anti-aliasing
at high frequencies is still maintained when using this DFF path. One local feedback
path across the first two integrators is used to generate two finite noise-transfer-
function (NTF) zeroes (at about 30 MHz) to further suppress the quantization noise
near the signal bandwidth. Although Gm-C integrators are popular in high-speed
CT DSM design, to achieve 12-bit linearity, both degeneration resistors and local-
feedback amplifiers would be needed for the transconductors [12], at the expense of
extra power. With active-RC integrators, more stable output nodes can be provided
to the feedback DAC, which improves its linearity. With the optimized loop filter
structure and coefficients, the amplifier GBW requirements for the second and third
integrators are only two times the sample frequency. Based on these considerations,
all integrators are realized as active-RC filters. The input resistors of the first integra-
tor are about 1.5 k which leads to a thermal noise contribution of about 80 dBFS. To
overcome the problem due to the large variation of the RC time constant, 4-bit binary
86 5 CT DSM ADCs with VCO-Based Quantization

Fig. 5.2 CT DSM SNDR DSM Performance versus OTA GBWs


performance as a function of 85
the scaled OTA GBWs
(normalized to 4, 2 and 80
2 GHz for loop filter
integrator one to three 75
respectively)

SNDR(dB)
70

65

60

55

50
0.2 0.4 0.6 0.8 1 1.2
Scale of the OTA GBWs(4−2−2GHz)

capacitor banks are connected in parallel with the fixed-value capacitors to adjust the
capacitor values in steps of 2.7% over 20% range. In CT DSM design, the OTA in the
first active-RC integrator largely determines the modulator performance: insufficient
gain over the whole signal bandwidth would result in a too large signal swing on
the amplifier input virtual ground, increasing the effect of integrator nonlinearity and
noise. A too small GBW of the OTA would induce an extra phase shift and gain atten-
uation, causing an increased in-band noise and distortion or even instability. Finally,
a too limited voltage swing of the amplifier would lead to nonlinearity of the loop
filter. Figure 5.2 shows the simulated modulator SNDR performance as a function of
the scaled GBW values (from the 1st to the 3rd OTA, the GBWs are normalized to 4,
2 and 2 GHz respectively) for fixed DC gains (55 dB) of the OTAs. A GBW of 4 GHz
for the first OTA is required for the modulator to keep the non-ideal OTA-induced
noise/distortion below 10% of the total budget.
The performance requirements for the feedback DACs are different because of
their different position in the modulator and their different inputs. Referring all non-
idealities to the modulator input directly, the performance requirement for the SC
DAC must be at least the same as the DSM specifications. On the other hand, the
requirements of the NRZ DAC are greatly relaxed because its noise and distortion
are suppressed by the 3rd-order analog filter. For the RZ DAC, in addition to the
suppression provided by the loop filter, two facts further relax the required accuracy.
Firstly, because of the integrating operation of the VCO, the input swing of the
VCO-based quantizer is small, and so is its digital frequency output. Although the
RZ DAC is 5-bit, it behaves as a lower-bit DAC most of the time (from behavioral
simulation at most 6 levels are used). With most of the DAC cells switched off, both
the nonlinearity and the output noise current are reduced. Secondly, in the ring VCO
the inverter stages change state continuously, so the frequency output of the quantizer
is intrinsically dynamic element matched, randomizing the error [13]. As they are
5.2 A 40 MHz-BW 12-Bit CT DSM with Digital Calibration and Shaped SC DAC 87

Fig. 5.3 CT DSM SNDR DSM Performance versus DAC Delays


performance as a function of 100
the delay for the three
different feedback DACs 80 SC−DAC
NRZ−DAC
60 RZ−DAC

SNDR
40

20

−20
0 0.05 0.1 0.15 0.2
Delay(*Ts)

placed in feedback loops with different order, the speed requirements for the feedback
DACs are also different. The CT DSM SNDR performance as a function of the DAC
delay is presented in Fig. 5.3. It is clear that the SC DAC can tolerate as much as
20% sampling period delay, but the delay for the RZ DAC should not exceed 5% of
the period. The reason is that in CT DSMs the lower-order loops are responsible for
the higher-frequency feedback signal components, and for these higher-frequency
signals a small delay already corresponds to a large phase shift. In this work, a shaped
SC DAC with smaller peak current is proposed for the outermost DAC as it offers
several advantages. The noise contribution of the SC DAC is about:

kT
2
VSC_DAC = (5.1)
OSR · C
If the total capacitor value C is chosen large enough, this noise can be ignored
compared to the thermal noise of the input integrator resistor Rin:

VR2in = 8kTRin BW (5.2)

For comparison, if a NRZ DAC is used, then the DAC noise is about [14]:

16kTgm 2 32kTI
2
VNRZ_DAC = Rin BW =   R2 BW
3 3 Vgs − Vth in
32kT Vref
=   Rin BW (5.3)
3 Vgs − Vth

where Vref is the reference voltage of the modulator, which is 0.6 V in our design,
and the gate-source over-drive voltage Vgs − Vth of the current-source transistors in
88 5 CT DSM ADCs with VCO-Based Quantization

the NRZ DAC is usually set to 0.15 V or less in nanoscale CMOS circuit design. It
is clear from (2) and (3) that in this case the NRZ DAC noise is about 5 times the
noise contribution of the input resistors. Although a degeneration resistor could be
used to reduce the NRZ DAC noise, a higher supply voltage is needed because of the
voltage across the resistor [8]. In addition, since the output current returns to zero
every period, the SC DAC is more immune to inter-symbol interference than a NRZ
DAC, which is reported to be a main performance limitation in high-bandwidth CT
DSMs [10]. Since it is not suppressed by any analog filter, the overall nonlinearity
of the DSM is limited by that of the SC DAC, which is discussed in detail in the
next subsection. With the selected architectural parameters, the DSM NTF has been
synthesized by Schreier toolbox [15] and the coefficients have been optimized by
behavioral-level parameter explorations.

5.2.1.3 LUT-Based Digital Calibration

As the nonlinearity of the 5-bit main DAC is not suppressed by any loop filter, it needs
to be solved by other methods. If explicit DEM is applied between the quantizer and
the resample flip-flop, the delay of the digital block that implements the DEM can not
be larger than half a sampling period [10]. Circuit-level simulations in 90 nm CMOS
show that if implemented in static logic, the delay of a 5-bit DEM block is about
740 ps, failing to meet the timing requirement. If realized in pseudo-NMOS logic,
simulations show that the delay is about 470 ps, with 18.2 mW power dissipation.
Considering the extra delay caused by any parasitic wiring capacitance and resistance,
there is almost no design margin on the timing. Obviously also the associated large
power consumption is a significant drawback.
Another solution to the nonlinearity of the main feedback DAC is to use digital
calibration [16]. This concept is illustrated in Fig. 5.4. Here, for each DAC level, an
accurate digital representation is stored in a digital LUT. For our case of a 5-bit DAC,
32 DAC levels must be stored. During the normal DAC operation, the corresponding
calibrated output, Dcal, is then retrieved from the LUT.
Obviously the accuracy of such a calibration depends on how accurately the
calibration coefficients in the LUT correspond to the actual DAC levels. To find
the optimal calibration coefficients, in this work a least-mean-square (LMS)-based
offline calibration approach is used.1 A complete mathematical discussion of this
technique is out of the scope of this book, but the basic idea can simply be under-
stood. It consists of applying a known analog input signal Ain to the ADC. Then
the calibration values in the LUT are tuned until the digital output Dcal matches the
input signal Ain in a LMS sense (i.e. the mean square of Dcal − Ain is minimized).
In our implementation we used a spectrally pure sinusoidal input signal, which is
mathematically perfectly known. For our approach it is not needed that the amplitude
or the frequency are known in advance. In this way such an input stimulus signal can

1 Thistechnique was developed by Pieter Rombouts and Maarten De Bock. It was applied in this
design as part of a collaboration within a FWO project.
5.2 A 40 MHz-BW 12-Bit CT DSM with Digital Calibration and Shaped SC DAC 89

Dout Dcal
31 31
30 29.998
Ain CT DSM Dout (5 bit) Dcal (16 bit) Decimation
(fig. 1) Filter
2 1.9993
1 1.0009
0 0
LUT

Fig. 5.4 Block diagram of the digital calibration of the DAC. The values in the look-up table are
arbitrary values, for illustrative purpose

readily be obtained in practice through bandpass filtering. Clearly this approach is


very well suited for prototype debugging and testing. Moreover, the implementation
of this LMS algorithm we used requires only 1024 baseband samples and hence is
also a viable approach for factory-level calibration.

5.2.2 Delta-Sigma Modulator Building Blocks Design

5.2.2.1 4-Stage Feedforward Compensated OTA

For OTAs with traditional structures, there are different bottlenecks to realize the
specification of 55 dB gain and 4 GHz GBW. With a single-stage amplifier, it is
difficult to reach a gain of over 30 dB in 90 nm CMOS. In cascoded and folded-
cascoded OTAs, either the supply voltage or the output swing is limited. For a Miller-
compensated topology, too much power is required to push all non-dominant poles
far beyond the GBW. Therefore, a 4-stage full-feedforward compensated OTA is
proposed in this design to achieve a high gain, a high GBW and a large signal
swing. The block diagram is shown in Fig. 5.5, and all transconductors are realized
by NMOS differential pairs, with PMOS current sources as loads. The 4 cascaded
stages generate 4 poles. The stability is guaranteed by introducing 3 left-half-plane
(LHP) zeros to cancel or compensate 3 poles. Here each feedforward path generates
one LHP zero [17].

Vout Gm1 R1 Gm2 R2 Gm3 R3 Gm4 + sGmf 3 C1 R1 R4


= (5.4)
Vin 1 + sC1 R1 1 + sCL R4

1 1 Gm1 Gm2 R2 Gm3 R3 Gm4


P1 = − P4 = − Z3 = − (5.5)
C1 R1 CL R4 Gmf 3 C1
90 5 CT DSM ADCs with VCO-Based Quantization

High-Speed Path z3

Gmf3
z2

Gmf2
z1
L=1.5*Lmin L=1*Lmin
Gmf1
L=4*Lmin L=2*Lmin High-Gain Path
p1 p2 p3 p4

Vi1 Vi2 Vi3


Vin Gm1 Gm2 Gm3 Gm4 Vout

R1 C1 R2 C2 R3 C3 R4 CL

Fig. 5.5 Block diagram of the 4-stage full-feedforward compensated operational transconductance
amplifier

In our design, because of the large output resistance R1 and the small output
resistance R4 , P1 is the dominant pole (about 7.7 MHz) and P4 is the non-dominant
pole (about 600 MHz); Z3 (at about 200 MHz) is placed between P1 and P4 to increase
the GBW and the phase margin (PM). For the 1st- and 2nd-stage cascaded amplifiers,
the output signal swings are small, so PMOS cascode transistors are added to increase
the gain and to reduce the output parasitic capacitors. The transistor gate lengths of
the 4 stages are 4, 2, 1.5 and 1 the minimum gate lengths respectively. For the internal
high-impedance nodes, capacitors are added intentionally to outweigh the transistor
parasitic capacitors, leading to controllable poles and zeroes. A Miller-compensated
common-mode feedback loop (CMFB) with about 500 MHz GBW is arranged for
every stage to get a stable CM voltage. In CMOS implementation, because of the
technology parameter variations, the above cancellation of the pole-zero pairs may
not be exact. In Fig. 5.6, the modulator SNDR performance with OTA doublets is
plotted. In these simulations, three fixed poles (30/200/600 MHz) and three zeroes
with variable positions are considered in all OTAs. It is shown that 90% normalized
doublet frequency variation can be tolerated by the proposed design, and there is
no SNDR performance penalty if the doublet relative position is within this range.
In Fig. 5.7, the frequency response of the first OTA is presented. Also the internal
gains of the three cascaded stages are shown. The proposed OTA reaches 56 dB DC
gain, 3.9 GHz GBW and 82◦ phase margin with a 1.3 pF load capacitor, dissipating
14.4 mW power. Because of the large power consumption, the input noise (from
1 kHz to 40 MHz) is only 16 µV. As their non-idealities are shaped by the preceding
integrators, the second- and third-integrator amplifiers have the same structure, but
with about 40% reduced size and power.
5.2 A 40 MHz-BW 12-Bit CT DSM with Digital Calibration and Shaped SC DAC 91

Fig. 5.6 SNDR DSM Performances with OTA Doublets


performance of the CT DSM 100
as a function of the relative
position of the poles and 80
zeros in the doublets

60

SNDR
40

20

−20
−1 −0.5 0 0.5 1 1.5
Doublet Relative position ( (WZ−WP)/WP )

5.2.2.2 Shaped Switched-Capacitor DAC with Reduced Peak Current

For the SC DAC design, three parameters are critical if applied in a CT DSM:
the peak current Ipeak , the tail current Itail and the delivered charge Q. The peak
current determines the slew-rate requirement in the first OTA; the tail current decides
the modulator clock jitter sensitivity, and the delivered charge sets the feedback
coefficient:
 
Vref Vref −0.5T
Ipeak = Itail = exp (5.6)
Ron Ron τ
  
−0.5T
Q = Vref C 1 − exp τ = Ron Cu (5.7)
τ

where Vref is the reference voltage, Ron is the on-resistance of the switches, Cu is
the capacitor of the SC DAC cell and τ is the discharge time constant. To relax
the OTA slew rate, one can reduce the peak discharge current, by increasing the
on-resistance Ron or by reducing the reference voltage Vref . However, to keep the
feedback coefficient (the charge Q) the same, a larger discharge time constant is
needed, which will result in a larger discharge tail current and impair the DSM clock
jitter performance. We solve this problem using a shaped SC DAC cell.
The shaped SC DAC cell and its non-overlapping driving circuit are shown in
Fig. 5.8: the capacitors are implemented as metal-insulator-metal (MIM) capacitors
and the switches are realized as transmission gates. The capacitor of each SC DAC cell
Cu is 200 fF, so the SC DAC noise is about −95 dBFS. In the charging phase, the two
capacitors are charged to the positive and negative reference voltages respectively.
In the discharging phase, the charges are injected onto the OTA virtual grounds and
92 5 CT DSM ADCs with VCO-Based Quantization

AC Response of the 1st OTA

50

150

Phase (Deg)
Gain (dB)

OTA Gain
0 1 Stage Gain
2 Stages Gain
3 Stages Gain
Phase
100

1 2 3 4 5 6 7 8 9 10
10 10 10 10 10 10 10 10 10 10
Frequency (Hz)

Fig. 5.7 AC simulation result of (internal) gain of the operational tranconductance amplifier used
in the first integrator

integrated on the first integration capacitors. The discharge polarity is determined by


the DAC input, and the shape of the discharge current depends on the rising/falling
edge slope of the DAC driving signal, the reference voltage Vref , the on-resistance
of the switches Ron and the discharge time constants. Compared with a traditional
SC DAC, our proposed shaped SC DAC introduces the driving signal slope to shape
the DAC discharge current and breaks down the trade-off between the OTA slew rate
requirement and the modulator clock jitter sensitivity. It is observed that if the rise
time of the discharge current is enlarged, the peak current is reduced, and so is the
slew rate requirement in the first OTA.
In our design, the rising edge of the DAC driving signal is low-pass filtered, which
can be realized by adjusting the transistor drive capability of the preceding gates. As
shown in Fig. 5.8, in the driving circuit for the transmission gates NMOS part, the
PMOS transistor width is reduced to realize an extended driving signal rising edge;
in the driving gate for the PMOS transmission transistor, the NMOS transistor size
is scaled down for a low-passed filtered falling edge. As a result, the rising edge of
the DAC discharge current gets one extra time constant, resulting in an extended rise
time (about 72 ps) and a reduction of about half of the current peak value, as shown
in Fig. 5.9. Since this effect only occurs during the switching-on of the switches, the
tail of the DAC discharge current is not affected by this shaping, and the current
amplitude at the end of the discharging phase is almost the same (increased only 8%
compared to the ideal discharge current). With this shaped SC DAC current pulse,
the slew-rate requirement of the first OTA can greatly be relaxed, reducing its power
consumption by about 30%. Because of the SC DAC incompletely settling in CT
5.2 A 40 MHz-BW 12-Bit CT DSM with Digital Calibration and Shaped SC DAC 93

CLK
DIN
VDCHN
Unbalanced Driving
DIN

VCH
VVGN1 VVGP1
DINB
VDCHN VDCHP

VDCHP

VDCHP

VDCHN VDCHP
VDCHN

CLK

VCH VCH

VREFN VREFP
Balanced Driving
Shared by All cells
VCH VCM

Fig. 5.8 Circuit implementation of the shaped SC DAC (right) and its driving circuit (left)

Fig. 5.9 The output current Shaped SC DAC Current


pulse of the shaped SC DAC, 60
compared to the traditional Shaped
pulse shape, as used in our Ideal
40
design
SC DAC Current (A)

20

−20

−40

−60
5.3 5.4 5.5 5.6 5.7 5.8 5.9
Time (nS)

DSMs, the matching of the SC DAC cells does not only depend on the matching of the
capacitors, but also on that of the on-resistance of the transmission gates (assuming
no mismatch on the driving gates):
     
ΔQ mTexp(−mT /τ )/τ 2 2 ΔRon
σ2 = σ
Q 1 − exp(−mT /τ ) Ron
   
1 − exp(−mT /τ ) − mTexp(−mT /τ )/τ 2 2 ΔC
+ σ (5.8)
1 − exp(−mT /τ ) C

Here mT is the fall time of the shaped SC DAC discharge current, which is about
410 ps in our design. With the component sizing and mismatch information, the
mismatch of the redistributed charge is about 1%, which could be solved by the
digital calibration described in the last subsection.
94 5 CT DSM ADCs with VCO-Based Quantization

CM Parts Vdd
Vdd

0.9V
0.9V
0.95V
0.95V
0.65V

0.65V
0.7V
1.2V 1.2V
0.6V

1.2V 1.2V 0.6V


0.12um/0.08um
1.2V 1.2V 1.2V
0.24um/0.08um
0.5V

3.4um/0.16um 0.5V
0.5V
0.5V 7um/0.16um
0.2V 0.2V
0.3V 3.2um/0.9um 3.4um/0.2um
0.3V

NRZ DAC Cell RZ DAC Cell

Fig. 5.10 Schematic of the current-steering DAC cells

5.2.2.3 Current-Steering DACs

The schematics of the NRZ and RZ DAC cells and their biasing voltages are shown
in Fig. 5.10: cascode transistors with about 8× gain are used to shield the current-
source transistors from the noisy switching transistors. Overlapping drive signals
(high crossing for the NMOS part and low crossing for the PMOS part) are applied
to minimize the glitch energy [18]. Because full swing is used for the driving signals,
the sizes of the switch transistors are small, leading to a high-speed operation.

