0% found this document useful (0 votes)
49 views160 pages

Tronson Dissertation

Uploaded by

suliman
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)
49 views160 pages

Tronson Dissertation

Uploaded by

suliman
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/ 160

SINGLE-CHIP VLF MAGNETIC FIELD RECEIVER

A DISSERTATION
SUBMITTED TO THE DEPARTMENT OF ELECTRICAL
ENGINEERING
AND THE COMMITTEE ON GRADUATE STUDIES
OF STANFORD UNIVERSITY
IN PARTIAL FULFILLMENT OF THE REQUIREMENTS
FOR THE DEGREE OF
DOCTOR OF PHILOSOPHY

Andrew Tronson
May 2014
© 2014 by Andrew Henning Tronson. All Rights Reserved.
Re-distributed by Stanford University under license with the author.

This work is licensed under a Creative Commons Attribution-


Noncommercial 3.0 United States License.
http://creativecommons.org/licenses/by-nc/3.0/us/

This dissertation is online at: http://purl.stanford.edu/rt607vn7087

ii
I certify that I have read this dissertation and that, in my opinion, it is fully adequate
in scope and quality as a dissertation for the degree of Doctor of Philosophy.

Umran Inan, Primary Adviser

I certify that I have read this dissertation and that, in my opinion, it is fully adequate
in scope and quality as a dissertation for the degree of Doctor of Philosophy.

Ivan Linscott

I certify that I have read this dissertation and that, in my opinion, it is fully adequate
in scope and quality as a dissertation for the degree of Doctor of Philosophy.

Boris Murmann

Approved for the Stanford University Committee on Graduate Studies.


Patricia J. Gumport, Vice Provost for Graduate Education

This signature page was generated electronically upon submission of this dissertation in
electronic format. An original signed hard copy of the signature page is on file in
University Archives.

iii
Abstract

Electromagnetic waves in the VLF frequency band from 300 Hz to 30 kHz are used
to study the atmosphere, geolocate lightning, map subterranean features and pro-
vide navigation and timing signals. These waves can be generated by man-made
sources, such as the VLF transmitters operated by the US Navy to communicate
with submarines while they are submerged. They can also be generated by natural
phenomena. For example, a lightning strike generates an impulsive signal known as
a radio atmospheric. The interaction of these natural and man-made signals with
the ionosphere, the magnetosphere, and the earth’s electrical environment provides
valuable data for scientific research.
To facilitate this research, the VLF signals must be received and stored for further
analysis. Due to the unique nature of these signals, specialized receiver hardware is
required to receive them. Currently the most widely deployed VLF receiver is the
AWESOME receiver developed at Stanford University. This receiver offers excellent
data quality, but it has a power dissipation of around 60 Watts. The high power
dissipation can be a problem because some of the most desirable locations to deploy
receivers are remote locations, which are far away from power-lines and other sources
of electromagnetic interference. In these cases, the receiver must have a very low
power dissipation as it must operate on battery power for long periods of time. Low-
power VLF receivers for remote deployments have been developed, but their lower
power dissipation comes at the expense of data quality. The goal and challenge of
this research is to design a low-power receiver without sacrificing data quality.
To accomplish this goal, the first single-chip broadband VLF magnetic field re-
ceiver has been developed. The receiver consists of a low-noise amplifier (LNA) and

iv
an analog-to-digital converter (ADC) integrated on a single-chip. The LNA is im-
plemented using a low-impedance bipolar input stage followed by a variable gain
differential instrumentation amplifier. The ADC is implemented with a third-order
continuous-time delta-sigma modulator, which was selected for its implicit anti-alias
filtering capability and its robustness to mismatch and other non-ideal effects. The
receiver also includes an automatic biasing system that compensates for the large
temperature variations that are often encountered at remote deployment sites. The
receiver was fabricated in a 0.13 µm BiCMOS process and has a die area of 2.56 mm2 .
The LNA achieves a sensitivity of better than 1 fT/Hz1/2 using a standard six-
turn 4.9 meter square loop antenna. It also has a peak spurious-free dynamic range
of up to 104.02 dB and a 3 dB bandwidth that extends from 170 Hz to over 100 kHz.
The power dissipation of the LNA is approximately 908 µW. The on-chip delta-sigma
ADC has an effective resolution of 12.40 bits and a spurious-free dynamic range of
over 93 dB. The power dissipation of the ADC is roughly 640 µW. The full receiver
consists of the combination of the LNA and the ADC and has a total power dissipation
of approximately 1.55 mW.
A side-by-side comparison of field data from the single-chip receiver and the AWE-
SOME receiver reveals that the data quality of the single-chip receiver is at least as
good. Further, the single-chip receiver has a power dissipation that is over 30 times
less than the current low-power receiver. The high data quality, low power dissipation
and small size make the single-chip VLF magnetic field receiver especially well suited
for remote VLF data collection.

v
Acknowledgements

Although there is only one person listed on the title page of this dissertation, there
are many more who deserve recognition for helping to make this project a success.
First and foremost, I would like to thank my research adviser Professor Umran Inan.
Throughout my time at Stanford he has provided guidance, encouragement and unwa-
vering support. I’d also like to thank Dr. Ivan Linscott who served as the co-adviser
for this project. His technical contributions and feedback were invaluable. Addi-
tionally, I’d like to thank Professor Boris Murmann who reviewed this dissertation
and provided many helpful suggestions. I would also like to thank Ahmad Bahai and
Martin Lee at Texas Instruments. Not only did Texas Instruments generously provide
funding to support this research, but they also invited us to use their well-equipped
labs in Santa Clara to test the single-chip receiver described in this dissertation. Last
but not least, I would like to thank my family and friends for their support and
encouragement throughout my academic career.

Andrew Tronson
Stanford, California
May 30, 2014

vi
Contents

Abstract iv

Acknowledgements vi

1 Introduction 1
1.1 VLF Research . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Organization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

2 Background 7
2.1 VLF Receiver Front-End . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.1.1 Antenna Design . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.1.2 Receiver Design . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.1.3 Transformer Design . . . . . . . . . . . . . . . . . . . . . . . . 13
2.1.4 Amplifier Design . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.1.5 Receiver Testing Methods . . . . . . . . . . . . . . . . . . . . 17
2.2 Delta-Sigma Analog-to-Digital Conversion . . . . . . . . . . . . . . . 19
2.2.1 Analog-to-Digital Conversion . . . . . . . . . . . . . . . . . . 19
2.2.2 Oversampling . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.2.3 Delta-Sigma Modulators . . . . . . . . . . . . . . . . . . . . . 25
2.2.4 Continuous-Time Delta-Sigma Modulators . . . . . . . . . . . 30

3 Receiver Overview 34
3.1 Performance Metrics . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

vii
3.1.1 Bandwidth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.1.2 Gain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.1.3 Sensitivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.1.4 Spurious-Free Dynamic Range . . . . . . . . . . . . . . . . . . 36
3.2 Existing VLF Receivers . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.2.1 AWESOME Receiver . . . . . . . . . . . . . . . . . . . . . . . 38
3.2.2 Penguin Receiver . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.3 Receiver Architecture . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.3.1 Traditional Architecture . . . . . . . . . . . . . . . . . . . . . 40
3.3.2 Single-Chip Architecture . . . . . . . . . . . . . . . . . . . . . 42
3.4 Design Specifications . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

4 Circuit Implementation 47
4.1 LNA Implementation . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
4.1.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
4.1.2 Input Stage . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.1.3 Gain Stage . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
4.1.4 Auto-Bias Circuit . . . . . . . . . . . . . . . . . . . . . . . . . 59
4.2 ADC Implementation . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
4.2.1 Loop Filter Design . . . . . . . . . . . . . . . . . . . . . . . . 63
4.2.2 Modulator Architecture . . . . . . . . . . . . . . . . . . . . . 67
4.2.3 Circuit Blocks . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
4.3 OTA Implementation . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
4.3.1 Architecture Comparison . . . . . . . . . . . . . . . . . . . . . 77
4.3.2 Two-Stage OTA . . . . . . . . . . . . . . . . . . . . . . . . . . 78
4.3.3 Design Methodology . . . . . . . . . . . . . . . . . . . . . . . 85
4.4 Receiver Implementation . . . . . . . . . . . . . . . . . . . . . . . . . 87
4.4.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
4.4.2 Biasing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
4.4.3 Noise Coupling . . . . . . . . . . . . . . . . . . . . . . . . . . 89
4.5 Chip Layout . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

viii
5 Measurement Results 93
5.1 LNA Measurement Results . . . . . . . . . . . . . . . . . . . . . . . . 94
5.1.1 Test Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
5.1.2 Frequency Response . . . . . . . . . . . . . . . . . . . . . . . 96
5.1.3 Sensitivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
5.1.4 Linearity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
5.1.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
5.2 ADC Measurement Results . . . . . . . . . . . . . . . . . . . . . . . . 105
5.2.1 Test Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
5.2.2 Output Spectrum . . . . . . . . . . . . . . . . . . . . . . . . . 107
5.2.3 Dynamic Performance . . . . . . . . . . . . . . . . . . . . . . 108
5.2.4 Anti-Alias Filtering . . . . . . . . . . . . . . . . . . . . . . . . 112
5.2.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
5.3 Receiver Measurement Results . . . . . . . . . . . . . . . . . . . . . . 114
5.3.1 Test Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
5.3.2 Dynamic Performance . . . . . . . . . . . . . . . . . . . . . . 115
5.3.3 Noise Coupling . . . . . . . . . . . . . . . . . . . . . . . . . . 121
5.3.4 Anti-Alias Filtering . . . . . . . . . . . . . . . . . . . . . . . . 126
5.3.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
5.4 Field Test Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
5.4.1 Test Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
5.4.2 Stanford Field Test . . . . . . . . . . . . . . . . . . . . . . . . 131
5.4.3 Santa Cruz Field Test . . . . . . . . . . . . . . . . . . . . . . 132
5.4.4 Receiver Comparison . . . . . . . . . . . . . . . . . . . . . . . 133

6 Conclusion 138
6.1 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139

Bibliography 141

ix
List of Tables

2.1 VLF Loop Antenna Designs . . . . . . . . . . . . . . . . . . . . . . . 11

3.1 Single-Chip Receiver Design Specifications . . . . . . . . . . . . . . . 45

4.1 OTA Architecture Performance Comparison . . . . . . . . . . . . . . 77

5.1 LNA Measurement Results Summary . . . . . . . . . . . . . . . . . . 105


5.2 ADC Measurement Results Summary . . . . . . . . . . . . . . . . . . 113
5.3 Receiver Measurement Results Summary . . . . . . . . . . . . . . . . 129

x
List of Figures

1.1 Lightning Example . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2


1.2 VLF Data Example . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 VLF Receiver Map . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

2.1 Loop Antenna Model . . . . . . . . . . . . . . . . . . . . . . . . . . . 8


2.2 Receiver Front-End Model . . . . . . . . . . . . . . . . . . . . . . . . 12
2.3 Transformer Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.4 Receiver Frequency Response . . . . . . . . . . . . . . . . . . . . . . 15
2.5 Receiver Current Injection Circuit . . . . . . . . . . . . . . . . . . . . 18
2.6 ADC Quantizer Transfer Curve . . . . . . . . . . . . . . . . . . . . . 21
2.7 ADC Quantization Error . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.8 ADC Quantization Spectrum Example . . . . . . . . . . . . . . . . . 23
2.9 ADC Oversampling Example . . . . . . . . . . . . . . . . . . . . . . . 24
2.10 Delta-Sigma Modulator Block Diagram . . . . . . . . . . . . . . . . . 25
2.11 Delta-Sigma Modulator Linear Model . . . . . . . . . . . . . . . . . . 26
2.12 Delta-Sigma Modulator Time-Domain Simulation . . . . . . . . . . . 29
2.13 Delta-Sigma Modulator Output Spectrum . . . . . . . . . . . . . . . 30
2.14 Continuous-Time Delta-Sigma Modulator Block Diagram . . . . . . . 31
2.15 Continuous-Time Delta-Sigma Modulator AAF Response . . . . . . . 32

3.1 Spurious-Free Dynamic Range Example . . . . . . . . . . . . . . . . . 37


3.2 Traditional Receiver Architecture . . . . . . . . . . . . . . . . . . . . 40
3.3 Single-Chip Receiver Architecture . . . . . . . . . . . . . . . . . . . . 43

xi
4.1 LNA Top-Level Schematic . . . . . . . . . . . . . . . . . . . . . . . . 48
4.2 LNA Input Stage Schematic . . . . . . . . . . . . . . . . . . . . . . . 50
4.3 LNA Input Stage Noise Model . . . . . . . . . . . . . . . . . . . . . . 53
4.4 LNA Gain Stage Schematic . . . . . . . . . . . . . . . . . . . . . . . 55
4.5 Ideal Feedback Block Diagram . . . . . . . . . . . . . . . . . . . . . . 59
4.6 LNA Auto-Bias Schematic . . . . . . . . . . . . . . . . . . . . . . . . 61
4.7 LNA Auto-Bias Error Amplifier Schematic . . . . . . . . . . . . . . . 62
4.8 LNA Temperature Sweep Simulation . . . . . . . . . . . . . . . . . . 63
4.9 ADC Modulator Block Diagram . . . . . . . . . . . . . . . . . . . . . 64
4.10 ADC Modulator NTF and STF . . . . . . . . . . . . . . . . . . . . . 65
4.11 ADC Modulator Schematic . . . . . . . . . . . . . . . . . . . . . . . . 67
4.12 ADC Modulator Noise Model . . . . . . . . . . . . . . . . . . . . . . 70
4.13 ADC Comparator Schematic . . . . . . . . . . . . . . . . . . . . . . . 72
4.14 ADC Flipflop Schematic . . . . . . . . . . . . . . . . . . . . . . . . . 74
4.15 ADC Feedback DAC Schematic . . . . . . . . . . . . . . . . . . . . . 75
4.16 ADC Tunable Capacitor Schematic . . . . . . . . . . . . . . . . . . . 76
4.17 OTA Fully-Differential Schematic . . . . . . . . . . . . . . . . . . . . 79
4.18 OTA Noise Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
4.19 OTA Common-Mode Feedback Schematic . . . . . . . . . . . . . . . 84
4.20 OTA Single-Ended Schematic . . . . . . . . . . . . . . . . . . . . . . 86
4.21 Receiver Top-Level Schematic . . . . . . . . . . . . . . . . . . . . . . 88
4.22 Current Source Schematic . . . . . . . . . . . . . . . . . . . . . . . . 89
4.23 Single-Chip Receiver Die Photo . . . . . . . . . . . . . . . . . . . . . 92

5.1 Single-Chip Receiver Test Board . . . . . . . . . . . . . . . . . . . . . 94


5.2 LNA Test Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
5.3 LNA Frequency Response . . . . . . . . . . . . . . . . . . . . . . . . 96
5.4 LNA Output Noise Spectrum . . . . . . . . . . . . . . . . . . . . . . 98
5.5 LNA Sensitivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
5.6 LNA SFDR vs. Input Amplitude . . . . . . . . . . . . . . . . . . . . 101
5.7 LNA Output Spectrum Max SFDR Mode 0 . . . . . . . . . . . . . . 102

xii
5.8 LNA Output Spectrum Max SFDR Mode 1 . . . . . . . . . . . . . . 102
5.9 LNA Output Spectrum Max SFDR Mode 2 . . . . . . . . . . . . . . 103
5.10 LNA Output Spectrum Max SFDR Mode 3 . . . . . . . . . . . . . . 103
5.11 LNA SFDR vs. Input Frequency . . . . . . . . . . . . . . . . . . . . 104
5.12 ADC Test Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
5.13 ADC Output Spectrum . . . . . . . . . . . . . . . . . . . . . . . . . . 108
5.14 ADC Performance vs. Input Amplitude . . . . . . . . . . . . . . . . . 109
5.15 ADC Output Spectrum In-Band Max SFDR . . . . . . . . . . . . . . 111
5.16 ADC Performance vs. Input Frequency . . . . . . . . . . . . . . . . . 111
5.17 ADC Anti-Alias Filtering . . . . . . . . . . . . . . . . . . . . . . . . . 112
5.18 Receiver Test Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
5.19 Receiver Performance vs. Input Amplitude Mode 0 . . . . . . . . . . 116
5.20 Receiver Performance vs. Input Amplitude Mode 1 . . . . . . . . . . 116
5.21 Receiver Performance vs. Input Amplitude Mode 2 . . . . . . . . . . 117
5.22 Receiver Performance vs. Input Amplitude Mode 3 . . . . . . . . . . 117
5.23 Receiver Output Spectrum Max SFDR Mode 0 . . . . . . . . . . . . 119
5.24 Receiver Output Spectrum Max SFDR Mode 1 . . . . . . . . . . . . 119
5.25 Receiver Output Spectrum Max SFDR Mode 2 . . . . . . . . . . . . 120
5.26 Receiver Output Spectrum Max SFDR Mode 3 . . . . . . . . . . . . 120
5.27 Receiver Performance vs. Input Frequency Mode 0 . . . . . . . . . . 121
5.28 Receiver Performance vs. Input Frequency Mode 1 . . . . . . . . . . 122
5.29 Receiver Performance vs. Input Frequency Mode 2 . . . . . . . . . . 122
5.30 Receiver Performance vs. Input Frequency Mode 3 . . . . . . . . . . 123
5.31 Receiver Noise Coupling Mode 0 . . . . . . . . . . . . . . . . . . . . . 124
5.32 Receiver Noise Coupling Mode 1 . . . . . . . . . . . . . . . . . . . . . 124
5.33 Receiver Noise Coupling Mode 2 . . . . . . . . . . . . . . . . . . . . . 125
5.34 Receiver Noise Coupling Mode 3 . . . . . . . . . . . . . . . . . . . . . 125
5.35 Receiver Anti-Alias Filtering Mode 0 . . . . . . . . . . . . . . . . . . 127
5.36 Receiver Anti-Alias Filtering Mode 1 . . . . . . . . . . . . . . . . . . 127
5.37 Receiver Anti-Alias Filtering Mode 2 . . . . . . . . . . . . . . . . . . 128
5.38 Receiver Anti-Alias Filtering Mode 3 . . . . . . . . . . . . . . . . . . 128

xiii
5.39 Field Test Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
5.40 Stanford Field Test Spectrogram . . . . . . . . . . . . . . . . . . . . 132
5.41 Santa Cruz Field Test Spectrogram . . . . . . . . . . . . . . . . . . . 134
5.42 Receiver Spectrogram Comparison . . . . . . . . . . . . . . . . . . . . 135
5.43 Receiver Average Spectrum Comparison . . . . . . . . . . . . . . . . 136
5.44 Receiver Average Spectrum Comparison Zoomed . . . . . . . . . . . 137

xiv
Chapter 1

Introduction

1.1 VLF Research


Very low frequency (VLF) research covers a wide range of topics that includes the
study of lightning, the atmosphere and other topics related to very low frequency
electromagnetic waves [43, 53, 24, 35]. VLF refers to a band of frequencies that
extends from 3 kHz to 30 kHz. Often in the course of these studies, signals that fall
slightly above and below the VLF frequency band are also of interest. This extended
frequency range includes the ELF (extremely low frequency) band that covers 300
Hz to 3 kHz and the LF (low frequency) band that covers 30 kHz to 300 kHz. For
simplicity, in this dissertation the term VLF is used to refer to the frequency range
from 300 Hz to 50 kHz, which includes the ELF frequency band, the VLF frequency
band and the low end of the LF frequency band.
There are many sources of VLF radiation. These sources can be divided into
two basic groups: man-made sources and naturally occurring sources. The man-
made signals are generated by transmitters that broadcast in the VLF band. For
example, VLF transmitters are used extensively by the US Navy to communicate
with submarines while they are submerged. VLF signals are particularly well suited
for this application because their low frequency allows them to penetrate into the
water. The most common source of naturally occurring VLF radiation is lightning.
In addition to thunder and a flash of light, a lightning strike also emits an impulsive

1
CHAPTER 1. INTRODUCTION 2

Figure 1.1: Lightning example showing VLF wave propagation in the Earth-
ionosphere waveguide (red) and a wave escaping into the magnetosphere (green).

electromagnetic signal known as a radio atmospheric, or sferic for short [48]. A sferic
has frequency components that reside primarily in the VLF frequency band.
An example of VLF wave propagation that results from a lightning strike is shown
in Figure 1.1. In this example, a lightning strike occurs near the north pole and
generates a sferic, the propagation direction of which is shown with the red line in
the illustration. Due to the low frequency of the signal, it reflects off the earth and the
ionosphere. As a result, the signal can travel extremely long distances in the so-called
Earth-ionosphere waveguide. It is not uncommon to detect sferics from a lightning
strike that occurred over 10,000 kilometers away [20]. In some cases the VLF radiation
from a lightning strike can escape out of the ionosphere and into the magnetosphere,
as shown with the green arrow in the figure. In this case, the wave travels along the
magnetic field lines and can be detected at a geomagnetically conjugate point in the
southern hemisphere.
Figure 1.2 shows an example spectrogram of VLF data collected by the Stanford
VLF research group in Chistochina, Alaska [19]. A spectrogram is a plot which shows
time on the horizontal axis and frequency on the vertical axis. The color represents
the amplitude of the signal, where red corresponds to the largest signals and blue
corresponds to the smallest signals. Essentially, a spectrogram shows the temporal
variation of the dynamic spectrum of the signal.
The top plot in the figure shows 60 seconds of data over the frequency range of
0 Hz to 50 kHz, which is referred to as broadband VLF data because it covers the
CHAPTER 1. INTRODUCTION 3

Figure 1.2: Example VLF data collected in Chistochina, Alaska in 2007. The top
spectrogram shows broadband VLF data, which includes VLF transmitters and sfer-
ics. The bottom spectrogram zooms in on a whistler.

entire VLF frequency band. There are many different VLF signals visible in this
spectrogram. Several strong VLF transmitters are visible between 20 kHz and 25
kHz. The transmitters broadcast in a small bandwidth around a constant frequency
so they are easy to identify as horizontal lines in a spectrogram. The vertical lines
that extend from roughly 5 kHz to 30 kHz are sferics, which are the impulsive signals
generated by lightning strikes. Hundreds of sferics are visible in this data sample.
This large number of sferics is not uncommon, as the average global lighting flash
rate is approximately 50 flashes per second [16].
There is also an orange patch of signals in the lower right corner of the top
spectrogram. The bottom plot zooms in on this area, showing only the data from 50
to 60 seconds over the frequency range of 0 Hz to 6 kHz. The curved signals in the
middle of this zoomed spectrogram are called whistlers. These signals are detected
when a VLF wave from a lightning strike escapes through the ionosphere into the
magnetosphere, travels along the magnetic field lines and is received at a conjugate
point in the opposite hemisphere. The signal has a curved shape due to the fact that
CHAPTER 1. INTRODUCTION 4

Figure 1.3: Map of VLF receivers deployed by the Stanford VLF research group.

the phase and group velocities in the ionized magnetospheric medium are functions
of frequency. These signals are known as whistlers because they make a whistling
sound when played on an audio speaker [30].
To facilitate the study of VLF phenomena, the Stanford VLF research group has
developed several types of VLF receivers. These receivers are all custom designed for
the unique requirements of receiving VLF waves. The two current models are the
AWESOME receiver and the Penguin receiver. The AWESOME receiver is a high-
performance VLF receiver, which receives signals with the highest data quality [19].
The Penguin receiver is a low-power VLF receiver designed for remote deployments at
unmanned sites [32]. The Stanford VLF research group has deployed these receivers in
dozens of locations around the world to collect data for scientific research. Figure 1.3
shows a map of the VLF receiver locations [36]. The receivers are deployed on all
seven continents, including Antarctica. Many of the receivers are located in remote
areas to minimize the amount of interference from power-lines and other sources of
noise that are often found in urban environments.
The goal of this work is to develop the next-generation VLF receiver. This receiver
should combine the high data quality of the AWESOME receiver with the minimal
power dissipation of the Penguin receiver. At the same time, the next generation
receiver is to be packaged in a single chip, to drastically reduce the size of the receiver
CHAPTER 1. INTRODUCTION 5

and open up new potential applications.

1.2 Contributions
The primary contributions of this work are:

1. Design, implementation and validation of the first single-chip broadband VLF


magnetic field receiver.

2. Implementation of the first integrated VLF magnetic field low-noise amplifier


that simultaneously achieves over 90 dB spurious-free dynamic range and under
1 fT/Hz1/2 sensitivity.

3. Demonstration of an automatic biasing system that increases the robustness of


the low-noise amplifier to temperature variations.

4. Reduction of the signal path power consumption by over 30 times compared to


the next-best low-power VLF receiver design.