5.2.2.4 Adder

Both passive and active adders could be used to convert the CS DAC currents to
voltages and sum them together with the integrator output voltages. However, com-
pared with an active adder, a passive adder can only provide attenuation, especially
when there are several input branches. The VCO input capacitor is around 235 fF. If
a passive adder would be applied, small resistors are needed to push the poles formed
by the adder resistors and this capacitor to a high frequency to avoid an extra phase
shift, which means a large AC power. For these reasons, an active adder with about
5.2 A 40 MHz-BW 12-Bit CT DSM with Digital Calibration and Shaped SC DAC 95

7 mW power is applied in our design. One more advantage of the active adder is the
more stable outputs for the DACs, resulting in a better linearity of these DACs.

5.2.3 Measurement Setup and Experimental Results

5.2.3.1 Measurement Setup

The 40 MHz 12-bit CT DSM has been manufactured in UMC 90 nm standard-


performance (SP) CMOS. The chip photo is shown in Fig. 5.11. The area of the
core circuit is about 0.28 mm2 . The 31 thermometer bits are coded to 5-bit binary
outputs by a Wallace tree adder (the U2B block on the chip photo) and fed to low-
voltage differential signaling (LVDS) buffers to drive the measurement instruments.
The measurement setup for the designed CT DSM is shown in Fig. 5.12: a ceramic
substrate has been designed to mount the die and to provide electrical connections
with the instruments. All power supplies, bias voltages/currents and control bits are
generated by a FR4 printed circuit board (PCB). All supply voltages and bias nodes
are decoupled by capacitors on the ceramic substrate and the PCB. Passive bandpass
filters are used to suppress harmonics and noise of the analog input signals from the
AWGs. For the two-tone test a power combiner combines two filtered sine waveforms
without introducing intermodulation distortion (IMD). A balun converts the single-
ended analog waveforms into differential signals. The clock signal is generated by a
pulse generator with low jitter (typical value of 1 ps rms). The on-chip LVDS digital
output bits are captured by a logic analyser and then analysed by Matlab code. In our
setup the digital calibration described previously has been implemented in Matlab
and performed in an offline cycle after which the calibration coefficients are stored
in the memory. The calibration algorithm used 16 K samples (corresponding to 1365
baseband samples). The execution of the calibration script on a low-cost desktop
PC was nearly instantaneous (<0.2 s). Based on the observed calibration values, the
mismatch of the DAC unit elements was found to be in the order of 1%, most of
which is from the on-resistance of the transmission gates.

5.2.3.2 Result and Discussion

On the left of Fig. 5.13, the measured power spectral density is plotted for the uncal-
ibrated digital outputs of −4.5 dB single-tone inputs with frequencies of 938 kHz
and 7.5 MHz. On the right side, the corresponding calibrated results are shown. The
DAC mismatch distribution used for calibration is extracted from the measurement
with a 938 kHz input. When a 938 kHz input is applied, the SNDR without and with
calibration is 63.5 dB and 70.8 dB, respectively. As seen in the plot, after calibra-
tion, all harmonics are under the noise floor. In the 7.5 MHz input measurement, by
applying the digital calibration, the SNDR is improved from 63.1 to 70.3 dB. At this
input frequency, part of the 2nd harmonic remains. For both frequencies, the SNDR
96 5 CT DSM ADCs with VCO-Based Quantization

Fig. 5.11 Die photo of the 40 MHz 12-bit CT DSM

DC Supply
PCB

PoleZero Bandpass ParBERT


Filter Bank Agilent 81250
Matlab
DSM @PC
Signal Source Balun Mount Substrate
AWG 430 mini circuit
ADT1-1WT
Cascade bandpass filter
mini circuit NBP-10.7+
Pulse Generator
HP8133A
Signal Source
Agilent 33250A
PoleZero Bandpass
Filter Bank
Power Combiner
mini circuit ADP-2-1+
Signal Source
AWG 430

Fig. 5.12 Measurement setup for the 40 MHz 12-bit CT DSM

is improved by more than 7 dB, validating the effectiveness of the digital calibration.
The out-of-band peak in the measured spectrum of Fig. 5.13 is a result of the changed
NTF, and can be adjusted by means of the capacitor banks in the integrators and the
DAC biases. The measured modulator SNR/SNDR for increasing analog input pow-
ers is shown in Fig. 5.14. For the 938 kHz sine input, the peak SNR is 73.6 dB with a
−2 dB analog input, and the peak SNDR is 72.9 dB when the input is −2.5 dB. The
peak SNR/SNDR is 71.7/70.8 dB respectively when a 7.5 MHz −3/−3.5 dB analog
input is applied. The peak SFDR for the two measurements is 84.2 and 80.6 dB,
respectively, which is very large for a 40 MHz ADC.
5.2 A 40 MHz-BW 12-Bit CT DSM with Digital Calibration and Shaped SC DAC 97

0 0

−20 −20

−40 −40
PSD (dB)

PSD (dB)
−60 −60

−80 −80

−100 −100

−120 −120
5 6 7 8 5 6 7 8
10 10 10 10 10 10 10 10
Frequency (Hz) Frequency (Hz)

0 0

−20 −20

−40 −40
PSD (dB)

PSD (dB)
−60 −60

−80 −80

−100 −100

−120 −120

5 6 7 8 5 6 7 8
10 10 10 10 10 10 10 10
Frequency (Hz) Frequency (Hz)

Fig. 5.13 Measured spectra with (right) and without (left) digital calibration for 938 kHz and
7.5 MHz input signals

In Fig. 5.15, the spectrum of a −10.5 dB two-tone input (10 and 10.5 MHz) IMD
test is shown. Because of the digital calibration, the 3rd-order IMD is almost buried
in the noise floor, and the 2nd-order IMD is about −78 dB. The power dissipation of
the core circuit is 69.6 mW, of which about 45 mW is consumed by analog blocks,
including the integrators, adder and DACs. At low input frequencies (1 MHz) the
circuit achieves a FoM of 0.31 pJ/Step (Power/(2BW*2ENOB )) at the worst-case input
frequency. The overall performance of the implemented CT DSM is summarized in
Table 5.1, and will be compared to the state-of-the-art works at the end of Chap. 8,
after having presented the final ADC design.

5.3 A 40 MHz-BW 12-Bit CT DSM with Capacitive Local


Feedback and Current-Sharing OTA

In the design of the previous section, the foreground digital calibration can not trace
the temperature and other (drift, ageing, etc.) variations. To reduce the modulator
power consumption and apply the DSM without digital calibration, an improved
design of the 40-MHz 12-bit CT DSM is done and presented in this section.
98 5 CT DSM ADCs with VCO-Based Quantization

80 80
SNDR SNDR
70 SNR 70 SNR

60 60
SNR/SNDR (dB)

SNR/SNDR (dB)
50 50

40 40

30 30

20 20

10 10

0 0
−80 −60 −40 −20 0 −80 −60 −40 −20 0
Normalized Input Signal(dB) Normalized Input Signal(dB)

Fig. 5.14 SNR/SNDR versus input amplitude for 938 kHz input signals (left) and 7.5 MHz input
signals (right)

Fig. 5.15 Output spectrum for a −10.5 dB two-tone test with 10 and 10.5 MHz frequencies
5.3 A 40 MHz-BW 12-Bit CT DSM with Capacitive Local Feedback … 99

Loop Filter with


Capacitive Local Feedback
Cf

Cint1 Cint2 Cint3


VCO_Based_Phase_Quantizer
VINN
R1 R2 R3
D_OUT

VINP R1 R2 R3 1
z
First_Order
Cint1 Cint2 Cint3 Difference

CLK
Cf
LATCH
DAC3 Implicited
D Q
(HRZ) DEM

CLK
CLK_B
DAC2
(NRZ)
LATCH
Q D

DAC1
(NRZ) CLK_B

Fig. 5.16 Block diagram of the 40 MHz 12-bit CT DSM

5.3.1 System Design of the 40 MHz 12-Bit CT DSM

5.3.1.1 System Parameter Considerations

The noise budgets of wideband DSMs are typically designed such that the thermal
noise is dominant. To ensure this, the signal-to-quantization-noise ratio (SQNR) is
typically chosen to be 10–15 dB higher than the target specification [8]. In theory,
many combinations of OSR, number of quantization bits and noise shaping order
can reach the required SQNR. The sampling frequency is limited by the speed of
the digital gates and the OTA GBW. With a lower clock frequency, the modulator is
less sensitive to clock jitter. This is important from the perspective of the application
and measurements. For stability reasons, the order of noise shaping of a single-loop
low-pass DSM should not be over 5th [5]. Multi-stage noise shaping (MASH) on
the other hand can achieve a higher order of noise shaping and still maintain the
modulator stability. But the matching between the analog circuits and the digital
noise cancellation filters is a problem, especially for CT DSMs. A higher number
of quantization bits means less quantization noise, which improves the modulator
stability and the clock jitter sensitivity if NRZ feedback DACs are used. Normally
it should not exceed 5 bits because of the resulting digital complexity and power
consumption. Within these boundary constraints, the architecture with 12× OSR,
5-bit quantization and 4th-order noise shaping is chosen for its robust operation and
the relatively moderate requirements on the building blocks.
100

Table 5.1 Measurement summary of the CT DSM ADC with digital calibration and shaped SC DAC
Specification Value
Signal bandwidth 40 MHz
Sampling frequency 0.96 GHz
Peak SNR 71.7 dB
Peak SNDR 70.8 dB
Power 69.6 mW
FoM1 0.31 pJ/Step
FOM2 158.4 dB
Core area 0.28 mm2
Technology UMC 90 nm SP
5 CT DSM ADCs with VCO-Based Quantization
5.3 A 40 MHz-BW 12-Bit CT DSM with Capacitive Local Feedback … 101

5.3.1.2 Architecture of the 40 MHz 12-Bit CT DSM

The architecture of the CT DSM is shown in Fig. 5.16. To make use of the digital cir-
cuit potential in accuracy and power efficiency in advanced CMOS technologies, the
1st-order noise shaping and 5-bit quantization are implemented by a power-efficient
VCO-based quantizer [10], which is composed of a 31-stage VCO and 31 slices of
frequency and phase detectors. The remaining 3rd-order noise shaping is done by an
analog filter based on active-RC integrators, which have better linearity than Gm-C
integrators. In the feedforward loop filter, a noise-free capacitive local feedback is
used to generate a pair of complex NTF zeros to suppress the quantization noise
further. A direct-feedforward (DFF) path from the modulator input to the quantizer
input is applied to keep the output signal swing of each integrator within 100 mV. This
deteriorates the anti-aliasing filtering at high frequencies, but due to the integration
performed by the VCO, 1st-order anti-aliasing filtering is still maintained in this CT
DSM. The feedback DAC signals are delayed by half a clock period to absorb the
excess loop delay from the quantizer to the resample flip-flop (about 210 ps without
parasitics). An extra RZ DAC is added to feed back the VCO digital frequency. This
is needed to compensate for the explicit loop delay of half a clock cycle. Compared
to using a full clock cycle delay in the feedback path, in our configuration more phase
lag can be tolerated in the feedback DACs and the loop filter. As a result, a smaller
OTA GBW is required to reach the same performance. The half-period-delayed archi-
tecture also makes the modulator more robust to other nonidealities of the analog
parts. For instance, the 3rd DAC delay (6% of the period for the one-period-delayed
architecture versus 10% for the half-period-delayed one) and the integrator time
constant negative variation (−10% for the one-period-delayed architecture versus
−13% for the half-period-delayed one) are relaxed. As no analog filter is preceding
the 1st DAC, its matching should not be worse than the DSM linearity requirement.
For the 2nd DAC, as its nonidealities are suppressed by the 3rd-order analog filter,
5% mismatch can be tolerated for 12-bit linearity. The intrinsic dynamic element
matching (DEM) of the VCO-based quantizer and the low signal swing further relax
the accuracy requirements for the 3rd DAC. Behavioral simulations show that its
nonlinearity has almost no effect on the whole modulator performance.

5.3.2 Circuit Design of the DSM Building Blocks

5.3.2.1 Capacitive Local Feedback

The schematic of the loop filter with capacitive local feedback is shown in blue dashed
box in Fig. 5.16. The 1st OTA, the feedback capacitor Cf and the 1st integration
capacitor Cint1 form a capacitive attenuator with a gain of Cf /Cint1 . The output of
the 3rd integrator is attenuated by this gain and fed into the 2nd integrator to build a
resonator, generating two complex NTF zeros. Compared with the traditional local
feedback implementation, the noise-free capacitor Cf is proposed to replace a noisy
102 5 CT DSM ADCs with VCO-Based Quantization

resistor. Because of the small feedback coefficient, the value of the feedback capacitor
Cf is small compared to the integration capacitor Cint3 , requiring very limited area
and power consumption.

5.3.2.2 Current-Sharing Feedforward OTA

In DSMs with high-order noise shaping, the integrator OTAs are the most power-
hungry blocks. For example, they take 52% of the total power consumption in [10].
The thermal noise contribution of the OTAs is a small fraction of the noise bud-
get. So in order to implement low-power high-order DSMs, it is highly needed to
design power-efficient OTAs. In feedforward compensated OTAs [6], the amplifica-
tion is split into a high-gain path and a high-speed path to avoid the trade-off conflict
between gain and bandwidth. Since all poles are within the GBW and no compen-
sation capacitor is used, more power can be saved compared to the Miller OTA.
The schematic of the 4-stage current-sharing feedforward (CS FF) OTA is shown in
Fig. 5.17. At the outputs of the 4 cascade stages, 4 poles are generated. To ensure the
OTA stability, 3 FF paths are introduced from the input to the different stage outputs
to generate 3 left-half-plane zeros that cancel or compensate the poles. Compared to
previously published FF-compensated OTAs, the FF transconductances (Gms) are
implemented by PMOS differential pairs, which can share their bias currents with
the cascade Gms.
The CS technique improves the OTA design in three aspects. Because of the current
sharing between the FF Gms and the cascade Gms, less current is consumed to reach
the same GBW compared to the traditional FF OTA. The FF Gms and cascade Gms
load each other and less current is used in the last three branches. As a result, a larger
small-signal impedance is obtained in each stage and the amplifier gain is improved
by more than 10 dB. Finally, differential pairs but not simple PMOS current sources
connect the outputs and the power supply, so the isolations between these nodes are
better. Besides the higher gain, the power supply rejection ratio (PSRR) is improved
by 13 dB, making it more robust to noise coupled from the digital part. In the 3rd
stage the transconductance difference between the FF Gm and the cascade Gm is too
large, so extra NMOS bias transistors are added to provide more bias current for the
FF Gm, avoiding pushing the PMOS transistors into the subthreshold region. Thanks
to the DSM feedforward architecture and the VCO integration, the output voltage
swing of each OTA is small enough (<100 mV) to guarantee having no nonlinearity
penalty due to the stacked transistors. The SPICE simulation results in Fig. 5.18
show that, with 1.3 pF load and 8 mW power consumption from a 1.2 V supply, the
CS FF OTA reaches 66 dB DC gain and 3.47 GHz GBW. The phase margin and
input-referred noise are 66.2◦ and 18 Vrms (from 1 kHz to 40 MHz) respectively.
Since non-idealities are shaped by the preceding stages, the size and power of the
other OTAs are scaled by 50%.
5.3 A 40 MHz-BW 12-Bit CT DSM with Capacitive Local Feedback … 103

Vdd
Vdd Vdd
Feedforward Gm Vdd

VCMFB2 VCMFB3 VCMFB4


VCMFB1
VINP VINN VINP VINN

VBP1

VOUTN VOUTP

VINN VINP VBN

VBIAS VBIAS VBIAS VBIAS

Cascade Gm

Fig. 5.17 Schematic of the current-sharing feedforward OTA

AC Response of the 2nd OTA

50
150

Phase (Deg)
Gain (dB)

Gain
Phase
0
100

1 2 3 4 5 6 7 8 9 10
10 10 10 10 10 10 10 10 10 10
Frequency (Hz)

Fig. 5.18 AC simulation results of the current-sharing feedforward OTA

5.3.2.3 Feedback DACs

In this work, a current-mode NRZ DAC [18] consisting of an array of cascode current
sources is used as the main feedback DAC. For the current source transistors, the
overdrive voltage is set to a large value to reduce the noise current. A large area
(18 µm2 ) is used to achieve a good intrinsic matching. In the layout design, double-
centroid layout and tree-structure connections are adopted to cancel the mismatch
further. For the cascode transistors, a gate length of two times the minimum length
is used to reach a higher output impedance and an effective glitch attenuation from
the noisy switching node to the quiet current-source transistor. Low-swing (400 mV)
high-crossing driving signals are used to obtain a small glitch energy; they also put the
switch transistors in saturation, increasing the DAC impedance further. Simulations
show that the input-referred DAC noise (noise of the input resistors included) is
104 5 CT DSM ADCs with VCO-Based Quantization

Fig. 5.19 Die photo of the low-power 40 MHz 12-bit CT DSM

about 72 Vrms, and the output impedance and output parasitic device capacitor of
the whole DAC are about 21 k and 66 fF respectively. In the 2nd and 3rd DAC, the
sizes of the current-source transistors are scaled largely due to their relaxed impact
on the modulator noise and matching. Full-swing driving signals are adopted for
high-speed operation and small output parasitic capacitors.