1.3 Organization
This first chapter has provided a basic introduction to VLF science. It included
an overview of the current network of VLF receivers deployed by the Stanford VLF
research group and a description of some example VLF data that was collected with
one of these receivers. It also included a summary of the primary contributions of
this work. The remaining chapters are organized as follows.
Chapter 2 covers relevant background information and is divided into two parts.
The first part gives a high-level overview of the various components of a VLF magnetic
field receiver, which include the antenna, transformer and amplifier. This section also
describes the methods used to test the receiver in a lab environment. The second
part of this chapter reviews the basics of analog-to-digital conversion, with a focus
on oversampling and delta-sigma modulation.
CHAPTER 1. INTRODUCTION 6

Chapter 3 gives a more detailed description of the VLF magnetic field receiver.
It begins by discussing the key performance metrics, including the bandwidth, gain,
sensitivity and spurious-free dynamic range. This is followed by a summary of the
past VLF magnetic field receivers developed by the Stanford VLF research group.
This chapter also includes a detailed discussion of the proposed single-chip receiver
architecture and a comparison with the traditional receiver architecture. This chapter
concludes by outlining the design specifications for the single-chip receiver.
Chapter 4 covers the circuit implementation of the single-chip receiver. This
chapter describes the transistor-level design of all of the components of the receiver,
including the low-noise amplifier and the analog-to-digital converter. It also provides
a section dedicated to the design of the operational transconductance amplifier, which
is a block that is used throughout the receiver. In addition, this chapter discusses
several high-level receiver design issues, such as biasing and noise coupling.
Chapter 5 describes the measurement results of the single-chip receiver. It includes
lab measurements of the performance of the low-noise amplifier and analog-to-digital
converter, as well as measurements of the full receiver. Additionally, this chapter
describes the field tests that were performed at Stanford University and at a quiet
site in the Santa Cruz Mountains. This chapter concludes with a comparison of field
test data between the single-chip receiver and the existing high-performance VLF
receiver.
Chapter 6 summarizes the results of this work. This chapter provides a brief
summary of the receiver design and the measurement results. It also includes a
section that describes possible avenues for further research related to the single-chip
receiver.
Chapter 2

Background

2.1 VLF Receiver Front-End

2.1.1 Antenna Design


A receiver can use one of two methods to detect an electromagnetic wave. It can use
an electric field antenna, which is sensitive to the electric field of the received wave.
Alternately, it can use a magnetic field antenna, which is sensitive to the magnetic
field of the received wave. In the VLF frequency range the preferred method is to
use an antenna that is sensitive to the magnetic field of the received signal [39].
The primary reason that a magnetic field receiver is used is because it has better
noise performance at the low-end of the VLF frequency range when compared to a
similarly designed electric field receiver. Additionally, a magnetic field receiver is
simpler to calibrate as it is much less sensitive to nearby structures than its electric
field counterpart.
The magnetic field antenna used in a standard VLF receiver is an air-core loop
antenna. A model of the antenna is shown in Figure 2.1. It consists of a voltage source
Va that represents the antenna voltage induced by the incident wave, noise generator
va2 that represents the thermal noise of the antenna, resistor Ra that represents the
resistance of the loop, inductor La that represents the inductance of the loop, capacitor
Ca that represents the capacitance of the loop and resistor Rr that represents the

7
CHAPTER 2. BACKGROUND 8

Rr Ra La

va2
Ca
Va

Figure 2.1: Model of the air-core loop antenna.

radiation resistance of the antenna.


In the VLF frequency band the radiation resistance is negligible because the loop
is small compared to the wavelength of the received waves. The radiation resistance
of the loop can be approximated by
 2
4 NA
Rr = 320π (2.1)
λ2

where N is the number of turns, A is the loop area and λ is the wavelength of the
received wave [1]. With a frequency of 100 kHz and a standard six-turn 4.9 meter
square antenna, the radiation resistance is approximately 8×10-6 Ω. This value is
much smaller than the loop resistance of 1 Ω and can thus be safely ignored.
Similarly, the loop capacitance Ca is small and has a negligible effect on the
antenna performance within the target bandwidth of the receiver. With the radiation
resistance and the loop capacitance removed, the antenna impedance is given by

Za = Ra + jωLa (2.2)

which only depends on the loop resistance Ra and the loop inductance La .
A key metric of the loop antenna is the antenna turnover frequency, which is given
by
Ra
fa = (2.3)
2πLa
CHAPTER 2. BACKGROUND 9

At frequencies below fa the antenna impedance is resistive. At frequencies above fa


the impedance is inductive. All of the loop antennas deployed by the Stanford VLF
research group have a 1 Ω, 1 mH impedance, which results in a turnover frequency
of approximately 159 Hz.
The voltage induced in the loop antenna by an incident electromagnetic wave
can be calculated by using Faraday’s law of induction, which states that the induced
voltage is proportional to the rate of change of the magnetic field [11]. If a sinusoidal
signal is assumed, then the induced voltage is

Va = jωN ABω cos (Θ) (2.4)

where ω is the radian frequency of the wave, Bω is the amplitude of the wave and Θ is
the angle of incidence. Note that the induced voltage in the antenna is proportional
to frequency. In all of the following analysis it is assumed that the incident wave
is perpendicular to the antenna, so the cosine term is dropped. The magnetic field
amplitude can be converted to an equivalent electric field amplitude using

Eω = cBω (2.5)

where c is the speed of light (3×108 m/s).


The antenna has two regions of operation. The first is the region below the
antenna turnover frequency fa , where the antenna impedance is resistive. In this
region the output current from the antenna is proportional to frequency because the
impedance is constant and the induced voltage is proportional to frequency. The
second region of operation is above fa , where the antenna is inductive. In this region
the output current from the antenna is flat with frequency because the impedance is
proportional to frequency and the induced voltage is proportional to frequency. The
power of naturally occurring VLF signals is roughly constant over the entire VLF
band, so it is desirable to operate the receiver in this flat region of the antenna above
fa . The 1 Ω, 1 mH loop antenna is used primarily because its turnover frequency falls
below the start of the VLF band at 300 Hz.
Noise from the antenna originates from the thermal noise of the loop resistance
CHAPTER 2. BACKGROUND 10

Ra . The noise power of the antenna can be expressed as

va2 = 4kT Ra ∆f (2.6)

where k is the Boltzmann constant, T is the temperature and ∆f is the bandwidth


of the noise calculation [25]. This expression for the antenna noise can be used
to calculate the sensitivity of the antenna when combined with Equation 2.4 for
the induced voltage from an incident wave. Setting the noise voltage equal to the
induced voltage and solving for the magnetic field amplitude Bω results in an antenna
sensitivity of √
4kT Ra
Sa = (2.7)
ωN A
The antenna sensitivity is a measure of the minimum detectable signal and is defined
as the magnetic field amplitude that results in a 0 dB signal-to-noise ratio at the
output of the antenna when calculated in a 1 Hz bandwidth [39]. For a fixed antenna
impedance, the antenna sensitivity can be improved by increasing the area of the
loop. As a result, larger antennas generally have better sensitivities than smaller
antennas. The sensitivity of the antenna decreases with increasing frequency. It is
useful to remove this frequency dependence from the expression to aid in comparing
the performance of different antennas. For this purpose, the normalized antenna
sensitivity is defined as √
4kT Ra
S0 = (2.8)
2πN A
The actual sensitivity can be calculated from the normalized sensitivity by dividing
by the desired frequency.
The loop antennas deployed by the Stanford VLF research group have a 1 Ω, 1
mH impedance. This impedance was chosen because the antenna turnover frequency
falls well below the low-end of the VLF frequency band. An iterative approach can
be used to generate antenna designs of various sizes and shapes that fit the 1 Ω, 1
mH impedance. Table 2.1 shows a selection of example designs and their normalized
sensitivities [39]. The size of the square antennas refers to the length of each side of
the square. The triangle antennas are right isosceles triangles where the size refers to
CHAPTER 2. BACKGROUND 11

Table 2.1: VLF loop antenna designs with 1 Ω, 1 mH impedance.

Shape Size Wire N Ra La fa A S0


m AWG Ω mH Hz m2 V Hz1/2 m-1
Square 0.16 20 47 1.002 0.998 159.8 0.02563 5.03×10-3
0.567 18 21 1.006 0.994 160.9 0.3219 8.96×10-4
1.70 16 11 0.987 1.013 155.0 2.892 1.89×10-4
4.90 14 6 0.972 1.029 150.5 24.05 4.13×10-5
Triangle 2.60 16 12 0.994 1.005 157.5 1.695 2.97×10-4
8.39 14 6 1.004 0.996 160.3 17.59 5.74×10-5
27.3 12 3 1.035 0.967 170.3 187.0 1.10×10-5
60.7 10 2 0.959 1.043 146.3 920.9 3.22×10-6
202 8 1 1.005 0.995 160.9 10164 5.97×10-7

the size of the base (hypotenuse) of the triangle.


The antennas range from compact antennas that could be used with a hand-held
receiver to large antennas that would be constructed to achieve the best sensitivity.
Often the small antennas are used to survey an area for the optimal location to con-
struct a large permanent antenna. The right isosceles triangle antennas are preferred
for larger antenna deployments because they can be easily constructed with a single
tower supporting the apex of the triangle. In this dissertation, for the purpose of
calculating and comparing receiver performance, the 4.9 meter square antenna with
six turns is used.

2.1.2 Receiver Design


The loop antenna is combined with a transformer and a low-noise amplifier to form the
complete receiver front-end. Figure 2.2 shows the full model of the receiver front-end,
including the loop antenna, which is modeled with a voltage source Va that represents
the induced voltage from an incident wave, a noise source va2 , a loop resistance Ra and
a loop inductance La . The antenna is followed by a 1:m transformer, which is assumed
to be ideal in this analysis. The transformer is followed by a low-noise amplifier, which
has a transimpedance gain of G and an input resistance of Rin . The receiver model
also includes the input referred noise sources of the amplifier. The input referred
CHAPTER 2. BACKGROUND 12

vi2
Ra La
1:m

i2i Rin Vout


va2

Vout
Va Iin
=G

Figure 2.2: Model of the receiver front-end including the antenna, transformer, am-
plifier and noise sources.

voltage noise of the amplifier is represented by vi2 and the input referred current noise
of the amplifier is represented by i2i .
The output voltage as a function of the antenna voltage can be calculated directly
from the receiver model in Figure 2.2. The expression for the induced antenna voltage,
given in Equation 2.4, can then be substituted to derive the output voltage as a
function of the received magnetic field amplitude. The resulting relationship is
 
N ABω G f
Vout = (2.9)
mLa f − jfi

where fi is the input turnover frequency and is given by

Ra + Rin /m2
fi = (2.10)
2πLa

as shown in [39]. The input turnover frequency is the frequency at which the reactance
of the loop is equal to the total resistance of the input circuit. Below fi the output of
the receiver is proportional to frequency. Above fi the output of the receiver is flat
with frequency. It is desirable to operate in the flat region, which is why a low input
impedance amplifier is required. A typical design choice is to set Rin /m2 = Ra , which
results in fi = 2fa . In this case, the input turnover frequency is approximately 318
Hz with a 1 Ω, 1 mH antenna, which is right at the start of the VLF band.
CHAPTER 2. BACKGROUND 13

vp2 vs2
Rp L2 Rs
1:m

Lp Cs

Figure 2.3: Model of the transformer including relevant parasitics.

The sensitivity of the receiver front-end can be calculated from the three noise
sources included in the receiver model in Figure 2.2. The resulting sensitivity is
s
1 vi2
Ssys = va2 + + i2i m2 Za2 (2.11)
ωN A m2

At low frequencies the voltage noise sources dominate. However, over the major-
ity of the VLF band the current noise is dominant because the antenna impedance
Za ≈ jωLa . The current noise causes the sensitivity to flatten and reach a constant
minimum value. As a result, the input referred current noise of the amplifier is critical
to ensure good receiver sensitivity in the VLF band.

2.1.3 Transformer Design


The receiver front-end uses a hand-wound transformer to interface between the loop
antenna and the input amplifier. A model of the transformer is shown in Figure 2.3. In
addition to the ideal 1:m transformer, the model also includes the relevant parasitic
elements. It includes the resistance of the primary winding Rp and the resistance
of the secondary winding Rs , as well as their associated noise sources vp2 and vs2 .
The model also includes the inductance of the primary winding Lp and the leakage
inductance referred to the secondary L2 . Finally, it includes the winding capacitance
referred to the secondary Cs .
The transformer must be carefully designed to avoid degrading the performance
of the receiver. The output voltage of the receiver as a function of the magnetic field
CHAPTER 2. BACKGROUND 14

at the antenna with the full transformer model is given by


   
N ABω G f f −jfc
Vout = (2.12)
m (La + pL2 /m2 ) f − jft f − jfi f − jfc

where

(Ra + Rp ) k (Rs + Rin ) p/m2


ft = (2.13)
2π (La + Lp )
Ra + Rp + (Rs + Rin ) p/m2
fi = (2.14)
2π (La + pL2 /m2 )
1
fc = (2.15)
2πCs Rin
La
p=1+ (2.16)
Lp

as shown in [39]. The transformer parasitics affect the frequency response of the
receiver through the three corner frequencies shown above. The ft corner was not
present in the ideal transformer analysis and is the result of the effect of Lp , which
reduces the gain of the receiver at low frequencies. The high frequency corner fc is
also new and is caused by the parasitic capacitance at the output of the transformer,
which limits the bandwidth of the receiver. The input turnover frequency fi , which
was present before, is increased by the transformer parasitics.
All of the receivers deployed by the Stanford VLF research group use a hand-
wound 24:548 turns ratio center-tapped transformer (m = 22.83). This turns ratio
was selected because it provides a good balance between low frequency and high
frequency noise performance [39]. For compatibility, this transformer is also used in
the design of the single-chip receiver.
Figure 2.4 shows the frequency response of the receiver with the transformer para-
sitics included. A model of the 24:548 center-tapped transformer was used to generate
CHAPTER 2. BACKGROUND 15

Figure 2.4: Receiver frequency response with the full transformer model.

this plot. The three corner frequencies are highlighted in the figure and occur at

ft = 3.61 Hz
fi = 337 Hz
fc = 3.96 MHz

The receiver is operated in the flat region of the response, which extends from roughly
fi to fc or 337 Hz to 3.96 MHz. In a good design ft  fi and fc is larger than
maximum frequency of interest. These goals are generally accomplished by designing
the transformer to have a large Lp and a small Cs , as well as small parasitic resistances
Rp and Rs . These conditions are satisfied by the 24:548 center-tapped transformer
used by the single-chip receiver.
A center-tapped transformer offers two primary benefits. First, the center-tapped
CHAPTER 2. BACKGROUND 16

transformer allows for the use of differential signaling throughout the receiver front-
end, which rejects common-mode noise and interference, and reduces even-order har-
monic distortion. Second, the center-tapped secondary winding allows for bias cur-
rents from the amplifier to pass through the transformer windings. Running bias
currents in the windings is not a problem because the currents are balanced and cre-
ate no net DC field in the transformer core. As a result, the receiver is much less
sensitive to vibrations.

2.1.4 Amplifier Design


The frequency response and noise performance are the two primary design consider-
ations for the front-end amplifier. The receiver frequency response is affected by the
input resistance of the amplifier. Ideally, Rin /m2 would be zero, which would result
in the input turnover frequency being equal to the antenna turnover frequency. In
this case, the receiver bandwidth would be maximized. However, this is not practical.
A good compromise is to set Rin /m2 = Ra . Given the predefined transformer turns
ratio (m = 22.83), the necessary input impedance of the amplifier can be calculated.
This calculation results in an input impedance of a few hundred ohms.
The noise performance of the amplifier is the second key design consideration. The
amplifier noise can be characterized in terms of its input referred voltage noise and
input referred current noise. Because of the inductive antenna impedance, the input
referred current noise is dominant over most of the receiver bandwidth. As a result,
minimizing the input referred current noise is critical to ensure good sensitivity of
the overall receiver.
There are three basic types of amplifiers that can be used to achieve the impedance
and noise goals. The first is a standard voltage amplifier with a physical resistor of
value Rin terminating its input. While this meets the requirement for the input
impedance, it has the worst noise performance of the three options. A well designed
amplifier with an input impedance of Rin can have an input referred noise that is
much lower than the noise of a physical resistor of value Rin .
CHAPTER 2. BACKGROUND 17

The second amplifier option is a high impedance amplifier that uses current feed-
back to achieve a low input impedance. This architecture can achieve good noise
performance if the feedback resistor and the forward gain are large enough. However,
these requirements often lead to feedback stability issues that make this architecture
not practical.
The third amplifier option is a low impedance amplifier, such as a BJT common-
base amplifier or a CMOS common-gate amplifier. This is the preferred amplifier
architecture because it has an inherently low input impedance and good noise per-
formance. Generally the common-base amplifier is used due to its superior current
efficiency and the lower 1/f noise of bipolar transistors [34]. One downside of us-
ing this amplifier architecture is that current noise at the output of the amplifier is
referred directly to the input. This is because the common-base and common-gate
amplifiers have a current gain of one. As a result, any circuits following the amplifier
must be carefully designed to minimize their input referred current noise.
The front-end amplifier is implemented with a differential circuit architecture. As
discussed in the previous section, this improves the common-mode rejection of the
amplifier, balances the DC currents in the transformer windings and cancels even-
order harmonic distortion [25].

2.1.5 Receiver Testing Methods


There are two methods of testing the receiver. The first is the induced signal method,
where a second loop antenna is used to generate a known magnetic field that can then
be received by the receiver’s loop antenna. This method is problematic because it
requires exact knowledge of the size, shape, position and orientation of the two loop
antennas. The second method is the injected signal method, in which a known current
is injected parallel to the antenna to simulate a received signal.
The injected signal method is preferred due to its accuracy and simplicity. Fig-
ure 2.5 shows the current injection circuit. It consists of a resistor Rcal and a capacitor
Ccal . The figure also shows the antenna (Ra , La , Va ), the 1:m transformer, and the
front-end amplifier. If Rcal is much larger than the impedance seen at the antenna
CHAPTER 2. BACKGROUND 18

Rcal Ccal
+

Ical
Ra
1:m

Vcal La G Vout

Va

Rcal Ccal

Ical

Figure 2.5: Receiver current injection circuit with the antenna, transformer and am-
plifier included for reference.

terminals, then the injected current Ical only depends on the calibration resistor and
capacitor
Vcal
Ical = (2.17)
Rcal (1 + fcal /jf )
where
1
fcal = (2.18)
2πRcal Ccal
Essentially, the calibration resistor and capacitor convert the differential input voltage
into a differential current that is injected in parallel with the loop antenna.
The calibration voltage that is equivalent to a magnetic field amplitude of Bω at
the antenna can be calculated with

N ARcal Bω 1 + fcal /jf


Vcal = (2.19)
La 1 + fa /jf

If fcal = fa then the frequency dependence of this relationship cancels. This is


CHAPTER 2. BACKGROUND 19

accomplished by setting the calibration capacitor to

La
Ccal = (2.20)
Ra Rcal

With a properly selected calibration capacitor, the relationship between the calibra-
tion voltage and the magnetic field amplitude simplifies to

N ARcal Bω
Vcal = (2.21)
La

Notice that the relationship is independent of frequency. It also does not depend on
circuit parameters, such as the input impedance of the amplifier. This relationship
can be used to convert a given calibration voltage into an equivalent magnetic field
amplitude, assuming a particular antenna size and number of turns.
For testing purposes the loop antenna in Figure 2.5 can be replaced with equivalent
passive components. This so-called dummy loop consists of a 1 Ω resistor and a 1 mH
inductor. The dummy loop provides the same impedance as an actual antenna, but
greatly reduces the amount of noise and interference that couples into the receiver.
This is particularly important in a lab setting where an actual loop antenna would
pick up a large amount of interference from nearby measurement equipment.

2.2 Delta-Sigma Analog-to-Digital Conversion

2.2.1 Analog-to-Digital Conversion


An analog-to-digital converter (ADC) is a device that converts a continuous signal,
such as a voltage or current, into a digital value that represents the amplitude of
the signal. Real world signals are analog, which means that they are continuous in
time and amplitude. Analog-to-digital conversion is often used to take advantage of
the benefits of digital signal processing (DSP). DSP provides immunity to noise and
interference, enables error detection and data compression and allows for reconfigura-
bility. Further, digital circuits can typically be designed and manufactured at a much
lower cost than equivalent analog circuits.
CHAPTER 2. BACKGROUND 20

An ADC performs two operations: sampling and quantization. The majority of


ADCs use uniform sampling where the analog signal is sampled at a constant interval
in time. For example, an ADC with a 1 MHz clock rate will sample the input every 1
µs. There has been research into non-uniform sampling, but it will not be considered
here [6]. Assuming uniform sampling, the required sampling rate can be related to the
bandwidth of the input signal through the Nyquist sampling theorem, which states
that the original analog signal can be reconstructed from the digital values if

fs ≥ 2fB (2.22)

where fB is the bandwidth of the input signal and fs is the sampling frequency [38].
The Nyquist theorem states that the sampling frequency must be at least twice the
bandwidth of the input signal. Often the sampling rate is set higher than the minimum
value to allow room for the anti-alias filter to roll-off.
Quantization is the process of converting the amplitude of the input signal into
discrete digital values. Because there are a finite number of quantization levels and
the input signal is continuous, there will always be some error introduced in the
quantization process. Figure 2.6 shows the quantizer transfer curve for a 3-bit ADC.
The horizontal axis represents the input signal from −1 to +1 and the vertical axis
represents the 8 digital values. The blue curve shows the transfer characteristic of the
3-bit ADC and the red curve shows the characteristic of an ideal infinite resolution
ADC. Notice that error is introduced by the 3-bit ADC because there are only 8
digital output values but an infinite number of possible input values.
The full scale range of the ADC is defined as the range of the input signal that
doesn’t cause saturation. In this example the full scale range extends from −1 to +1.
The step size ∆ of the ADC is the width of the steps in the transfer curve. The step
size can be calculated with
FS
∆= (2.23)
2B
where FS is the full scale range and B is the resolution of the ADC. Therefore, higher
resolution ADCs will have smaller step sizes.
Figure 2.7 shows the error of the quantizer, which is related to the difference
CHAPTER 2. BACKGROUND 21

Figure 2.6: Quantizer transfer curve for a 3-bit ADC.

between the ideal infinite resolution ADC transfer curve and the 3-bit ADC transfer
curve. Notice that the error is bounded between − ∆2 and + ∆2 for input signals from
−1 to +1. Outside of this range the error increases without bound. This is caused
by the input signal saturating the ADC.
If the input signal is sufficiently active and exercises a large range of digital codes
then the quantization error can be assumed to be statistically random, which produces
a white noise spectrum. The total quantization noise power is

∆2
e2q = (2.24)
12

which is calculated by summing the noise components over the entire ADC band-
width [38]. The total quantization noise power depends only on the step size of the
ADC, which is related to its resolution, and not on the sampling rate of the ADC.
Figure 2.8 shows an example of the output spectrum of an 8-bit ADC with a
sinusoidal input signal. The horizontal axis corresponds to frequency from 0 to fs /2
and the vertical axis is the amplitude of the signal. The large tone in the spectrum is
CHAPTER 2. BACKGROUND 22

Figure 2.7: Quantization error for a 3-bit ADC.

the input sinusoid and the other small tones are the quantization noise components.
Notice that the quantization noise spectrum is roughly flat over the ADC bandwidth,
which agrees with the white noise approximation.
The performance of an ADC can be characterized in terms of its signal to quanti-
zation noise ratio (SQNR). The SQNR is calculated by taking the ratio of the signal to
the sum of the quantization noise components, which can be related to the resolution
of the ADC with  
Ps
SQNR = 10 log = 6.02B + 1.76 (2.25)
Pn
Therefore, the SQNR is directly related to the resolution, B, of the ADC. Higher
resolution gives a larger SQNR and lower resolution gives a smaller SQNR. In the
example spectrum in Figure 2.8 the SQNR is equal to 50.07 dB, which is very close to
the 49.92 dB value predicted by Equation 2.25 for an 8-bit ADC. This equation can
be inverted to calculate the resolution of the ADC from a given SQNR value. The
CHAPTER 2. BACKGROUND 23

Figure 2.8: ADC output spectrum illustrating the distribution of quantization noise.

effective number of bits is

SQNR − 1.76
ENOB = (2.26)
6.02

where the SQNR is given in dB. This relationship is particularly useful for oversampled
ADCs where the effective number of bits is not necessarily equal to the number of
bits that the ADC outputs.

2.2.2 Oversampling
Oversampling refers to using a sampling frequency that is higher than the Nyquist
rate. The oversampling ratio (OSR) is defined as

fs
OSR = (2.27)
2fB

where fs is the sampling frequency and fB is the bandwidth of the input signal.
Oversampling has two primary benefits. First, an oversampled ADC has relaxed
CHAPTER 2. BACKGROUND 24

Ne (f )
∆2 1 ∆2
12 OSR 12

f
fB fs /2

Figure 2.9: Example of oversampling. The red box is the desired ADC bandwidth.

requirements on the anti-alias filter because immediate roll-off is not needed. Second,
the effective number of bits of the ADC is increased by oversampling because the total
quantization noise is not dependent on the sampling rate. To illustrate this point,
Figure 2.9 shows the quantization noise spectrum of an ADC. The noise is spread
uniformly over the ADC bandwidth from 0 to fs /2, but only the spectrum up to fB is
of interest if oversampling is used. The total quantization noise of the ADC is ∆2 /12
and since it is spread over a larger bandwidth the quantization noise that falls below
fB is reduced. The in-band quantization noise is

∆2 1
e2q = (2.28)
12 OSR

The quantization noise that falls outside the signal bandwidth can be filtered using
a digital filter.
Oversampling improves the SQNR of the ADC, which results in a higher effective
number of bits. In fact, each doubling of the sampling frequency results in an increased
resolution of approximately 0.5 bits [45]. For example, consider an 8-bit ADC that
is oversampled to produce a 12-bit effective resolution. To increase the resolution
by 4 bits the sampling rate would have to be increased by a factor of 256. While
this is possible, it may not be practical. As a result, oversampling to increase the
effective resolution of an ADC is usually not done. However, oversampling to relax
the requirements on the anti-alias filter is common.
CHAPTER 2. BACKGROUND 25

X(z) A(z) ADC Y (z)


DAC

Figure 2.10: Block diagram of a discrete-time delta-sigma modulator.