5.3.3 Measurement Results and Discussions

This modulator has been fabricated in UMC 90 nm SP CMOS, with a 0.14 mm2 core
area. The micro photo of the die is shown in Fig. 5.19. The core is surrounded by
MOS decoupling capacitors. In Fig. 5.20 the power spectral density (PSD) of single-
tone measurements is presented. Because of the residual mismatch in the main DAC,
some harmonic distortions appear and limit the performance. The peak SFDR values
of the 938 kHz and 7.5 MHz measurements are 78.3 dB and 77.2 dB respectively.
Figure 5.21 shows the relationship between the input power and the SNR/SNDR.
The peak SNDR values are 69.7/68.4 dB and the DRs are 74.4/71.9 dB for a 938 kHz
and a 7.5 MHz input frequency. A two-tone test with a 10 MHz and a 10.5 MHz input
is given in Fig. 5.22. A peak IM3 of 75.1 dB is obtained with −9.5 dBFS input. The
power consumption of the CT DSM is 45 mW, 62% of which is used for the OTAs and
the DACs. The corresponding FoM is 0.22 and 0.26pJ/Step (Power/(2BW*2ENOB ))
5.3 A 40 MHz-BW 12-Bit CT DSM with Capacitive Local Feedback … 105

938KHz,−3dBFS 7.5MHz, −3dBFs


0 0

−20 −20

−40 −40
PSD (dB)

PSD (dB)
−60 −60

−80 −80

−100 −100

−120 −120
5 5
10 10
Frequency (Hz) Frequency (Hz)

Fig. 5.20 Measured spectra for a −3 dBFS 938 kHz (left) and a 7.5 MHz (right) single-tone mea-
surement

Fin=938KHz Fin=7.5MHz

70 SNDR 70 SNDR
SNR SNR
60 60
SNR/SNDR (dB)

SNR/SNDR (dB)

50 50

40 40

30 30

20 20

10 10

0 0
−80 −60 −40 −20 0 −80 −60 −40 −20 0
Normalized Input Signal(dB) Normalized Input Signal(dB)

Fig. 5.21 Measured SNR/SNDR versus input amplitude for a 938 kHz input signal (left) and a
7.5 MHz input signal (right)

respectively for the two input frequencies. The CT DSM performance is summarized
in Table 5.2. At the end of Chap. 8, after having presented the final ADC design, this
work is compared to the published state-of-the-art DSM ADCs.
106 5 CT DSM ADCs with VCO-Based Quantization

Fig. 5.22 Spectrum of


two-tone measurement
(−9.5 dBFS, 10 MHz and
10.5 MHz input signals)

Table 5.2 Measurement summary of the CT DSM ADC with capacitive local feedback
Specification Value
Signal bandwidth 40 MHz
Sampling frequency 0.96 GHz
Peak SNR 70.5 dB
Peak SNDR 68.4 dB
Power 45 mW
FoM1 0.26 pJ/Step
FoM2 157.9 dB
Core area 0.14 mm2
Technology UMC 90 nm SP

5.4 Conclusions

In this chapter, the full design procedures and measurements results of two 40 MHz-
bandwidth 12 bit CT DSM ADCs with phase-type VCO-based quantizer have been
presented. In the first design, the SC DAC peak current is reduced by a cost-efficient
shaping technology, and the DAC static and dynamic mismatches are cancelled by
a digital LUT-based calibration. 4-stage FF OTAs have been designed to meet the
stringent system requirements caused by the large bandwidth. The power consump-
tion of the second CT DSM is optimized by using a half period delay architecture
and a current-sharing FF OTA. A noise-free capacitive local feedback is proposed to
generate the non-zero NTF zeros.
5.4 Conclusions 107

The measurement results of the two prototypes in 90 nm CMOS show that with a
960 MHz sampling frequency the first CT DSM achieves 72.9 dB SNDR and 84.2 dB
SFDR with 69.6 mW power consumption, while the peak SNDR of the second CT
DSM is 68.5 dB with only 45 mW power. Their FoM is 0.31 and 0.26 pJ/Step, respec-
tively, comparing favourably to other published designs with a high signal bandwidth.
The second design has an improved energy efficiency, but it is still higher than other
DSM designs with lower bandwidth. The next chapter will therefore present a third
design that improves significantly the energy efficiency.

References

1. P.G.R. Silva, L.J. Breems, K.A.A. Makinwa, R. Roovers, J.H. Huijsing, An IF-to-baseband
Sigma-Delta modulator for AM/FM/IBOC radio receivers with a 118 dB dynamic range. IEEE
J. Solid-State Circuits 42(5), 1076–1089 (2007)
2. L.J. Breems, R. Rutten, R.H.M. van Veldhoven, G. van der Weide, A 56-mW continuous-
time quadrature cascaded modulator with 77-dB DR in a near zero-IF 20 MHz band. IEEE J.
Solid-State Circuits 42(12), 2696–2705 (2007)
3. G. Mitteregger, C. Ebner, S. Mechnig, T. Blon, C. Holuigue, E. Romani, A 20-mW 640-MHz
CMOS continuous-time ADC with 20-MHz signal bandwidth, 80-dB dynamic range and 12-bit
ENOB. IEEE J. Solid-State Circuits 41(12), 2641–2649 (2006)
4. F. Gerfers, M. Ortmanns, Continuous-Time Sigma-Delta A/D Conversion: Fundamentals, Per-
formance Limits and Robust Implementations (Springer, Berlin, 2006)
5. Y. Ke, Design of reconfigurable A/D converter for 4G radios. Ph.D. thesis, KU Leuven, Hev-
erlee, Belgium, 2011
6. X. Xing, M. De Bock, P. Rombouts, G. Gielen, A 40 MHz 12 bit 84.2 dB-SFDR continuous-time
Delta-Sigma modulator in 90 nm CMOS, in 2011 IEEE Asian Solid State Circuits Conference
(A-SSCC) (2011), pp. 249–252
7. X. Xing, G. Gielen, 40 MHz-BW continuous-time ΔΣ modulator with capacitive local feed-
back and current-sharing OTA. Electron. Lett. 49(9), 585–586 (2013)
8. M. Bolatkale, L.J. Breems, R. Rutten, K.A.A. Makinwa, A 4 GHz CT ADC with 70 dB DR and
−74 dBFS THD in 125 MHz BW, in 2011 IEEE International Solid-State Circuits Conference
Digest of Technical Papers (ISSCC) (2011), pp. 470–472
9. G. Cauwenberghs, G.C. Temes, Adaptive digital correction of analog errors in MASH ADCs I.
Off-line and blind on-line calibration. IEEE Trans. Circuits Syst. II Analog Digit. Sig. Process.
47(7), 621–628 (2000)
10. M. Park, M.H. Perrott, A 78 dB SNDR 87 mW 20 MHz bandwidth continuous-time ΔΣ ADC
with VCO-based integrator and quantizer implemented in 0.13 µm CMOS. IEEE J. Solid-State
Circuits 44(12), 3344–3358 (2009)
11. J. Silva, U. Moon, J. Steensgaard, G.C. Temes, Wideband low-distortion Delta-Sigma ADC
topology. Electron. Lett. 37(12), 737–738 (2001)
12. Z. Chang, D. Haspeslagh, J. Verfaillie, A highly linear CMOS Gm-C bandpass filter with
on-chip frequency tuning. IEEE J. Solid-State Circuits 32(3), 388–397 (1997)
13. M.Z. Straayer, M.H. Perrott, A 12-bits, 10-MHz bandwidth, continuous-time Delta-Sigma
ADC with a 5-bits, 950-MS/s VCO-based quantizer. IEEE J. Solid-State Circuits 43(4), 805–
814 (2008)
14. Z. Li, T.S. Fiez, A 14 bit continuous-time Delta-Sigma A/D modulator with 2.5 MHz signal
bandwidth. IEEE J. Solid-State Circuits 42(9), 1873–1883 (2007)
15. R. Schreier, G.C. Temes, Understanding Delta-Sigma Data Converters (Wiley, New York,
2005)
108 5 CT DSM ADCs with VCO-Based Quantization

16. M. De Bock, X. Xing, L. Weyten, G. Gielen, P. Rombouts, Calibration of DAC mismatch errors
in ΣΔ ADCs based on a sine-wave measurement. IEEE Trans. Circuits Syst. II Express Briefs
60(9), 567–571 (2013)
17. R.G.H. Eschauzier, L.P.T. Kerklaan, J.H. Huijsing, A 100-MHz 100-dB operational amplifier
with multipath nested Miller compensation structure. IEEE J. Solid-State Circuits 27(12),
1709–1717 (1992)
18. T. Wu, C. Jih, J. Chen, C. Wu, A low glitch 10-bit 75-MHz CMOS video D/A converter. IEEE
J. Solid-State Circuits 30(1), 68–72 (1995)
Chapter 6
Two-Step Open-Loop VCO-Based ADC

6.1 Introduction

As seen in Chap. 1, for both wireline and wireless communications, new standards
always require larger data rates and signal bandwidths, e.g. VDSL2 and WLAN
802.11n are expected to extend their analog bandwidths to 30 MHz and 40 MHz
respectively. In the meantime, more and more work is done with mobile devices
powered by battery. Tablets and smart phones with explosive shipments are good
examples. Although CMOS technology scaling is beneficial for the increasing band-
width requirement, more challenges are imposed to the analog and mixed-signal
designs, limiting the achievable accuracy and power efficiency. One solution for
ADC design is to shift more quantization and A/D conversion from the analog and
the voltage domains to the digital and the time domains. In Chap. 4, several new
VCO-based quantizer and ADC architectures (both closed-loop and open-loop ones)
have been presented. The main design challenge is on how to solve the VCO V-F
nonlinearity. In this chapter we explore an alternative new approach.
We combine the two-step conversion principle [1] with a frequency-type VCO-
based quantizer [2] to achieve 1st-order noise shaping and solve the VCO nonlinearity
issue. The nonlinearities of the coarse and the fine VCO-based quantizers are mit-
igated by distortion cancellation and input swing reduction schemes. At the same
time, the intrinsic DEM of the VCO-based quantizer output relaxes the DAC match-
ing requirement greatly. Because of the simple ADC architecture and the highly
digital building blocks, a 42fJ/step ADC figure-of-merit (FoM) is realized.
This chapter covers the full design procedure of this two-step open-loop VCO-
based ADC [3, 4]. First the proposed ADC architecture and the system-level design
are given. The main ADC nonidealities are considered and modelled to guarantee the
ADC performance robustness. Then the circuit design of the frequency-type VCO-
based quantizer and the current-steering DAC together with the passive subtractor
are presented. Finally the measurement results are shown and discussed.
110 6 Two-Step Open-Loop VCO-Based ADC

6.2 Architecture Design of Two-Step Open-Loop


VCO-Based ADC

In Fig. 4.2 in Chap. 4, the structure and behavioral model of a frequency-type VCO-
based quantizer [2] has been shown. First a ring VCO converts the input voltage into
phase information, which is illustrated by the rising/falling edges of the VCO outputs.
Then a sense-amplifier flip flop (SA FF) is used to synchronize each VCO output,
introducing a quantization error in the phase domain. Finally a digital differentiator
(constructed by a TSPC FF and an XOR gate) processes the sampled phase and
outputs digital frequency information. By applying such a simple, highly digital
circuit, not only quantization is done, but also 1st-order noise shaping is obtained.
In addition, there is an intrinsic dynamic element matching (DEM) for the digital
output. However, the VCO’s nonlinear transfer function limits its application in ADC
design. Although the 2nd-order harmonics can be cancelled by a pseudo differential
topology, the VCO 3rd-order nonlinearity is still a performance limitation for designs
with more than 6-bit linearity.
In ADC design, in order to break the exponential increase of the number of
comparators occurring in flash ADCs and to reach higher accuracies, the two-step
A/D conversion topology has been proposed [1]. The drawback of this two-step
architecture is that it finishes one conversion in two clock periods, hence introducing
a delay of one clock period. Here a DAC and an analog subtractor with full accuracy
are required to generate the residual signal. Also a full accuracy analog amplifier is
needed to amplifier the residual signal.
In this section, we combine the two-step conversion principle with a VCO-based
quantizer to achieve 1st-order noise shaping and to solve the VCO nonlinearity issue.
At the same time, the intrinsic DEM of the VCO-based quantizer output relaxes the
DAC matching requirement greatly.

6.2.1 A Two-Step Open-Loop VCO-Based ADC Architecture

6.2.1.1 Intrinsic VCO Nonlinearity Mitigation

The proposed two-step open-loop VCO-based ADC architecture and its illustrative
signal spectra are shown in Fig. 6.1 (VCO-based quantizers have been presented in
Fig. 4.2 in Chap. 4). Its structure and operation principle are similar to a two-step
flash ADC except for two differences. First, highly digital frequency-type VCO-
based quantizers with 1st-order noise shaping and intrinsic DEM are used instead
of simple flash quantizers. Secondly, there is no analog amplifier preceding the fine
quantizer.
6.2 Architecture Design of Two-Step Open-Loop VCO-Based ADC 111

coarse VCO-based quantizer (N bit)


Full input swing, nonlinear

Path 1

DAC Path 2 CLK


CLK (N bit)
D_OUT_C
D_OUT_F
A_IN D_OUT

fine VCO-based quantizer (M bit)


Small input swing, linear
|H(f)| |H(f)| |H(f)| |H(f)| |H(f)|

Fin Freq Fin 3Fin Freq Fin 3Fin Freq Fin 3Fin Freq Fin Freq

2 distortions
cancel each other

Fig. 6.1 Proposed two-step open-loop VCO-based ADC architecture and the VCO nonlinearity
mitigation concept

In the two-step open-loop ADC, the N-bit VCO-based coarse quantizer converts
the analog input into digital bits. With full-swing input voltage, the coarse quan-
tizer’s digital output includes not only the desired signal, but also 1st-order shaped
quantization error and harmonic distortions. A N-bit DAC is used to convert the
coarse quantizer output back to an analog signal, without introducing quantization
noise. By subtracting the DAC output from the ADC input, the residual error of the
coarse quantizer is obtained. Since the subtractor is linear, the noise and distortion
parts of the residual signal are almost the same as those of the coarse output. If the
number of coarse quantization bits is large enough and so the coarse quantization
step is small enough, the voltage swing of the residual signal is much smaller than
the whole ADC’s input signal swing. The analog residual signal is converted further
by the M-bit fine VCO-based quantizer. With its low input signal swing, the fine
VCO-based quantizer transfer function is very linear and its nonlinearity does not
limit the ADC performance. In the fine quantizer, although the input signal swing is
small, because of the integration provided by the VCO and the quantization in the
time domain, all 2 M − 1 quantization steps are used [5] and only a small quantization
error is introduced. Finally, the digital outputs of the coarse and the fine VCO-based
quantizers are summed as the output of the whole ADC.
In this design, the gain of the path 2 in Fig. 6.1 (the DAC, the subtractor and the fine
VCO-based quantizer) is designed to be 1 to match that of the path 1. By doing this, all
nonidealities (quantization error, thermal noise, distortion) of the coarse VCO-based
quantizer are cancelled, and the two-step open-loop VCO-based ADC is equivalent
112 6 Two-Step Open-Loop VCO-Based ADC

to a linear 1st-order Delta-Sigma modulator (DSM) with M-bit quantization. The


noise contributions of the DAC and the VCO in the fine quantizer refer to the ADC
input directly, and they need to be optimized in the circuit-level design.

6.2.1.2 Implicit DAC DEM

For the DAC in Fig. 6.1, its input swing is almost full and there is no gain preceding
it, so its mismatch refers to the ADC input directly. If a simple flash ADC is used as
the coarse quantizer, there is no implicit DEM for the D/A conversion. As a result,
either transistor intrinsic matching or an explicit DEM block is required to achieve
the required linearity. Intrinsic transistor matching means a larger area and a complex
switching scheme, consuming large chip area and power. With an external DEM, not
only extra power consumption and chip area are required, but also extra path delay
is introduced, which affects the two-step ADC performance and sets more stringent
requirements on the other blocks.
In our two-step open-loop ADC, a frequency-type VCO-based quantizer is applied
as coarse quantizer instead of a flash ADC. In such a quantizer, the different phase
outputs of the ring VCO change their state consecutively, and a state change corre-
sponds to a “1” in the frequency detector output. This means that there is an intrinsic
DEM with the quantizer digital outputs [2]. For the DAC design, because of the
intrinsic DEM of the digital input, the DAC cell matching requirement is relaxed
and can easily be reached. With this arrangement, the DAC nonlinearity is no longer
the ADC linearity bottleneck. Furthermore, there is no extra hardware or power
consumed, and no extra delay is introduced in the path.

6.2.1.3 Implicit AAF

Another benefit of using a VCO-based quantizer as the coarse and fine quantizers is
that intrinsic 1st-order anti-aliasing filtering (AAF) is kept in the proposed two-step
open-loop VCO-based ADC. It has been explained in Chap. 4 that there is intrinsic
1st-order AAF with the frequency-type VCO-based quantizer [6]. From the analog
input of the open-loop ADC to its digital output, there are in total three paths, and
in each path there is a VCO-based quantizer preceding another block (DAC, digital
adder). With the VCO-based quantizer as the front-end, the high-frequency signal
component is attenuated by 1st-order filtering in every path, so this AAF is done for
the whole ADC.
To verify the proposed two-step open-loop VCO-based ADC architecture, a
40 MHz-BW 10-bit ADC has been designed and implemented. In this design, both
the coarse and the fine VCO-based quantizers are 5-bit pseudo-differential structures
(composed by 2 4-bit single-end ones). The sampling frequency of the two-step open-
loop VCO-based ADC is 1.6 GHz, with an oversampling ratio (OSR) of 20.
6.2 Architecture Design of Two-Step Open-Loop VCO-Based ADC 113

6.2.2 Nonidealities of the Two-Step Open-Loop VCO-Based


ADC

In a real circuit implementation, there are several nonidealities in each block, such
as gain variation, delay and mismatch. In reality, the ADC clock signal can not be
jitter-free either. To obtain a stable ADC performance and to determine the design
specifications for each building block, important ADC nonidealities should be con-
sidered and included in the behavioral-level simulations.
As the most important error source and design objective in VCO-based ADC
design, the VCO V-F nonlinearity is modelled here by a 4th-order polynomial:

f = f o + kvco Ain + k2 Ain


2
+ k3 Ain
3
+ k4 Ain
4
(6.1)

where the coefficients are calculated reversely from the distortion levels in SPICE
simulation results:
1 k2 1 k3 2 1 k4 3
H D2 = Ain H D3 = Ain H D4 = A (6.2)
2 kvco 4 kvco 8 kvco in

6.2.2.1 Block Gain Variation

In the frequency-type VCO-based quantizer, the quantizer gain is determined by


the gain between the input control voltage and the oscillation frequency. Once the
delay cells are fixed, such gain is decided by the transconductance of the current
source transistor. The gain of a current-steering DAC is set by its bias current. In
a CMOS implementation, both the VCO-based quantizer gain and the DAC gain
can vary as much as [−20%, 20%]. The gain deviation of the coarse and the fine
VCO-based quantizers and the DAC have different mechanisms and effects on the
ADC performance.
From the ADC analog input to the fine quantizer input, there are two paths: one
analog path with a gain of unity, and another one constructed by the coarse quantizer
and the DAC. If the gain of the second path is not exact unity, then the voltage
swing of the residual signal will become larger. With a larger input signal swing, the
linearity of the fine quantizer will be worse. The nonlinearity of the fine quantizer
refers to the ADC input directly. From the coarse quantizer output to the whole ADC
output, there are also two paths: one digital path with unity gain (path 1 in Fig. 6.1),
and the path composed by the DAC and the fine quantizer (path 2 in Fig. 6.1). If
the gains of paths 1 and 2 are not the same, then part of the coarse quantizer output
appears in the ADC output. Although the coarse quantization is also 1st-order noise
shaped, its input swing is almost full swing and its distortion leakage degrades the
ADC linearity performance.
In Fig. 6.2, the sensitivity of the ADC SNDR performance to the block gain vari-
ations are shown. The ADC performance is robust to the coarse ADC gain variation:
114 6 Two-Step Open-Loop VCO-Based ADC

Fig. 6.2 ADC SNDR 64


performance as a function of
the normalized gain of the
62
coarse ADC, fine ADC and
DAC
60

SNDR (dB)
58

56
Coarse ADC
54 Fine ADC
DAC

52
0.9 0.95 1 1.05 1.1
Normalized Gain

within [−10%, 10%] gain variation, the gain variation has almost no effect on the
ADC performance. However, the ADC SNDR is much more sensitive to the gain
variations of the DAC and the fine ADC: smaller than ±5% gain variation windows
are required for the ADC to achieve 60 dB SNDR. Such simulation results show that
the second mechanism is much more severe than the first one. In circuit design, the
transconductances of current source transistors in the coarse and the fine VCOs are
digitally adjustable, and so are their gains. Also the DAC bias current can be tuned
continuously. With these discrete and continuous tunings, the gain of path 2 can be
tuned to be very close to 1.

6.2.2.2 Block Delay

In the frequency-type VCO-based quantizer, the input-to-output and the clock-to-


output delays of all digital gates contribute to the total quantizer digital delay. For
the NRZ current-steering DAC, the finite settling of the output current introduces
time delay.
For the two-step open-loop VCO-based ADC, its performance not only depends
on the block gains, but also on the delay of each block. Since even building block
delay causes time/phase mismatch between two paths, increasing the residual signal
swing and degrading the distortion cancellation. In Fig. 6.3, the relationships between
the ADC SNDR and the delays of the coarse ADC and the DAC are shown. For these
two blocks, less than 0.25·Ts delay is allowed for the ADC to reach 60 dB SNDR.
On the other hand, because a digital signal is robust to time skew, the fine ADC can
tolerate as much as one clock period delay. The reason why the coarse ADC delay
tolerance is smaller is that its digital output needs to be processed further by another
block to generate the analog residual signal, and the analog signal is much more
sensitive to time delay.
6.2 Architecture Design of Two-Step Open-Loop VCO-Based ADC 115

Fig. 6.3 ADC SNDR 65


performance as a function of
the delays of the coarse ADC
and the DAC Coarse ADC
DAC

SNDR (dB)
60

55
0 0.1 0.2 0.3 0.4 0.5
Delay (*Ts)

6.2.2.3 DAC Mismatch

In an N-bit DAC


 design, because of the systematic and random process variations,
the 2 N − 1 DAC cell values can not be exactly the same. This mismatch makes
the DAC transfer function nonlinear. In our design, the DAC nonlinearity is not
suppressed by any gain and appears in the ADC output directly. Fortunately, by
using a frequency-type VCO-based quantizer as the coarse quantizer, the DAC input
is intrinsic DEMed, which greatly relaxes the matching requirement on the DAC
cells. In Fig. 6.4, the ADC SNDR is plotted as a function of the DAC cell variation
standard deviation. For comparison, the ADC SNDR values are also given for the
case of no DEM. In our design with DEM, the DAC cell sigma can be as large as 0.15
unit, which can be realized easily in CMOS technology. If there is no DEM, then
the ADC linearity performance is only around 37 dB, and 0.005 standard deviation
is required to reach 60 dB performance. It is obvious that the proposed application
of the coarse VCO-based quantizer largely relaxes the DAC design.

6.2.2.4 Clock Jitter

Another important nonideality for ADC design is clock jitter. In the proposed two-
step open-loop VCO-based ADC, clock jitter introduces an error in all three blocks. In
the VCO-based quantizer, because the VCO integration is preceding the time-domain
sampling, the clock-jitter-induced noise is 1st-order shaped. On the other hand, the
error caused by the DAC clock jitter is referred to the ADC input directly. In Fig. 6.5,
the ADC performance degradation due to clock jitter is presented. Because of the
NRZ DAC topology and the 5-bit coarse quantization, as high as 15ps clock jitter
can be tolerated by our design to reach 60 dB SNDR.
116 6 Two-Step Open-Loop VCO-Based ADC

Fig. 6.4 ADC SNDR 65


performance as a function of
the standard deviation of the
60
current of the DAC cells. For
comparison, both cases with
and without DEM are 55

SNDR (dB)
showed
wi DEM
50 wo DEM

45

40

35
0 0.05 0.1 0.15
DAC Cell Standard Deviation

Fig. 6.5 ADC SNDR 63


performance as a function of
the clock jitter: around 15ps
62
jitter can be tolerated for
10-bit accuracy
61
SNDR (dB)

60

59

58

57
0 5 10 15 20 25 30 35
Jitter (Ps)

6.3 Circuit Implementation of the Two-Step Open-Loop


VCO-Based ADC

To design the two-step open-loop VCO-based ADC, two VCO-based quantizers, one
DAC and one subtractor are required. Theoretically speaking, since the nonidealities
of the coarse VCO-based quantizer are cancelled, its design (for example phase
noise) can be relaxed. However, in frequency-type VCO-based quantizers, most of
power consumption is from the digital gates and not from the VCO. So it is not so
beneficial to design two different VCO-based quantizers. Based on this fact and to
reduce the design time, the coarse and fine VCO-based quantizers are taken the same
in this design.
6.3 Circuit Implementation of the Two-Step Open-Loop VCO-Based ADC 117

6.3.1 VCO-Based Quantizer Design

To cancel the even-order harmonic distortions, the pseudo-differential 5-bit VCO-


based quantizers are composed of 2 4-bit single-ended quantizers. The schematic of
the 15-stage ring VCO of a 4-bit single-end quantizer is presented in Fig. 6.6, which
is composed of a ring of delay cells and a transconductor. The circuit of the delay cell
is shown in the blue dashed box [7]. It is a pair of current-starved inverters, together
with two more inverters to synchronize their outputs. The oscillation frequency of
the ring VCO is determined by the drain currents of the delay cells, which is gener-
ated by the voltage-controlled current source. To accommodate the corner variations
introduced by circuit fabrication, a 3-bit binary control scheme is adopted to tune
the transconductance of the current source, and hence the VCO gain.
With this topology, the VCO output can not be full swing (it can reach the power
supply but it can not reach the ground), especially when the oscillation frequency is
high. To detect the edge of the VCO phase output correctly, an SA FF with NMOS
input is used as the time-domain S/H. With a 500 mV common-mode voltage and
a 400 mVpp input swing, the maximum oscillation frequency of the ring VCO is
designed to be no larger than half of the sampling frequency (800 MHz) to avoid
overloading the quantizer in the time domain.
The simulated power supply sensitivity of the VCO-based quantizer gain is
13.1%/V, and its temperature sensitivity is 2.1 ‰ (per-mille). In Fig. 6.7, the transistor-
level simulations of the VCO-based quantizer with 0dBFS input power are shown.
Compared to the 4-bit single-ended topology (left), the 5-bit pseudo-differential
topology not only increases the quantization accuracy, but it also cancels the even-
order harmonic distortions. As presented, the SNDR of the 5-bit pseudo differential
VCO-based quantizer is about 37 dB. In Fig. 6.8, the simulation result of the VCO-

15 stages delay cells


Vdd

OUTB
IN INB

OUT

VCTRL
BIAS
WF WT 2WT 4WT
L L L L

D1 D2 D4

Fig. 6.6 Schematic of the 15-stage ring VCO in the quantizer (single-ended); the delay cell circuit
is shown in the blue dashed box
118 6 Two-Step Open-Loop VCO-Based ADC

10 10
Hann−windowed output Hann−windowed output
Integration Noise Integration Noise
0 Integration Noise Distortion 0 Integration Noise Distortion

−10 −10
Power spectrum density(dB)

Power spectrum density(dB)


−20 −20
BW=40.0 MHz BW=40.0 MHz
OSR=20 OSR=20
−30 SNDR=21.7 dB −30 SNDR=36.7 dB
SNR=57.3 dB SNR=61.1 dB
ENOB=3.31 bits ENOB=5.81 bits
−40 −40

−50 −50

−60 −60

−70 −70

−80 −80

−90 −90

−100 6 7 8
−100 6 7 8
10 10 10 10 10 10

Sampling frequency Fs =1600MHz Sampling frequency Fs =1600MHz

Fig. 6.7 Large-signal simulation of a 5-bit frequency-type VCO-based quantizer. Both the single-
ended result (left) and the pseudo-differential result (right) are presented

Frequency domain
0
Hann−windowed output
Integration Noise power
Integration Noise Distortion power
−20 BW=40.0 MHz
OSR=20
Power spectrum density(dB)

SNDR=43.5 dB
SNR=44.5 dB
−40 ENOB=6.93 bits

−20dB/Dec

−60

−80

−100

−120

5 6 7 8
10 10 10 10
Sampling frequency Fs =1600MHz

Fig. 6.8 Small-signal simulation of a 5-bit frequency-type VCO-based quantizer; with around
−20 dBFS input, the 3rd-order distortion is about −76 dB
6.3 Circuit Implementation of the Two-Step Open-Loop VCO-Based ADC 119

Fig. 6.9 The simulated −60


VCO phase noise; it is
−113 dBc/Hz at an offset −70
frequency of 10 MHz

Phase Noise (dBc)


−80

−90

−100

−110

−120

−130
5 6 7 8
10 10 10 10
Frequency (Hz)

based quantizer with small-signal input is presented. When the input is reduced to
−22 dBFS, the VCO-based quantizer is quite linear, and the 3rd-order distortion is
only around −76 dBFS (Fig. 6.8).
The simulated phase noise of the VCO is shown in Fig. 6.9, which is about
−113 dBc/Hz at 10 MHz offset frequency. According to the formula in [8], the cal-
culated signal-to-phase-noise-ratio is about 68 dB. It is clear that this VCO phase
noise does not limit the ADC targeted performance (10bit).

6.3.2 DAC and Subtractor Design

The 5-bit DAC used in the two-step open-loop VCO-based ADC is a NRZ current-
steering DAC, chosen for its high speed and robustness to clock jitter. The detailed
schematic of the DAC together with the subtractor is shown in Fig. 6.10. The basic
DAC cell is composed of a current-source transistor, a cascode transistor and a pair
of switches with full-swing digital input. For the current-source transistor, about
0.64 µm2 active area is used to achieve the required matching (around 8% standard
deviation). For the cascode transistor, 2× minimum gate length is used to reach about
7.5× gain. This gain not only provides sufficient shielding from the noisy switching
node to the quiet bias node, but also increases the DAC output impedance. Because
of the full-swing input, the size of the switch transistors can be minimized, reducing
the coupling effect. With proper transistor sizes and source terminal voltages, even
with a full-swing DAC input, the switch transistor is still in the saturation region,
increasing the DAC output impedance further. To keep the CM voltages for the coarse
and the fine VCO-based quantizers the same, two CM current sources are added to the
DAC, which are PMOS cascode constant current sources. The subtraction operation
120 6 Two-Step Open-Loop VCO-Based ADC

CM Part Vdd Bias

Vinp Voutp
To Fine Quantizer
Vinn Voutn

DAC Cell

Fig. 6.10 Schematic of the current-steering DAC; the resistor connected to the ADC input and the
DAC output forms a passive subtractor

Fig. 6.11 Noise simulation


of the feedback
current-steering DAC and
the passive subtractor
Noise Power (V )
2

−15
10

−16
10

5 6 7
10 10 10
Frequency (Hz)

between the positive and the negative parts of the coarse VCO-based quantizer are
realized by changing the polarity of the DAC’s negative part output.
To avoid active power consumption, a simple passive subtractor is applied in
this design, by connecting resistors between the ADC inputs and the DAC outputs.
The output currents of the current-steering DAC go through the resistors and cause
corresponding voltage drops across the resistors. The ADC differential input voltage
6.3 Circuit Implementation of the Two-Step Open-Loop VCO-Based ADC 121

Coarse Output Fine Output


0 0
SNR=65.1 dB SNR=42.9 dB
−20 −20
SNDR=37.3 dB SNDR=17.6 dB
PSD (dB)

PSD (dB)
−40 −40
−60
−60
−80
−80
−100
−100
−120
6 7 8 6 7 8
10 10 10 10 10 10
Frequency (Hz) Frequency (Hz)

Total Output
0
SNR=66.8 dB
−20
SNDR=61.0 dB
PSD (dB)

−40

−60

−80

−100
6 7 8
10 10 10
Frequency (Hz)

Fig. 6.12 Transistor-level simulation result of the two-step open-loop VCO-based ADC. Top left
PSD plot of the coarse quantizer output; top right PSD plot of the fine quantizer output; bottom left
PSD plot of the overall ADC output

minus the voltage drop results in the residual signal for the fine VCO-based quantizer.
As the load of the current-steering DAC, the resistor value is 500 , which means that
the DAC full-scale current is 0.4 mA, and the DAC cell current is around 26.7 µA.
Noise simulation results are presented in Fig. 6.11, showing that the RMS (root-
mean-square) output noise voltage of the current-steering DAC and the subtractor is
about 54 V, which corresponds to about −77 dBFS.
With the above designed building blocks, the two-step open-loop VCO-based
ADC is constructed and simulated in Cadence APS. The simulation results are shown
in Fig. 6.12. In the top left, the output PSD of the coarse quantizer is shown: 65.1 dB
SNR is achieved, but the SNDR is only 37.3 dB. The PSD of the fine quantizer output
is given in the top right: compared to the coarse quantizer output, the signal energy
is 20 dB lower and the 3rd-order distortion energy is the same. In the whole ADC
output (PSD shown in bottom left), the distortions of the coarse and the fine quantizer
outputs cancel each other and 61dB SNDR is obtained.
122 6 Two-Step Open-Loop VCO-Based ADC

6.4 Experimental Results and Discussions

The proposed two-step open-loop VCO-based ADC has been fabricated in TSMC
40 nm general purpose (GP) CMOS. The die photo is given in Fig. 6.13; the build-
ing blocks are marked by red rectangles. Because of the simple and highly digital
architecture, only 0.017 mm2 core area is utilized. To clearly illustrate the two-step
conversion idea, both the digital outputs of the coarse and the fine VCO-based quan-
tizers are converted into binary codes and fed to digital buffers and a low-voltage
differential signal (LVDS) interface. The remaining space is occupied by on-chip
decoupling capacitors (for power supplies and bias voltages) and bonding pads.
To measure the two-step open-loop VCO-based ADC, analog sine voltages with
12 MHz frequency, 500 mV offset voltage and variable amplitude are used as the ADC
input. The measured results with −1 dBFS input power are shown in Fig. 6.14. In the
top left of Fig. 6.14, the power spectral density (PSD) of the coarse quantizer output
is plotted, the 20 dB/decade roll-off characteristic provided by the VCO inherent
integration is evident. With almost full input swing and the nonlinear VCO transfer
function, the 3rd-order distortion appears, and the linearity of the quantizer is about
5-bit. The actual matching between the positive and the negative parts of the pseudo
differential quantizer is limited in practice, so part of the 2nd-order distortion still
remains. In the top right of Fig. 6.14, the PSD of the fine quantizer output is shown.
Here the signal power is about 20dB lower than that of the coarse quantizer output,
and the distortions are accurately the same as is the coarse quantizer output. As
presented in the simulation result of Fig. 6.8, with such a small input power, the
transfer function of the fine VCO-based quantizer is very linear. So the distortion
of the fine quantizer output is from the residual signal itself and not from the fine

Fig. 6.13 Die photo of the


two-step open-loop
VCO-based ADC
6.4 Experimental Results and Discussions 123

Coarse Output Fine Output


0 0

−10
−20 −20

−30
SNR=36.3 dB
SNR=59.8 dB −40
−40 SNDR=11.1 dB
SNDR=32.2 dB

PSD (dB)
PSD (dB)

−50
−60
−60
−70
−80
−80
−90
−100
−100
−110
−120
5 6 7 8 5 6 7 8
10 10 10 10 10 10 10 10
Frequency (Hz) Frequency (Hz)

Total Output
0
−10
−20
−30
SNR=60.7 dB
−40
SNDR=59.5 dB
PSD (dB)

−50
−60
−70
−80
−90
−100
−110
5 6 7 8
10 10 10 10

Frequency (Hz)

Fig. 6.14 Measured results of the two-step 1st-order VCO-based DSM with −1 dBFS 12 MHz sine
input. Top left PSD plot of the coarse quantizer output; top right PSD plot of the fine quantizer
output; bottom left PSD plot of the overall ADC output

62

SNDR
60 SNR

58
SNR/SNDR (dB)

56

54

52

50

48

46

44

−18 −16 −14 −12 −10 −8 −6 −4 −2

Normalized Input Signal(dB)

Fig. 6.15 The measured ADC SNR/SNDR as a function of the analog input power

quantizer. Finally, by adding the coarse and fine quantizer outputs together, their
distortions cancel each other. As a result, 59.5 dB SNDR and 67.6 dB spurious-free
dynamic range (SFDR) are achieved for the whole ADC, as shown in the bottom
left of Fig. 6.14. The relationship between the measured ADC SNR/SNDR and the
analog input power is given in Fig. 6.15.
The power consumption of the two-step VCO-based ADC is about 2.57 mW,
most of which is from the digital gates (FF, XOR). Because of the simple open-loop
architecture and the highly digital building blocks, an excellent FoM of 42fJ/Step
124 6 Two-Step Open-Loop VCO-Based ADC

Table 6.1 Measurement Specification Value


summary of the two-step
open-loop VCO-based ADC Signal bandwidth 40 MHz
Sampling frequency 1.6 GHz
Peak SNR 60.7 dB
Peak SNDR 59.5 dB
Power 2.57 mW
FoM1 42fJ/Step
FoM2 161.4 dB
Core area 0.017 mm2
Technology TSMC 40 nm GP

((Power/(2BW∗2 E N O B ))) or 161.4 dB (SNDR+10log(BW/Power)) is achieved, indi-


cating the high energy efficiency of this structure. Since the sampling frequency and
the power consumption are limited by digital blocks, it is expected that with a more
advanced CMOS technology, a higher ADC bandwidth (or higher ADC accuracy)
and an even lower FoM can be achieved. The measurement results of the two-step
open-loop VCO-based ADC are summarized in Table 6.1, and will be compared to
the state-of-the-art works at the end of Chap. 8, after having presented the final ADC
design.

6.5 Conclusions

A frequency-type VCO-based quantizer has several advantages and is very suitable


for ADC design, offering high power efficiency, intrinsic 1st-order noise shaping and
implicit DEM. To utilize these merits and to mitigate the VCO nonlinearity, a two-
step open-loop VCO-based ADC architecture has been proposed and implemented
in this chapter. With compact open-loop structure and highly-digital building blocks,
a stable performance, a high bandwidth, a very low FoM and a small area can easily
be achieved. The nonlinearities of the coarse and the fine VCO-based quantizers are
solved by distortion cancellation and input swing reduction schemes respectively.
The DAC matching is relaxed greatly as its input has intrinsic DEM.
In this chapter the effects of the different nonidealities have been investigated in
detail, based on which the design specifications for each building block have been
determined. Then the full circuit design and the implementation of the two-step open-
loop VCO-based ADC in 40 nm CMOS has been presented. Simulation results of all
building blocks and the whole ADC have been given in detail. Measurement results
show that the proposed ADC achieves a SNDR of 59.5 dB and a SFDR of 67.6 dB
for a 40 MHz bandwidth, with a 1.6 GHz sampling frequency and only 2.57 mW
power consumption. The corresponding excellent FoM (42fJ/Step) is one of the best
among high-bandwidth (≥20 MHz) DSM ADCs. This VCO-based ADC design has
6.5 Conclusions 125

achieved a high-digital architecture and thus a outstanding power efficiency, however,


the first-order noise-shaping limits its accuracy performance. Therefore, in the next
chapter, we will present a second-order VCO-based ADC design, targetting better
accuracy and even better power efficiency.