2.2.3 Delta-Sigma Modulators


A delta-sigma modulator is an oversampling ADC architecture that operates on the
principle of shaping the quantization noise spectrum away from the signal band so
that it can later be removed by a digital filter [27]. This enables higher resolutions
to be achieved while simultaneously relaxing the requirements on the analog com-
ponents. In the previous section it was shown that basic oversampling can increase
the resolution of an ADC, but significant increases in resolution require impractical
sampling rates. The delta-sigma modulator overcomes this limitation by shaping the
noise spectrum.
A block diagram of a delta-sigma modulator is shown in Figure 2.10. The mod-
ulator includes a discrete-time loop filter A(z), a low resolution ADC that quantizes
the signal and produces the digital output value, and a DAC that converts the digital
output value to an analog value and feeds it back to the input of the modulator. The
input to the modulator is a discrete-time analog signal, which means that X(z) is a
sampled version of the original input signal. The output signal Y (z) is a discrete-time
digital signal.
To understand the operation of the delta-sigma modulator it can be converted to
a linear model, which is shown in Figure 2.11. The ADC is replaced by a summation
node that adds noise from a noise source E(z). This noise source corresponds to the
quantization noise generated by the ADC at this point in the modulator. The DAC
is assumed to be ideal in this analysis and is replaced by a direct connection from
the output to the input summation node. Calculating the output as a function of the
CHAPTER 2. BACKGROUND 26

E(z)

X(z) A(z) Y (z)


Figure 2.11: Linear model of a delta-sigma modulator.

input signal X(s) and the quantization noise E(z) yields

A(z) 1
Y (z) = X(z) + E(z) (2.29)
1 + A(z) 1 + A(z)

The term preceding the input signal is referred to as the signal transfer function (STF)
and the term preceding the quantization noise is called the noise transfer function
(NTF).
The fact that the STF and the NTF are different can be exploited to shape the
quantization noise spectrum. For example, consider the case in which the loop filter,
A(z), has a large gain at low frequencies. In this case, at low frequencies the STF
would approach one while the NTF would approach zero. This means that a low
frequency signal would pass through the modulator unchanged while the quantization
noise at low frequencies would be significantly attenuated.
A common loop filter implementation is a discrete-time integrator. An integrator
produces the desired transfer function, with a very high gain at low frequencies. The
transfer function of a discrete-time integrator is

z −1
A(z) = (2.30)
1 − z −1

with this loop filter transfer function the output of the delta-sigma modulator becomes

Y (z) = z −1 X(z) + 1 − z −1 E(z)



(2.31)
CHAPTER 2. BACKGROUND 27

In this case, the STF is simply a one sample delay and the NTF is a high-pass
filter. Assuming that the delta-sigma modulator is oversampling the input signal,
then signals within the signal band (0 to fB ) are passed unchanged while the noise
in the signal band is filtered and pushed to higher frequencies. The configuration
with a single integrator as the loop filter is called a first-order modulator and results
in first-order noise shaping. It can be shown that first-order noise shaping results
in an increase in the effective resolution of the ADC of 1.5 bits for each doubling of
the sampling rate. For example, if the internal ADC has a resolution of 4 bits and
an oversampling ratio of 16 is used, then the effective resolution of the first-order
modulator is increased by 6 bits for a total resolution of 10 bits.
In addition to using a higher oversampling ratio to increase the effective resolution,
there are several other design strategies that can be used to increase the resolution
of the delta-sigma modulator [45]. One common modification is to use a higher order
loop filter. The previous example used a first-order loop filter, but higher order loop
filters are possible. If a second-order loop filter is used then the transfer function of
the modulator becomes

2
Y (z) = z −2 X(z) + 1 − z −1 E(z) (2.32)

The input signal is still transfered to the output with only a delay, but now the
quantization noise is filtered with a second-order high-pass filter. The higher order
filter provides more quantization noise attenuation in the signal band and results in
a higher SQNR. In fact, with a second-order loop filter the effective resolution of the
ADC is increased by 2.5 bits for each doubling of the oversampling ratio.
This concept can be extended to an N th-order loop filter, which result in an
increased resolution of N + 0.5 bits for each doubling of the oversampling ratio [45].
However, there is a trade-off with increasing the filter order. Higher order modulators
are more prone to instability, so precautions must be taken in their design. As a result,
there are diminishing returns to increasing the modulator order above a certain point,
which typically occurs around 5th or 6th order loop filters.
Another parameter of the delta-sigma modulator is the resolution of the internal
CHAPTER 2. BACKGROUND 28

ADC. Typically only low resolution quantizers are used for the internal ADC, which
have resolutions in the range of 1 to 4 bits. Using a resolution larger than 1 bit for
the internal quantizer has two benefits. First, it provides a higher starting point for
the resolution of the modulator. As a result, a lower oversampling ratio can be used
to achieve the same resolution. Second, a higher resolution internal ADC improves
the stability of higher order modulators.
However, there is one major drawback to using an internal ADC with a resolution
greater than 1 bit. The problem arises because the feedback DAC, which uses the
same resolution as the internal ADC, has to have the same accuracy as the overall
modulator [38]. This is because the signal from the DAC is added directly to the input
signal. For example, if a 4-bit internal ADC is used to build a 16-bit modulator, then
the 4-bit DAC needs to have 16-bit accuracy. This requirement severely limits the
performance of the modulator because it is difficult to build a DAC with linearity
that matches the high resolution of the modulator. There are techniques to overcome
this problem, such as dynamic element randomization and mismatch error shaping,
but these methods add complexity to the design and consume additional power [8, 2].
As a result, single-bit internal ADCs are often used. This is because a 1-bit DAC is
perfectly linear as it has only two levels.
To illustrate the operation of a delta-sigma modulator, a time-domain simulation
of a single-bit third-order modulator is shown in Figure 2.12. The red curve is a
sinusoidal input signal with an amplitude of −6 dB relative to full-scale (dBFS). The
blue curve is the 1-bit digital output signal from the modulator. Notice that the
output signal has a pulse-width modulation characteristic, in which the output signal
has more +1 values when the input signal is high and more −1 values when the input
signal is low.
The output spectrum can be calculated using the fast Fourier transformer (FFT)
to visualize the effect of the quantization noise shaping. Figure 2.13 shows the output
spectrum of the single-bit third-order delta-sigma modulator. An 8192 point FFT
was computed with a Hann window to prevent spectral smearing. The tone from the
sinusoidal input signal is clearly visible near the middle of the plot with an amplitude
of −6 dBFS. The remainder of the spectrum is the quantization noise. Notice that
CHAPTER 2. BACKGROUND 29

Figure 2.12: Time-domain simulation of the delta-sigma modulator.

the quantization noise is high-pass filtered, with the majority of the noise occurring
at higher frequencies. If an oversampling ratio of 64 is assumed, then the SQNR is
82.03 dB, which corresponds to an effective resolution of 13.33 bits.
One question that remains is how to extract the high-resolution data from the
low-resolution modulator output. For instance, the third-order modulator described
above has a 1-bit output bitstream, but an effective resolution of over 13 bits. This
conversion is accomplished by using a decimation filter. A decimation filter is a digital
filter that performs two functions. First, it applies a low-pass filter to remove the
spectral components that fall above the desired signal bandwidth while simultaneously
increasing the width of the data to match the resolution of the full ADC. This filtering
will remove the quantization noise that has been shaped to higher frequencies. Second,
the decimation filter downsamples the data to the Nyquist sampling rate. Decimation
filters for delta-sigma modulators are typically implemented with low-pass sinc filters
where the order of the sinc filter is one higher than the order of the modulator loop
filter [7].
CHAPTER 2. BACKGROUND 30

Figure 2.13: Output spectrum of the delta-sigma modulator.

2.2.4 Continuous-Time Delta-Sigma Modulators


Historically, the majority of delta-sigma modulators have been implemented with
discrete-time loop filters. These filters are built with switched-capacitor circuits that
exhibit good accuracy and linearity [26]. However, discrete-time modulators are lim-
ited in terms of their maximum speed and the requirement that sampling be performed
at the input of the modulator. An alternative is to implement the modulator with a
continuous-time loop filter.
Figure 2.14 shows the block diagram of a continuous-time delta-sigma modulator.
There are two differences when comparing the continuous-time modulator with the
discrete-time modulator shown in Figure 2.10. First, the discrete-time loop filter,
A(z), is replaced with a continuous-time loop filter, A(s). Second, the input signal is
now a continuous-time signal, while in the discrete-time modulator the input was a
sampled signal.
There are two primary benefits of using a continuous-time loop filter in a delta-
sigma modulator. The first is related to the speed of the modulator. In a discrete-time
CHAPTER 2. BACKGROUND 31

X(t) A(s) ADC Y (z)


DAC

Figure 2.14: Block diagram of a continuous-time delta-sigma modulator.

modulator the signal is a series of fast changing pulses. The maximum clock rate of
the modulator is therefore limited by the settling time of the integrators, which is
determined by the bandwidth of the amplifiers used in the loop filter. In a continuous-
time modulator the signals are continuous, which greatly reduces the requirements
on the amplifiers. As a result, a continuous-time modulator can be clocked two to
four times faster than an equivalent discrete-time modulator [45].
The second benefit is the implicit anti-alias filter [38]. In a discrete-time modulator
the input signal must be sampled at the input of the modulator. This places stringent
requirements on the sample and hold circuit and mandates the use of an anti-alias
filter to prevent out of band signals from aliasing into the signal band. In a continuous-
time modulator sampling doesn’t occur until the signal reaches the internal ADC,
which is after the loop filter. As a result, the continuous-time modulator has an
implicit anti-alias filtering characteristic, which is easy to understand because the
signal passes through the loop filter before being sampled. The implicit anti-alias
filter response for a third-order single-bit continuous-time modulator is shown in
Figure 2.15. The filter has a sinc characteristic with nulls at multiples of the sampling
frequency. These nulls are conveniently located at frequencies that would alias into
the signal band.
While these benefits make continuous-time modulators an attractive option, there
are a few drawbacks to using them. The first is that they are more sensitive to clock
jitter than their discrete-time counterparts [12]. This is because the feedback signal
from the DAC is typically a constant current pulse that lasts for the duration of the
clock period, which means that the total charge deposited on the capacitor used in
CHAPTER 2. BACKGROUND 32

Figure 2.15: Response of the implicit anti-alias filter of a third-order continuous-time


delta-sigma modulator.

the continuous-time integrator is

QDAC = IDAC Ts (2.33)

where IDAC is the DAC current and Ts is the sampling period. Any error in the clock
period will result in a proportional error in the total feedback charge. The primary
method to mitigate this problem is to use non-rectangular DAC pulse shapes, but
this approach places additional strain on the amplifiers in the loop filter. Clock jitter
becomes particularly problematic at high sampling rates, where the magnitude of the
jitter is a significant fraction of the sampling period. At low sampling rates it can
often be disregarded.
The other main issue that affects continuous-time modulators is excess loop de-
lay [14]. In a discrete-time modulator the signals in the modulator only need to be
valid at the sampling instants, which gives them a full clock period to settle to an
CHAPTER 2. BACKGROUND 33

accurate value. As a result, a small amount of delay in the feedback path isn’t prob-
lematic as long as there is enough settling time. In a continuous-time modulator the
signals have to be valid at all times, so any delay causes inaccuracies. This delay
could be caused by the finite decision time of the internal quantizer or the response
time of the DAC. There are several methods to overcome this problem. If the amount
of excess loop delay is fixed, then the loop filter coefficients can be adjusted to com-
pensate for this delay. If the delay is signal dependent, which is often the case for
quantizer delay, then a fixed delay can be added to the loop to give the quantizer
additional decision time. The loop filter can then be modified to compensate for this
fixed delay.
The design process for implementing a continuous-time delta-sigma modulator
starts with the design of a discrete-time modulator. Due to the wide availability of
design tools for discrete-time modulators, it is generally straightforward to design a
discrete-time loop filter. From there, the discrete-time loop filter is converted into
an equivalent continuous-time loop filter. The two primary methods for accomplish-
ing this are the impulse-invariant transform and the modified z-transform [38]. The
impulse-invariant transform attempts to find the continuous-time loop filter coeffi-
cients that result in an equivalent impulse response to the discrete-time loop filter.
The modified z-transform computes the equivalent discrete-time loop filter from the
continuous-time loop filter and then compares it with the original discrete-time loop
filter to determine the coefficients. Both methods produce similar results in most
cases.
In a continuous-time modulator there are many different circuit architectures that
can be used to implement the integrators, each with its own benefits. Two of the
most common are the active-RC integrator and the gm-C integrator. Active-RC
integrators generally have the highest linearity at the expense the power consumption
and tunability, while gm-C integrators generally have the highest speed and lowest
power consumption at the expense of linearity. Other architectures include active gm-
C integrators, active MOS-C integrators and log integrators. The choice of integrator
architecture is application specific and often multiple integrator architectures are used
in the same ADC design.
Chapter 3

Receiver Overview

3.1 Performance Metrics


Specialized receiver hardware is required to collect broadband VLF magnetic field
data with the accuracy and resolution necessary for scientific research due to the
unique nature of these signals [39]. To receive high quality data, an example of
which was shown in the spectrogram in Figure 1.2, the receiver must have certain
performance characteristics. The four most important performance characteristics
for a VLF magnetic field receiver are described in this section. These include the
receiver bandwidth, gain, sensitivity and spurious-free dynamic range. A detailed list
that summarizes all of the design goals and specifications is given in Section 3.4.

3.1.1 Bandwidth
The first key specification is the receiver bandwidth, which is defined as the 3 dB
cutoff from the low-end of the frequency band to the high-end of the frequency band.
A flat frequency response is desired due to the roughly constant power spectral density
of naturally occurring VLF signals. The goal for the receiver is to have a flat response
that extends from 300 Hz to 50 kHz.
At the low-end of the frequency range the receiver response is limited by the
impedance of the antenna. A standard 1 Ω, 1 mH antenna is used by all of the VLF

34
CHAPTER 3. RECEIVER OVERVIEW 35

receivers deployed by the Stanford VLF group, so there is minimal control over the
low frequency cutoff. At the high-end of the frequency range the bandwidth is limited
by the bandwidth of the low-noise amplifier used in the receiver.
The goal for the high frequency cutoff of the receiver is 50 kHz, which is higher
than the normal limit of the VLF band of 30 kHz. The reason for this extended
bandwidth is because several signals of interest occur just above the standard 30 kHz
VLF limit [18]. One example is the NAU transmitter in Puetro Rico that transmits
at 40.75 kHz.
It is important to point out that a broadband VLF receiver is very different from
a standard RF receiver. In a standard RF receiver a mixer is used to frequency
shift a narrow-band signal from a high RF frequency to a low frequency where it can
be digitized and processed. In contrast, in a broadband VLF receiver there is no
frequency shifting because the signal is already at a low frequency. In this case, the
entire VLF band can be directly digitized. Therefore, the bandwidth of the receiver
must cover the entire VLF band.

3.1.2 Gain
The gain of the receiver is defined as the transimpedance gain (current-to-voltage)
from the input of the receiver to the output of the receiver. The goal for the receiver
is to have a transimpedance gain of approximately 100 dB. In terms of the received
magnetic field, this amount of gain corresponds to 0.63 mV/pT when using a six-turn
4.9 meter square antenna. This is the standard antenna size for comparing receiver
performance.
The large amount of transimpedance gain is necessary because naturally occurring
VLF signals are often very small. As a result, a significant amount of gain is required
to amplify the signals enough to be captured by the ADC. This amount of gain is
consistent with previous VLF receiver designs deployed by the Stanford VLF research
group when accounting for differences in the ADC full-scale range and supply voltage.
To support a variety of applications, the single-chip receiver described in this dis-
sertation implements multiple gain levels. The single-chip receiver includes four gain
CHAPTER 3. RECEIVER OVERVIEW 36

modes with transimpedance gain values of approximately 90 dB, 100 dB, 110 dB and
120 dB. This enables the user to select the appropriate amount of gain depending on
the application and location of the receiver. For example, in a noisy urban environ-
ment the lower gain modes may be needed to prevent the receiver from saturating.
On the other hand, in a remote area with little noise the higher gain modes could be
used to further amplify the small VLF signals.

3.1.3 Sensitivity
The sensitivity of the receiver is defined as the magnetic field amplitude at the antenna
that results in a 0 dB signal-to-noise ratio at the output of the receiver when calculated
in a 1 Hz bandwidth [39]. Essentially, the sensitivity is the minimum detectable signal.
It is a signal with an amplitude that is ultimately equal to the output noise floor of
the receiver. A smaller sensitivity is better. The desired sensitivity of the receiver is
below 1 fT/Hz1/2 over the entire receiver bandwidth, which is from 300 Hz to 50 kHz.
The target sensitivity is 1 fT/Hz1/2 because the environmental noise in the VLF
band is between 1 fT/Hz1/2 and 2 fT/Hz1/2 at the quietest locations [15]. Designing
the receiver to have a sensitivity below the environmental noise floor ensures that the
noise contributed by the receiver does not degrade the received data.
It is important to note that the sensitivity of the receiver depends on the size of the
antenna. For the purposes of design and testing, a six-turn 4.9 meter square antenna
is assumed. A larger antenna will result in better sensitivity, while a smaller antenna
will result in worse sensitivity. The size of the antenna deployed will depend on the
desired sensitivity, as well as the amount of expected interference at the receiver site.

3.1.4 Spurious-Free Dynamic Range


The spurious-free dynamic range (SFDR) is a measure of the linearity of the receiver.
When a sinusoidal signal is applied to the input of the receiver, the SFDR is defined as
the ratio of the magnitude of the input tone to the magnitude of the largest spurious
tone in the output spectrum of the receiver [38]. Typically the largest spurious tone
is a harmonic of the input signal. An example of the SFDR calculation is shown in
CHAPTER 3. RECEIVER OVERVIEW 37

Figure 3.1: Example of the calculation of the spurious-free dynamic range.

Figure 3.1. This example shows an output spectrum with a 10 kHz sinusoidal input
signal. The largest spurious tone is the third harmonic at 30 kHz. Therefore, in this
example the SFDR is the difference between the fundamental and the third harmonic,
which is 75 dB.
The goal for the SFDR is 90 dB at all frequencies in the receiver bandwidth.
A large SFDR is necessary because strong transmitters, as well as weak naturally
occurring VLF signals, fall within the VLF band. Poor linearity would result in the
harmonic distortion from the strong transmitters corrupting the smaller signals. The
difference between the largest transmitters and the smallest VLF signals can be as
large as 90 dB.

3.2 Existing VLF Receivers


Over the past several decades the Stanford VLF research group has deployed many
different VLF receivers. The two most recent and widely used are discussed in this
section. The first is the AWESOME receiver, which is the high-performance receiver
CHAPTER 3. RECEIVER OVERVIEW 38

currently deployed around the world. The second is the Penguin receiver, which is
the low-power receiver that is used for remote deployments.

3.2.1 AWESOME Receiver


The primary receiver used by the Stanford VLF research group is the AWESOME
receiver [19]. AWESOME is an acronym which stands for Atmospheric Weather
Electromagnetic System for Observation, Modeling and Education. The AWESOME
receiver is deployed around the world in locations including Japan, Australia, Hawaii,
Europe and Africa. The global network of receivers constantly collects, processes and
stores the VLF data that is received. In some cases the data is automatically uploaded
to servers at Stanford for further analysis.
The AWESOME receiver is the gold-standard VLF receiver. It has the highest
data quality of any of the VLF receivers developed at Stanford. It is a broadband
receiver that covers the entire VLF range from approximately 300 Hz to 50 kHz. It has
a sensitivity of around 1 fT/Hz1/2 and a spurious-free dynamic range of approximately
90 dB. Due to its high data quality, the AWESOME receiver is often used as a
benchmark to compare new receiver designs.
Despite its good performance, the AWESOME receiver has several weaknesses
that prevent it from being used in some applications. The first is its power consump-
tion. The signal path power consumption of the receiver is approximately 10 Watts.
Additional power is consumed by support circuitry, which includes a GPS time ref-
erence, calibration circuitry and digital communication interfaces. The receiver also
requires a computer, either a desktop or a laptop, to process and store the received
data. The total power consumption of the AWESOME receiver is around 60 W when
combined with a low power laptop. A typical implementation uses a standard desk-
top computer, which likely pushes the total power dissipation over 200 W. The high
power dissipation prevents the AWESOME receiver from being used in remote lo-
cations where it would be required to operate on battery power for long periods of
time.
The other weakness of the AWESOME receiver is its size. The receiver consists of
CHAPTER 3. RECEIVER OVERVIEW 39

two metal enclosures. One enclosure contains the digital portion of the receiver and
the other holds the analog components. Each enclosure is approximately 12 inches
wide by 12 inches long by 8 inches tall. This does not include the size of the desktop
or laptop computer. The large size would preclude the AWESOME receiver from use
in a mobile or hand-held application.

3.2.2 Penguin Receiver


The Penguin receiver is a broadband VLF receiver that is designed for remote de-
ployments [32]. Penguin is an acronym which stands for Polar Experiment Network
for Geophysical Upper-atmosphere Investigations. The Penguin receiver is deployed
in remote locations where AC power is not available. The receiver is optimized for
low power dissipation and can run on battery power for long periods of time. It uses
a custom designed low-noise amplifier ASIC that was also developed by the Stanford
VLF research group [28]. The receiver is designed to have enough data storage and
battery power to run autonomously for up to one year without maintenance. The
Penguin receiver was successfully deployed in Antarctica at Amundsen-Scott South
Pole Station in 2009.
The signal path power consumption of the Penguin receiver is approximately 50
mW, which is considerably less than that of the AWESOME receiver. Additional
power is consumed by support circuitry, but the Penguin receiver does not require
a separate desktop or laptop computer. The data quality of the Penguin receiver is
not as good as the AWESOME receiver’s data quality. While both receivers have a
sensitivity of approximately 1 fT/Hz1/2 , the Penguin receiver has a maximum signal
frequency of 30 kHz and a spurious-free dynamic range of roughly 65 dB. The size of
the Penguin receiver is similar to that of the AWESOME receiver. It consists of two
metal enclosures with dimensions of approximately 12 inches wide by 12 inches long
by 8 inches tall.
One issue that was uncovered after the deployment of the Penguin receiver was its
sensitivity to temperature variations. Namely, it was discovered that the biasing of
the low-noise amplifier was not robust to large temperature swings. As a result, the
CHAPTER 3. RECEIVER OVERVIEW 40

Loop
Antenna

Line Line
Driver Receiver

XFMR LNA AAF ADC DSP

Clock Storage

Cable
(up to 1000m) Time
Ref.

Figure 3.2: Block diagram of the traditional receiver architecture.

noise performance of the receiver was reduced if the temperature changed significantly
because the biasing could only be optimized for one specific temperature.

3.3 Receiver Architecture

3.3.1 Traditional Architecture


The AWESOME receiver and the Penguin receiver, as well as previous VLF magnetic
field receivers designed by the Stanford VLF group, use a similar receiver architecture.
Figure 3.2 shows this traditional receiver architecture. While this discussion focuses
on a magnetic field receiver architecture, a similar architecture has been used for
an electric field receiver [37]. The traditional magnetic field receiver architecture is
divided into two sections: the analog front-end and the digital back-end.
The magnetic field antenna is a loop antenna. The size of the loop and the number
of turns can vary depending on the desired receiver sensitivity and the amount of
interference at the receiver location. All of the antennas have a 1 Ω, 1 mH impedance
so that different antennas can be used interchangeably with the same receiver. The
constant antenna impedance also ensures that different receiver designs can be used
with the same antenna.
CHAPTER 3. RECEIVER OVERVIEW 41

The antenna connects directly to the analog front-end. The analog front-end
consists of a transformer, a low-noise amplifier and a line driver. The transformer is
used to reduce the resistance looking into the receiver in order to ensure a low input
turnover frequency. Ideally the input resistance is much less than the resistance
of the antenna, but for practical considerations it is often set equal to the antenna
resistance of 1 Ω. The standard transformer is a 24:548 turns ratio hand-wound center-
tapped transformer. After the transformer, a low-noise amplifier (LNA) is employed
to increase the amplitude of the received signal while adding as little noise as possible.
The signal then passes through a line driver circuit, which further amplifies the signal
and drives a low impedance cable that connects the analog front-end to the digital
back-end.
The signal from the analog front-end is processed by the digital back-end, which
consists of a line receiver, an anti-alias filter (AAF), an analog-to-digital converter
(ADC), as well as several auxiliary components. The line receiver terminates the
cable from the analog front-end and provides any additional gain that is needed to
match the signal swing to the full scale range of the ADC. An anti-alias filter follows
the line receiver and provides a sharp low-pass filter with a cut-off frequency that
falls just above the receiver bandwidth. The AAF can be implemented with an active
circuit or a passive filter and is necessary to prevent aliasing caused by the sampling
operation of the ADC. Finally, the ADC converts the analog signal into digital values
for further processing and storage.
The digital back-end includes several additional subsystems that perform func-
tions outside the signal path of the receiver. These systems include the time refer-
ence, clock generator, digital signal processing and data storage. The time reference
is particularly important and is typically implemented with a GPS receiver and ad-
ditional logic that uses the one pulse per second GPS signal to ensure an accurate
sampling frequency [17]. Accurate sampling is necessary for interferometric measure-
ments in which a VLF event is captured simultaneously at multiple receiver locations
and precise timestamps are needed to compare the data. The time reference includes
a feedback system which adjusts the frequency of the clock generator based on the
one pulse per second GPS signal. This compensates for inaccuracies in the clock
CHAPTER 3. RECEIVER OVERVIEW 42

generator frequency. Depending on the application and the accuracy of the clock
generator, the time reference may correct the clock frequency every second or only
periodically.
Additionally, the digital back-end includes digital-signal processing (DSP) and
data storage. The DSP may provide real-time spectral analysis or data compression.
It may also be used to implement the digital interfaces with the data storage system.
The type of data storage is application specific. For example, in the AWESOME
receiver an external hard-drive connected to a computer is used for storage. In the
Penguin receiver a CompactFlash card is used to eliminate the need for a computer.
The analog front-end and the digital back-end are often separated by a long cable.
The enclosure that holds the analog front-end is typically placed very close to the
antenna to prevent resistive losses in the wires. The digital back-end is placed indoors
where it can be near a computer and a source of AC power. The primary reason for
having a long cable between the analog front-end and the digital back-end is to prevent
the antenna from picking up any interference radiated by the potentially noisy digital
circuits. This also allows for the antenna to be placed a long distance from any other
sources of noise, such as power lines or other electronic equipment. Typically a 100
Ω cable is used and separation distances of up to 1000 meters have been successfully
deployed. Due to its low-power design and self-contained data storage, the Penguin
receiver has minimal radiated interference. As a result, it is not strictly necessary to
separate the digital and analog portions of the Penguin receiver. However, the analog
front-end and the digital back-end are still placed in separate enclosures with a cable
between them for legacy reasons.