References

1. H. van der Ploeg, R. Remmers, A 3.3-V, 10-b, 25-MSample/s two-step ADC in 0.35-µm CMOS.
Solid-State Circuits, IEEE J. 34(12), 1803–1811 (1989)
2. M.Z. Straayer, M.H. Perrott, A 12-bits, 10-MHz bandwidth, continuous-time Delta-Sigma ADC
with a 5-bits, 950-MS/s VCO-based quantizer. IEEE J. Solid-State Circuits 43(4), 805–814
(2008)
3. X. Xing, P. Gao, G. Gielen, A 40 MHz-BW two-step open-loop VCO-based ADC with 42fJ/step
FoM in 40 nm CMOS, in Proceedings of IEEE ESSCIRC, (Bucharest, Romania, 2013), pp. 327–
330
4. X. Xing, G.G.E. Gielen, A 42fJ/Step-FoM two-step VCO-based Delta-Sigma ADC in 40 nm
CMOS. Solid-State Circuits, IEEE J. 50(3), 714–723 (2015)
5. M. Park, M.H. Perrott, A 78 dB SNDR 87 mW 20 MHz bandwidth continuous-time ΔΣ ADC
with VCO-based integrator and quantizer implemented in 0.13 µm CMOS. Solid-State Circuits,
IEEE J. 44(12), 3344–3358 (2009)
6. K. Reddy, S. Rao, R. Inti, B. Young, A. Elshazly, M. Talegaonkar, P.K. Hanumolu, A 16-mW
78-dB SNDR 10-MHz BW CT ΔΣ ADC using residue-cancelling VCO-based quantizer. Solid-
State Circuits, IEEE J. 47(12), 2916–2927 (2012)
7. G. Taylor, I. Galton, A mostly-digital variable-rate continuous-time Delta-Sigma modulator
ADC. Solid-State Circuits, IEEE J. 45(12), 2634–2646 (2010)
8. J. Kim, T.K. Jang, Y.G. Yoon, S.H. Cho, Analysis and design of voltage-controlled oscillator
based analog-to-digital converter. IEEE Trans. Circuits Syst. I Regul. Pap. 57(1), 18–30 (2010)
Chapter 7
VCO-Based 0-ΔΣ MASH ADC

7.1 Introduction

As discussed in Chap. 4, to address the issue of limited voltage headroom and to


take advantage of the reduced CMOS gate delay in scaled CMOS technologies,
VCO-based highly-digital ADCs have recently emerged in recent years. However,
the performance of VCO-based ADCs is limited by the inherent nonlinearity of the
VCOs. As shown in Fig. 7.1, a VCO-based ADC can be utilized as a quantizer in
a single-loop ΔΣ modulator. The preceding loop filter suppresses both the quanti-
zation noise and the nonlinearity, however, a high loop gain results in a high power
consumption and potential instability problems. As discussed in Chap. 3, in tradi-
tional ΔΣ ADC design, compared to single-loop structures, MASH architectures
can offer a higher-order noise shaping with less instability concerns. This structure
can also be implemented with VCO-based quantizers as will be demonstrated.
In this chapter, the VCO-based MASH ΔΣ ADC architecture will be investigated.
Specially, an open-loop quantizer without preceding filters, or so-called a zero-order
quantizer, can be adopted in the stages of the MASH modulator. With VCO-based
quantizers, the 0-ΔΣ MASH topology can overcome the VCO nonlinearity problem
due to its inherent operation principle, being more power efficient than the single-loop
structure. Section 7.2 will analysis the architecture and the nonlinearity-cancellation
technique, with a 0–2 MASH VCO-based ADC as an example [1, 2]. The system-
level design and the nonlinearity-cancellation robustness against gain variations will
be investigated. Furthermore, the delay matching technique for the nonlinearity can-
cellation will be presented. Section 7.3 will cover the circuit design of the 0–2 MASH
VCO-based ADC building blocks, including the three-input adder, the VCO-based
quantizer, the analog integrator, the feedback DAC and the digital interface. The
measurement results in 40-nm CMOS and discussions will be given in Sect. 7.4. At
last, the conclusions will be provided in Sect. 7.5.
128 7 VCO-Based 0-ΔΣ MASH ADC

L0
L1

Fig. 7.1 VCO-based quantizer in a single-loop ΔΣ modulator

7.2 Architecture Analysis of the 0-ΔΣ MASH VCO-Based


ADC

7.2.1 0-ΔΣ MASH VCO-Based ADC

A 0-order stage is an open-loop quantizer without preceding filters. By replacing


flash quantizers with VCO-based quantizers, a 0-ΔΣ VCO-based MASH structure
is shown in Fig. 7.2. Several advantages are obtained by adopting this topology.

Fig. 7.2 Block diagram of a 0-ΔΣ MASH VCO-based ADC


7.2 Architecture Analysis of the 0-ΔΣ MASH VCO-Based ADC 129

Firstly, the VCO-based quantizer has an intrinsic 1st-order noise shaping, as evi-
denced by the 1st-order difference function shown in the behavioral model. Different
from other 0-ΔΣ MASH ADCs, e.g., [3–5], which use a flash or a SAR ADC as
the 1st-stage quantizer, the first stage of the VCO-based MASH ADC is no longer
of zero order. Therefore, any noise leakage due to unperfect matching is 1st-order
shaped. As the 1st-order first stage can not provide additional noise shaping for the
whole ADC, we still call it 0-ΔΣ MASH structure for simplicity.
Secondly, due to its mostly-digital structure, the VCO-based quantizer is extremely
power-efficient. Although the final NTF order is only determined by the second stage,
the power consumption of the first stage is reduced.
Thirdly, the VCO-based quantizer has an implicit dynamic element matching
(DEM) which relaxes the matching requirements of the DAC cells, and an intrinsic
1st-order anti-aliasing filter, which is lost in the traditional zero-order first stage.

(a)

(b)

Fig. 7.3 The nonlinearity-cancellation principle in a 0-ΔΣ MASH VCO-based ADC


130 7 VCO-Based 0-ΔΣ MASH ADC

Most importantly, the VCO nonlinearity problem can easily be solved in this
structure. Similar to the quantization noise, the nonlinearity of the first stage can also
be cancelled in the cascaded structure [6].
Figure 7.3 illustrates the nonlinearity-cancellation principle, where node A in
Fig. 7.3b corresponds to the distortion input position in Fig. 7.3a. For modeling sim-
plicity, H(s) represents the transfer function for the loop filter, while the quantizer
and the DACs are represented by blocks with unity gain.
In order to cancel out the distortion, a negative path with gain equal to −1
needs to be created. Particularly, an input-feedforward topology in the second stage
can be used to ensure this requirement. The negative path is a conventional input-
feedforward ΔΣ loop of which the STF is equal to 1. When the outputs from the
positive and the negative paths are summed together, the input distortion is cancelled
out thoroughly.
In addition, as the second stage only processes the noise-shaped quantization
noise, the input-voltage swing is greatly reduced, resulting in a linear second-stage
quantizer [7, 8]. Therefore, the overall output can be expressed as:

V = H1 · U + (H1 − ST F2 ) · [(1 − z −1 )E 1 + D1 ] − N T F2 · E 2 (7.1)

where H refers to the stage transfer function, E indicates the quantization noise and
D represents all the nonlinear harmonics. Clearly, if the digital filter H1 = ST F2
is adopted, then both the quantization noise and the nonlinearity of the first-stage
modulator can be eliminated. As discussed above, the simplest choice to satisfy the
condition of H1 = ST F2 is to set ST F2 = 1.
Figure 7.4 illustrates the signal model of the 0-ΔΣ MASH VCO-based ADC,
where E 1 and E 2 represent the corresponding quantization noise of the two quan-
tizers, K Q1 and K Q2 indicate the quantizer gains, D Q1 represents all the nonlinear
harmonics generated in the first quantizer due to VCO’s nonlinear behavior, H(s)
represents the transfer function of the analog integrator, K D AC,I indicates the inter-
nal DAC in the feedforward path and K D AC,G indicates the global DACs connected
to the virtual ground of the integrator. Under perfect matching of H1 = ST F2 , E 1

Fig. 7.4 Equivalent signal model of the proposed 0-ΔΣ MASH VCO-based ADC
7.2 Architecture Analysis of the 0-ΔΣ MASH VCO-Based ADC 131

and D Q1 from the 1st-stage quantizer are fully cancelled out at the output, while the
quantization noise of the second quantizer is shaped by the analog integrator and the
digital differentiator.

7.2.2 Nonlinearity-Cancellation Robustness Against PVT


Variations

Unfortunately, in practice, due to the process-voltage-temperature (PVT) variations,


it is difficult to satisfy the matching condition of H1 = ST F2 , which then degrades
the performance of the nonlinearity cancellation.
Observed from the signal model in Fig. 7.4, all the nonlinear harmonics appearing
at the final ADC output can be expressed as:

K D AC,G K Q2 H (s) + K D AC,I K Q2


D Q1 K Q1 (1 − )
1 + K D AC,G K Q2 H (s)
1 − K D AC,I K Q2
=D Q1 K Q1 (7.2)
1 + K D AC,G K Q2 H (s)

This shows that the cancellation is mainly dependent on the product of the internal
DAC gain and the 2nd-stage quantizer gain, and is insensitive to the gain variation
of the first VCO and the global DAC. Significantly, the variation of K D AC,I and K Q2
is shaped by the loop filter H(s).
MATLAB behavioral simulations have been performed to compare the perfor-
mance degradation due to any stage-gain mismatch in the 0–1 MASH and 0–2 MASH
structures respectively. The results are shown in Fig. 7.5. In the simulations, both 5-
bit/5-bit quantizers clocked at 1.6 GHz and a 40-MHz-bandwidth input are used. It
shows that the 0–2 MASH VCO-based structure is much more insensitive to gain
mismatch than the 0–1 MASH structure. In addition, since the 0–1 MASH structure
merely offers 1st-order noise shaping, only a moderate SNDR can be achieved with
a reasonable sampling frequency. The power consumption may increase due to the
additional integrator and feedback DAC compared to the 0-1 MASH structure. How-
ever, by using a low-power design technique at the circuit level, designers can gain
more from the higher resolution and robustness, resulting in a better overall FoM.
Furthermore, to investigate the mismatch effect of different blocks, Fig. 7.6 illus-
trates the SFDR performance of a 0–2 MASH VCO-based ADC as a function of the
normalized gain of the quantizers and DACs. The ±10% gain variations of the first
quantizer and the global DAC have no evident effect on the SFDR performance. In
contrast, the SFDR degrades with the gain variations of the second quantizer and the
internal DAC. Thanks to the 1st-order high-pass shaping, the proposed ADC is still
robust and achieves 78-dB SFDR with ±10% gain variations.
132 7 VCO-Based 0-ΔΣ MASH ADC

Fig. 7.5 SNDR of the 0–1 MASH and 0–2 MASH VCO-based ADCs as a function of the stage-gain
mismatch

Fig. 7.6 ADC SFDR performance as a function of the normalized gain of the VCOs and DACs

7.2.3 System Architecture of a 0–2 MASH VCO-Based


ΔΣ ADC

The proposed VCO-based 0–2 MASH ΔΣ ADC is shown in Fig. 7.7. The digital filter
H1 will be illustrated in detail in Fig. 7.10 later on. Nonlinearity exists in different
blocks, but it is mitigated in this design at both architecture and circuit levels.
7.2 Architecture Analysis of the 0-ΔΣ MASH VCO-Based ADC 133

Fig. 7.7 The presented 0–2 MASH VCO-based ΔΣ ADC

Inevitably, the first-stage VCO’s nonlinear V-to-F function can introduce harmon-
ics. The low-swing voltage input results in a linear second-stage VCO. As a result, the
output of the second quantizer contains harmonics derived from the previous stage
rather than from the quantizer itself. The harmonics have the same amplitude levels
as the first-stage output but with the opposite phases. When the two stage outputs are
added together, these undesired distortions are neutralized. Compared to this design,
although the second-stage output in [9] is free from distortion, no cancellation exists
to improve the first-stage quantizer’s linearity.
Secondly, the integrator opamp only processes quantization noise. In addition, the
feedforward path can also reduce the output swing of the opamp. The distortion of
the loop filter is therefore not the ADC performance bottleneck.
Finally, the intrinsic DEM in the VCO quantizer for the digital output greatly
improves the DAC’s linearity. Consequently, considering that the VCO-based quan-
tizer has an inherent 1st-order noise shaping capability, the proposed cascaded VCO-
based ADC is equivalent to a linear 2nd-order MASH ΔΣ modulator.
In the practical design implementation, different circuit non-idealities exist and
have an impact on the modulator performance. The gain mismatch due to PVT varia-
tions has been discussed in Sect. 7.2. As discussed in Chap. 3, clock jitter introduces
a sampling error in the VCO-based quantizers. Under perfect matching of H1 ST F2 ,
the jitter-induced sampling noise of the 1st-stage quantizer is cancelled out at the
output, while the jitter-induced sampling noise of the 2nd-stage quantizer is reduced
by the preceding integrator. However, the clock jitter of the global DACs generates
errors which are referred to the input directly. According to the behavioral-level sim-
ulations, the distortion-cancellation performance is robust to clock jitter, while the
134 7 VCO-Based 0-ΔΣ MASH ADC

Fig. 7.8 SNDR performance as a function of the clock jitter

ADC noise floor degrades because of the jitter-induced noise in the global DACs. As
shown in Fig. 7.8, to reach a 12-bit linearity, the clock jitter must be limited to 4 ps.

7.2.4 Delay Matching Technique

The impact of the circuit delay will be investigated in this subsection. As discussed in
Sect. 7.2, the nonlinearity cancellation requires H1 = ST F2 , which can be simplified
to ST F2 = 1. However, for a VCO-based quantizer, the VCO’s integration operation
requires one clock period, meaning that the delay of this quantizer is always larger
than the sampling period. Taking this delay into account, ST F2 is no longer flat.
The delay matching is therefore crucial for the first-stage nonideality cancellation,
especially for wide-bandwidth applications.

Fig. 7.9 Block diagram of a


simplified CT loop to
approximately represent the
2nd stage in a MASH ADC
7.2 Architecture Analysis of the 0-ΔΣ MASH VCO-Based ADC

Fig. 7.10 Impulse invariance transformation of the 2nd stage in the presented 0–2 MASH VCO-based ADC
135
136 7 VCO-Based 0-ΔΣ MASH ADC

To improve the delay matching, Fig. 7.9 illustrates a simplified continuous-time


(CT) loop with different additional delays which approximately represents the 2nd
stage. In the presence of these delays, the transfer function of this loop is obtained
as:
Y 1 + H (s)e−s(T1 −T2 )
(s) = e−s(T2 +T3 ) (7.3)
X 1 + H (s)e−sT3

Specificly, if T1 = T2 + T3 is ensured, this transfer function can be simplified as


a delay:
Y
(s) = e−sT1 (7.4)
X

In this case, if we also add the delay e−sT1 to the first-stage output, the cancellation
condition is satisfied.
The schematic from the 1st-stage output to the 2nd-stage output is re-drawn in
Fig. 7.10. To study the delay matching, an impulse input is applied at the 1st-stage
output. The impulse invariance transformation is shown. A quarter clock delay, intro-
duced between the first quantizer and the feedforward internal DAC, is used to absorb
the quantizer’s propagation delay and to re-synchronize the output signals before
entering into the DAC. This clock delay refers to the delay T2 shown in Fig. 7.9.
Similarly, a flip-flop (FF) succeeding the second quantizer determines the delay T3 .
Then we can introduce the delay T1 = T2 + T3 = 1.5TS preceding the second-stage
input. Therefore, the delay matching is satisfied as needed to improve the nonlinearity
cancellation.

7.3 Circuit Implementation of the 0–2 MASH VCO-Based


ΔΣ ADC

7.3.1 Three-Input Adder

As shown in Fig. 7.7, a three-input adder is needed to sum up all the branches pre-
ceding the second quantizer. Normally, two kinds of adder, active and passive, can
be employed. To reduce the power consumption, it is preferred to remove the power-
hungry active components from the adder. Compared to the active adder, a passive
adder is preferred as it has no power-hungry active components and thus no extra
DC power is wasted. However, the summed signal is also attenuated, which is not
suitable for the nonlinearity cancellation.
In this work, a novel dual-input VCO-based quantizer is proposed in the 2nd
stage to solve this design challenge. As shown in Fig. 7.11, a two-input passive
adder combined with an implicit current adder realizes the summation function.
In this circuit, an inherent current adder is implemented in the dual-input VCO
structure, where the sum of the two current-based signals represents the adder output,
7.3 Circuit Implementation of the 0–2 MASH VCO-Based ΔΣ ADC 137

Fig. 7.11 Topology of the dual-input VCO-based fine quantizer (single-ended) to realize the three-
input adder

introducing no penalty in terms of DC power or latency. The extra common-mode


(CM) current is subtracted through a cascode current mirror. Thanks to the shaping
from the preceding loop filter, this three-input adder does not require a 12-bit accuracy
in this design. Since both voltage swings of the two VCO inputs are small enough, the
proposed dual-input VCO-based quantizer does not cause any nonlinearity problem.

7.3.2 VCO-Based Quantizer

The schematic of the VCO-based quantizer [10] is also shown in Fig. 7.11. To cancel
out the even-order harmonic distortion [11], the pseudo-differential 5-bit quantizers
clocked at 1.6 GHz are composed of two 4-bit single-ended VCO-based quantizers.
Each ring VCO contains 15 current-starved delay cells, nominally operating at about
400 MHz, with a VCO gain K V of about 2 GHz/V. A sense-amplifier-based flip-flop
(SAFF) regenerates and samples the converted phase signal. Its schematic is given
in Fig. 7.12. At the end, a subsequent true-single-phase clock (TSPC) register and an
XOR gate fulfill the 1st-order differentiation function.
The phase noise of the first VCO is also canceled in the cascade structure, while the
phase noise of the second VCO can be regarded as 1st-order high-pass shaped noise
and is added to the 2nd-order-shaped quantization noise E2. As shown in Fig. 7.13,
the simulated phase noise is −115 dBc/Hz at 10-MHz offset frequency. Following
the formula in [12] and considering the one extra order of shaping from the analog
138 7 VCO-Based 0-ΔΣ MASH ADC

Fig. 7.12 Schematic of the sense-amplifier-based flip-flop

Fig. 7.13 Simulated phase noise performance of the VCO


7.3 Circuit Implementation of the 0–2 MASH VCO-Based ΔΣ ADC 139

integrator, the corresponding calculated signal-to-phase-noise-ratio is 102 dB. It is


clear that this phase noise does not limit the ADC performance.
For PVT variations, simulations show that the power-supply sensitivity of the
VCO gain is 13.6%/V while the temperature sensitivity is 2.2/◦ C. Considering the
operation conditions in practice (e.g., ±20% power supply voltage variation and
50-◦ C temperature range), the maximum allowed ±10% gain accuracy requirement
of the second VCO-based quantizer is sufficient for the 12-bit ADC accuracy in
this design. In addition, 3-bit binary-controlled current-trimming cells are utilized
to tune the VCO’s biasing current, therefore the impact of process variations can be
accommodated.