3.3.2 Single-Chip Architecture


The single-chip VLF magnetic field receiver presented in this dissertation uses an
optimized receiver architecture. Figure 3.3 shows the single-chip receiver architec-
ture. The signal path of the receiver consists of a loop antenna, transformer, LNA
and ADC. The architecture also includes the same auxiliary subsystems seen in the
traditional receiver architecture, including the time reference, clock generator, digital
CHAPTER 3. RECEIVER OVERVIEW 43

Loop
Antenna
Single-Chip Receiver

XFMR LNA ADC DSP

Storage
Clock

Time
Ref.

Figure 3.3: Block diagram of the single-chip receiver architecture

signal processing and data storage.


The loop antenna in the single-chip receiver architecture is a 1 Ω, 1 mH antenna
that is identical to the one used by both the AWESOME receiver and Penguin receiver.
This ensures compatibility with already installed antennas. Similarly, the receiver
uses the same 24:548 center-tapped transformer for impedance matching the antenna
to the input of the receiver. The single-chip receiver uses a low-noise amplifier at its
input, followed immediately by an analog-to-digital converter.
There are two key differences between the single-chip architecture and the tra-
ditional architecture. The first is that the single-chip architecture does not include
an explicit anti-alias filter (AAF). The AAF prevents signals from aliasing into the
signal band as a result of the sampling operation performed by the ADC. The AAF is
eliminated by using an ADC architecture that does not require an AAF. Eliminating
the AAF results in a significant power savings. In the Penguin receiver, for example,
the AAF uses approximately 40% of the total signal path power consumption. In
addition to burning a large amount of power, the AAF can be challenging to design
as it can degrade the linearity of the receiver.
The second primary difference is that the single-chip architecture does not separate
the analog front-end from the digital back-end with a long cable. This simplifies the
CHAPTER 3. RECEIVER OVERVIEW 44

architecture as the line driver and line receiver are no longer necessary. This is possible
because the single-chip receiver targets low power applications. As a result, all of the
circuits run on battery power and exhibit minimal radiated interference.
The goal of this work is to integrate the entire signal path of the receiver on
a single chip. The signal path includes the low-noise amplifier and the analog-to-
digital converter, as highlighted in Figure 3.3. The auxiliary subsystems, such as
the time reference and digital signal processing, are not included on-chip as they are
application specific.
There are three primary benefits to implementing the receiver on a single chip.
The first is that it reduces the size of the receiver. Significantly reducing the size opens
up new applications, such as the possibility of a hand-held VLF receiver. The second
benefit is reduced cost. The AWESOME receiver, for example, costs approximately
$3,000. A fabrication run of the single-chip receiver would result in thousands of chips
with a unit cost orders of magnitude less than the cost of an AWESOME receiver.
The third benefit is simplicity. Currently, deploying a VLF receiver requires detailed
knowledge of the various receiver components. A single-chip receiver would simplify
the deployment process by requiring only a single component. The single-chip receiver
would interface with the antenna, produce digital output data and guarantee a certain
level of performance. While a single-chip implementation of a VLF receiver provides
a number of benefits, it also presents design challenges. The most critical challenge
is noise coupling between the potentially noisy digital circuits in the ADC and the
sensitive analog circuits in the LNA, which both reside on the same die.

3.4 Design Specifications


The goal for the single-chip VLF magnetic field receiver is to achieve the high data
quality of the AWESOME receiver while having a low power dissipation similar to that
of the Penguin receiver. The full design specifications are summarized in Table 3.1.
The objective is to achieve the design specifications while minimizing the power dis-
sipation. The specifications can be broken down into the various components of the
design, which include the low-noise amplifier, the analog-to-digital converter, the full
CHAPTER 3. RECEIVER OVERVIEW 45

Table 3.1: Design specifications for the single-chip receiver.

Component Specification Goal


Low-Noise Amplifier Bandwidth 300 Hz to 50 kHz
Gain 100 dB
Sensitivity 1 fT/Hz1/2
SFDR 90 dB
Analog-to-Digital Converter Bandwidth 50 kHz
Resolution 12 bits
SFDR 90 dB
Receiver Alias Rejection 80 dB
Temperature −70 C to 125◦ C

Antenna 1 Ω / 1 mH
Transformer 24:548 CT
Technology Process 0.13 µm BiCMOS
Supply Voltage 1.2 V

receiver and the implementation technology.


The design goals for the LNA follow directly from the key performance metrics
described in Section 3.1. The bandwidth of the LNA must extend from 300 Hz to
50 kHz with a flat frequency response over this range. The transimpedance gain
of the receiver needs to be approximately 100 dB. To achieve this goal, the single-
chip receiver implements multiple gain modes that span a range of 90 dB to 120 dB.
The sensitivity of the receiver needs to be less than 1 fT/Hz1/2 and the spurious-free
dynamic range should be at least 90 dB.
The goal for the ADC is to digitize the received signal without degrading the
quality of the data. The ADC requires a bandwidth of 50 kHz to capture the full
VLF band and a spurious-free dynamic range of 90 dB. Previous work by the Stanford
VLF research group has shown that in order to achieve 90 dB SFDR, a resolution
of at least 10 bits is necessary given the desired frequency resolution of the output
spectrum [49]. To achieve the SFDR goal and to ensure that the noise from the ADC
is negligible compared to the LNA, the target resolution for the ADC is 12 bits.
In addition to the design goals for the LNA and ADC, there are also specifications
CHAPTER 3. RECEIVER OVERVIEW 46

for the receiver as a whole. The first is related to aliasing. Although the single-
chip receiver architecture does not include an explicit anti-alias filter, the receiver
still requires at least 80 dB of alias suppression to prevent out of band signals from
aliasing into the VLF band and corrupting the received data. The receiver also must
be able to operate over a wide range of temperatures, which extends from −70◦ C
to 125◦ C, to account for the conditions at the remote deployment locations. As
discussed previously, the single-chip receiver is designed to use the standard 1 Ω, 1
mH loop antenna and the standard 24:548 center-tapped transformer. In terms of
implementing the integrated circuit, the single-chip receiver uses a 0.13 µm BiCMOS
process, which has a nominal supply voltage of 1.2 V.
Chapter 4

Circuit Implementation

This chapter covers the circuit implementation of the single-chip VLF magnetic field
receiver. The primary challenge in the design of the receiver is achieving the high
level of performance outlined in Section 3.4, while minimizing the power dissipation.
To reduce the power dissipation, the receiver uses a 1.2 V supply voltage. As a
result, low-voltage design techniques are required to achieve the performance goals
with reduced headroom [51]. To further complicate the design, the receiver is required
to be capable of operating in remote environments, such as at the South Pole where
temperatures can be below −70◦ C in the winter [47]. To overcome this challenge,
steps are taken in the design of the receiver to ensure that it can operate properly
over the wide range of conditions that may be encountered in the field.
This chapter is divided into five parts. The first part covers the design of the
low-noise amplifier (LNA). This section includes an overview of the LNA architecture
and the design details for the input stage, gain stage and automatic biasing circuits.
The second part covers the design of the analog-to-digital converter (ADC). This part
includes the top-level design of the continuous-time delta-sigma modulator and the
transistor-level design of the RC integrator stages, the comparator, the flipflop and
the feedback digital-to-analog converters (DACs). The third part covers the design
of the two-stage Miller compensated operational transconductance amplifier (OTA),
which is used throughout the LNA and ADC. The fourth part covers the design of the
full receiver, which includes the top-level architecture and the strategies employed to

47
CHAPTER 4. CIRCUIT IMPLEMENTATION 48

+ +
Input Gain
Iin Vout
Stage Stage
− −

Gain
Mode
Vbias Autobias

Figure 4.1: Top-level block diagram of the LNA.

reduce noise coupling between the LNA and ADC. Finally, the fifth part details the
layout of the test chip.

4.1 LNA Implementation

4.1.1 Overview
The top-level architecture of the LNA is shown in Figure 4.1. The LNA is a two-stage
amplifier. The first amplifier stage, the input stage, is a transimpedance amplifier
that converts the input signal from a current to a voltage. The input stage also
provides a low input impedance to properly match the antenna and provide an optimal
receiver frequency response. According to Frii’s formula for noise, the first stage
of a multi-stage amplifier contributes the most noise to the system assuming that
subsequent stages provide considerable gain [22]. As a result, the noise performance
is a primary design consideration for the LNA input stage to ensure good overall
receiver sensitivity.
The second amplifier stage, the gain stage, is a variable-gain voltage amplifier with
four discrete gain modes. The gain stage produces the final output of the LNA, which
means that its output will experience the largest voltage swings. For this reason, the
CHAPTER 4. CIRCUIT IMPLEMENTATION 49

linearity of the gain stage is a primary concern in order to meet the spurious-free
dynamic range design goal. Noise is also a consideration, but it is not as critical
as in the input stage. The gain stage has four discrete gain modes, where the gain
mode is selected by a two-bit digital signal that is controlled external to the chip. A
standard decoder with NAND logic is used to convert the two-bit digital signal into
four one-hot control lines.
A bias voltage is required to set the DC current in the input stage. There are
two methods to set this bias voltage. The first is with an external voltage that is
set outside the test chip. This voltage could be set with a potentiometer or provided
by a DC power supply. While this method may be useful for basic testing, it can
run into problems in the field due to temperature variations and component drift.
To overcome these problems, a second method to set the input stage bias voltage is
included. The second method is an automatic biasing circuit, which is more robust.
This method uses a feedback circuit to automatically set the optimal bias voltage
regardless of temperature changes or component drift. An external control signal
is used to select the biasing mode using an analog multiplexer that consists of two
CMOS transmission gates.

4.1.2 Input Stage


There are several competing goals in the design of the LNA input stage. The input
stage must provide a large transimpedance gain (current-to-voltage) while adding as
little noise as possible to the signal to ensure good receiver sensitivity. It also needs to
have a low input impedance to avoid adversely affecting the input turnover frequency,
which is the frequency above which the receiver response is flat. Finally, the input
stage must not degrade the linearity of the receiver or have a large offset voltage.
A number of different amplifier architectures could be used in the input stage, each
with different performance characteristics. For example, previous work has explored
using a resistor to terminate the antenna followed by a voltage amplifier. However,
this architecture does not have optimal noise performance. Another idea might be
to use a voltage amplifier with feedback to reduce the input impedance, but stability
CHAPTER 4. CIRCUIT IMPLEMENTATION 50

RL RL

+ Vout −

Vbias
Q1 Q2

+ Iin −

Figure 4.2: Schematic of the common-base input stage.

problems are encountered. A thorough analysis of the different architectures revealed


that a low-impedance amplifier produces the best overall performance [39]. For this
reason, a differential bipolar common-base amplifier is selected.
A common-gate amplifier, the CMOS equivalent, would also be suitable for this
application. However, the bipolar implementation was selected for two reasons. First,
the flicker noise (1/f noise) performance of the bipolar transistors was superior in the
target process. Second, bipolar transistors have better current efficiency, meaning
that they can generate the same transconductance with lower bias current. The
better current efficiency translates to an input stage that uses less power for equivalent
performance.
The common-base input stage is shown in Figure 4.2. The amplifier consists of
two bipolar transistors, Q1 and Q2. The collectors of the two transistors are each
connected to a load resistor, RL . The emitters serve as the differential input port and
the base is connected to a DC bias voltage.
The gain of the input stage can be calculated by considering the differential half-
circuit and applying small-signal analysis. The transimpedance gain, which is the
ratio of the output voltage to the input current, is

Vout
AZ = = αRL (4.1)
Iin
CHAPTER 4. CIRCUIT IMPLEMENTATION 51

where
β
α= (4.2)
1+β
Typical β values for NPN transistors in integrated circuits range from 50 to 500,
which means that α is approximately one [25]. Therefore, the transimpedance gain
of the input stage can be approximated as RL , the value of the load resistor. This
result makes sense, because under ideal conditions the current gain of a common-base
amplifier is one.
The input impedance of the input stage of the LNA directly affects the frequency
response of the receiver. Recall from Section 2.1.2 that the input turnover frequency
is
Ra + Rin /m2
fi = (4.3)
2πLa
At frequencies below fi , the magnitude of the receiver output is proportional to
frequency. At frequencies above fi , the output is flat with frequency. Ideally Rin /m2
would be zero to minimize the input turnover frequency, maximizing the flat region
of the frequency response. However, this is not practical, so typically the input
impedance is set such that Rin /m2 ≈ Ra . With a standard 1 Ω, 1 mH antenna
and the 24:548 transformer (m = 22.83), this gives an input turnover frequency of
approximately 318 Hz, which matches up nicely with the start of the VLF band (300
Hz).
The input impedance of the differential common-base amplifier can be calculated
using small-signal analysis, resulting in

2
Rin = 1 (4.4)
gm + rπ

The factor of two is due to the fact that the amplifier is differential. Assuming that the
β of the transistor is much greater than one, then rπ is large and the input impedance
simplifies to
2
Rin = (4.5)
gm
This result shows that the input impedance depends primarily on the transconduc-
tance of the NPN transistors, which in turn depends on the bias current because
CHAPTER 4. CIRCUIT IMPLEMENTATION 52

gm = IC /VT . Increasing the current will decrease the input impedance, while decreas-
ing the current will increase the input impedance. To satisfy Rin /m2 ≈ Ra , the bias
current in each side of the amplifier is set to 100 µA.
The noise performance of the input stage is critical to the sensitivity of the re-
ceiver. This is because, according to Frii’s formula for noise, the first amplifying
stage of a receiver primarily sets the noise performance of the full receiver. The noise
performance of the common-base input stage can be analyzed by adding the relevant
noise sources. Figure 4.3 shows the differential half-circuit with the noise sources
added. The load resistor contributes mean-square thermal noise of

1
i2L = 4kT ∆f (4.6)
RL

where k is the Boltzmann constant, T is the temperature and ∆f is the bandwidth


over which the noise is being calculated [25]. The bipolar transistor contributes shot
noise and flicker noise through two noise sources with mean-square values of

i2c = 2qIC ∆f (4.7)


IBa
i2b = 2qIB ∆f + K1 ∆f (4.8)
f

where IC is the collector current, IB is the base current, q is the charge of an electron
(1.6×10-19 ), and K1 and a are constants that capture the 1/f noise characteristics of
a particular device [25].
To analyze the effect of these noise sources on the sensitivity of the receiver, the
noise sources can be referred to the input of the amplifier. This process results in an
input voltage noise source and input current noise source that are equivalent to the
internal noise sources of the amplifier. The input referred mean-square noise sources
are
 
1
vi2 = 4kT rb + ∆f (4.9)
2gm
IBa
 
2 IC 1
ii = 2q IB + K1 + 2
∆f + 4kT ∆f (4.10)
f β(jω) RL
CHAPTER 4. CIRCUIT IMPLEMENTATION 53

RL i2L

Vout

Vbias Q1 i2c

i2b

Iin

Figure 4.3: Simplified schematic of the input stage for noise analysis.

Notice that the noise from the load resistor is unattenuated when referred to the
input. This is due to the fact that the current gain of the amplifier is one.
The receiver sensitivity depends on the noise from the antenna and the input
referred noise from the LNA. Recall, from Section 2.1.2, that the receiver sensitivity
can be expressed as s
1 vi2
Ssys = va2 + + i2i m2 Za2 (4.11)
ωN A m2
In this expression the voltage noise terms are dominant at low frequencies, while the
input referred current noise dominates in the receiver bandwidth because the antenna
impedance Za increases with increasing frequency.
As a result, there are several design insights that can be gained from the input
referred current noise expression in Equation 4.10. First, the load resistance should
be made as large as possible to reduce its contribution to the noise. Second, the bias
current should be as small as possible to reduce the noise from the NPN transistor.
Both of these insights are consistent with other goals for the input stage. For exam-
ple, the resistor is already set to the maximum possible value to give the maximum
gain while still providing a suitable DC output voltage. Similarly, the bias current
CHAPTER 4. CIRCUIT IMPLEMENTATION 54

is already set to the minimum value that produces the required input impedance.
Finally, the 1/f noise is made negligible by using physically large NPN devices. This
is accomplished by using many identical transistors connected in parallel.
Offset is another important design criteria for the input stage. This is because any
offset is amplified by subsequent stages of the LNA, which could potentially degrade
the performance of the receiver. Offset is caused by mismatch between components
that are nominally identical. For example, the load resistors in the input stage are
both designed to have the value RL . However, random variations in the fabrication
process will cause these resistors to not be exactly identical. Slightly different load
resistor values will lead to slightly different DC output voltages for the two halves of
the differential circuit. Mismatch between two components can be calculated with

AX
σ∆X = √ (4.12)
WL

where σ∆X is the standard deviation of the mismatch between two nearby components,
AX is a constant that characterizes the mismatch of a particular feature of the device,
and W and L represent the width and length of the device [41].
To reduce the offset to an acceptable level, large areas are used for the both the
NPN transistors and the load resistors. For the transistors, many unit sized devices
are connected in parallel to increase the effective area. For the resistors, a large width
is used to increase the total resistor area. Using large devices also helps reduce flicker
noise, which is also inversely related to the device area.
The distortion of the input stage is another design consideration. However, since
the signals at the output of the input stage are small compared to the signals in the
subsequent stages of the LNA, it is expected that any distortion is minor. At any
rate, the transimpedance gain of the input stage is simply RL , which means that the
linearity depends primarily on the linearity of the load resistor itself. There are several
different types of resistors that can be built on integrated circuits. For this design, a
high resistivity polysilicon resistor was used because it offered the best linearity when
compared with the diffused resistors and well resistors in the target process.
CHAPTER 4. CIRCUIT IMPLEMENTATION 55

Vin− + R1 R2
OTA Vout+

R3

+−
R4a R4b R4c R4d OTA
−+

R3

OTA Vout−
Vin+ + R1 R2

Figure 4.4: Schematic of the gain stage, which is implemented with a fully-differential
instrumentation amplifier.

4.1.3 Gain Stage


The goal for the gain stage is to provide variable voltage gain while contributing as
little noise and distortion as possible to the receiver. The gain stage must also pro-
vide a large input impedance to prevent loading the input stage. The gain stage is
implemented with a fully-differential adaptation of an instrumentation amplifier [9].
An instrumentation amplifier is a common OPAMP circuit that is known for having
a large input impedance, low noise and low distortion. The architecture for the gain
stage is shown in Figure 4.4. The circuit consists of three operational transconduc-
tance amplifiers (OTAs) and a resistive network that sets the gain. The gain stage
has fully-differential input and output ports. A switch network is used to select the
R4 resistor from a set of four possible values, which implements the four gain modes.
The gain of the gain stage can be divided into two parts. The first half consists
of the two input OTAs, resistor R3 and switchable resistor R4 . The gain of the first
CHAPTER 4. CIRCUIT IMPLEMENTATION 56

half is  
2R3
Av1 = 1+ (4.13)
R4
The values of the switchable resistor R4 are chosen to implement 10 dB gain steps.
Small values of R4 give large gain, while large values of R4 give small gain. The
second half of the gain stage consists of the output OTA and resistors R1 and R2 .
The second half has a gain of
R2
Av2 = (4.14)
R1
Combining these equations gives a total gain for the gain stage of
 
R2 2R3
Av = Av1 Av2 = 1+ (4.15)
R1 R4

The total gain of the gain stage is selected to give the appropriate range of gain values
when combined with the gain of the input stage.
The input impedance of the gain stage is large because the differential input port
is connected directly to the input of the OTAs. The CMOS transistors in the input
differential pair of the OTA provide a capacitive impedance, which won’t adversely
affect the input stage.
The noise performance of the gain stage is important. It must be designed such
that any noise contribution is small when referred back to the input of the receiver.
The noise from the gain stage can be represented by an equivalent noise voltage and
2
noise current at the input of the gain stage, vi,GS and i2i,GS respectively. Referring
these quantities to an effective noise current at the input of the receiver results in

2
vi,GS
i2i = i2i,GS + (4.16)
RL2

Notice that the gain stage current noise is unattenuated when referred to the input
of the receiver. This is because the current gain of the input stage is one. As a result,
the current noise of the gain stage is particularly important. On the other hand,
the voltage noise is reduced by the gain of the input stage, which relaxes its design
requirement.
CHAPTER 4. CIRCUIT IMPLEMENTATION 57

The equivalent voltage noise and current noise sources at the input of the gain
stage depend on the noise from the OTAs and the noise from the resistors. The
equivalent noise sources are
   
2 2 R4 R4
vi,GS = vi,OTA + i2i,OTA R3 k + 4kT R3 k ∆f (4.17)
2 2
i2i,GS = i2i,OTA (4.18)

2
where vi,OTA is the voltage noise from the OTA and i2i,OTA is the current noise from the
OTA. These expressions show that the current noise of the OTA must be minimized
because it determines the current noise of the whole gain stage. Similarly, to reduce
the voltage noise, steps must be taken to keep the OTA voltage noise small and
the resistor values must be kept small. The dependence of the voltage noise on the
resistor values will lead to slightly different noise levels in each of the gain modes.
The highest gain mode (small R4 ) will have the best noise performance, while the
lowest gain mode (large R4 ) will have the worst noise performance.
A two-stage CMOS architecture is used in the design of the OTAs. The equivalent
noise sources for the OTA are
 
2 γp γn gm3
vi,OTA = 8kT 1+ ∆f (4.19)
gm1 γp gm1
 
2 2 γp γn gm3
i2i,OTA = 2qIG ∆f + 4kT ω Cgs 1+ ∆f (4.20)
gm1 γp gm1

where γp is a coefficient that characterizes the noise of the PMOS transistors, γn is a


coefficient that characterizes the noise of the NMOS transistors, gm1 is the transcon-
ductance of the input differential pair, gm3 is the transconductance of input stage load
transistors, and IG is the gate current of the input differential pair transistors. The
CMOS architecture was chosen specifically for its good current noise performance, as
shown in Equation 4.20. The first term in the current noise expression depends on
the gate current (typically less than 1×10-15 A for CMOS transistors) and the second
term is proportional to frequency squared. As a result, generally the current noise
of a CMOS OTA is only significant at frequencies that are well above the maximum
CHAPTER 4. CIRCUIT IMPLEMENTATION 58

frequency of interest for a VLF receiver. The voltage noise from the OTA depends on
the transconductance of the input differential pair. The voltage noise can be reduced
by increasing gm1 , which corresponds to increasing the bias current. The OTA noise
expressions shown above include thermal noise and shot noise. A detailed analysis of
the OTA noise performance is given in Section 4.3.2, which also includes the effect of
1/f noise.
The distortion of the gain stage is important because it processes large signals
with considerable gain. In particular, the amplifier that drives the final output will
limit the distortion performance because it handles the largest signal swings. The
linearity is maximized through the use of feedback. A basic negative feedback block
diagram is shown in Figure 4.5. The circuit consists of a forward amplifier with gain
a and a feedback network with gain f . In the LNA gain stage, the forward amplifier
is an OTA and the feedback network is a resistive network. The gain of the basic
feedback network is
vo a
= (4.21)
vi 1 + af
The loop gain is defined as T = af . If the loop gain is much greater than one then
the gain of the feedback circuit simplifies to

vo 1
= (4.22)
vi f

Feedback has a number of benefits, including the fact that the gain of the system
depends only on the feedback factor f if the loop gain is large. Feedback also has an
effect on the distortion performance of the system [40]. The harmonic distortion can
be represented in terms of the HD 2 and HD 3 , which are the ratio of the amplitude
of the 2nd and 3rd harmonic to the amplitude of the fundamental. When feedback is
applied the HD 2 and HD 3 become

1
HD 2,FB = HD 2 (4.23)
(1 + a1 f )2
2a22 f
1− a3 (1+a1 f )
HD 3,FB = HD 3 (4.24)
(1 + a1 f )3
CHAPTER 4. CIRCUIT IMPLEMENTATION 59

vi a vo

Figure 4.5: Block diagram of the ideal feedback configuration.

Applying feedback improves the HD 2 by a factor of (1 + T )2 and the HD 3 by a factor


of (1 + T )3 . Clearly, increasing the loop gain can lead to significant increases in the
linearity of the amplifier.
Several strategies have been applied to ensure the best possible linearity of the gain
stage. First, the OTAs were designed to have a large DC gain and maximum possible
signal swing. Given the low frequency of the receiver, the bandwidth of the amplifiers
is not a significant concern so the OTAs are optimized for high DC gain. Second, the
gain in the gain stage is split into two parts, which results in a larger feedback factor
for each section and therefore a larger loop gain for each section. Third, it is expected
that the switches used to select the R4 resistor will introduce some nonlinearity. As
a result, the switchable resistor is placed in the first section of the gain stage where
the signal swings are smaller and the switch nonlinearity will have a smaller effect.
Fourth, since the output amplifier will contribute most significantly to the distortion
performance, a relatively small gain of three is implemented with R1 and R2 to ensure
that the loop gain is as large as possible in this section of the gain stage.

4.1.4 Auto-Bias Circuit


Due to the unique nature of the circuits used in the LNA, biasing can be a challenge.
In the gain stage a constant-gm current source is used to give consistent performance
over variations in process, voltage and temperature. However, the input stage of the
LNA is more difficult to properly bias due to the fact that its output DC voltage
depends on the NPN collector current and on the load resistor value, both of which
CHAPTER 4. CIRCUIT IMPLEMENTATION 60

can vary significantly. The output DC voltage is given by

VOUT = VDD − IC RL (4.25)

If the DC output voltage varies too much, it may go out of the suitable input range
of the gain stage causing the receiver to not function properly. In past designs,
this problem has been overcome by setting the NPN bias voltage externally using a
potentiometer. Using this method the voltage can be tuned to produce the proper
bias point. However, this approach has a serious drawback. While the bias point
can be properly tuned when the receiver is initially deployed, the bias point may shift
over time due to changes in temperature or component drift. Recall that the collector
current of the NPN is  
Vbias
IC = IS exp (4.26)
kT /q
which depends directly on temperature [25].
To overcome this problem an automatic biasing system has been implemented.
The auto-bias circuit is shown in Figure 4.6. The circuit operates by measuring
the DC voltage at the output of the input stage using a common-mode detector.
The common-mode detector consists of two large resistors that calculate the average
of the differential outputs, which is effectively the DC output voltage. The DC
output voltage is then compared with the desired output voltage, Vref . The difference
between the actual DC output voltage and the desired voltage is amplified with an
error amplifier and fed back to the input stage transistors. Essentially, this negative
feedback loop automatically adjusts the bias voltage to ensure that the output DC
voltage is equal to the reference voltage. In this design, the reference voltage is set
to VDD /2 in order to keep the outputs centered between the two supply rails. This
auto-bias strategy is similar to how a common-mode feedback circuit controls the DC
output voltage of a fully-differential OPAMP.
There are two important considerations for the design of the auto-bias circuit.
First, the resistors in the common-mode detector must be very large to ensure that
they don’t reduce the output impedance of the input stage and so that they don’t
adversely affect the noise performance of the input stage. To satisfy this requirement
CHAPTER 4. CIRCUIT IMPLEMENTATION 61

RL RL

Vbias
Q1 Q2
Common
Rx Rx Mode
Iin+ Iin− Detector

+
A
− Vref
Error Amplifier

Figure 4.6: Schematic of the auto-bias circuit and its connection to the input stage.

the resistors are each 1 MΩ. Second, the error amplifier itself must be robust to
changes in temperature and component drift, otherwise the auto-bias circuit will fail
to operate properly. This is accomplished by using a constant-gm current source to
bias the error amplifier.
The architecture for the error amplifier is shown in Figure 4.7. It is a single-stage
differential pair amplifier with a current mirror load. The single-stage architecture
provides the benefit of easing the requirements for stabilizing the feedback loop be-
cause it only produces a single dominant pole, while still giving more than enough
DC gain for the feedback loop to operate correctly.
Transistor-level simulations were performed to validate the auto-bias circuit. Fig-
ure 4.8 shows the LNA transimpedance gain as a function of the temperature of the
LNA over a range of −80◦ C to 120◦ C. This simulation includes transistor-level cir-
cuits for the input stage, gain stage, auto-bias circuit and all other relevant circuit
blocks. The red curve represents the performance when the LNA is biased using a
manual bias voltage that is set external to the chip. In this case, the manual bias
CHAPTER 4. CIRCUIT IMPLEMENTATION 62

VDD

M6 M5

Vin+ M1 M2 Vin−

Vout

Ibias M3 M4

Figure 4.7: Schematic of the error amplifier used in the auto-bias circuit.

voltage was set for optimal performance at around 25◦ C. Notice how the gain is
consistent over a limited range of roughly ±10◦ C. Outside of this range the gain of
the amplifier falls dramatically, which indicates that the LNA is no longer operating
properly. Temperature variations of this amount could easily be encountered between
day and night and between different seasons when the receiver is deployed. The blue
curve represents the performance when the auto-bias circuit is enabled. With the
auto-bias circuit the gain is consistent over the entire range of temperatures, with a
variation of less than 1 dB.

4.2 ADC Implementation


The ADC was implemented with a fully-differential continuous-time delta-sigma mod-
ulator. The primary reason this ADC architecture was selected was that it has an
implicit anti-alias filter. In previous VLF receiver designs the AAF has used a large
percentage of the total power consumption of the signal path blocks (LNA, AAF
and ADC). For example, in the Penguin receiver the AAF used 47% of the total
CHAPTER 4. CIRCUIT IMPLEMENTATION 63

Figure 4.8: LNA temperature sweep simulation.

signal path power [32]. Eliminating the AAF can result in significant power savings.
Additionally, delta-sigma modulators are robust to mismatch in their components.
Mismatch is often a key contributor to distortion in differential circuits. In fact,
past ADC designs for VLF receivers have required calibration to achieve over 90 dB
SFDR [49]. Delta-sigma modulators provide a notable advantage in terms of their
insensitivity to mismatch when compared with other ADC architectures. Together,
these qualities make the continuous-time delta-sigma modulator an ideal candidate
for the single-chip VLF receiver.

4.2.1 Loop Filter Design


A third-order modulator with a single-bit quantizer was used in this design. The
third-order loop filter provides good quantization noise suppression at a reasonable
oversampling ratio without running into the stability problems seen with higher order
loop filters. The oversampling ratio is set to 150, which at a 100 kHz Nyquist sampling
rate results in an ADC clock frequency of 15 MHz. The single-bit quantizer was
CHAPTER 4. CIRCUIT IMPLEMENTATION 64

X(t) 1 c1 1 c2 1 c3 Y (n)
b1
− s − s − s −

a1 a2 a3 a4

Figure 4.9: Block diagram of the third-order continuous-time delta-sigma modulator.

selected for its perfect linearity, which is due to it only having two levels. Using a
multi-bit architecture would have added unnecessary complexity and made achieving
the stringent linearity goal even more challenging.
The modulator block diagram is shown in Figure 4.9. A distributed feedback
architecture was used, which is evident because of the multiple feedback paths that
connect to the loop filter. The distributed feedback architecture was selected for its
superior anti-alias filtering characteristic. The third-order loop filter consists of three
continuous-time integrators that are represented by the 1/s blocks. The modulator
has several gain coefficients. The input signal is scaled with an input coefficient, b1 .
The output of each integrator is scaled with its own scaling coefficient, cn . Similarly,
each feedback path has its own gain coefficient, an . Finally, a single-bit compara-
tor with a half-sample delay was used for the quantizer in this design. Although a
typical third-order distributed feedback modulator would only have three feedback
coefficients, this design includes a fourth feedback coefficient, a4 , which is used to
compensate for the effect of the half-sample delay.
The loop filter design begins with the design of a discrete-time loop filter, which
was later converted to a continuous-time filter. The discrete-time loop filter was de-
signed using the Delta Sigma Toolbox [44]. This toolbox includes dozens of functions
for the design, synthesis and simulation of delta-sigma modulators in the MATLAB
environment. Here it was used to generate optimal gain coefficients for the discrete-
time loop filter given the filter order, oversampling ratio, out-of-band gain and filter
type. Figure 4.10 shows the resulting discrete-time loop filter noise transfer function
and signal transfer function. The horizontal axis in this plot is frequency and the
vertical axis is the response in dB. The blue curve shows the NTF. At low frequencies
the NTF rolls off at a rate of 60 dB/decade, which is consistent with a third-order
CHAPTER 4. CIRCUIT IMPLEMENTATION 65

Figure 4.10: Noise transfer function and signal transfer function of the modulator.

filter. At high frequencies the maximum gain of the NTF is set to 1.5 in order to
satisfy the modified Lee criterion for single-bit modulator stability [45, 10]. The STF
is flat over the signal band, as expected.
A common modification to the standard loop filter is to optimize the placement
of the NTF zeros [45]. Spreading the zeros throughout the signal band can further
reduce the in-band quantization noise and therefore increase the SQNR at the same
oversampling ratio. For example, a third-order loop filter with optimized zeros gives
an SQNR improvement of 8 dB. Zero optimization is accomplished in a third-order
loop filter by adding an additional feedback path from the output of the third integra-
tor to the input of the second integrator. This additional feedback path is generally
implemented with a resistor. Unfortunately, in this design the additional feedback
path would require a resistor with a value of over 50 MΩ, which was not practical to
implement in an integrated circuit. For this reason, non-optimized zeros are used in
this design.
The discrete-time loop filter was converted to a continuous-time loop filter using
the impulse-invariant transform [13]. This method determines the continuous-time
CHAPTER 4. CIRCUIT IMPLEMENTATION 66

loop filter coefficients that result in an equivalent impulse response when measured at
the sampling instants. This is done by setting the impulse response of the discrete-
time loop filter equal to the impulse response of the continuous-time loop filter. This
relationship can be expressed as

Z −1 {H(z)} = L−1 {RDAC (s)H(s)} |t=nTs (4.27)

where Z −1 is the inverse z-transform, L−1 is the inverse Laplace transform and
RDAC (s) is the DAC transfer function used in the continuous-time modulator im-
plementation.
After conversion, the outputs of the integrators are scaled to limit their signal
swing. This process is referred to as dynamic range scaling and is accomplished by
adjusting the cn coefficients in the loop filter. It is important to note that changes
to the cn coefficients will also require changes to the other coefficients in order to
maintain the same loop filter response. With a continuous-time modulator, behavioral
simulations are needed to determine the scaling coefficients because the discrete-time
model in MATLAB only predicts the integrator outputs at the sampling instants. The
integrator outputs in a continuous-time modulator may have larger values between
samples.
The signal swing at the output of the integrators is directly linked to the dis-
tortion performance of the ADC. Small signal swings will result in better distortion
performance because the OTAs have less non-linearity with smaller signal swings.
Unfortunately, the scaling also affects the size of the capacitors in the integrators.
More scaling results in larger capacitances, which will negatively affect the frequency
response of the OTAs. Therefore, a design trade-off exists where the amount of scal-
ing has be chosen to give the desired distortion performance while not impacting the
OTA performance. The most important integrator for distortion performance is the
first stage, because any errors in the first stage are referred directly to the input.
Distortion caused by the later integrator stages will have a smaller effect because the
error occurs at less sensitive points in the modulator. For this reason, the first inte-
grator is scaled to have a maximum signal swing of 40% full scale, while the second
CHAPTER 4. CIRCUIT IMPLEMENTATION 67

C1 C2 C3 Rx
R1 R2 R3 R4
Vin OTA OTA OTA OTA FF Dout
R1 R2 R3 R4
C1 C2 C3 Rx
DAC

DAC

DAC

DAC
Figure 4.11: Schematic of the third-order continuous-time delta-sigma modulator.

and third integrators are scaled to have a swing of 50% full scale.

4.2.2 Modulator Architecture


The top-level schematic of the continuous-time delta-sigma modulator is shown in
Figure 4.11. The modulator consists of three continuous-time integrator stages, fol-
lowed by a summation amplifier, a comparator, a flip-flop and four feedback DACs.
The input signal is a differential voltage with a full scale range that extends from
ground to the supply rail (1.2 V). The output of the modulator is a 1-bit digital
signal. Both the non-inverted and inverted output signal are used by the modulator,
but only the non-inverted signal is buffered and sent off-chip.
The filter stages are implemented with active-RC integrators. There are many
different continuous-time integrator architectures, but the active-RC circuit was se-
lected because it offers the highest linearity. The transfer function of the active-RC
integrator is
1
H(s) = (4.28)
sRC
where R is the input resistor and C is the integrating capacitor [38]. This relationship
can be used to calculate the resistor and capacitor values that implement the correct
loop filter gain coefficients. In many cases, the value of the resistor is set by the noise
requirement, and the gain coefficient is used to determine the appropriate capacitor
CHAPTER 4. CIRCUIT IMPLEMENTATION 68

value.
A dynamic comparator is used to implement the single-bit quantizer in this mod-
ulator. A flip-flop follows the comparator to capture its output value and add a
half-sample delay. The half-sample delay serves two purposes. First, it gives the
comparator a half-sample to settle to its final value before the value is captured by
the flip-flop. This reduces the speed requirement on the comparator. Second, it
eliminates the effect of signal-dependent delay. A common problem with dynamic
comparators is that the settling time of the output depends on the input value. This
causes errors in the modulator feedback path and adversely affects the noise shaping
performance. By adding a fixed half-sample delay before the flip-flop captures the
comparator output, the signal-dependent delay problem of the quantizer is eliminated.
Unfortunately, adding a half-sample delay to the modulator changes the transfer
function and itself reduces the noise shaping performance. This negative effect can
be exactly canceled by adding an additional feedback path at the input of the com-
parator, which is implemented with the fourth DAC (a4 coefficient). The value of a4
is selected to match the impulse response of the half-sample delayed loop filter to the
ideal discrete-time loop filter that the design is based upon [3].
The overall performance of the modulator is largely determined by the noise per-
formance. There are two sources of noise in the modulator: quantization noise and
circuit noise. The quantization noise is caused by the error introduced in the quan-
tizer. This noise is shaped by the delta-sigma modulator. As a result, the amount of
quantization noise in the signal band is determined by the loop filter design, which
was discussed in the previous section. The circuit noise is caused by thermal noise,
flicker noise and other noise sources in the passive and active components that are
used to implement the modulator. Typically the quantization noise is designed to be
10 dB to 20 dB below the circuit noise in order to allocate as much of the noise budget
as possible to the circuit noise. This design strategy has the added benefit of allowing
extra margin for the quantization noise shaping to be degraded by mismatch or other
non-ideal effects in the circuit without adversely affecting the overall performance of
the ADC.
The total allowable input referred noise power as a function of the resolution of
CHAPTER 4. CIRCUIT IMPLEMENTATION 69

the ADC is given by


2
VFS
vn2 = (4.29)
12 (2B − 1)2
where VFS is the full-scale voltage of the ADC, and B is the desired resolution. In
this design the quantization noise is set to be about 20 dB below the circuit noise.
Therefore, the contribution of quantization noise is negligible and the majority of the
noise power can be allocated to circuit noise.
The modulator circuit noise performance is limited by the circuit components
connected to the input of the modulator. Later stages of the modulator don’t con-
tribute significant noise because of the gain that precedes them. Figure 4.12 shows a
simplified schematic of the first stage of the modulator for noise analysis. The circuit
includes the first active-RC integrator stage and the resistor from the first feedback
DAC. There are three noise sources: the input resistor R, the DAC resistor RDAC and
the OTA. The values of the noise sources are

vR2 = 4kT R∆f (4.30)


2
vDAC = 4kT RDAC ∆f (4.31)
2
   
2 γp γn gm3 Kfp 1 Kfn 1 gm3 ∆f
vOTA = 4kT 1+ ∆f + + 2
(4.32)
gm1 γp gm1 Cox W1 L1 Cox W3 L3 gm1 f

The resistors contribute thermal noise, while the OTA contributes both thermal and
1/f noise. A detail analysis of the OTA noise performance is given in Section 4.3.2.
All of these noise sources can be referred to the input of the ADC to determine the
total input referred noise. The value of each noise source when referred to the input
is given by

2
vi,R = vR2 (4.33)
 2
2 2 R
vi,DAC = vDAC (4.34)
RDAC
 2
2 2 R
vi,OTA = vOTA 1+ (4.35)
RDAC
CHAPTER 4. CIRCUIT IMPLEMENTATION 70

RDAC C

2
vDAC
vR2 2
vOTA
R
+ −
Vin OTA
− +
R
vR2
2
vDAC

RDAC C

Figure 4.12: Simplified schematic of the first stage of the modulator for noise analysis.

The noise from the input resistor is unchanged, but the other noise sources are mod-
ified when referred to the input. The total input referred noise of the delta-sigma
modulator can be calculated by summing these values, which gives
 2  2 !
R R
vi2 = 2 vR2 + vDAC
2 2
+ vOTA 1 + (4.36)
RDAC RDAC

where the factor of two is due to the fact that the modulator is a differential circuit. In
this design, the loop filter input coefficient b1 is equal to the first feedback coefficient
a1 , which results in R = RDAC . This relationship can be used to simplify the total
input referred noise expression above.
There are two competing goals when designing the first stage of the modulator.
The first goal is meeting the noise requirement, which mandates that small resistance
values be used to stay within the noise budget set by Equation 4.29. The second goal is
minimizing the distortion, which requires that large resistance values be used because
they lead to better linearity from the modulator [5]. To satisfy both requirements, the
CHAPTER 4. CIRCUIT IMPLEMENTATION 71

resistance values are set to the maximum value that still meets the noise requirement.
Further, since it is desirable to make the resistors as large as possible, the majority
of the noise in the first stage is allocated to the resistors. In this design 80% of the
noise budget is allocated to the resistors and 20% is allocated to the OTA. Reducing
the noise contribution from the OTA requires increasing the power dissipation. The
slightly higher power dissipation will have to be tolerated in order to achieve the
stringent linearity goals.
The later integrator stages are not as critical to the overall noise and distortion
performance of the modulator, which means that the design constraints on the passive
components and on the OTA in these stages are relaxed. The noise requirements are
not as high for the later stages because of the gain of the stages that precede them.
The distortion requirement is also not as aggressive because errors introduced by the
later stages are rejected by the noise shaping effect of the modulator. These reduced
requirements lead to larger resistors, smaller integrating capacitors and lower-power
OTAs. As a result, the power dissipation of the ADC can be optimized by scaling
back the power dissipation of the later integrator stages.
There are four OTAs in the modulator, each with a different design objective.
The most stringent requirements are in the design of the input stage OTA, which
requires very low noise, high gain and minimal distortion. The requirements are
reduced for the OTAs in the second and third integrators. The OTA used in the
summation amplifier is designed for maximum bandwidth with lower gain. A two-
stage Miller compensated OTA architecture is used to implement all of the OTAs.
The same architecture is also used to implement the OTAs in the LNA portion of the
receiver. The design of the two-stage Miller compensated OTA is discussed separately
in Section 4.3.

4.2.3 Circuit Blocks


The modulator uses a single-bit quantizer that is implemented with a regenerative
latch comparator [4]. Figure 4.13 shows the schematic of the comparator circuit.
The circuit consists of two cross coupled transistor pairs, one PMOS and one NMOS,
CHAPTER 4. CIRCUIT IMPLEMENTATION 72

VDD

Vin+ M1 M2 M2 M1 Vin−

CLK M3 M3 CLK

Vout− Vout+

CLK M5 M4 M4 M5 CLK

Figure 4.13: Schematic of the regenerative latch comparator.

which create a positive feedback loop. It also has a differential input and differential
output, as well as a clock signal which controls the latch. The comparator is a dynamic
circuit, which means that it ideally has no DC power consumption.
The basic operation of the comparator has two phases. In the first clock phase,
when the clock is high, the comparator is reset by the two M5 transistors that pull
the outputs to ground. The positive feedback loop is disconnected in the reset mode
by disabling the two M3 transistors. The second phase occurs when the clock is low.
As soon as the clock goes low the regenerative latch is enabled and the comparator
makes a decision based on the difference between the two inputs. This decision is
latched and doesn’t change until the comparator is reset again.
The comparator is the least critical block in the continuous-time delta-sigma mod-
ulator because any error introduced by the comparator is rejected by the full noise
shaping effect of the modulator. As a result, the offset and hysteresis of the com-
parator are not critical design constraints. Instead, the primary design constraint for
the comparator is the timing constraint. Namely, the output of the comparator must
settle to its final value within a half sample. After a half sample the value is captured
by the flipflop. The regenerative latch settling time is only a small fraction of the half
sample requirement, so the transistors are sized to reduce power consumption while
CHAPTER 4. CIRCUIT IMPLEMENTATION 73

still comfortably meeting the timing constraint.


The flipflop is implemented with a differential semistatic flipflop [52]. The flipflop
uses a standard master-slave architecture, which consists of two latches where the
output of the master latch is connected to the input of the slave latch. The master
latch is shown in Figure 4.14a. It is a dynamic latch that has high speed and low
power consumption. It consists of a cross-coupled PMOS pair and has a differential
input and output. The slave latch is shown in Figure 4.14b. The slave latch is a static
latch, which ensures that the outputs of the flipflop are firmly set to the positive and
negative rails. A differential flipflop is used because the feedback DACs each require
the non-inverted and inverted output values. By using a differential flipflop it is not
necessary to use an inverter to generate the complementary signal, which could lead
to problems with unequal delay.
Both the master latch and the slave latch run on the same clock phase, which is
referred to as single-phase clocking. The comparator is also designed to run with the
same clock phase. Therefore, all of the clocked circuits in the ADC run on the same
clock phase. This greatly simplifies the clock generation and distribution network on
the test chip as there is no need to generate and synchronize an inverted clock signal.
The feedback DACs are implemented with a current-steering DAC that uses a
resistance to set the value of the feedback current. Figure 4.15 shows the schematic
of the DAC. The DAC is a differential circuit and consists of two resistors, with value
RDAC , that are connected to either the positive reference voltage or the negative
reference voltage depending on the value of the modulator output. The inputs to
the DAC are a complementary digital signal, which means that one of the resistors
is connected to the positive reference while the other is connected to the negative
reference. In this design the reference voltages are set to ground and VDD , so the
value of each DAC resistance can be calculated with

VDD
RDAC = (4.37)
an fs Cn

where an is the loop filter feedback coefficient, fs is the sampling rate and Cn is the
value of the integrating capacitor.
CHAPTER 4. CIRCUIT IMPLEMENTATION 74

VDD

CLK M1

M2 M2

Vout− Vout+

Vin+ M3 M3 Vin−

(a) Master latch.

VDD

M6 M6

Vout− Vout+

Vin+ M5 M7 M7 M5 Vin−

CLK M4

(b) Slave latch.

Figure 4.14: Schematic of the semistatic flipflop.


CHAPTER 4. CIRCUIT IMPLEMENTATION 75

Vrefp

M1 M1
RDAC RDAC
Vin+ + OUT − Vin−

M2 M2

Vrefn

Figure 4.15: Schematic of the current-steering feedback DAC.

The single-bit DAC is inherently linear, so the primary concern in the DAC design
is matching the positive and negative slope of the feedback current to prevent inter-
symbol interference. This is accomplished by properly setting the ratio of the width
of the PMOS to NMOS transistors in the DAC. In this design, the PMOS transistors
are three times wider than the NMOS transistors.
In a discrete-time delta-sigma modulator the loop filter gain coefficients are imple-
mented with ratios of capacitors, which are typically accurate to within 1% if proper
layout techniques are applied [29]. On the other hand, in a continuous-time delta-
sigma modulator the gain coefficients are set by the RC products in the active-RC
integrators. Process variations of 10% to 20% are not uncommon for integrated re-
sistors and capacitors, which can result in an RC product variation of over 30% [38].
This RC product variation changes the loop filter coefficients and can lead to reduced
noise shaping, and in severe cases it can cause modulator instability.
To overcome this problem the integrating capacitors (C1 -C3 ) are tunable. This
allows for the values of the capacitors to be changed to compensate for the process
variations of the RC products. Figure 4.16 shows the schematic of the tunable ca-
pacitor [50]. The circuit uses a single fixed capacitor plus three binary weighted
unit capacitors. The three control bits give a total of eight capacitance values. The
switches are implemented with CMOS transmission gates. The total capacitance of
CHAPTER 4. CIRCUIT IMPLEMENTATION 76

b0 Cu

b1 2Cu

b2 4Cu

Cfix

Figure 4.16: Schematic of the tunable capacitor.

the bank is given by


2
X
Cbank = Cfix + Cu bi 2i (4.38)
i=0

where [b2 b1 b0 ] is the 3-bit control signal. The fixed capacitance, Cfix , and the unit
capacitance, Cu , are selected to implement the desired capacitance value with a tuning
range of −30% to +40% with 10% steps. In this application, the capacitors are
tuned manually using external control signals. The appropriate tuning value can be
determined by trial and error or by measuring the provided on-chip test structures.

4.3 OTA Implementation


Operational transconductance amplifiers (OTAs) are used throughout the receiver
design. The LNA uses three OTAs: two single-ended OTAs at the input of the
gain stage and one fully-differential OTA at the output of the gain stage. The ADC
uses four OTAs: three fully-differential OTAs in the RC integrator stages and one
fully-differential OTA in the summation amplifier.
This section covers the design of the OTA circuit that is employed throughout
the receiver. A common architecture has been selected for simplicity, which is then
optimized to fit the requirements of each individual amplifier. For example, the OTA
in the first RC integrator stage in the ADC requires a low-noise OTA with high gain,
CHAPTER 4. CIRCUIT IMPLEMENTATION 77

Table 4.1: OTA architecture performance comparison.

Architecture Gain Output Swing Speed Power Noise


Telescopic Medium Medium Highest Low Low
Folded-Cascode Medium Medium High Medium Medium
Two-Stage High Highest Low Medium Low
Gain-Boosted High Medium Medium High Medium

while the ADC summation amplifier OTA requires high-bandwidth and low-power.
These are very different design goals, but both of these OTAs are successfully imple-
mented with the same OTA architecture using different design parameters (transistors
sizes and bias currents).
It is important to point out the distinction between OTAs and OPAMPs. The dif-
ference between an OTA and an OPAMP lies in its output stage. An OTA has a large
output impedance, which generally makes it only useful for driving high impedance
loads. On the other hand, an OPAMP includes an output buffer that has a low
output impedance and is capable of driving low impedance loads. While the circuit
presented in this section may be commonly referred to as an OPAMP, technically it
is an OTA due to its high output impedance.

4.3.1 Architecture Comparison


There are a number of different amplifier architectures that could be used in the
receiver, each with its own advantages and disadvantages. Table 4.1 shows a compar-
ison of four popular OTA architectures in terms of their gain, output swing, speed,
power dissipation and noise [42]. The four architectures are the telescopic OTA, the
folded-cascode OTA, the two-stage OTA and the gain-boosted OTA.
There are four main requirements for the OTA used in this design. First, it must
have a large DC gain to ensure good linearity of the system. Recall that the harmonic
distortion is directly related to the loop gain, which itself depends on the DC gain
of the amplifiers. Second, the OTA noise must be low to give good overall receiver
sensitivity without consuming unnecessary power. Third, the amplifier must have a
CHAPTER 4. CIRCUIT IMPLEMENTATION 78

large available output swing to accommodate large signal swings without degrading
the performance. Finally, the receiver runs on a 1.2 V supply voltage, which limits
the available headroom. As a result, the amplifier must have a minimum number of
vertically stacked transistors.
The two-stage OTA architecture most closely fits the requirements outlined above
and is selected for use in this design. This architecture has high gain, the highest
output swing, and low noise. The one downside of the two-stage architecture is
its speed, but the bandwidth of the VLF receiver is low so speed is not a critical
requirement.