7.3.3 Integrator

The structure of the analog integrator used is shown in Fig. 7.14. Due to PVT vari-
ations, the RC time constant can vary as much as ±30%. To accmodate this effect,
a 4-bit binary-weighted capacitor array, shown in Fig. 7.15, is used to tune the time
constant. The total equivalent capacitance is:

Ctotal = 8C + kC, k = 0, 1, ..., 15 (7.5)

The minimum and maximum available capacitor values are Cmin = 8C and Cmax
= 23C. So the tuning range of the tunable capacitor is:

Cmax
= 2.88 (7.6)
Cmin

and the tuning resolution is given by:

C
= 6.25% (7.7)
16C
where the 16 C is the nominal value of the capacitor bank, which can accomodate
the effect of PVT variations.

Fig. 7.14 Schematic of the


integrator
140 7 VCO-Based 0-ΔΣ MASH ADC

Fig. 7.15 Discretely tunable


capacitor for the loop filter

In the layout, shielding is carefully placed at some critical locations, e.g., the vir-
tual ground, to avoid unwanted crosstalk with high-frequency digital signals. In the
measurements, an off-line manual capacitance tuning targeted at the 6.25% tuning
resolution is adopted. After tuning, the capacitance value remains fixed during the
ongoing measurements. An RC time constant auto-tuning structure might be imple-
mented in the future. It only needs to detect the noise floor of the second stage output
since the first stage is open loop.
The operational amplifier used in the integrator is implemented with a fully-
differential topology as shown in Fig. 7.16. In order to save power consumption,
the Miller compensation scheme is eliminated. Instead, a left-half-plane zero intro-
duced by a feed-forward path is used to compensate the phase shift. The simplified
small-signal equivalent block diagram of this amplifier is illustrated in Fig. 7.17. The
transfer function of the amplifier is given by:

1 + s gm1 RR1 1gCm21 gRm3 R2


2 +gm3 R2
H (s) = (gm1 R1 gm2 R2 + gm3 R2 ) · (7.8)
(1 + s R1 C1 )(1 + s R2 C2 )

Clearly, a left-half-plane zero is generated as:


gm1 gm2
z≈− (7.9)
gm3 C1
7.3 Circuit Implementation of the 0–2 MASH VCO-Based ΔΣ ADC 141

Fig. 7.16 Schematic of the operational amplifier in the integrator

Fig. 7.17 Small-signal


equivalent block diagram of
the operational amplifier

The DC levels at the two-stage outputs are controlled by two common-mode feed-
back (CMFB) blocks. Opamp with high DC gain are usually required in a traditional
MASH structure to maintain N T F1 = H2 and to reduce the noise leakage. However,
as discussed in Sect. 7.3, thanks to the open-loop first stage, the matching require-
ment is greatly relaxed in this novel structure. In other words, the noise leakage is
mitigated even with a finite-gain opamp. Transistor-level simulations predict that
31-dB DC gain and 1-GHz GBW are achieved with 1.5-mW DC power and 1.5-pF
load, which satisfies the requirements needed according to the MATLAB behavioral
simulations.

7.3.4 DACs

According to the behavioral simulations in Matlab, Fig. 7.18 illustrates the impact of
the DAC cells mismatch. The inherent DEM provided by the VCO-based quantizer
reduces the linearity degradation due to the DAC cells mismatch. To achieve a 12-bit
accuracy (74-dB SNDR), the standard deviation of the global DACs needs to be
limited to 8%, while the matching requirement of the internal DAC is greatly relaxed
because it is further shaped by the analog integrator. The explicit DEM is therefore
avoided, saving power and area.
142 7 VCO-Based 0-ΔΣ MASH ADC

Fig. 7.18 ADC SNDR performance as a function of the standard deviation of the current of the
global and internal DAC cells

The DACs in the main feedback path and in the internal feedback path are
implemented using a non-return-to-zero (NRZ) current-steering structure. The global
current-feedback DACs are the dominant noise sources in this design. The current-
source transistors are designed with large dimensions and high overdrive voltage
to reduce the noise contribution. Noise simulations show that the root-mean-square
(RMS) output noise voltage of the global DAC is 36 µV, which corresponds to about
−78 dBFS.

7.3.5 Interface

The output of the single-ended quantizer is a 4-bit thermometer-coded signal. A


unary-to-binary encoder, as shown in Fig. 7.19, is integrated on chip to transfer the
output codes from unary to binary type, and then to sum them up. The function of the
encoder is to count the number of ones in the thermometer-coded input. A Wallace
tree topology is adopted considering the trade-off of speed and power.
The implemented mirror adder in the encoder is presented in Fig. 7.20. The NMOS
and PMOS chains are completely symmetrical. A maximum of two series transistors
appear in the carry generation path.
7.4 Experimental Results 143

Fig. 7.19 Block diagram of the Wallace tree encoder

7.4 Experimental Results

The prototype design of the presented 0–2 MASH VCO-based ADC has been fab-
ricated in a 40-nm CMOS process with 0.9-V supply voltage. The chip micropho-
tograph is shown in Fig. 7.21. The core circuit occupies an active area of 0.16 mm2 .
The modulator output bitstreams are fed to digital buffers and to a low-voltage dif-
ferential signal (LVDS) interface. As the main purpose of this chip is to verify the
proposed architecture and to investigate the critical design considerations, the dec-
imation filter has not been implemented in this design. A multi-stage configuration
might be adopted for this in future work.
The measurement setup is shown in Fig. 7.22. A microcontroller board is imple-
mented to program the on-chip shift register which can adjust the digital-controlled
bits of the biasing current bank in the VCOs and the capacitor bank in the integrator.
The averaged 8-length 16384-point measured PSD is shown in Fig. 7.23, with an
8-MHz −2.5-dBFS 0.5-V P P input signal. The top of Fig. 7.23 presents the output of
the 1st stage. A strong 3rd-order harmonic with 1st-order noise shaping is observed.
The signal level in the middle of Fig. 7.23 is about 20 dB lower, providing that
the 2nd-stage quantizer only processes the residual signal. Its harmonic has the same
amplitude level as for the first quantizer. Finally, when adding the two paths together,
144 7 VCO-Based 0-ΔΣ MASH ADC

Fig. 7.20 Schematic of the mirror adder implemented in the unary-to-binary encoder

as shown in the bottom of Fig. 7.23, the distortion is largely cancelled and a 77.1-dB
SFDR is obtained for the complete ADC. Figures 7.24 and 7.25 present the measured
spectrum with a 4-MHz and 1-MHz input respectively.
The measured SNDR performance as a function of the input level for different
input frequencies is shown in Fig. 7.26. The peak SNDR for a 1-MHz, 4-MHz and 8-
MHz input are 68.9 dB, 67.8 dB and 66.8 dB, respectively. As indicated in Eq. (7.2),
the distortion at the final output is frequency-dependent due to the high-pass shaping.
Therefore the ADC prototype with a lower-frequency input achieves a higher peak
SNDR. The 8-MHz frequency of the input signal represents the worst case because
it is the highest frequency for which the 5th-order harmonics fall within the 40-MHz
7.4 Experimental Results 145

BW. In addition, a feedforward path is added in the second stage to maintain ST F2


flat at high frequencies. Therefore, the ADC performance can be maintained over the
whole signal bandwidth. With an 8-MHz input, the ADC achieves a peak SNR/SNDR
of 68.7/66.8 dB over a 40-MHz signal bandwidth, resulting in an ENOB of 10.81
bit.
A two-tone intermodulation test has been conducted with a 10-MHz and 11-MHz
input. The signal levels are both −8.2 dBFS. In Fig. 7.27, it is shown again that the
IM3 is frequency-dependent because of the 1st-order high-pass shaping. The low-
frequency 3rd-order IMD located at 9 and 12 MHz are −90 dBFS, while the high-
frequency 3rd-order IMD at 31 and 32 MHz are about 8 dB higher. The corresponding
IM3 is 73 dB, which is nearly the worst case because the high-frequency 3rd-order
IMD falls within the 40-MHz BW.
The total power consumption is 4.98 mW under 0.9-V supply, resulting in
an excellent FoM of 165.8 dB (SNDR+10log(BW/Power)) or 35 fJ/step (Power/
(2BW*2 E N O B )). The performance summary of the 0–2 MASH VCO-based ΔΣ
ADC is shown in Table 7.1, and will be compared to the state-of-the-art works at the
end of Chap. 8, after having presented the final ADC design. The presented design
achieves the best FoM among all the state-of-art wide-bandwidth (>20 MHz) ΔΣ
ADCs, as far as we know.

Fig. 7.21 Chip microphotograph of the 0–2 MASH ΔΣ ADC prototype


146

PoleZero Filter Bank


ParBERT 81250
ADT1-1WT

AWG430 HP8133A PIC18F4620

Fig. 7.22 Measurement setup for the 0–2 MASH ΔΣ ADC prototype
7 VCO-Based 0-ΔΣ MASH ADC
7.4 Experimental Results 147

First−Stage Output
0
Hann−windowed output
Integration Noise power
−10 Integration Noise Distortion power

−20

−30
Magnitude(dB)

−40
20dB/Dec

−50

−60

−70

−80

−90

−100
10 6 10 7 10 8
Frequency(Hz)

Second−Stage Output
0
Hann−windowed output
Integration Noise power
−10 Integration Noise Distortion power

−20

−30
Magnitude(dB)

−40
40dB/Dec

−50

−60

−70

−80

−90

−100
10 6 10 7 10 8
Frequency(Hz)

Complete ADC Output


0
Hann−windowed output
Integration Noise power
−10 Integration Noise Distortion power

−20

−30
Magnitude(dB)

−40
40dB/Dec

−50

−60

−70

−80

−90

−100
106 107 108
Frequency(Hz)

Fig. 7.23 Measured PSD results of the 0–2 MASH ΔΣ ADC chip prototype with an 8-MHz input.
Top first stage; middle second stage; bottom complete ADC output
148 7 VCO-Based 0-ΔΣ MASH ADC

4−MHz Input
0
Hann−windowed output
−10 Integration Noise power
Integration Noise Distortion power
−20

−30
Magnitude(dB)

−40

−50

−60

−70

−80

−90

−100

−110
10 6 10 7 10 8
Frequency(Hz)

Fig. 7.24 Measured PSD results of the 0–2 MASH ΔΣ ADC chip prototype with a 4-MHz input

1−MHz Input
0
Hann−windowed output
−10 Integration Noise power
Integration Noise Distortion power
−20

−30
Magnitude(dB)

−40

−50

−60

−70

−80

−90

−100

−110
106 107 108
Frequency(Hz)

Fig. 7.25 Measured PSD results of the 0–2 MASH ΔΣ ADC chip prototype with a 1-MHz input
7.4 Experimental Results 149

Fig. 7.26 Measured SNDR versus the input amplitude for different input frequencies

Two−tone Measurement
0

−10

−20

−30
Magnitude (dB)

−40

−50

−60

−70

−80

−90

−100

5 10 15 20 25 30 35 40 45
Frequency (MHz)

Fig. 7.27 Output spectrum of the two-tone measurement (at 10 and 11 MHz)
150 7 VCO-Based 0-ΔΣ MASH ADC

Table 7.1 Measurement Specification Value


summary of the 0–2 MASH
VCO-based ΔΣ ADC Signal bandwidth 40 MHz
Sampling frequency 1.6 GHz
Peak SNR 68.7 dB
Peak SNDR 66.8 dB
Power 4.98 mW
FoM1 35 fJ/Step
FoM2 165.8 dB
Core area 0.16 mm2
Technology TSMC 40 nm GP

7.5 Conclusions

VCO nonlinearity is the bottleneck of VCO-based ΔΣ ADCs with moderate accu-


racies. This chapter has presented a nonlinearity-cancellation technique in a power-
efficient 0-ΔΣ MASH ADC structure with dual-input VCO-based ADC. A 0–2
MASH VCO-based ΔΣ ADC prototype implemented in 40-nm CMOS confirms
the theoretical analysis and achieves a SNDR of 66.8 dB over a 40-MHz BW, con-
suming only 4.98 mW. By taking full advantage of the low-power VCO-based circuit
and overcoming its inherent nonlinearity problem, the FoM of the state-of-art high-
bandwidth ΔΣ ADCs has been extended to 166 dB or 35 fJ/step. Compared with the
design presented in Chap. 6, this 0–2 MASH VCO-based ADC design has achieved
better accuracy performance and even better power efficiency, an analog integrator
is still needed however. To implement a highly-digital architecture and high-order
noise shaping simultaneously, a fully-VCO-based 0–2 MASH ADC will be shown
in the next Chapter, without any analog amplifier and with an excellent FoM.

References

1. P. Zhu, X. Xing, G. Gielen, A 40MHz-BW 35fJ/step-FoM nonlinearity-cancelling two-step


ADC with dual-input VCO-based quantizer, in IEEE European Solid-State Circuits Conference
(ESSCIRC), 2014, (Venice, Italy, 2014), pp. 63–66
2. P. Zhu, X. Xing, G. Gielen, A 40-MHz bandwidth 0–2 MASH VCO-based Delta-Sigma ADC
with 35-fJ/step FoM. IEEE Trans. Circuits Syst. II Express Briefs 62(10), 952–956 (2015)
3. A. Gharbiya, D.A. Johns, A 12-bit 3.125 MHz bandwidth 0–3 MASH Delta-Sigma modulator.
Solid-State Circuits, IEEE J. 44(7), 2010–2018 (2009)
4. Y. Dong, R. Schreier, W. Yang, S. Korrapati, A. Sheikholeslami, A 235 mw CT 0–3 MASH
ADC achieving −67 dbBFS/Hz NSD with 53 MHz BW, in Solid-State Circuits Conference
Digest of Technical Papers (ISSCC), 2014 IEEE International, (IEEE, 2014), pp. 480–481
5. A. Sanyal, K. Ragab, L. Chen, T. Viswanathan, S. Yan, and N. Sun, A hybrid SAR-VCO Delta-
Sigma ADC with first-order noise shaping, in 2011 IEEE Symposium on CICC, (2014), pp.
1–4
References 151

6. X. Xing, G.G.E. Gielen, A 42 fJ/step-FoM two-step VCO-based Delta-Sigma ADC in 40 nm


CMOS. Solid-State Circuits, IEEE J. 50(3), 714–723 (2015)
7. K. Reddy, S. Rao, R. Inti, B. Young, A. Elshazly, M. Talegaonkar, P.K. Hanumolu, A 16-
mW 78-dB SNDR 10-MHz BW CT ΔΣ ADC using residue-cancelling VCO-based quantizer.
Solid-State Circuits, IEEE J. 47(12), 2916–2927 (2012)
8. A. Gupta, K. Nagaraj, T. Viswanathan, A two-stage ADC architecture with VCO-based second
stage. IEEE Trans. Circuits Syst. II Express Briefs 58(11), 734–738 (2011)
9. W. Yu, J. Kim, K. Kin, S. Cho, A time-domain higher-order MASH Delta-Sigma ADC using
voltage-controlled gated-ring oscillator. IEEE Trans. Circuits Syst. I: Regul. Pap. 59(8), 856–
866 (2012)
10. M.Z. Straayer, M.H. Perrott, A 12-bits, 10-MHz bandwidth, continuous-time Delta-Sigma
ADC with a 5-bits, 950-MS/s VCO-based quantizer. IEEE J. Solid-State Circuits 43(4), 805–
814 (2008)
11. S. Rao, K. Reddy, B. Young, P. Hanumolu, A deterministic digital background calibration
technique for VCO-based ADCs. IEEE J. Solid-State Circuits 49(4), 950–960 (2014)
12. J. Kim, T.K. Jang, Y.G. Yoon, S.H. Cho, Analysis and design of voltage-controlled oscillator
based analog-to-digital converter. IEEE Trans. Circuits Syst. I Regul. Pap. 57(1), 18–30 (2010)
Chapter 8
Fully-VCO-Based High-Order ΔΣ ADC

8.1 Introduction

Scaling friendly in the light of the reducing voltage headroom and increasing time
resolution of the nanometer CMOS technology, VCO-based quantizers have been
exploited in CT ΔΣ ADCs. As discussed in Chap. 7, the 0-ΔΣ MASH topology
overcomes the VCO nonlinearity problem due to its inherent operation principle,
being more power-efficient than a single-loop structure. The analog integrator in
the ΔΣ loop, however, remains difficult to be implemented with technology scaling
and consumes a large amount of the power as a high OTA DC gain is required.
Conceptually, a VCO accumulates the excess phase like an ideal integrator. It is
therefore beneficial to shift the analog integrator to the time and the digital domains
as well.
In this chapter, we will report a 0-2 MASH CT ΔΣ ADC with fully-VCO-based
integrator and quantizers, without any analog integrators. Firstly the principle of
a VCO-based integrator will be presented in Sect. 8.2. The circuit design details
and silicon implementation results will be given and discussed in Sect. 8.3. Taking
advantage of the fully-digital high-order structure and relieving the VCO nonlinearity
bottleneck, a 74-dB SFDR and a FoM of 162 dB or 52 fJ/step over a 40-MHz
bandwidth are achieved. Finally, the conclusion will be drawn in Sect. 8.4.

8.2 Integrators

8.2.1 Traditional Analog Integrator

In a traditional ΔΣ ADC, the first integrator of the loop filter needs to satisfy a
severe gain and speed requirement due to the absence of the noise shaping, therefore
it consumes most of the power. Compared to the amplifers implemented in switched-
154 8 Fully-VCO-Based High-Order ΔΣ ADC

R-C Integrator Gm-C Integrator

Fig. 8.1 Traditional analog R-C and G m -C integrator

Fig. 8.2 Simplified diagram


of a R-C integrator, in which
the opamp has a finite DC
gain and a finite GBW

capacitor (SC)-based DT ΔΣ ADCs, the opamps in CT ΔΣ ADCs have a relaxed


speed requirement. However, the non-idealities of the integrator, such as the finite
GBW and finite DC gain, also degrade the modulator performance.
Conventionally, as shown in Fig. 8.1, the integrators are implemented by R-C
or G m -C structures. Compared to the active R-C integrator, the open-loop G m -C
structure consumes less power, but has a worse linearity performance.
As an example, a simplified diagram of an R-C integrator, in which the opamp
has a finite DC gain A0 and a finite bandwidth ω0 , is shown in Fig. 8.2. The transfer
function of the integrator is obtained as:
A0
H (s) ≈ − (8.1)
1 + s(A0 RC + 1
ω0
) + RC 2
ω0
s

Assuming that the DC gain is high enough to satisfy A0  1, the above transfer
function can be simplified as:
8.2 Integrators 155

Ideal integrator Single-pole roll-off filter

Fig. 8.3 Equivalent model of an R-C integrator, in which the opamp has a finite gain and GBW

1
H (s) ≈ −
s(RC + 1
G BW0
) + RC
G BW0
s2
1 1
=− · (8.2)
s RC 1 + 1
G BW0 RC
+ 1
G BW0
s
1 1 − G err or
=− ·
s RC 1 + ωs2

where GW B0 is the GBW of the opamp. G err or and ω2 refer to the gain error and an
additional pole caused by the opamp’s finite GBW, which are defined by:
1
G err or = RC
(8.3)
G BW0 + 1
RC

and
1
ω2 = G BW0 + (8.4)
RC
Therefore, the finite GBW results in both a gain error and a time delay. The gain
error can be compensated by tuning the integrator coefficient. However, the additional
delay leads to a decreased noise suppression, or even loop instability. As shown in
Fig. 8.3, the transfer function of an R-C integrator is modeled as an ideal integrator
along with a 1st-order roll-off filter. Moreover, the finite DC gain also results in a
reduced noise shaping ability of the integrator. Therefore, to increase the DC gain
and the GBW especially in a nanometer CMOS process, a penalty of considerable
power consumption needs to be paid.