4.3.2 Two-Stage OTA


This section covers the design and analysis of the fully-differential version of the two-
stage Miller compensated OTA architecture [25]. A single-ended version of the circuit
is also used in the receiver and many of the results for the fully-differential version
also apply to the single-ended version. The schematic of the fully-differential version
of the two-stage OTA architecture is shown in Figure 4.17. The circuit uses only
CMOS transistors and consists of two amplification stages.
The gain of the two-stage OTA can be calculated by considering the two stages
separately. The first stage consists of the input transistor, M1, and the current source
loads, M3. The gain of the first stage is

Av1 = gm1 (ro1 k ro3 ) (4.39)

where gm1 is the transconductance of M1, ro1 is the output impedance of M1 and ro3
is the output impedance of M3. The second stage, which consists of transistor M2
and current source load M4, has a gain of

Av2 = gm2 (ro2 k ro4 ) (4.40)

where gm2 is the transconductance of M2, ro2 is the output impedance of M2 and
ro4 is the output impedance of M4. The total gain of the OTA can be calculated by
CHAPTER 4. CIRCUIT IMPLEMENTATION 79

VDD

M4 M5
M6 M4

Vin+ M1 M1 Vin−
Vout+ Vout−
VB VB
Cc Cc

M7 VCMFB M7
Ibias M2 M3 M3 M2

Figure 4.17: Schematic of the fully-differential two-stage OTA.

combining the gain from the first stage and the gain from the second stage, which
gives a total gain of

Av = Av1 Av2 = gm1 (ro1 k ro3 ) gm2 (ro2 k ro4 ) (4.41)

The gain of the amplifier can be increased by increasing the channel length of the
transistors, which results in a larger output impedance and therefore higher gain.
The input impedance of the amplifier is essentially infinite due to the PMOS
differential pair at the input. The PMOS transistors present a capacitive load to the
driving circuit. The output impedance of the amplifier depends on the output stage
transistors and is
Rout = ro2 k ro4 (4.42)

The large output impedance gives a large gain for the output stage, but limits the
types of loads that the amplifier can drive. For example, a small resistive load will
reduce the gain of the amplifier because it is in parallel with the output impedance.
CHAPTER 4. CIRCUIT IMPLEMENTATION 80

One of the key advantages of the two-stage OTA architecture is the output range.
Because there are no cascoded transistors in the output stage, the output voltage can
swing within one overdrive voltage of the supply rails. Therefore, the output range is

Vov2 <= Vo <= VDD − Vov4 (4.43)

where Vov2 is the overdrive voltage of M2 and Vov4 is the overdrive voltage of M4. The
output range can be maximized by reducing these overdrive voltages by increasing
the W/L (width over length) ratios of the output transistors.
The input offset voltage of a fully-differential amplifier is defined as the differen-
tial input voltage that results in zero differential output voltage. The offset can be
calculated as a function of the mismatch between pairs of components in the amplifier
circuit. The input offset voltage of the two-stage OTA is
   
gm3 ∆Vov1 ∆(W/L)3 ∆(W/L)1
Voffset = ∆Vt1 + ∆Vt3 + − (4.44)
gm1 2 (W/L)3 (W/L)1

where ∆Vt1 is the difference between the threshold voltage of the two M1 transistors,
∆(W/L)3 is the difference between the width over length ratio of the two M3 tran-
sistors, and so forth. The offset voltage can be reduced by using a longer channel
length for M3 than M1, which gives a small gm3 /gm1 ratio. It can also be reduced by
operating M1 with a small overdrive voltage, which reduces the ∆Vov1 term.
The frequency response of the two-stage OTA is compensated using the Miller
capacitance Cc . The compensation capacitor splits the poles generating a dominant
pole [25]. However, the compensation capacitor also creates a right half-plane zero
that can degrade the frequency response and cause the amplifier to have a negative
phase margin and become unstable. Transistor M7 corrects this problem, creating
a resistance that can be used to move the zero to higher frequencies. With the
resistance, the location of the zero is

1
z=  (4.45)
1
gm2
− Rz Cc
CHAPTER 4. CIRCUIT IMPLEMENTATION 81

By operating M7 in the linear region, as a resistor, the zero can be moved to infinity
if the resistance is equal to 1/gm2 . This is accomplished by making M7 identical to
M2 and operating it with the same gate-source voltage. By doing this, the negative
effect of the zero on the frequency response and phase margin can be eliminated.
After compensation and zero cancellation the poles of the two-stage OTA are

−1
p1 = (4.46)
R1 (C1 + Cc ) + R2 (C2 + Cc ) + gm2 R2 R1 Cc
−gm2 Cc
p2 = (4.47)
C1 C2 + Cc (C1 + C2 )
−1
p3 = (4.48)
Rz C1

where R1 is the output resistance of the first stage, R2 is the output resistance of the
second stage, C1 is the total capacitance connected to the gate of M2 and C2 is the
total capacitance connected to the output node. The first two poles determine the
majority of the frequency response and set the phase margin, while the third pole is
at a very high frequency and typically has little effect.
The noise of the two-stage OTA is determined primarily by the transistors in the
first stage of the amplifier. The noise analysis can be simplified to the circuit shown
in Figure 4.18. The circuit is a differential half-circuit that includes only the input
transistor M1 and the first stage current source load transistor M3. Together the noise
contributions of these two devices dominate the noise performance of the two-stage
OTA.
The noise generated by a MOSFET can be divided into two parts. The first is
thermal noise and the second is 1/f noise. The full expression for the drain current
noise of a MOSFET is
2
Kf gm ∆f
i2d = 4kT γgm ∆f + (4.49)
Cox W L f

where the first term represents the thermal noise and the second term represents the
1/f noise [42]. In the thermal noise k is the Boltzmann constant, T is temperature,
γ is a device dependent coefficient, gm is the transconductance of the device and ∆f
CHAPTER 4. CIRCUIT IMPLEMENTATION 82

VDD

Vin M1 i2d1

VCMFB M3 i2d3

Figure 4.18: Simplified schematic of the OTA for noise analysis.

is the bandwidth over which the noise is being computed. In the 1/f noise term Kf
is a technology dependent coefficient, Cox is the oxide capacitance, W is the width of
the device, L is the length of the device and f is frequency.
Using the expression above for the drain current noise power of a MOSFET, the
input referred noise voltage and input referred noise current of the two-stage OTA
can be calculated. The analysis is performed separately for thermal noise and 1/f
noise. The input referred thermal noise voltage is
 
2 γp γn gm3
vi,thermal = 8kT 1+ ∆f (4.50)
gm1 γp gm1

The thermal noise voltage can be reduced by increasing the transconductance of M1,
which is accomplished by increasing the bias current. It can also be reduced by
ensuring that the gm3 /gm1 ratio is small. This is done by making the channel length
of M3 longer than M1. The additional factor of two in the expression comes from the
fact that the OTA is a differential circuit. The input referred 1/f noise voltage can
be calculated using the same method, which results in

2
 
2 Kfp 1 Kfn 1 gm3 ∆f
vi,1/f =2 + 2
(4.51)
Cox W1 L1 Cox W3 L3 gm1 f
CHAPTER 4. CIRCUIT IMPLEMENTATION 83

The 1/f noise voltage can be reduced by increasing the area of M1 and M3. Similar to
the thermal noise, the 1/f noise voltage can also be reduced by reducing the gm3 /gm1
ratio.
The input referred current noise can be calculated for the two-stage OTA using a
similar procedure. The input referred thermal noise current is
 
2 2 γp γn gm3
i2i,thermal = 2qIG1 ∆f + 4kT ω Cgs1 1+ ∆f (4.52)
gm1 γp gm1

where q is the charge of an electron (1.6×10-19 ), IG1 is the gate current of M1, ω is
the frequency in radians per second and Cgs1 is the gate source capacitance of M1.
The first term in the expression is technically shot noise, not thermal noise, but it
is included here for completeness. The shot noise of a CMOS amplifier is typically
negligible because the gate current of a MOSFET is very small (typically less than
1×10-15 A). The second term in the input referred current noise is the result of thermal
noise. It is usually only significant at very high frequencies because it scales with the
square of the frequency. In the event that it is significant in this design, it can be
reduced by increasing the transconductance of M1 or by reducing the gm3 /gm1 ratio.
It is important to note that there is no additional factor of two when calculating the
input referred current noise of a differential circuit. The input referred 1/f noise
current is
2
 
2 2 Kfp 1 Kfn 1 gm3 ∆f
i2i,1/f =ω Cgs1 + 2
(4.53)
Cox W1 L1 Cox W3 L3 gm1 f
The input referred 1/f noise current also scales with frequency. Similar to the input
referred 1/f noise voltage, it can be reduced by using large devices and by reducing
the gm3 /gm1 ratio.
The DC output voltage of the fully-differential two-stage OTA is not well defined,
which creates the need for a circuit to regulate this voltage. A common-mode feedback
circuit is used for this purpose. The common-mode feedback circuit used in this design
is shown in Figure 4.19. The circuit can be divided into two parts. The first part is
the common-mode detector, which measures the DC output voltage of the OTA. The
common-mode detector is implemented with the two RCM resistors. These resistors
CHAPTER 4. CIRCUIT IMPLEMENTATION 84

VDD

Vbias M10

RCM
Vout+
M8 M8 Vref
Vout−
RCM

VCMFB
M9 M9

Figure 4.19: Schematic of the common-mode feedback circuit.

essentially calculate the average of the positive and negative outputs of the OTA.
The resistors are given a value that is much larger than the output impedance of the
OTA to ensure that they don’t affect the gain. The second part of the common-mode
feedback circuit is the error amplifier, which compares the DC output voltage of the
OTA with the desired value, Vref , and amplifies the difference. The output of the
error amplifier is fed back to the OTA at the gate of the M3 transistors. This process
creates a negative feedback loop that keeps the DC output voltage fixed at the desired
value (Vref ). This particular common-mode feedback circuit was selected primarily
because it doesn’t limit the output swing of the OTA, which can be problematic with
other common-mode feedback circuits.
The slew rate of the two-stage OTA depends on both stages of the amplifier circuit.
As a result, the analysis can be divided into the internal slew rate and the external
slew rate. The internal slew rate is

2ID1
SR int = (4.54)
Cc
CHAPTER 4. CIRCUIT IMPLEMENTATION 85

where ID1 is the bias current of M1 and Cc is the compensation capacitor. The
internal slew rate arises because of the need to charge and discharge the compensation
capacitor. The external slew rate is

2ID4
SR ext = (4.55)
(CL + Cc )

where ID4 is the bias current of M4 and CL is the load capacitance at the output of
the OTA. The external slew rate is the slew rate of the output stage of the OTA and
is caused by the need to charge and discharge the load capacitance and compensation
capacitance. The overall slew rate of the two-stage OTA is the minimum of the
internal and external slew rates. It can be increased by increasing the bias currents,
at the expense of power dissipation.
The preceeding analysis covers the fully-differential two-stage OTA, which is used
in both the LNA and ADC. Additionally, a single-ended version of the circuit is
used in the gain stage of the LNA. Figure 4.20 shows the single-ended version of the
two-stage Miller compensated OTA circuit. The primary difference is that only one
output stage is included and no common-mode feedback circuit is necessary. Most
of the design equations for the fully-differential version of the amplifier also apply to
the single-ended amplifier.

4.3.3 Design Methodology


The design equations outlined in the previous section are used to guide the design of
the two-stage OTAs that are used in the receiver. Each of the OTAs in the receiver
has a different set of design goals. For example, some amplifiers require high gain,
some require high speed, and others require low noise. Given the number of unique
OTAs that need to be designed, an efficient design methodology is needed to quickly
arrive at designs that achieve the performance goals.
One common issue with directly using the design equations derived from the small-
signal model of the amplifier is that the hand calculations often give very different
results than a circuit simulator. This is caused by the fact that the square law model
for MOSFETs doesn’t always match well with the complex simulation models. This
CHAPTER 4. CIRCUIT IMPLEMENTATION 86

VDD

M5
M6 M4

Vin− M1 M1 Vin+
Vout
VB
Cc

M7
Ibias M3 M3 M2

Figure 4.20: Schematic of the single-ended two-stage OTA.

problem is particularly pronounced for short-channel devices. To overcome this prob-


lem the gm /ID design methodology has been applied [33]. To use this method, first
the NMOS and PMOS transistors are characterized. This is accomplished by creat-
ing a test-bench with a single PMOS and NMOS transistor and sweeping them over
their full range of operating points. At each operating point the relevant transistor
parameters are logged in design tables. These parameters include things like the
transconductance, bias current, threshold voltage, output resistance, parasitic capac-
itances and noise coefficients. There are several ways to organize this data, but one
popular method is to organize it as a function of the ratio gm /ID . Although it’s not
obvious, this is a beneficial organization because the gm /ID ratio is directly related
to the transistor characteristics. For instance, a small gm /ID ratio leads to fast tran-
sistors, while a large gm /ID ratio leads to high efficiency or low-power transistors.
Now when designing, instead of calculating parameters using the square law models,
the parameters can be found by looking them up in the design tables. The matching
between predicted performance and simulation results is greatly improved by using
parameters from the design tables.
CHAPTER 4. CIRCUIT IMPLEMENTATION 87

The basic methodology used to design the OTAs involves two steps. First, the de-
sign is iterated in MATLAB to arrive at a set of amplifier parameters (transistor sizes
and bias currents) that meet the design specifications for that particular amplifier.
The basic inputs for the iterative design are the transconductance of each transistor,
the channel length of each transistor and the gm /ID value for each transistor. The
amplifier performance is predicted using the design equations derived in Section 4.3.2
and the design tables are used to look up the transistor parameters. Once a design
that meets the specifications is found, the second step involves simulating the design
at the transistor-level to verify that the predicted performance matches with simu-
lation results. The simulations verify that the approximations made in the design
equations are valid and that the amplifier performs as expected.
This design approach allows for much smarter and more efficient design because
the design equations show the most important knobs that can be tuned to achieve
the desired performance goals, rather than blindly changing parameters in the circuit
until the design works. Further, the gm /ID design methodology ensures good matching
between hand calculations and simulation results.

4.4 Receiver Implementation

4.4.1 Overview
The top-level receiver architecture consists of the interconnection of the LNA and the
ADC. Thanks to the implicit anti-alias filtering of the continuous-time delta-sigma
modulator, no additional filtering is required. The full receiver schematic is shown
in Figure 4.21. The input of the receiver is a fully-differential signal that originates
at the antenna and is connected to the receiver with the transformer. The output of
the receiver consists of two digital signals from the ADC, which are the 1-bit digital
output data stream and the 15 MHz clock output.
In addition to the primary input and output signals, the receiver has a number
of control signals. The LNA has three control signals: an external bias voltage that
is used to optionally bias the LNA input stage, a bias mode selection signal that
CHAPTER 4. CIRCUIT IMPLEMENTATION 88

+
IIN LNA ADC Dout

CLK out

SEL
Gain[1:0]

Tune[2:0]
Vbias

RESET
CLK
Figure 4.21: Block diagram of the receiver.

selects between either automatic biasing or manual biasing, and a 2-bit gain selection
signal that selects between the four LNA gain modes. In addition, the ADC has three
control signals: a 15 MHz square wave clock signal, a reset signal that clears the RC
integrators, and a 3-bit RC tuning signal that adjusts the capacitors to compensate
for process variation.

4.4.2 Biasing
For the purpose of validating the performance of the receiver in the lab, a simplistic
biasing architecture could have been employed. For example, a lab current source
could have been used to provide the reference current. However, the goal for the
single-chip receiver was to design a system that was robust enough to be deployed in
the field without further modification. To satisfy this goal an on-chip biasing network
was designed that could withstand changes in process, voltage and temperature.
A constant-gm current source was selected to provide the bias currents for the
OTAs and the other active circuits in the receiver [34]. The schematic of the constant-
gm current source is shown in Figure 4.22. The output bias current of the constant-gm
current source is  
Vov1 1
Ibias = 1− √ (4.56)
Rref m
where Vov1 is the overdrive voltage of M1, Rref is an off-chip reference resistor and
m is the scaling ratio between transistors M2 and M1. While this may not seem
CHAPTER 4. CIRCUIT IMPLEMENTATION 89

VDD

Rref

m:1 M1
M2 M5

Ibias
M6

M4 M3

Figure 4.22: Schematic of the constant-gm current source.

very useful, when the transconductance is calculated it becomes more clear. The
transconductance of transistor M1 is
 
2 1
gm1 = 1− √ (4.57)
Rref m

The transconductance depends only on the reference resistor and the scaling ratio.
This means that any transistor biased with this reference current will have the same
transconductance independent of changes in process, voltage or temperature. Given
that the noise of the OTAs depends directly on the transconductance of the input
devices, the constant transconductance characteristic will provide predictable noise
performance regardless of the operating conditions. Transistor M6 in the circuit is a
diode connected MOSFET that ensures that the self-biased feedback loop starts up
correctly when power is applied to the current source.

4.4.3 Noise Coupling


The coupling of noise between the various circuits on an integrated circuit is often a
challenging design problem [23, 46]. In this receiver design it is particularly important
CHAPTER 4. CIRCUIT IMPLEMENTATION 90

as noise generated by the ADC could corrupt the sensitive LNA. In fact, this is part
of the reason that in past VLF receiver designs the LNA was separated from the
noisy digital portion of the receiver by placing them in separate enclosures with a
long cable between them. Three different strategies have been used to reduce the
amount of on-chip noise coupling in the single-chip receiver.
First, differential signaling has been used throughout the receiver design. All
signal path connections are fully-differential, from the LNA input through to the
delta-sigma ADC output. Differential signaling is much more robust to noise than
single-ended signals assuming that the circuits are matched and have a good common-
mode rejection ratio (CMRR).
Second, separate supplies are used for the various circuit blocks in the design. This
ensures that noise does not couple through the power or ground rails. There are five
different supplies used in the single-chip receiver: LNA analog supply, ADC analog
supply, ADC digital supply, ADC reference supply and ESD protection supply. Each
supply consists of a power and ground connection that are routed to dedicated pins
on the chip. Each supply has its own decoupling network on the test board. There
are no on-chip connections between the supplies. Further, the LNA and the ADC
each have their own reference current source to prevent noise from coupling through
the bias network.
Third, the layout is optimized to minimize noise coupling. The two main circuit
blocks, the LNA and the ADC, are physically separated on the die by approximately
60 µm to prevent noise coupling through the substrate. Within the ADC, the noisy
digital circuits are placed away from sensitive analog circuits and sensitive intercon-
nect is not run near noisy signals. The pin locations have been carefully selected
to ensure that noisy digital signals are not near the sensitive LNA input signals.
Guard rings are also used throughout the layout to locally isolate sensitive circuits
and prevent noise from coupling out of noisy circuits [29].
CHAPTER 4. CIRCUIT IMPLEMENTATION 91

4.5 Chip Layout


The single-chip receiver was fabricated in a 0.13 µm BiCMOS technology from Texas
Instruments. A photograph of the die is shown in Figure 4.23. The die area is 2.56
mm2 (1.6 mm × 1.6 mm). The large rectangular block on the right half of the die is
the ADC. The large rectangular block on the left half of the die is the LNA. There
are also two smaller test circuit blocks visible just above the LNA.
CHAPTER 4. CIRCUIT IMPLEMENTATION 92

Figure 4.23: Die photo of the single-chip receiver.


Chapter 5

Measurement Results

This chapter presents the measurement results for the single-chip VLF magnetic field
receiver. In addition to testing the full receiver, the LNA and ADC are each tested
individually. The results are divided into four sections with each section describ-
ing the test setup and corresponding measurements. The first section examines the
performance of the LNA, which includes measurements of the frequency response,
sensitivity, linearity and power dissipation. The second section covers the ADC per-
formance, which includes the dynamic performance, anti-alias filtering and power
dissipation. The third section presents the results of the full receiver, which includes
the dynamic performance, noise coupling, anti-alias filtering and power dissipation.
Finally, the fourth section details field test results that were collected at two VLF
antenna sites. This final section also compares the field test data with data from the
AWESOME receiver.
A custom printed circuit board (PCB) was designed to test the single-chip receiver.
A picture of the finished PCB is shown in Figure 5.1. The single-chip receiver is
visible near the center of the board. It was packaged in a 48-pin quad flat no-lead
(QFN) package that is roughly 7 mm × 7 mm. The hand-wound 24:548 center-tapped
transformer is also visible on the test board. To prevent coupling between traces on
the PCB, care was taken during design to physically separate the noisy digital signals
from the sensitive analog signals. Further, all of the differential signals are routed
side-by-side with length-matched traces to ensure that any noise that does couple

93
CHAPTER 5. MEASUREMENT RESULTS 94

Figure 5.1: Test board for the single-chip receiver.

into these lines will appear as a common-mode signal, which is then rejected by the
differential circuits in the receiver.

5.1 LNA Measurement Results

5.1.1 Test Setup


The test setup used to measure the LNA performance is shown in Figure 5.2. The
input signal for the LNA was generated with a Stanford Research Systems SR-1 Audio
Analyzer. The SR-1 was selected primarily for its ability to generate low-distortion
signals. The instrument is capable of generating a sinusoidal signal with a SFDR of
approximately 105 dB over the entire receiver bandwidth (300 Hz to 50 kHz). This
level of distortion is well below the 90 dB SFDR performance target of the receiver,
which means that the input signal from the SR-1 will not limit the SFDR of the
system and no additional filtering of the input signal is required. The SR-1 generates
a fully-differential signal. The analog output port on the SR-1 was connected to the
injection circuit on the test board with a shielded XLR cable to reduce the amount
CHAPTER 5. MEASUREMENT RESULTS 95

SRS SR-1 Output Signal


Audio Analyzer

DUT
Injection
Circuit
Input Signal Agilent E3630A
LNA ADC
Power Supply

Figure 5.2: Test setup for LNA measurements.

of noise coupling into the input of the LNA.


The injection circuit interfaces between the signal generator and the LNA and
is used to inject a test current in parallel with a dummy antenna as described in
Section 2.1.5. It consists of three main components. The first component converts
the differential voltage from the SR-1 into a current using a passive RC network. The
second component is a dummy antenna that is implemented with a 1 Ω resistor and
a 1 mH inductor. The dummy antenna mimics the impedance of an actual antenna.
The third component is the 24:548 center-tapped input transformer, which matches
the input impedance of the LNA to the impedance of the antenna. The injection
circuit components are represented with a single block in the test setup diagram.
The fully-differential output of the LNA was connected to the analog input port on
the SR-1 via a shielded XLR cable. The SR-1 was configured to compute the spectrum
of the signal, which enabled monitoring of the signal amplitude, noise floor, harmonics
and other performance metrics in real-time. Additionally, the SR-1 provided a 100 kΩ
load impedance. This impedance is roughly equivalent to the ADC input impedance,
which ensures that there is a good match between the LNA measurements and the
performance of the LNA when it is used in the full receiver system.
Power was supplied to the device under test (DUT) by an Agilent E3630A Triple
Output DC Power Supply. Separate supply and ground connections were used to
power the LNA and the ESD protection circuitry. Separate supplies are not necessary
for the LNA measurements, but are important for ADC and receiver measurements
CHAPTER 5. MEASUREMENT RESULTS 96

Figure 5.3: LNA frequency response.

to reduce noise coupling between the various circuit blocks. All components of the
receiver run on 1.2 V. Unless otherwise noted, all of the measurements use the auto-
bias setting for the LNA. In this setting the input stage bias voltage is set by the
automatic biasing system, not by an external reference voltage.

5.1.2 Frequency Response


The frequency response of the LNA was measured by configuring the signal generator
to sweep over its entire frequency range (10 Hz to 200 kHz) and calculate the ratio
of the output amplitude to the input amplitude at each test frequency. The results
are shown in Figure 5.3. The horizontal axis corresponds to the frequency of the
test signal and the vertical axis is the gain in dB. The gain includes the effect of the
input transformer, which means that it is the gain from the current generated by the
loop antenna to the voltage at the LNA output. Therefore, it is the gain of the full
front-end of the receiver.
The frequency response figure includes curves for all four LNA gain modes. The
CHAPTER 5. MEASUREMENT RESULTS 97

modes are separated by roughly 10 dB. Mode 0, the highest gain mode, has a peak
gain of 92.25 dB. Mode 1 has a peak gain of 82.93 dB. Mode 2 has a peak gain of
73.26 dB. Mode 3, the lowest gain mode, has a peak gain of 63.44 dB. The gain can
also be referred to the magnetic field at the antenna if the antenna parameters are
known. With a six-turn 4.9 meter square antenna, the gain of the front-end is 5.91
mV/pT in mode 0, 2.02 mV/pT in mode 1, 0.66 mV/pT in mode 2 and 0.21 mV/pT
in mode 3.
The transimpedance of the LNA can be calculated by adjusting for the effect of
the transformer. The current gain of the 24:548 transformer is approximately −27.17
dB. The effect of the transformer can be removed by adding this value back to the
peak front-end gain values from Figure 5.3. The resulting transimpedance values for
the four gain modes are 119.42 dB, 110.10 dB, 100.43 dB and 90.61 dB. The goal for
the LNA transimpedance was 100 dB, which means that the four gain modes enable
the end-user to select from a range of gain values that extends above and below the
target value.
The bandwidth of the LNA is also an important performance metric. The de-
sign goal was for the receiver bandwidth to extend from 300 Hz to 50 kHz. The 3
dB bandwidth of the LNA is calculated from the frequency response. The antenna
impedance causes the gain to roll-off at low frequencies. The antenna is unchanged
between the gain modes so the roll-off should be the same in all cases. As expected,
the 3 dB point is approximately 170 Hz in all of the gain modes. Roll-off at the
high-end of the frequency range is caused by the poles of the LNA, which are not
necessarily the same between gain modes. For this reason, it is expected that the
high frequency roll-off will occur at a different frequency in each gain mode. In the
highest gain mode, mode 0, the 3 dB point is at 117.18 kHz. In mode 1 the 3 dB
point is at 182.03 kHz. In mode 2 it’s at 191.94 kHz and in mode 3 it’s at 199.30
kHz. Although the bandwidth depends on the gain mode, in all cases it comfortably
exceeds the 50 kHz design goal.
CHAPTER 5. MEASUREMENT RESULTS 98

Figure 5.4: LNA output noise spectrum.

5.1.3 Sensitivity
The noise performance of a VLF receiver is characterized in terms of its sensitivity.
The sensitivity is defined as the magnetic field amplitude at the antenna that will
result in a 0 dB signal-to-noise ratio at the LNA output when measured in a 1 Hz
bandwidth [39]. Essentially, the sensitivity is the input amplitude that is equivalent
to the noise floor at the LNA output at any given frequency. Smaller values for the
sensitivity are better.
The sensitivity can be calculated from the output noise spectrum. The output
noise spectrum is measured by removing the input signal from the LNA and capturing
the output with the SR-1 Audio Analyzer. The SR-1 then performs an FFT to
generate the output spectrum. Figure 5.4 shows the LNA output noise spectrum
from 10 Hz to 50 kHz for each gain mode. A data length of 32 thousand samples was
used and the result was averaged 16 times to smooth the spectrum. As expected,
the noise floor increases as the gain increases because the input stage dominates the
noise performance of the LNA.
CHAPTER 5. MEASUREMENT RESULTS 99

The spurious tones visible between 10 Hz and 2 kHz are the result of power line
noise coupling into the system. The largest tones are visible at 60 Hz, 180 Hz, 300
Hz and 420 Hz, which are the fundamental, 3rd, 5th and 7th harmonic respectively.
There are also smaller tones visible at 50 Hz and its harmonics, as well as the even
harmonics of 60 Hz. This power line noise is likely coupling into the system at two
points. First, through the power supply rails from the E3630A DC Power Supply,
which runs on AC power. Second, through radiated electromagnetic interference
(EMI) from nearby electronic equipment, which is being captured by the large coil in
the input transformer. Fortunately, the power line noise is small enough and at low
enough frequency that it does not adversely affect the ability to measure the LNA
performance. It is also important to point out that in a real deployment the receiver
would run on battery power and would be located at a quiet site far away from any
other electronics. Under these conditions power line noise would not be a concern.
For the purpose of calculating the sensitivity of the LNA, the power line noise is
filtered out in post-processing and a best-fit line is computed.
The first step to compute the sensitivity of the LNA is to calculate the input
referred noise. This is done by dividing the output noise spectrum by the gain of
the LNA. The input referred noise is then converted into an equivalent magnetic field
amplitude at the antenna to determine the sensitivity. A six-turn 4.9 meter square
antenna was used for this calculation. The resulting sensitivity is shown in Figure 5.5
for each gain mode.
The sensitivity is below 1 fT/Hz1/2 over the entire target bandwidth (300 Hz to 50
kHz) in all four gain modes. The LNA has a minimum sensitivity of 0.240 fT/Hz1/2
in mode 0, 0.250 fT/Hz1/2 in mode 1, 0.282 fT/Hz1/2 in mode 2 and 0.398 fT/Hz1/2
in mode 3. These results are consistent with Equation 4.17, which predicted that the
noise would increase slightly in the lower gain modes due to the larger resistors that
are switched into the gain stage.
CHAPTER 5. MEASUREMENT RESULTS 100

Figure 5.5: LNA sensitivity.

5.1.4 Linearity
The linearity of the LNA is characterized by the peak spurious-free dynamic range
(SFDR). The SFDR is defined as the ratio of the input signal to the strongest spu-
rious signal, which is typically a harmonic of the input signal. For the purpose of
determining the SFDR, the power line noise seen in Figure 5.4 is not included in the
calculation.
To determine the peak SFDR the input signal is swept over a range of amplitude
values. Figure 5.6 shows the amplitude sweep for each gain mode. The horizontal
axis represents the peak-to-peak voltage applied to the injection circuit. The vertical
axis represents the calculated SFDR in dB at the output of the LNA. A 10 kHz input
frequency was used because it lies near the middle of the VLF band and because the
most important harmonics, the 2nd through the 5th, also fall in-band.
At small input signal amplitudes the SFDR increases linearly with the input am-
plitude. This is because the harmonics are below the noise floor so the SFDR is simply
the ratio between the input signal and the noise floor (the noise floor is constant).
CHAPTER 5. MEASUREMENT RESULTS 101

Figure 5.6: LNA SFDR versus input amplitude.

However, as the input signal amplitude increases, eventually it becomes large enough
that the harmonics in the LNA output spectrum exceed the noise floor and the plot
levels off and ultimately starts decreasing due to the harmonics growing at a faster
rate than the fundamental.
The peak SFDR varies depending on the gain mode. The peak SFDR is 87.81 dB
in mode 0, 95.01 dB in mode 1, 100.30 dB in mode 2 and 104.02 dB in mode 3. The
gain stage in the LNA limits the linearity. The reason that the peak SFDR is different
in each mode is because the loop gain of the feedback circuits changes depending on
which gain mode is selected. When the gain stage is in a low gain mode the loop gain
is high, which results in a higher SFDR. In contrast, when the gain stage is in a high
gain mode the loop gain is low, which results in a lower SFDR.
To further illustrate the peak SFDR, an output spectrum for each gain mode was
generated using a 10 kHz input signal. The amplitude of the input signal was chosen
to be the amplitude that resulted in the maximum SFDR from the corresponding
amplitude sweep. The output spectrum that illustrates the maximum SFDR for each
gain mode is shown in Figures 5.7 to 5.10.
CHAPTER 5. MEASUREMENT RESULTS 102

Figure 5.7: LNA output spectrum with the peak SFDR in mode 0.

Figure 5.8: LNA output spectrum with the peak SFDR in mode 1.
CHAPTER 5. MEASUREMENT RESULTS 103

Figure 5.9: LNA output spectrum with the peak SFDR in mode 2.

Figure 5.10: LNA output spectrum with the peak SFDR in mode 3.
CHAPTER 5. MEASUREMENT RESULTS 104

Figure 5.11: LNA SFDR versus input frequency.

Measurements were also taken to verify the linearity of the LNA as a function of
the input frequency. To accomplish this the input frequency was swept over the entire
amplifier bandwidth, from 300 Hz to 50 kHz. The input amplitude used for each gain
mode was the input amplitude that produced the maximum SFDR in the amplitude
sweep (Figure 5.6). The results of the frequency sweep are shown in Figure 5.11.
The linearity performance is consistent over the full range of frequencies. In mode
0 the SFDR hovers around 87 dB. In mode 1 the SFDR hovers around 96 dB. In
mode 2 the SFDR hovers around 100 dB. In mode 3 the SFDR hovers around 105
dB. There is a small decrease in the SFDR below 1 kHz, but this can be attributed
to the roll-off of the amplifier gain at the low-end of the frequency band.

5.1.5 Summary
The LNA includes four modes with gain values ranging from 92.25 dB to 63.44 dB
when the effect of the transformer is included. The 3 dB bandwidth extends from
approximately 170 Hz to over 100 kHz, with the high frequency cutoff depending on
CHAPTER 5. MEASUREMENT RESULTS 105

Table 5.1: LNA measurement results summary.

Measurement Mode 0 Mode 1 Mode 2 Mode 3


Front-End Gain (dB) 92.25 82.93 73.26 63.44
System Gain (mV/pT) 5.91 2.02 0.66 0.21
Bandwidth Start (Hz) 168.87 170.37 171.39 171.71
Bandwidth End (kHz) 117.18 182.03 191.94 199.30
Sensitivity (fT/Hz1/2 ) 0.240 0.250 0.282 0.398
Peak SFDR (dB) 87.81 95.01 100.30 104.02
Power Dissipation (µW) 907.88 908.60 908.48 908.72

the selected gain mode. The sensitivity is well below the 1 fT/Hz1/2 design goal, with
a range of 0.240 fT/Hz1/2 to 0.398 fT/Hz1/2 . The linearity of the LNA, which was
measured in terms of the peak SFDR, varies from 87.81 dB to 104.02 dB. The power
dissipation of the LNA was approximately 908 µW, which was measured using a BK
Precision 5492 5 1/2 Digit Multimeter.
Table 5.1 shows a full summary of the LNA measurement results for each gain
mode. It is important to point out that a six-turn 4.9 meter square antenna was used
in the calculation of the system gain and the sensitivity. Additionally, the peak SFDR
values are the maximum values from the amplitude sweep, which was performed with
a 10 kHz input signal.

5.2 ADC Measurement Results

5.2.1 Test Setup


The test setup for the ADC measurements is shown in Figure 5.12. Similar to the
LNA test setup, the input signal to the ADC is generated by the Stanford Research
Systems SR-1 Audio Analyzer. This signal generator was selected for its good dis-
tortion performance, which doesn’t require any additional filtering. It outputs a
fully-differential signal that is fed directly to the ADC input via a shielded XLR
cable.
The clock signal for the ADC is provided by an Agilent 81110A Pulse Generator.
CHAPTER 5. MEASUREMENT RESULTS 106

SRS SR-1 Input Signal


Audio Analyzer

DUT

Data
LNA ADC
Phase Agilent 16750A
Lock Input Clock Logic Analyzer
Clock

Agilent 81110A Agilent E3630A


Pulse Generator Power Supply

Figure 5.12: Test setup for ADC measurements.

This clock generator was selected for its good jitter performance of roughly 15 psRMS.
A clock with a large amount of jitter can degrade the performance of the ADC [12].
The clock generator was configured to output a 1.2 V single-ended square wave at 15
MHz.
In order to produce a clean output spectrum from the ADC it is necessary to pre-
vent spectral smearing. Smearing occurs when the input signal doesn’t fall directly
into one of the FFT bins. Although FFT windowing can be used to alleviate the
problem, it is possible to nearly completely eliminate smearing in a lab environment
when precise control of the input signal and clock source is available. This is accom-
plished by phase-locking the input signal to the clock source [31]. In the ADC test
setup this is done by connecting a copy of the 15 MHz clock to the external reference
port on the SR-1.
The ADC outputs two digital signals. The first is the 1-bit data stream at 15 MHz.
The second is the corresponding 15 MHz clock signal. The digital output signals are
captured with an Agilent 16750A Logic Analyzer. The logic analyzer is run in state
mode, which means that the data is collected synchronously with the 15 MHz clock
CHAPTER 5. MEASUREMENT RESULTS 107

output from the ADC. The resulting data files are stored on the logic analyzer and
then transferred to a PC for analysis in MATLAB.
Power for the device under test is provided by an Agilent E3630A DC Power
Supply. The setup uses four separate supply connections to power the test chip: ADC
analog supply, ADC digital supply, ADC reference voltage and ESD protection supply.
Each supply has a separate connection on the test board and its own dedicated ground
connection. The supply connections also each have a separate decoupling network to
prevent noise coupling. All of the supplies run at 1.2 V.

5.2.2 Output Spectrum


A typical output spectrum from the ADC is shown in Figure 5.13. The horizontal axis
represents frequency, which extends from DC to 7.5 MHz (fs /2). The vertical axis
is the magnitude of the signal in units of dBFS/NBW. In this particular example a
−10 dBFS sinusoidal signal at 10 kHz was input to the ADC. This signal is visible as
the large spike near the middle of the output spectrum. In order to prevent spectral
smearing, the exact input frequency is actually 10.070801 kHz, which falls exactly
in an FFT bin. The 1-bit ADC output bitstream was processed in MATLAB using
an appropriately scaled FFT with a Hann window. The data length was 2,097,152
samples.
The noise shaping from the delta-sigma modulator is clearly visible in the spec-
trum. The majority of the noise in the spectrum is pushed to higher frequencies
where it can be subsequently filtered with a digital decimation filter. This is possible
because the ADC oversamples the input signal, causing the majority of the noise to
be outside the desired 50 kHz band.
There are a few other relevant features of the spectrum that are worth pointing
out. The first is the spike at the far left of the spectrum. This peak corresponds
to the DC offset of the ADC. The second is the small peak around 60 Hz, which
can be attributed to power line noise. Both the DC offset and the 60 Hz noise are
not included in the performance calculations of the ADC, such as the spurious-free
dynamic range. This is because both of these signals fall below the lower limit of the
CHAPTER 5. MEASUREMENT RESULTS 108

Figure 5.13: ADC output spectrum.

receiver bandwidth (300 Hz). Finally, the 3rd and 5th harmonic of the 10 kHz input
signal are also visible in the spectrum at 30 kHz and 50 kHz, respectively.

5.2.3 Dynamic Performance


The performance of an ADC can be characterized using either static performance
metrics or dynamic performance metrics. Static performance performance metrics
include the gain error, offset, differential nonlinearity (DNL) and integral nonlinearity
(INL). Dynamic performance metrics include the signal-to-noise ratio (SNR), signal-
to-noise-and-distortion ratio (SNDR) and the spurious-free dynamic range (SFDR).
In general, delta-sigma modulators are characterized in terms of dynamic performance
metrics [38]. Therefore, dynamic performance metrics are used in this analysis.
By sweeping the input signal amplitude the peak value of the SNR, SNDR and
SFDR can be determined. Figure 5.14 shows the amplitude sweep from −80 dBFS
to 0 dBFS using a 10 kHz input frequency. The horizontal axis in this plot represents
the input amplitude in dB relative to the full-scale range of the ADC. The vertical
CHAPTER 5. MEASUREMENT RESULTS 109

Figure 5.14: ADC performance versus input amplitude.

axis represents the corresponding metric in dB.


As the input amplitude approaches 0 dBFS a large drop-off is observed in the
SNR, SNDR and SFDR. The metrics fall from around 80 dB at −4 dBFS to around
40 dB at −3 dBFS. This is normal behavior for a delta-sigma modulator and is the
result of the quantizer becoming overloaded, causing the feedback system to become
unstable. For a single-bit modulator this typically occurs a few dB below the full-scale
range. In the ADC in this design, the maximum stable amplitude is −4 dBFS.
The SNR, which measures the ratio of the signal to the sum of the in-band noise
components, is expected to increase monotonically as the input amplitude is increased.
This is because the noise floor is constant and the signal harmonics are excluded. This
trend is observed in the amplitude sweep and the peak SNR of the ADC is measured
to be 78.93 dB.
The SNDR measures the ratio between the signal and the sum of the in-band
noise including the harmonic distortion components. It is expected that the SNDR
will increase linearly until the harmonics exceed the noise floor, at which point the
curve will flatten out and then begin to fall. This expectation is consistent with the
CHAPTER 5. MEASUREMENT RESULTS 110

measurement results and the peak SNDR in the amplitude sweep is 76.43 dB. The
SNDR is used to determine the effective number of bits (ENOB) of the ADC using
Equation 2.26. The measured SNDR corresponds to an effective resolution of 12.40
bits.
The SFDR, which measures the ratio between the signal and the largest spurious
tone, is the primary metric used to characterize the linearity of the receiver. Similar
to the SNDR, it is expected that in the amplitude sweep the SFDR will increase
linearly until the harmonics exceed the noise floor, at which point it will level off and
start decreasing. From the amplitude sweep the peak SFDR of the ADC is 93.60 dB.
To further illustrate the maximum SFDR, Figure 5.15 shows the output spectrum
of the ADC with a −10 dBFS input signal at 10 kHz. This corresponds to the
case where the peak SFDR was achieved in the amplitude sweep in Figure 5.14.
Additionally, the plot is zoomed in on the 50 kHz ADC bandwidth to more clearly
show the relevant portion of the spectrum. The input signal is clearly visible at 10
kHz with an amplitude of −10 dBFS. The largest spurious signal is the 3rd harmonic
with an amplitude of −103.60 dB at approximately 30 kHz. Recall that the input
frequency isn’t precisely 10 kHz, which is why the 3rd harmonic is seen slightly above
30 kHz.
The SNR, SNDR and SFDR were also tested over the full range of possible fre-
quencies, from 300 Hz to 50 kHz. Figure 5.16 shows the frequency sweep performance.
A −10 dBFS input amplitude was used in this test because that was the amplitude
which resulted in the maximum SFDR in the amplitude sweep. The ADC perfor-
mance is consistent over the entire frequency range with the exception of the SFDR,
which increases for frequencies above 20 kHz. This is caused by the fact that as the
input frequency approaches the upper-end of the range, the harmonics start to move
out of the ADC bandwidth. For example, when the input frequency is 30 kHz, the
2nd harmonic is at 60 kHz and the 3rd harmonic is at 90 kHz, both of which are out
of band.
CHAPTER 5. MEASUREMENT RESULTS 111

Figure 5.15: ADC output spectrum with the peak SFDR.

Figure 5.16: ADC performance versus input frequency.


CHAPTER 5. MEASUREMENT RESULTS 112

Figure 5.17: ADC output spectrum with a 14.95 MHz input signal.

5.2.4 Anti-Alias Filtering


The implicit anti-alias filter of the continuous-time delta-sigma modulator is measured
by inputting a signal with a frequency that is close to the clock frequency, such that
it will alias to within the signal band. The aliased signal can then be measured
to determine the amount of alias rejection provided by the ADC. The SR-1 signal
generator is only capable of producing signals up to 200 kHz, so a different input
signal source is needed. For this purpose, a BK Precision 4086 80 MHz Function
Generator is used in place of the SR-1. The rest of the ADC test setup remains
unchanged.
The input signal is a −20 dBFS sinusoidal signal at 14.95 MHz, which is the worst
case frequency for aliasing. It is important to point out that the 4086 function gen-
erator is only capable of a single-ended output. The ADC expects a fully-differential
signal so the input signal is adjusted accordingly. Figure 5.17 shows the in-band
output spectrum from the aliasing test.
The 14.95 MHz signal aliases to approximately 50 kHz and is visible at the far
CHAPTER 5. MEASUREMENT RESULTS 113

Table 5.2: ADC measurement results summary.

Measurement Value
Peak SNR (dB) 78.93
Peak SNDR (dB) 76.43
Peak SFDR (dB) 93.60
ENOB (bits) 12.40
Alias Rejection (dB) 63.63
Power Dissipation (µW) 641.52

right of the spectrum. The strength of the signal is −83.63 dBFS, which corresponds
to an alias rejection of 63.63 dB from the −20 dBFS input signal. It is important to
note that this test is the worst case aliasing. Any other frequency will either see more
alias rejection, or will alias to a frequency that is out of band. In addition to the alias
tone, it would appear that there are several other small spurious signals visible in the
output spectrum. This is likely caused by the single-ended input signal used in this
test, which is more susceptible to interference.

5.2.5 Summary
The ADC was characterized in terms of its dynamic performance. The peak SNR was
78.93 dB. The peak SNDR was 76.43 dB, which corresponds to an effective resolution
of 12.40 bits. The ADC linearity was measured in terms of its peak SFDR, which
was 93.60 dB. The implicit anti-alias filter was measured to have a worst-case alias
rejection of 63.63 dB. Finally, the power dissipation of the ADC was 641.52 µW,
which was measured using a BK Precision 5492 5 1/2 Digit Multimeter.
Table 5.2 shows a summary of the ADC measurement results. The peak SNR,
peak SNDR and peak SFDR values are all derived from the amplitude sweep, which
was performed with an input frequency of 10 kHz. The alias rejection measurement
was performed with a 14.95 MHz input signal.
CHAPTER 5. MEASUREMENT RESULTS 114

SRS SR-1
Audio Analyzer

DUT
Injection
Circuit
Input Signal Data
LNA ADC
Phase Agilent 16750A
Lock Input Clock Logic Analyzer
Clock

Agilent 81110A Agilent E3630A


Pulse Generator Power Supply

Figure 5.18: Test setup for receiver measurements.

5.3 Receiver Measurement Results

5.3.1 Test Setup


The test setup for the receiver measurements is shown in Figure 5.18. The setup is
very similar to the ADC test setup. The input signal is generated by the Stanford Re-
search System SR-1 Audio Analyzer, which is used for its good distortion performance
that doesn’t require any additional filtering of the input signal. The SR-1 outputs
a fully-differential signal and is connected to the injection circuit with a shielded
XLR cable. The injection circuit consists of a passive voltage to current converter, a
dummy 1 Ω, 1 mH antenna and an input transformer. The fully-differential output
of the LNA is fed directly to the input of the ADC.
The clock signal for the ADC is generated by the Agilent 81110A Pulse Gener-
ator. This clock source was selected for its good jitter performance. The 81110A is
configured to output a single-ended 1.2 V square wave at 15 MHz. In addition to
connecting the clock signal to the ADC, it is also connected to the external reference
port on the SR-1 to facilitate phase-locking between the signal source and the clock
CHAPTER 5. MEASUREMENT RESULTS 115

source. This is done to prevent spectral smearing in the output spectrum.


The receiver output is captured by an Agilent 16750A Logic Analyzer. The single-
chip receiver outputs two digital signals. The first is the 1-bit data stream. The
second is the corresponding 15 MHz clock. The logic analyzer is operated in state
mode, which means that it captures data synchronously with the 15 MHz clock from
the receiver. The data is saved to files on the logic analyzer and then transferred to
a PC for analysis in MATLAB.
The receiver is powered by an Agilent E3630A DC Power Supply. The power
supply is configured to output 1.2 V. Separate power and ground connections are used
for each portion of the receiver to prevent noise coupling. There are five separate
power connections: LNA analog supply, ADC analog supply, ADC digital supply,
ADC reference voltage and ESD protection supply. Each power connection has its
own dedicated decoupling network.

5.3.2 Dynamic Performance


The full receiver can be characterized using the same dynamic performance metrics
as the ADC, which are the SNR, SNDR and SFDR. The peak values of these three
metrics are determined by sweeping the amplitude of the input signal and analyzing
the digital signal at the output of the receiver. A separate sweep is performed for
each LNA gain mode. The receiver amplitude sweeps for each gain mode are shown
in Figures 5.19 to 5.22. In some cases, particularly in mode 0, the SNR and SNDR
are difficult to differentiate because they are equal over much of the amplitude sweep.
A 10.070801 kHz input frequency was used and the amplitude was swept from −18
dBFS to −3 dBFS. The value of the amplitude corresponds to the amplitude of the
signal at the input of the ADC, which is not the same as the amplitude at the input of
the LNA. By specifying the amplitude in this way the results are consistent with the
ADC results shown in the previous section. The receiver output data was analyzed
in MATLAB using a 2,097,152 point FFT with a Hann window.
The peak SNR is different in each gain mode. The peak SNR in mode 0, mode
1, mode 2 and mode 3 are 58.31 dB, 66.71 dB, 74.29 dB and 77.98 dB respectively.
CHAPTER 5. MEASUREMENT RESULTS 116

Figure 5.19: Receiver performance versus input amplitude in mode 0.

Figure 5.20: Receiver performance versus input amplitude in mode 1.


CHAPTER 5. MEASUREMENT RESULTS 117

Figure 5.21: Receiver performance versus input amplitude in mode 2.

Figure 5.22: Receiver performance versus input amplitude in mode 3.


CHAPTER 5. MEASUREMENT RESULTS 118

The change in signal-to-noise ratio is due to the fact that the noise from the LNA
sets the noise floor of the output spectrum in all but the lowest gain mode. As shown
in Figure 5.4, the noise floor at the output of the LNA roughly scales with the gain.
Only in mode 3, the lowest gain mode, is the noise floor slightly below the noise floor
of the ADC. The SNDR follows a similar trend as the SNR except when the signal
amplitude becomes large, in which case the distortion terms cause it to level off. The
peak SNDR in mode 0, mode 1, mode 2 and mode 3 are 58.19 dB, 66.02 dB, 71.14
dB and 73.33 dB respectively.
The SFDR also varies in the different LNA gain modes. The peak SFDR in mode
0, mode 1, mode 2 and mode 3 are 78.98 dB, 84.50 dB, 90.08 dB and 91.91 dB
respectively. The amplitude where the peak SFDR occurs is also different in each
more. In the higher gain modes, mode 0 and mode 1, the peak occurs at −6 dBFS
and −12 dBFS. In contrast, in the lower gain modes, mode 2 and mode 3, the peak
occurs at −10 dBFS in both cases. This difference is likely due to the fact that the
LNA limits the SFDR in the higher gain modes, while the ADC limits in the SFDR
in the lower gain modes. Recall that the peak SFDR of the ADC also occurred at
−10 dBFS.
To further illustrate the linearity of the receiver, the output spectrum is shown for
each gain mode with the input amplitude that produced the peak SFDR. Figures 5.23
to 5.26 show the output spectrum of the receiver in each gain mode. There are
two features to point out in these figures. First, the spurious tones at the far left
of the spectrum are due to the 60 Hz power line noise that was first encountered
when measuring the LNA. Second, notice that the noise floor decreases as the gain is
reduced, which agrees with peak SNR trend discussed above.
In addition to performing the amplitude sweep, it is also important to sweep
the frequency range to uncover any frequency dependent effects in the receiver. To
accomplish this, the amplitude of the input signal is set to the value that resulted in
the maximum SFDR in the amplitude sweep, then the frequency is swept from 300
Hz to 50 kHz. A separate frequency sweep is performed for each LNA gain mode.
Figures 5.27 to 5.30 show the frequency sweeps for each gain mode.
The SNR and SNDR both have consistent performance over the entire frequency
CHAPTER 5. MEASUREMENT RESULTS 119

Figure 5.23: Receiver output spectrum with the peak SFDR in mode 0.

Figure 5.24: Receiver output spectrum with the peak SFDR in mode 1.
CHAPTER 5. MEASUREMENT RESULTS 120

Figure 5.25: Receiver output spectrum with the peak SFDR in mode 2.

Figure 5.26: Receiver output spectrum with the peak SFDR in mode 3.
CHAPTER 5. MEASUREMENT RESULTS 121

Figure 5.27: Receiver performance versus input frequency in mode 0.

range in all of the gain modes. There is a slight decrease in the SNR and SNDR at
300 Hz, but this can be attributed to the roll-off of the LNA gain at low frequencies.
The SFDR also maintains approximately equivalent performance over the frequency
sweeps, which confirms the peak SFDR results measured in the amplitude sweeps.
There is a small increase in the SFDR at the highest test frequencies (30 kHz and
50 kHz). This increase is the result of the harmonics falling outside the receiver
bandwidth. Interestingly, this increase is visible in all of the gain modes except mode
1. In this mode the input amplitude is the smallest of any test case (−12 dBFS) and
the SFDR is determined by the noise floor, not the largest harmonic, which explains
why there is no jump in the SFDR at the highest input frequencies.