8.2.2 VCO-Based Integrator

To take advantage of the technology scaling that has greatly decreased the power
dissipation of the digital parts, it is beneficial to shift the analog integrator to the
digital domain as well. As discussed in Chap. 3, a VCO acts like an integrator for the
output phase. The frequency-domain transfer function is depicted in Fig. 8.4. A pole
156 8 Fully-VCO-Based High-Order ΔΣ ADC

Fig. 8.4 Frequency-domain transfer function of a VCO

Fig. 8.5 Block diagram of a N-stage VCO-based 1st-order integrator

located at DC is provided with a unity gain frequency K V C O . Clearly, an advantage


of a VCO-based integrator is that a VCO theoretically has an infinite DC gain which
is power-efficient and independent of the supply voltage.
As can be seen in Fig. 8.5, in the frequency domain, the N-stage VCO modulates
the input f sig with a carrier frequency f C which is defined by:

fC = N · f V C O (8.5)

where f V C O is the free-running frequency of the VCO. The VCO output phases are
converted to duty-cycled PWM pulses by phase detectors (PD), and then transferred
back to voltage/current form by DACs. Interestingly, the barrel-shifting property of
the multi-level PWM signals offers an implicit DEM of the DAC cells.
In summary, the total transfer function of this integrator HI N T is given by:
8.2 Integrators 157

Fig. 8.6 Block diagram of a single-loop ΔΣ ADC with a VCO-based integrator

K V C O · K P D · K D AC
HI N T (s) = (8.6)
s
where K V C O , K P D and K D AC are the gain of the VCO, PD and DAC respectively.
Moreover, the VCO-based integrator can be extended to a higher order by cascading
the basic structures. Reference [1] presents a silicon prototype of a VCO-based 4th-
order Butterworth filter by cascading two biquads.
However, as discussed in Chap. 4, the VCO-based integrator suffers from the
VCO nonlinearity. The techniques presented in Chap. 4 can be used to mitigate this
nonideality. A design example will be shown in Sect. 8.3.

8.2.3 Fully-VCO-Based ΔΣ ADC Structure

Replacing the conventional analog integrator by a VCO, the VCO-based integra-


tor can be utilized in a ΔΣ loop. Figure 8.6 illustrates the block diagram. For the
VCO-based integrator, since the sampling operation does not take place here, no
quantization error is added. However, f C needs to be sufficiently high to avoid spurs
folding back into the baseband due to the later sampling. The VCO input swing is
reduced due to the feedback and the integration. The VCO nonlinearity is therefore
mitigated.
As shown in Fig. 8.7, an architecture of a VCO-based ΔΣ ADC with a VCO-
based integrator is proposed in [2], with a 2nd-order noise shaping coming from the
VCOs in the integrator and in the quantizer. This purely-VCO-based ADC is power-
efficient and scaling friendly, benefiting from the mostly-digital structures. However,
the ADC performance is still limited by the nonlinearity of the VCO-based quantizer.
The 0-ΔΣ MASH topology discussed in Chap. 7 can overcome the VCO’s non-
linearity problem due to its inherent operation principle, being more power-efficient
than the single-loop structure. By replacing the analog integrator, the architecture of
a fully-VCO-based 0-ΔΣ MASH ADC is proposed in Fig. 8.8. This design will now
be elaborated in detail.
158 8 Fully-VCO-Based High-Order ΔΣ ADC

Fig. 8.7 Block diagram of a 2nd-order VCO-based ΔΣ ADC with a VCO-based integrator [2]

Fig. 8.8 Block diagram of


the proposed 0-ΔΣ MASH
ADC with VCO-based
integrators

8.3 Design Example: A Fully-VCO-Based 0-2 MASH


VCO-Based ΔΣ ADC

8.3.1 System Architecture

A design example of a fully-VCO-based 0-2 MASH VCO-based ΔΣ ADC is pre-


sented in this section. Figure 8.9 shows the block diagram of the proposed two-stage
ADC. According to the analysis in Chap. 7, the nonlinearity of the first stage is can-
celled if the STF of the 2nd stage is equal to 1. An input-feedforward structure is
adopted in the second stage to improve the cancellation. As the second stage only
processes the quantization noise of the first stage, its input voltage swing is greatly
reduced. Therefore, the VCOs implemented in the integrator and in the quantizer of
the second stage do not limit the ADC nonlinearity performance.
The clock timing of the proposed ADC is shown in Fig. 8.10. The details con-
cerning delay matching have been discussed in Chap. 7 and [3].
Figure 8.11 depicts the presented MASH ADC and its corresponding signal model,
where D Q1 represents all the nonlinear harmonics generated in the first quantizer due
to the VCO’s nonlinear behavior, E 1 and E 2 represent the corresponding quantization
noise from the two quantizers, K Q1 and K Q2 indicate the quantizer gains, K D AC,I
indicates the internal DAC in the feedforward path and K D AC,G indicates the global
8.3 Design Example: A Fully-VCO-Based 0-2 MASH VCO-Based ΔΣ ADC 159

Fig. 8.9 Block diagram of the presented fully-VCO-based 0-2 MASH ΔΣ ADC

Fig. 8.10 Clock timing of the presented fully-VCO-based ADC

+
– –
+ +

Fig. 8.11 Signal model of the presented fully-VCO-based 0-2 MASH VCO-based ADC

DACs connected to the virtual ground of the integrator. HI N T (s) represents the
transfer function of the VCO-based integrator, which is defined as Eq. (8.6).
As derived from the signal model, all the nonlinear harmonics appearing at the
final ADC output can be expressed as:
1 − K D AC,I · K Q2
D Q1 K Q1 K D AC,G ·K Q2 ·K V C O ·K P D ·K D AC,I N T
(8.7)
1+ s
160 8 Fully-VCO-Based High-Order ΔΣ ADC

This shows that the nonlinearity cancellation is mainly dependent on the product
of K D AC,I K Q2 , which is shaped by the integrator HI N T (s).
In behavioral simulations, the proposed ADC achieves 78-dB SFDR within ±10%
gain variations of K D AC,I K Q2 and 81-dB SFDR within ±10% gain variations of
K V C O,I N T . Simulations show that the power-supply sensitivity and the temperature
sensitivity of the VCO’s gain both in the integrator and in the quantizers are robust
enough for the targeted 12-bit ADC accuracy in this design. In addition, 3-bit binary-
controlled current-trimming cells are utilized to accommodate the impact of process
variations.

8.3.2 Circuit Implementation

As shown in Fig. 8.9, the VCO-based integrator transfers the input signal to the
phase domain, and then recovers it back to the voltage domain after completing the
integration process. The pseudo-differential integrator’s schematic is presented in
Fig. 8.12, consisting of inverter-based ring oscillators, phase detectors and DACs.
The ring oscillator modulates the input with a carrier frequency f C , which needs
to be sufficiently higher than the sampling frequency to avoid spurs folding back into
the baseband. In this design, an f C of 3.5 GHz is realized with 24 stages and a 145-
MHz oscillation frequency f V C O , and is filtered out by the subsequent VCO-based
quantizer due to its inherent anti-aliasing property. 3-bit binary-controlled current-
trimming cells are utilized to tune the VCO’s biasing current. The current-source
transistors are designed with large dimensions and high overdrive voltage to reduce
the noise current.
In addition, the phase noise of the VCOs in the front-end integrator also adds
directly at the ADC input and therefore the VCOs should be carefully designed.
Non-minimum-size transistors are utilized in the delay cells to reduce the flicker
noise. As shown in Fig. 8.13, the simulated VCO phase noise is −125 dBc/Hz at
a 10-MHz offset frequency, which does not limit the ADC performance according
to the simulation. To accomodate the parasitic delay in the VCO, an extra delay
compensation path is added. A buffer and a feedforward capacitor C F couple the
negative VCO input to the biasing node which introduces a negative zero [4].
A SA-based buffer shown in Fig. 8.14 isolates the VCO delay cell and recov-
ers its output to rail-to-rail. The subsequent differential phase detector compares
the phases of the pseudo-differential oscillators and generates the duty-cycled out-
put, eliminating the external reference signal. The duty cycle is proportional to the
phase difference between the inputs. Moreover, in this differential-PD structure, the
impact of the free-running frequency mismatch on the common-mode signal can be
mitigated.
8.3 Design Example: A Fully-VCO-Based 0-2 MASH VCO-Based ΔΣ ADC 161

Fig. 8.12 Schematic of the pseudo-differential VCO-based integrator

Fig. 8.13 Simulated phase noise performance of the VCO used in the integrator
162 8 Fully-VCO-Based High-Order ΔΣ ADC

Fig. 8.14 Schematic of the


SA-based buffer in the
VCO-based integrator

There are three main structures that can be selected as phase detectors:
1. XOR gate as PD:
The XOR circuit produces error pulses on both the rising and falling edges. When
the two inputs are completely in-phase, the output has a constant level of zero.
The linear input range is π .
2. Two-state PD:
The two-state PD generates one error pulse for one comparison. The linear input
range is 2π .
3. Three-state phase-frequency detectors (PFD):
The three-state PFD can produce an output even when the two inputs differ not
only in phase but also in frequency. However, the mismatch at the subsequent
current subtraction introduces extra nonlinearity [1].
To avoid PD output overflow and to reduce the nonlinearity, the two-state PD is
preferred over the XOR gate and the three-state PFD. Therefore, the two-state PD
illustrated in Fig. 8.15 is implemented in this design. The transfer function of the
two-state PD is shown in Fig. 8.16. The PD outputs a 50%-duty-cycled pulse when
the input phase difference is zero, which indicates the common-mode condition.
In this design, since the input swing of the second stage is very small, the phase
differences always stay in the linear range, and the PD output pulses are duty cycled
around 50%. Therefore, the nonlinearity caused by the finite rise/fall time of the
PD-generated PWM signals is greatly reduced [1].
As observed from Eq. (8.6), the loop-filter coefficient can be tuned conveniently
by adjusting the DAC current. The DACs are implemented using a source-switched
structure, as shown in Fig. 8.17. The implicit DEM relaxes the matching requirement
of the DAC cells.
8.3 Design Example: A Fully-VCO-Based 0-2 MASH VCO-Based ΔΣ ADC 163

Fig. 8.15 Schematic of the


two-state phase detector

Fig. 8.16 Transfer function


of the two-state phase
detector

The pseudo-differential 5-bit quantizers in the two stages are clocked at 1.6 GHz.
Each single-ended ring VCO contains 15 current-starved delay cells, nominally oper-
ating at about 400 MHz. All the details of these building blocks have already been
discussed in Chap. 4.

8.3.3 Experimental Results

The prototype design of the presented fully-VCO-based 0-2 MASH ΔΣ ADC has
been fabricated in a 40-nm CMOS technology with 0.9-V supply voltage. Figure 8.18
shows the chip micrograph, the core area is 0.17 mm2 . The measurement setup has
been shown in the last chapter.
The averaged measured output power spectral density is shown in Fig. 8.19, with
an 8-MHz 0.3-V P−P input signal. A strong 3rd-order harmonic exists at the 1st-stage
output. When adding the two stages together, the distortion is significantly cancelled
and the SFDR is improved from 42 to 74 dB. Folded back by intermodulations,
164 8 Fully-VCO-Based High-Order ΔΣ ADC

Fig. 8.17 Schematic of the source-switched DAC cells in the VCO-based integrator

Fig. 8.18 Chip microphotograph of the fully-VCO-based 0-2 MASH ΔΣ ADC prototype
8.3 Design Example: A Fully-VCO-Based 0-2 MASH VCO-Based ΔΣ ADC 165

−10

−20

−30
Magnitude(dBFS)

−40 SFDR=74dB

−50 40dB/Dec
−60

−70

−80

−90

−100
106 107 108
Frequency(Hz)

Fig. 8.19 Measured PSD results of the fully-VCO-based 0-2 MASH ΔΣ ADC chip prototype
with an 8-MHz input

−10

−20

−30
Magnitude(dBFS)

−40

−50

−60

−70

−80

−90

−100

106 107 108


Frequency(Hz)

Fig. 8.20 Measured PSD results of the fully-VCO-based 0-2 MASH ΔΣ ADC chip prototype
with a 4-MHz input

some spurious tones are observed within the signal band in the PSD plot due to
the imperfect cancelation of tones around the oscillation frequency of the multi-
level VCOs. A 2nd-order noise shaping is observed, which validates the VCO’s
integration function. A peak SNR/SNDR of 65.5/62.7 dB is achieved over a 40-MHz
signal bandwidth. In comparison, Figs. 8.20 and 8.21 present the measured spectrum
with a 4-MHz and 2-MHz input, respectively. The measured SNDR performance as
a function of the input level is shown in Fig. 8.22.
166 8 Fully-VCO-Based High-Order ΔΣ ADC

0
−10
−20
−30
Magnitude(dBFS)

−40
−50
−60
−70
−80
−90
−100
−110
6 7 8
10 10 10
Frequency(Hz)

Fig. 8.21 Measured PSD results of the fully-VCO-based 0-2 MASH ΔΣ ADC chip prototype
with a 2-MHz input

Fig. 8.22 Measured ADC SNDR versus input amplitude

A two-tone intermodulation test has been conducted with a 10 and 11-MHz input.
The signal levels are both −8.3 dBFS. In Fig. 8.23, it is shown that the IM3 is
frequency dependent because of the 1st-order high-pass shaping. The low-frequency
3rd-order IMD located at 9 and 12 MHz are −86 dBFS, while the high-frequency
3rd-order IMD at 31 and 32 MHz are about 8 dB higher. The corresponding IM3 is
69 dB.
The total power consumption is 4.6 mW, resulting in a FoM of 162 dB (SNDR+10
log(BW/Power)) or 52 fJ/Step (Power/(2BW*2 E N O B )). This design realizes fully-
8.3 Design Example: A Fully-VCO-Based 0-2 MASH VCO-Based ΔΣ ADC 167

Two−tone Measurement
0
−10
−20
−30
Magnitude (dB)

−40
−50
−60
−70
−80
−90
−100
−110
5 10 15 20 25 30 35 40 45 50
Frequency (MHz)

Fig. 8.23 Output spectrum of the two-tone measurement (at 10 and 11 MHz)

Table 8.1 Measurement Specification Value


summary of the
fully-VCO-based 0-2 MASH Signal bandwidth 40 MHz
ΔΣ ADC Sampling frequency 1.6 GHz
Peak SNR 65.5 dB
Peak SNDR 62.7 dB
Power 4.6 mW
FoM1 52fJ/Step
FoM2 162.0 dB
Core area 0.17 mm2
Technology TSMC 40 nm GP

VCO-based 2nd-order noise shaping, achieves an excellent FoM and scales more
effectively with future CMOS technology.
The measurements summary of the fully-VCO-based 0-2 MASH ADC are shown
in Table 8.1. Compared to the design presented in the last chapter which achieves
the best FoM among all the state-of-art wide-bandwidth (>20 MHz) ΔΣ ADCs, this
fully-VCO-based ADC has a lower power consumption, but with a degraded SNDR
performance.
This result is worse than expected. To understand the cause of the linearity degra-
dation, Fig. 8.24 illustrates a digital calibration for the chip prototype, in which a
two-tap FIR filter is implemented on the chip’s output in Matlab. The additive dig-
ital filter, of which the coefficients can be determined through adaptive off-line or
on-line calibration [5], aims to reduce the nonlinearity leakage due to the stage-gain
mismatch. As shown in Fig. 8.25, after applying this digital correction, the 3rd-order
168 8 Fully-VCO-Based High-Order ΔΣ ADC

On Chip

By Matlab

Fig. 8.24 Digital correction setup of the fully-VCO-based 0-2 MASH ΔΣ ADC chip prototype

0
Before Digital Correction
−10 After Digital Correction

−20
−30
Magnitude(dBFS)

−40
−50
−60
−70
−80
−90
−100
−110
106 107 108
Frequency(Hz)

Fig. 8.25 Digitally corrected PSD results of the fully-VCO-based 0-2 MASH ΔΣ ADC chip
prototype

harmonic is reduced below the noise floor, while the total harmonic distortion (THD)
is improved from 72 to 81 dB, meaning that nonlinearity leakage exists due to PVT
variations. Moreover, apart from the nonideal matching, the excess delay resulting
form the VCO-based integrator degrades the NTF of the second stage. In future work,
local-feedback delay compensation paths can be added in the ΔΣ loop to mitigate
this nonideality.
The five ADC designs presented from Chaps. 5–8 are compared to the state-of-the-
art oversampling ADCs [6] with BW≥8MHz in Figs. 8.26 and 8.27. In the FoM-BW
comparison, as far as we know, the FoM of our 0-2 MASH VCO-based ΔΣ ADC
design presented in Chap. 7 is until now the best among all published high-bandwidth
(≥20MHz) ΔΣ ADCs, as seen in Fig. 8.27.
85
2006&CT
2006&DT
2007&CT
80 2007&DT
2008&CT
2008&DT
2009&CT
75 2009&DT
2010&CT
2010&VCO
2011&CT
70 2011&DT
2011&VCO
2012&CT

SNDR(dB)
2012&DT
65 2012&VCO
2013&CT
2014&CT
2014&VCO
60 2015&CT
2015&VCO
2016&CT
This work
55
8.3 Design Example: A Fully-VCO-Based 0-2 MASH VCO-Based ΔΣ ADC

50 1 2
10 10
Bandwidth(MHz)

Fig. 8.26 Comparison of the SNDR and the bandwidth of the presented work with the state-of-the-art DSM ADC designs (BW ≥8MHz)
169
170

0
10

2006&CT
2006&DT
2007&CT
2007&DT
2008&CT
2008&DT
2009&CT
2009&DT
2010&CT
−1 2010&VCO
10

FoM(pJ/Conv)
2011&CT
2011&DT
2011&VCO
2012&CT
2012&DT
2012&VCO
2013&CT
2014&CT
2014&VCO
2015&CT
2015&VCO
2016&CT
This Work
−2
10
1 2
10 10
Bandwidth(MHz)

Fig. 8.27 Comparison of the FoM and the bandwidth of the presented work with the state-of-the-art DSM ADC designs (BW ≥8MHz)
8 Fully-VCO-Based High-Order ΔΣ ADC
8.4 Conclusions 171

8.4 Conclusions

Power consumption is one important design challenge of ADCs. The analog inte-
grator in the ΔΣ loop remains difficult to be implemented with CMOS technology
scaling and consumes a large amount of power as a high OTA DC gain is required.
Accumulating the excess phase, a VCO-based integrator is more digitally oriented
and power efficient. Another advantage is that a VCO theoretically has an infinite
DC gain which is power-efficient and independent of the supply voltage.
In this chapter, we have presented a 0-2 MASH CT ΔΣ ADC with fully-VCO-
based integrator and quantizers in 40-nm CMOS, without any analog integrators.
Benefiting from the 2nd-order noise shaping from the VCOs and relieving the VCO
nonlinearity bottleneck by the MASH structure, a SFDR of 74 dB is achieved over a
40-MHz signal bandwidth, with a peak SNR/SNDR of 65.5/62.7 dB. Taking advan-
tage of the fully-digital structure, the power consumption is 4.6 mW, resulting in
a FoM of 162 dB or 52 fJ/step. This result is degraded than expected due to the
nonlinearity leakage and the excess loop delay.

References

1. B. Drost, M. Talegaonkar, P. Hanumolu, Analog filter design using ring oscillator integrators.
IEEE J. Solid State Circuits 47(12), 3120–3129 (Dec. 2012)
2. Y. Yoon, K. Lee, P. Wang, N. Sun, A purely-VCO-based single-loop high-order continuous-
time ΣΔ ADC, in Proceedings of the IEEE International Symposium on Circuits and Systems
(ISCAS) (June 2014)
3. P. Zhu, X. Xing, G. Gielen, A 40-MHz bandwidth 0–2 MASH VCO-based Delta-Sigma ADC
with 35-fJ/step FoM. IEEE Trans. Circuits Syst. II Express Briefs 62(10), 952–956 (Oct. 2015)
4. K. Reddy, Rao S., R. Inti, B. Young, A. Elshazly, M. Talegaonkar, P.K. Hanumolu, A 16-mW
78-dB SNDR 10-MHz BW CT ΔΣ ADC using residue-cancelling VCO-based quantizer. IEEE
J. Solid State Circuits 47(12), 2916–2927 (Dec. 2012)
5. P. Kiss, J. Silva, A. Wiesbauer, T. Sun, U. Moon, J. Stonick, G. Temes, Adaptive digital
correction of analog errors in MASH ADCs i. off-line and blind on-line calibration. IEEE
Trans. Circuits Systems II: Analog Digit. Signal Process. 47(7), 621–628 (Dec. 2010)
6. B. Murmann, ADC performance survey 1997-2016 (2017), http://www.stanford.edu/
murmann/adcsurvey.html
Chapter 9
Conclusions

9.1 Summary and Conclusions

The ongoing data explosion requires larger data rates for almost all kinds of com-
munication standards. Another trend for electronic products is that more and more
portable devices are powered by a battery. As a result, power consumption is the key
design specification. Besides the pressure from the product definition, the receiver
architecture and the CMOS technology used also impose several design challenges
for low-power ADCs. This book has investigated and demonstrated the design of
highly-digital power-efficient DSM ADCs, covering both closed-loop and open-loop
architectures. The research work can be summarized as follows.

• In Chap. 2, the basic operation principle of different ADC topologies have shortly
been introduced. For each architecture, the typical specifications have been listed
and compared. Also, the application of ADC in receiver systems has briefly been
presented, and design factors affecting the ADC specifications have been given
and discussed.
• Design issues of traditional CT DSMs have been covered in Chap. 3. First the
basic techniques of DSM–oversampling and noise shaping–have been introduced.
Then different DSM architectures and their advantages/disadvantages have been
presented. Finally, important nonidealities of a CT DSM and their modelling have
been shown to set the requirements for the design of the building blocks. Corre-
sponding solutions for circuit nonidealities have also been discussed.
• Highly-digital quantization and conversion based on a VCO and digital gates have
been presented in Chap. 4. The main design challenge of VCO-based quantizers
is the VCO nonlinearity. Several existing closed-loop DSM ADCs and open-loop
oversampling ADCs have been introduced that solve the VCO nonlinearity.
• Chapter 5 has discussed the first two designs of the research work, which are
40 MHz-BW 12-bit DSM ADCs in 90 nm CMOS with 4th-order noise-shaping
and a 5-bit phase-type VCO-based quantizer. The linearity of the multi-bit feed-
174 9 Conclusions

back DAC has been solved by digital LUT-based calibration or transistor intrinsic
matching.
• A two-step open-loop VCO-based ADC architecture has been proposed and imple-
mented in Chap. 6. Only two frequency-type VCO-based quantizers and one
current-steering DAC are needed for the ADC design. The VCO nonlinearities
have been mitigated by distortion cancellation and input-swing reduction tech-
niques. The DAC matching requirement is released by the implicit DEM of the
quantizer. Because of the compact architecture and the highly-digital building
blocks with relaxed performance requirements, the proposed ADC prototype in
40 nm CMOS reaches an excellent FoM of 42fJ/Step.
• Chapter 7 has investigated the VCO-based MASH ΔΣ ADC architecture. A
power-efficient nonlinearity-cancellation technique in a 0-ΔΣ MASH structure
has been analyzed. This chapter has covered the full design procedure of a 0–
2 MASH VCO-based ADC prototype. The measurement results in 40 nm CMOS
achieves a FoM of 35fJ/Step, which is the best among high-bandwidth(BW≥20 MHz)
ΔΣ ADCs.
• Conceptually, a VCO accumulates the excess phase like an ideal integrator. It is
therefore beneficial to shift the analog integrator to the time domain as well. The
principle of a fully-VCO-based integrator has been discussed in Chap. 8. Based on
this technique, a 0–2 MASH CT ΔΣ ADC with fully-VCO-based integrator and
quantizers has been presented, with circuit design details and silicon implementa-
tion results. This high-order noise shaping ADC is entirely digital and hence can
benefit from CMOS technology scaling.
Based on this research work, several conclusions can be draw for the design of
power-efficient highly-digital ADCs.
• In future communications, the RF and analog parts will be simpler and the digital
core will become more important and more powerful. As a result, ADCs will be
closer to the receiver antenna and play a more critical role in the overall perfor-
mance of the whole system.
• With technology scaling, analog circuits suffering from the reduced voltage
dynamic range cannot fully benefit from this scaling. So digitization is one impor-
tant design trend for ADCs, for both speed and power considerations. There will
be less analog blocks, more digital blocks (e.g. digital calibration) and semi-digital
blocks (e.g. comparator, DAC and VCO) used in future ADC design.
• Using an advanced technology is always beneficial for high-bandwidth low-power
DSM ADC designs, especially for highly-digital architectures. At the same time
several problems become more severe when using these technologies, including
a low supply voltage, a low amplifier gain and bulk/gate leakage currents. To
resolve such challenges, innovations in both ADC architectures and circuit blocks
are required, as presented in this book.
• VCO-based ADC is one of the most promising candidates due to its simple struc-
ture and inherent 1st-order noise-shaping characteristic. However, the performance
of VCO-based ADCs is severely limited by the inherent nonlinearity of the VCOs,
which needs to be solved by design innovations.
9.1 Summary and Conclusions 175

coarse VCO-based quantizer (N bit) DAGC


Full input swing, nonlinear

DVGA
Path 1

DAC Path 2 CLK


CLK (N bit) D_OUT_C Distortion
Detector

D_OUT_F
A_IN D_OUT

fine VCO-based quantizer (M bit)


Small input swing, linear

Fig. 9.1 Block diagram of the digital calibration for the two-step open-loop VCO-based ADC

• A power-efficient two-step open-loop VCO-based ADC architecture has been pro-


posed in this book to mitigate the VCO-based ADC nonlinearity.
• A power-efficient nonlinearity-cancellation technique in a 0-ΔΣ MASH ADC
structure has been proposed in this book to solve the performance bottleneck of
the VCO-based ADCs.
• The power-hungry analog integrators can also be replaced by VCOs. Therefore,
high-order CT ΔΣ ADCs can be implemented with fully-VCO-based integrators
and quantizers, without any analog amplifier.
• In this work, low-power (FoM as low as 35fJ/Step) highly-digital ΔΣ ADCs with
moderate accuracy (10–12 bit) have been demonstrated in 90 and 40 nm CMOS
technologies. They are suitable for next-generation communication systems with
performance-relaxed RF and analog front-ends.

9.2 Suggestions for Future Work

As presented in this book, research has been carried out on power-efficient highly-
digitally ΔΣ ADC designs, using tens of MHz bandwidth as example in this book.
However, this is only one brick of the whole wall. Several research directions can be
explored further in the future to improve the ADC specifications and to develop the
design closer to a real product.
• Digital calibration for the two-step open-loop VCO-based ADC. In the two-
step open-loop VCO-based ADC presented in Chap. 6, the linearity performance
of the whole ADC depends on the matching of two paths. Now the gain of one path
is tuned manually, by adjusting the VCO gain (coarse tuning) and the DAC bias
current (fine tuning). Since the outputs of both paths are digital, digital calibra-
tion can easily be included to solve the path matching automatically, as shown in
Fig. 9.1. First, a digital harmonic detector is used to sense the ADC output distor-
tion. Then a simple decision unit is assigned to determine whether to increase or
decrease the gain of the DAC fine quantizer path. For the gain tuning, two methods
176 9 Conclusions

Traditional MASH

VCO-based MASH

Fig. 9.2 Topology comparison between the traditional MASH ADC and the presented VCO-based
MASH ADC

Fig. 9.3 Improved 1–1 MASH structure for VCO-based ADC

Fig. 9.4 Block diagram of a purely-time-domain ΔΣ ADC

can be applied. One option is to add some digital control bits to the DAC and the
fine VCO, which are controlled by the decision unit outputs. Another choice is
to insert a digital reconfigurable-gain stage between the coarse quantizer and the
adder, and the gain is set by the decision unit. This digital calibration scheme is
similar to the noise cancellation filter used in CT cascade DSM ADCs [1] and the
complex image rejection circuit used in a low-IF receiver [2].
9.2 Suggestions for Future Work 177

• Bandpass/Quadrature DSM ADC for low-IF, IF and RF signal A/D conver-


sions. Today most DSMs are designed for baseband signal A/D conversions.
With the increasing requirement for system programmability and the advances
in receiver architectures, more and more A/D conversions are done in the IF or
even the RF in the receiver signal chain. Among all ADC topologies, the DSM
is the only architecture which can do bandpass signal A/D conversion, making
it a perfect choice for IF and RF signal conversions. If the central frequency is
not so high, the basic resonator can still be implemented by local feedback across
two active-RC or gm-C integrators. If the IF/RF signal frequency is much higher,
then an active-LC resonator is needed in the design. In such a design, high-quality
on-chip inductors is a design challenge. VCO-based bandpass ADC for RF sam-
pling has also been reported [3] without feedback loop, showing design merit of
time-domain quantization in bandpass ADCs. One important difference between
the bandpass and the baseband A/D converters is that the bandpass ADCs are much
more sensitive to clock jitter. A lot of research is still needed to achieve bandpass
A/D conversion with a high central frequency and a high conversion accuracy.
• Improved MASH structure for VCO-based ADC. Considering the ADC designs
presented from Chaps. 6 to 8, although the first VCO-based stage is 1st-order noise
shaped, it cannot provide additional noise shaping for the whole ADC, therefore
these prototypes are regarded as 0–1 or 0–2 MASH ADCs. The reason is sum-
marized in Fig. 9.2. For the presented VCO-based MASH ADC, the input of the
second stage is E 1 (1 − z −1 ) instead of E 1 for the traditional MASH structure. To
cancel out E 1 at the final output, the digital filter H2 should be set to 1 instead of
N T F1 . As a result, for the quantization noise E 2 of the second stage, only N T F2
provides a shaping capability. To improve the ADC performance, an improved
VCO-based MASH structure as shown in Fig. 9.3 can be implemented. If the
phase error of the first VCO-based stage can be extracted, the new structure is
similar to a conventional 1–1 MASH ADC, where the digital filter H2 is set as
N T F1 . The final output V can be expressed as:

V = ST F1 · ST F2 · U + N T F1 · N T F2 · E 2 (9.1)

As observed from the above analysis, an important advantage of this topology is


that a higher-order noise shaping can be achieved by cascading open-loop VCO-
based stages. Therefore, the power-hungry analog integrators can be removed,
which is beneficial for high-speed and low-power applications. Reference [4] pro-
vides a potential design example of an improved 2nd-order 1–1 MASH VCO-based
ADC, in which the front-end PWM pre-coding can be removed.
• Purely-time-domain ΔΣ ADCs. Reviewing the fully-VCO-based ADC design
presented in Chap. 8, the VCO-based integrator transfers the voltage-based signal
into the phase domain, and then recovers it back to the voltage domain by a current-
steering DAC. Therefore, the signal is switched from the voltage domain to the
time domain back and forth, causing extra power to be dissipated in the transfer.
If the signal can be processed purely in the time domain, the whole ADC will be
more digital and power-efficient. The structure is illustrated in Fig. 9.4. After the
178 9 Conclusions

voltage-to-time conversion, a ΔΣ closed loop is implemented in the time domain.


References [5, 6] provide two potential design examples, both of which utilize a
VCO as the voltage-to-time converter (VTC). To realize the ΔΣ function entirely
in the time domain, [5] implements a phase interpolation, while [6] adopts a up-
down counter. In this kind of structure, the ADC performance is limited by the
nonidealities of the front-end VTC.

This research work has achieved and demonstrated power-efficient highly-digital


DSM A/D conversions in nanoscale CMOS technologies.

References

1. L.J. Breems, R. Rutten, G. Wetzker, A cascaded continuous-time ΔΣ modulator with 67-dB


dynamic range in 10-MHz bandwidth. IEEE J. Solid State Circuits 39(12), 2152–2160 (2004)
2. S. Lerstaveesin, Bang-Sup song, a complex image rejection circuit with sign detection only.
IEEE J. Solid State Circuits 41(12), 2693–2702 (2006)
3. Y.G. Yoon, S. Cho, A 1.5-GHz 63dB SNR 20mW direct RF sampling bandpass VCO-based
ADC in 65nm CMOS, in Symposium on VLSI Circuits, vol. 2009 (Kyoto, Japan, 2009), pp.
270–271
4. P. Gao, X. Xing, J. Craninckx, G. Gielen, Design of an intrinsically-linear double-VCO-based
ADC with 2nd-order noise shaping, in design. Autom. Test Eur. Confer. Exhib. (DATE) 2012,
1215–1220 (2012)
5. M. Amin, B. Leung, Design techniques for linearity in time-based SD analog-to-digital converter.
IEEE Trans. Circuits Syst. II Expr. Briefs 63(5), 433–437 (2016)
6. A. Babaie-Fishani, P. Rombouts, True high-order VCO-based ADC. Electr. Lett. 51(1), 23–25
(2015)
Index

A MASH, 45
Active adder, 94 multi-bit, 49
ADC, 13 multi-bit DAC mismatch, 62
AAF, 16 noise leakage, 46
accuracy, 14 noise shaping, 38
dynamic, 15 nonidealities, 53
static, 15 NRZ DAC, 41, 87, 94, 103, 119
applications, 28 NTF, 38
comparison, 26 NTF zero, 52, 85, 101
delta-sigma, 24, 38 OSR, 25, 40
active-RC integrator, 41, 85, 101 OTA parasitic, 56
capacitive local feedback, 101 out-of-band gain, 44
continuous-time, 25, 41 oversampling, 38
DAC clock jitter, 59 quantization bit, 25, 40
DAC ISI, 62 quantizer parasitic, 63
DAC mismatch, 50, 115, 141 RZ DAC, 41, 73, 94
decimation, 38 SC DAC, 41, 87
DEM, 50, 74, 75 SC integrator, 41
discrete-time, 25, 41 shaped SC DAC, 91
dithering, 44, 76 shaping order, 25, 40
DNCF, 45 single-bit, 48
ELD, 42, 63, 88 stability, 44, 76
feedback topology, 52 STF, 38, 130
feedforward topology, 51, 85 time constant variation, 58, 86
feedin path, 53, 85 toolbox, 45, 88
finite OTA gain, 54 zero order feedback, 42, 63, 72
finite OTA GBW, 55 digital inferface, 142
gm-C integrator, 41 encoder, 18
high order, 44 flash, 19, 110
hybrid topology, 52 averaging, 20
limit cycle, 44 folding, 20
limited OTA swing, 57 interpolation, 20
LUT calibration, 50, 78, 88 FoM, 16
180 Index

Nyquist, 24 N
oversampling, 18 Nonlinearity-cancellation robustness, 113,
pipelined, 22 131
MDAC, 22
power spectral density, 18
quantization error, 18
O
quantizer, 17
OTA, 89
SAR, 23
CMFB, 90
S/H, 17
current-sharing, 102
speed, 14
doublet, 90
time interleaving, 22, 24
feedforward compensated, 89, 102, 140
tradeoff, 27
LHP zero, 89, 140
two-step, 21, 110
over-ranger, 21 (folding) Cascode, 89
AFE, 28 Miller-compensated, 89, 102
AGC, 83 single-stage, 89

C P
Capacitor bank, 43, 58, 86, 139 Passive subtractor, 119, 136
CMOS technology, 3 Phase detector, 69, 162
amplifier gain degradation, 5 PVT variation, 117, 131, 139
headroom voltage reduction, 5 PWM, 79
scaling, 5 modulation depth, 79
Communications, 1 OCR, 79
low-power, 2
wireless, 2
wireline, 2
R
Receiver, 6
D BER, 30
DAC, 13, 21, 23, 24 block signal, 30
Delay matching, 114, 134 budget, 28
DSP, 28 crest factor, 29
DC offset, 29
direct-sampling, 7
I fading margin, 30
ICO, 76 heterodyne, 6
Implementations homodyne, 7
0–2 MASH VCO-based ADC, 143 low-IF, 7
DSM ADC with time-domain quantiza- noise figure, 29
tion, 95, 104 sensitivity, 30
fully-VCO-based 0-2 MASH ADC, 163 signal modulation, 30
two-step VCO-based ADC, 122 subsampling, 7
VGA, 31

L
LFSR, 76
S
LMS, 88
SA FF, 69, 137
LVDS, 95, 122, 143

M T
MIM capacitor, 91 TSPC FF, 69, 137
Index 181

V frequency-type, 69, 73, 74, 84, 110, 117,


VCO, 69, 117, 160 137
phase noise, 119, 137 implicit AAF, 85, 112
VCO-based ADC implicit DEM, 70, 75, 112, 141
closed-loop, 72 inherent noise shaping, 70
0- MASH, 128, 158 multi-phase counting, 68
Park, 74 nonlinearity, 71
residual-cancelling, 74 phase noise, 71
Straayer, 73 phase-type, 72, 74, 101
open-loop, 76 pseudo differential, 117
digital calibration, 76, 77 single-phase counting, 68
PWM precoding, 78 VGA, 83
two-step, 110
VCO-based integrator, 155
VCO-based quantizer, 68
clock jitter, 71 X
dual-input, 136 XOR, 69, 137

You might also like