5.3.3 Noise Coupling


The noise coupling between the ADC and LNA can be characterized by measuring the
output spectrum of the LNA with the ADC off and comparing it with a measurement
of the LNA output spectrum with the ADC on. By doing this, any noise that couples
CHAPTER 5. MEASUREMENT RESULTS 122

Figure 5.28: Receiver performance versus input frequency in mode 1.

Figure 5.29: Receiver performance versus input frequency in mode 2.


CHAPTER 5. MEASUREMENT RESULTS 123

Figure 5.30: Receiver performance versus input frequency in mode 3.

from the ADC can be differentiated from the normal LNA output spectrum. To
ensure a realistic comparison, when the ADC is on it is connected such that it is also
processing the output signal of the LNA, rather than sitting idle. A 10 kHz input
signal is used for the comparison and the measurements are repeated for each LNA
gain mode. A larger −6 dBFS input amplitude is used in this test to ensure that
the distortion terms are visible in the output spectrum. Figures 5.31 to 5.34 show
the noise coupling comparison for each gain mode. In each of the figures, the top
spectrum is with the ADC off and the bottom spectrum is with the ADC on.
Given the proximity of the LNA and the ADC on the die, noise coupling could
be a serious issue that could significantly degrade the performance of the receiver.
However, the LNA output spectrum comparisons show only minor differences when
the ADC is off versus when the ADC is on. In all of the gain modes, the noise floor is
unchanged when the ADC is turned on. Further, the amplitude of the 10 kHz input
tone is the same and the amount of power line noise (60 Hz and its harmonics) is the
same. The only noticeable difference is the amplitude of the distortion terms. The
worst case is mode 0, where the 3rd harmonic (30 kHz) increases slightly causing a
CHAPTER 5. MEASUREMENT RESULTS 124

Figure 5.31: LNA output spectrum with ADC off (top) and on (bottom) in mode 0.

Figure 5.32: LNA output spectrum with ADC off (top) and on (bottom) in mode 1.
CHAPTER 5. MEASUREMENT RESULTS 125

Figure 5.33: LNA output spectrum with ADC off (top) and on (bottom) in mode 2.

Figure 5.34: LNA output spectrum with ADC off (top) and on (bottom) in mode 3.
CHAPTER 5. MEASUREMENT RESULTS 126

reduction of the SFDR by approximately 3 dB. This reduction in SFDR is only visible
in the two highest gain modes. In the two lower gain modes the SFDR is unchanged.
It is likely that the higher gain modes have slightly reduced performance due to the
fact that any coupled noise is amplified by a larger amount. Overall these results
validate the precautions taken to reduce noise coupling between the noisy ADC and
the sensitive LNA, as only minor differences in the output spectrum are visible.

5.3.4 Anti-Alias Filtering


The anti-alias filtering performance of the receiver is determined by inputting a signal
at a frequency that will alias to within the signal band and then measuring the
strength of the aliased signal. The worst-case aliasing occurs when an interfering
signal is 50 kHz below the clock rate of 15 MHz, which is 14.95 MHz. The SR-1 signal
generator is not capable of generating signals above 200 kHz, so a BK Precision 4086
80 MHz Function Generator is used in its place. The rest of the receiver test setup
is unchanged. For this test a 0 dBFS single-ended sinusoidal signal at 14.95 MHz is
connected to the input of the receiver. Figures 5.35 to 5.38 show the output spectrum
of the receiver in each gain mode.
The 14.95 MHz input signal aliases to approximately 50 kHz at the output of the
receiver. The aliased signal is visible in each output spectrum, although it is difficult
to see in mode 0 due to the higher noise floor. In mode 1, for example, the aliased
signal has an amplitude of −82.69 dBFS. With an input amplitude of 0 dBFS, this
corresponds to an alias rejection of 82.69 dB. The alias rejection in mode 0, mode 1,
mode 2 and mode 3 are 82.34 dB, 82.69 dB, 84.47 dB and 86.61 dB respectively.
The alias rejection of the ADC was measured to be 63.63 dB. This means that
the alias rejection of the full receiver is approximately 20 dB higher. This additional
rejection can be attributed to the frequency response of the LNA, which rolls off at
high frequencies, and to the parasitics of the input transformer. It should also be
pointed out that because the 4086 function generator outputs a single-ended signal,
additional noise can couple into the input of the receiver and is likely the cause of
the small spurious tones and additional noise seen in some of the output spectra.
CHAPTER 5. MEASUREMENT RESULTS 127

Figure 5.35: Receiver output spectrum with a 14.95 MHz input signal in mode 0.

Figure 5.36: Receiver output spectrum with a 14.95 MHz input signal in mode 1.
CHAPTER 5. MEASUREMENT RESULTS 128

Figure 5.37: Receiver output spectrum with a 14.95 MHz input signal in mode 2.

Figure 5.38: Receiver output spectrum with a 14.95 MHz input signal in mode 3.
CHAPTER 5. MEASUREMENT RESULTS 129

Table 5.3: Receiver measurement results summary.

Measurement Mode 0 Mode 1 Mode 2 Mode 3


Peak SNR (dB) 58.31 66.71 74.29 77.98
Peak SNDR (dB) 58.19 66.02 71.14 73.33
Peak SFDR (dB) 78.98 84.50 90.08 91.91
Alias Rejection (dB) 82.34 82.69 84.47 86.61
Power Dissipation (mW) 1.5494 1.5512 1.5506 1.5516

5.3.5 Summary
The receiver was characterized in terms of the same dynamic performance metrics that
were used to characterize the ADC (SNR, SNDR, SDFR). The receiver performance
was different in each gain mode, with the peak SNR varying from 58.31 dB to 77.98
dB. The peak SNDR varied over a range of 58.19 dB to 73.33 dB. The peak SFDR
varied from 78.98 dB to 91.91 dB. The anti-alias filtering capability of the receiver
exceeded the 80 dB design goal, with values ranging from 82.34 dB to 86.61 dB. The
power dissipation of the full receiver was approximately 1.55 mW in all of the gain
modes, which is roughly equivalent to the sum of the LNA power dissipation and the
ADC power dissipation. The power dissipation was measured with a BK Precision
5492 5 1/2 Digit Multimeter.
Table 5.3 shows a summary of the receiver measurement results. The peak SNR,
peak SNDR and peak SFDR values were all derived from the amplitude sweeps, which
were performed with an input frequency of 10 kHz. The alias rejection measurement
used a 14.95 MHz test signal.

5.4 Field Test Results

5.4.1 Test Setup


The test setup used for field measurements is shown in Figure 5.39. The input signal
was generated by a loop antenna already installed at the test sites. The antenna was
connected directly to the input transformer. The passive voltage to current converter
CHAPTER 5. MEASUREMENT RESULTS 130

Loop DUT
Antenna
Transformer
Data
LNA ADC
Xilinx FPGA
Input Clock Spartan-3
Clock

BK 4086 Agilent E3630A Computer


Function Generator Power Supply MATLAB

Figure 5.39: Test setup for field measurements.

and the dummy antenna are not used for field measurements. While the VLF loop
antennas vary in size and shape, they all have a 1 Ω, 1 mH impedance, which is
connected to the input of the LNA with the 24:548 center-tapped transformer. The
output of the LNA is fed directly to the input of the ADC.
The clock signal for the ADC was generated with a BK Precision 4086 80 MHz
Function Generator. The 4086 has slightly worse jitter performance than the Agilent
81110A Pulse Generator used in the lab measurements. However, it was tested in the
lab and did not noticeably degrade the performance of the ADC. The clock source
was configured to generate a 1.2 V square wave at 15 MHz. The 4086 was selected
for the field test measurements due to its smaller size and weight when compared to
the 81110A.
The receiver was powered by an Agilent E3630A DC Power Supply, which was
configured to output 1.2 V. Separate power and ground connections were made to
the test board to prevent noise coupling between the various parts of the test chip.
There were five separate power connections: LNA analog supply, ADC analog supply,
ADC digital supply, ADC reference voltage and ESD protection supply. Each of the
connections had its own dedicated decoupling network on the test board.
The digital output data from the receiver was captured by a FPGA and streamed
CHAPTER 5. MEASUREMENT RESULTS 131

to a laptop computer in real-time to monitor the received signal. A TXB0104 level-


shifter (not shown) was placed between the test board and the FPGA to convert the
1.2 V digital output data from the receiver to 3.3 V digital data, which is expected
by the FPGA input pins. The level-shifter also provided an extra layer of isolation
between the receiver and the FPGA to prevent any coupling of noise between the two
systems.
A Xilinx Spartan-3 XC3S1500 FPGA was used in this application. The FPGA was
programmed to collect the two input signals from the receiver: the 1-bit data stream
and the 15 MHz clock. The clock signal from the chip was used to synchronously
capture the receiver output data on the FGPA. The data was then buffered in the
FPGA’s onboard RAM and streamed over USB in real-time to a nearby laptop.
Finally, the data was processed in MATLAB to produce a spectrogram of the received
signal.

5.4.2 Stanford Field Test


The first field test was performed at Stanford University on the roof of the Packard
Electrical Engineering building. This test used an existing loop antenna that was
installed by the Stanford VLF research group. This antenna consists of two perpen-
dicular loops. One loop faces in the north-south direction and the other faces in the
east-west direction. The north-south loop was used in this experiment. Each loop
is a square that is roughly 2 meters on each side with 11 turns of wire. Given that
this test site is in an urban environment with many nearby sources of interference, it
was expected that the output data would be very noisy. As a result, the lowest gain
mode of the receiver was used to prevent the output from saturating.
Figure 5.40 shows 60 seconds of data collected using the north-south antenna
and the lowest gain mode. The spectrogram shows how the spectral content of the
received signal changes over time. The horizontal axis represents time, the vertical
axis represents frequency and the color represents the strength of the signal.
The horizontal lines in the spectrogram are VLF transmitters that output at a
constant frequency. The two strongest transmitters in this spectrogram are at 24.8
CHAPTER 5. MEASUREMENT RESULTS 132

Figure 5.40: Spectrogram showing 60 seconds of data collected at Stanford University.

kHz and 25.2 kHz. These are the NLK transmitter in Jim Creek, Washington (24.8
kHz) and the NML transmitter in LaMoure, North Dakota (25.2 kHz) [18]. These
transmitters are used by the US Navy for communication with submerged submarines.
The vertical lines in the spectrogram that extend from roughly 5 kHz to 20 kHz
are radio atmospherics, or sferics for short. These are impulsive signals generated by
lightning strikes, which can propagate thousands of kilometers from their source in
the Earth-ionosphere waveguide.
There is also a large amount of interference visible in the spectrogram, particularly
at frequencies below 5 kHz. In fact, the strongest frequency component in the received
signal is the 60 Hz power line noise. This amount of interference is expected given
the urban environment where the antenna was installed.

5.4.3 Santa Cruz Field Test


The second field test was performed at a site operated by Lockheed Martin in the
Santa Cruz Mountains near Santa Cruz, California. The test site is in a remote area,
CHAPTER 5. MEASUREMENT RESULTS 133

far away from any cities, and the antenna is installed over 500 feet from the nearest
building. Under these conditions it is expected that there will be much less noise
and interference, which will result in a much cleaner received signal. The antenna
at this site is a right isosceles triangle with a base of approximately 10 meters and a
height of approximately 5 meters. The antenna has 5 turns of wire. This is one of the
largest and most sensitive antennas deployed by the Stanford VLF research group.
This antenna has two separate loops, one in the north-south direction and one in the
east-west direction. The east-west loop was used in this experiment.
Figure 5.41 shows 60 seconds of data collected at the Santa Cruz test site. The
LNA gain was set to the lowest gain mode in order to facilitate a direct comparison
with the Stanford field test data. There are several noticeable differences between
the data collected at Stanford and the data collected at the Santa Cruz site. As
expected, there is far less interference visible in the spectrogram and the power line
noise below 5 kHz is significantly reduced. There is also an increase in the number
of sferics visible in the data, which is due to the better noise environment and the
increased sensitivity of the larger antenna.
Many of the same VLF transmitters are visible in the Santa Cruz spectrogram,
including the NLK transmitter in Washington at 24.8 kHz and the NML transmitter
in North Dakota at 25.2 kHz. One new feature that was not seen in the Stanford data
is the set of horizontal dashed lines occurring between 11 kHz and 15 kHz. These
pulsating signals are generated by the Russian Alpha navigation system, which is
used to determine the position of ships and airplanes. It is likely that these Russian
signals are only seen in the data at the Santa Cruz test site due to the east-west
orientation of the antenna and the better sensitivity of the large antenna.

5.4.4 Receiver Comparison


The primary receiver currently used by the Stanford VLF research group is the AWE-
SOME receiver. The AWESOME receiver is deployed around the world and is the
gold-standard for VLF receiver systems with the highest data quality. An AWE-
SOME receiver was available at the Stanford field test site so a direct comparison
CHAPTER 5. MEASUREMENT RESULTS 134

Figure 5.41: Spectrogram showing 60 seconds of data collected in the Santa Cruz
Mountains.

can be made between the performance of the single-chip receiver and the AWESOME
receiver.
Figure 5.42 shows a side-by-side spectrogram of data received with the AWESOME
receiver and the single-chip receiver. The AWESOME receiver data is on the left and
the single-chip receiver data is on the right. Each spectrogram shows eight seconds
of data and was taken using the same north-south antenna at the Packard Electrical
Engineering building. It is important to point out that the gain of the two receivers
is not the same. This was corrected by shifting the AWESOME data to compensate
for this difference.
The two data samples were taken approximately 10 minutes apart, so the spec-
trograms are not expected to be identical. However, given the small time differential,
it is expected that the same transmitters will be visible and the same interfering
signals will be present in both samples. The data collected is consistent with these
expectations. Aside from differences in the sferics and the high frequency roll-off of
the AWESOME receiver, it is difficult to discern any significant difference between
CHAPTER 5. MEASUREMENT RESULTS 135

Figure 5.42: Side-by-side comparison of data collected at Stanford University with


the AWESOME receiver (left) and the single-chip receiver (right).

the two spectrograms.


The average spectrum was calculated to more accurately compare the two data
samples. This was accomplished by taking the average at each frequency over the
full eight seconds of data. Using this method a small random variation, such as a
sferic, won’t significantly affect the result due to the large number of data points
in the sample. The average spectrum for each receiver is shown in Figure 5.43. The
horizontal axis represents frequency and the vertical axis represents the average signal
strength. The AWESOME receiver data is the blue curve and the single-chip receiver
data is the red curve.
There is good agreement between the two data samples. All of the same trans-
mitters are present with similar amplitudes and the interference also matches well.
The most noticeable difference is at frequencies above approximately 47 kHz. Above
this frequency the AWESOME data is suppressed. This is due to the roll-off of the
anti-alias filter in the AWESOME receiver. Although it is difficult to tell from the
plot, it also appears that the single-chip receiver has a slightly lower noise floor than
CHAPTER 5. MEASUREMENT RESULTS 136

Figure 5.43: Comparison of the average spectrum.

the AWESOME receiver.


One metric that is often used to compare VLF receiver performance is to look
at the signal-to-noise ratio of the spectrum around a strong transmitter. The NLK
and NML transmitters can be used for this purpose. Figure 5.44 shows the average
spectrum of the data zoomed in on the frequency range from 22 kHz to 28 kHz.
Both transmitters are clearly visible when zooming in on the average spectrum.
NLK is at 24.8 kHz and NML is at 25.2 kHz. A third transmitter is visible at 24 kHz,
which is the NAA transmitter in Cutler, Maine. The strength of the transmitters is
not exactly equal in the two data samples, but since the data was taken approximately
10 minutes apart this is not unexpected as the propagation conditions change over
time. However, what is important is the lower noise floor in the spectrum of the
single-chip receiver. The lower noise floor results in a larger signal-to-noise ratio.
As a result, in this side-by-side comparison the single-chip receiver outperforms the
AWESOME receiver.
CHAPTER 5. MEASUREMENT RESULTS 137

Figure 5.44: Comparison of the average spectrum zoomed in around the NLK and
NML transmitters.
Chapter 6

Conclusion

This dissertation covered the design, implementation and validation of the first single-
chip broadband VLF magnetic field receiver. The single-chip receiver implements the
full signal path of a traditional VLF receiver, which includes interfacing with the 1
Ω, 1 mH loop antenna, processing the received signal and generating digital output
data. To achieve this functionality, the single-chip receiver includes a low-impedance
low-noise amplifier and a continuous-time delta-sigma analog-to-digital converter.
The low-noise amplifier was implemented with a two-stage architecture. The first
stage consists of a low-impedance common-base amplifier, which uses NPN transistors
to achieve good 1/f noise performance. The second stage uses a differential instru-
mentation amplifier that employs high gain OTAs and feedback to ensure minimal
harmonic distortion. The LNA has four gain modes that enable the performance
of the receiver to be optimized based on the requirements of the application. This
LNA is the first integrated VLF magnetic field low-noise amplifier to simultaneously
achieve better than 1 fT/Hz1/2 sensitivity and over 90 dB spurious-free dynamic
range, which is a result that was thought to be unlikely to be achieved just a few
years ago [32]. The LNA also demonstrates the use of an automatic biasing system
that increases the robustness of the receiver to variations in temperature. The LNA
consumes approximately 908 µW.
The analog-to-digital converter was implemented with a continuous-time delta-
sigma modulator. The delta-sigma ADC architecture combines oversampling and

138
CHAPTER 6. CONCLUSION 139

noise shaping to create a high resolution ADC from lower resolution components. The
delta-sigma architecture is relatively immune to mismatch and other non-ideal effects.
The primary reason for selecting this ADC architecture was that the continuous-time
loop filter provides a free anti-alias filter, which leads to a significant power savings.
The ADC uses a clock frequency of 15 MHz, which corresponds to an oversampling
ratio of 150. It achieves a resolution of 12.40 bits and a spurious-free dynamic range
of 93.60 dB. The power dissipation of the ADC is roughly 640 µW.
The single-chip receiver was fabricated in a 0.13 µm BiCMOS process. It has a
total power consumption of approximately 1.55 mW and uses a 1.2 V supply voltage.
This power dissipation corresponds to a reduction of over 30 times compared to the
signal path power consumption of the Penguin receiver, which is the current low-
power VLF receiver. In addition to measuring the receiver performance in the lab,
the receiver was also tested in the field at Stanford University and at a quiet site
in the Santa Cruz Mountains. The field test demonstrated that the data quality of
the single-chip receiver was just as good, if not better, than that of the AWESOME
receiver, which is the current high performance VLF receiver deployed by the Stanford
VLF research group.

6.1 Future Work


The successful implementation of the single-chip VLF magnetic field receiver opens
up several possibilities for further research. The research ideas can be divided into
two categories: improvements to the receiver and applications for the receiver.
In terms of improving the receiver, one attractive idea would be to implement
the entire receiver in CMOS to take advantage of the advanced short-channel CMOS
processes that are currently available. The reason this was not done in the current
design was because CMOS devices typically have inferior 1/f noise performance. As
a result, bipolar transistors were used in the input stage of the LNA to meet the noise
requirements. There are several strategies that could be employed to overcome the
1/f noise disadvantage of CMOS transistors. One method of particular interest is
chopping, which would not only solve the 1/f noise problem, but would also have the
CHAPTER 6. CONCLUSION 140

added benefit of canceling any offset in the LNA [21].


There are also several possibilities for further work at the system-level of the re-
ceiver. One possible improvement would be to implement an automatic gain control
system, which automatically changes the gain of the LNA depending on the charac-
teristics of the received signal. For example, if large interfering signals are detected,
the system would reduce the gain to prevent the receiver from clipping. Another idea
would be to automatically scale the power dissipation of the receiver depending on the
noise and dynamic range requirements. In many cases, the full dynamic range of the
receiver is not required and the power dissipation could be reduced. This optimization
could significantly increase the battery life of a remotely deployed receiver.
While the single-chip receiver implements the full signal path functionality of
a VLF magnetic field receiver, further work would be needed to build a complete
receiver system based on the single-chip receiver. In addition to the signal path,
a full receiver would need to add a time-reference, clock generator, digital signal
processing, data storage and power management. This could be accomplished using
separate components to implement these functions, or the receiver could be further
optimized by integrating some of this functionality on-chip.
There are many possible applications for the single-chip receiver. Due to its small
size, one idea of particular interest would be to build a hand-held VLF receiver. This
hand-held receiver could be used to survey locations to find quiet sites where large
permanent antennas could be constructed. It could also be used more widely as
a method to detect and monitor power-line interference. The implementation of a
hand-held receiver would require further innovation to reduce the size of the antenna
and input transformer, as both of these components are currently much too large for
a hand-held device.
The Stanford VLF research group has deployed a number of satellite-based instru-
ments. Future work on the single-chip receiver could include hardening the receiver to
the effects of radiation, which is encountered in space. This could be accomplished by
redesigning the circuits using radiation-hard techniques, or by appropriately shielding
the chip. The single-chip receiver would be particularly well suited for a satellite-based
application, because size is often a significant constraint.
Bibliography

[1] Andrew Alford and A. G. Kandoian. Ultrahigh-frequency loop antennas. AIEE


Transactions, 59(12):843–848, Feb 1940.

[2] Rex T. Baird and Terri S. Fiez. Improved ∆Σ DAC linearity using data weighted
averaging. IEEE International Symposium on Circuits and Systems, 1:13–16,
May 1995.

[3] Philippe Benabes, Mansour Keramat, and Richard Kielbasa. A methodology for
designing continuous-time sigma-delta modulators. Proceedings of the European
Design and Test Conference, pages 46–50, Mar 1997.

[4] Brian P. Brandt, Drew E. Wingard, and Bruce A. Wooley. Second-order sigma-
delta modulation for digital-audio signal acquisition. IEEE Journal of Solid-State
Circuits, 26(4):618–627, Apr 1991.

[5] Lucian J. Breems, Eric J. van der Zwan, and Johan H. Huijsing. Design for
optimum performance-to-power ratio of a continuous-time Σ∆ modulator. Pro-
ceedings of the European Solid-State Circuits Conference, pages 318–321, Sep
1999.

[6] Emmanuel J. Candès and Michael B. Wakin. An introduction to compressive


sampling. IEEE Signal Processing Magazine, 25(2):21–30, Mar 2008.

[7] James C. Candy. Decimation for sigma delta modulation. IEEE Transactions
on Communications, COM-34(1):72–76, Jan 1986.

141
BIBLIOGRAPHY 142

[8] L. Richard Carley. A noise-shaping coder topology for 15+ bit converters. IEEE
Journal of Solid-State Circuits, 24(2):267–273, Apr 1989.

[9] Bruce Carter. A differential op-amp circuit collection. Application report, Texas
Instruments, Dallas, TX, Jul 2001. SLOA064.

[10] Kirk C.-H. Chao, Shujaat Nadeem, Wai L. Lee, and Charles G. Sodini. A higher
order topology for interpolative modulators for oversampling A/D converters.
IEEE Transactions on Circuits and Systems, 37(3):309–318, Mar 1990.

[11] David K. Cheng. Fundamentals of Engineering Electromagnetics. Prentice-Hall,


1993.

[12] James A. Cherry and W. Martin Snelgrove. Clock jitter and quantizer metasta-
bility in continuous-time delta-sigma modulators. IEEE Transactions on Circuits
and Systems II, 46(6):661–676, Jun 1999.

[13] James A. Cherry and W. Martin Snelgrove. Continuous-Time Delta-Sigma Mod-


ulators for High-Speed A/D Conversion. Kluwer, 1999.

[14] James A. Cherry and W. Martin Snelgrove. Excess loop delay in continuous-
time delta-sigma modulators. IEEE Transactions on Circuits and Systems II,
46(4):376–389, Apr 1999.

[15] D. A. Chrissan and A. C. Fraser-Smith. Seasonal variations of globally measured


ELF/VLF radio noise. Radio Science, 31(5):1141–1152, Sep-Oct 1996.

[16] Hugh J. Christian, Richard J. Blakeslee, Dennis J. Boccippio, William L. Boeck,


Dennis E. Buechler, Kevin T. Driscoll, Steven J. Goodman, John M. Hall,
William J. Koshak, Douglas M. Mach, and Michael F. Stewart. Global frequency
and distribution of lightning as observed from space by the optical transient de-
tector. Journal of Geophysical Research, 108(D1), Jan 2003.

[17] Morris B. Cohen. ELF/VLF Phased Array Generation via Frequency-Matched


Steering of a Continuous HF Ionospheric Heating Beam. Ph.D. Dissertation,
Stanford University, Oct 2009.
BIBLIOGRAPHY 143

[18] Morris B. Cohen. The Stanford University ELF/VLF receiver: Narrowband


transmitter guide. Technical report, Stanford University, Stanford, CA, Jul 2009.

[19] Morris B. Cohen, Umran S. Inan, and Evans W. Pascal. Sensitive broadband
ELF/VLF radio reception with the AWESOME instrument. IEEE Transactions
on Geoscience and Remote Sensing, 48(1):3–17, Jan 2010.

[20] Kenneth Davies. Ionospheric Radio Propagation. National Bureau of Standards,


1965.

[21] Christian C. Enz and Gabor C. Temes. Circuit techniques for reducing the effects
of op-amp imperfections: Autozeroing, correlated double sampling, and chopper
stabilization. Proceedings of the IEEE, 84(11):1584–1614, Nov 1996.

[22] H. T. Friis. Noise figures of radio receivers. Proceedings of the IRE, 32(7):419–
422, Jul 1944.

[23] Ranjit Gharpurey and Robert G. Meyer. Modeling and analysis of substrate
coupling in integrated circuits. IEEE Journal of Solid-State Circuits, 31(3):344–
352, Mar 1996.

[24] K. L. Graf, M. Spasojevic, R. A. Marshall, N. G. Lehtinen, F. R. Foust, and U. S.


Inan. Extended lateral heating of the nighttime ionosphere by ground-based VLF
transmitters. Journal of Geophysical Research: Space Physics, 118:7783–7797,
Dec 2013.

[25] Paul R. Gray, Paul J. Hurst, Stephen H. Lewis, and Robert G. Meyer. Analysis
and Design of Analog Integrated Circuits. Wiley, 5th edition, 2009.

[26] Roubik Gregorian, Kenneth W. Martin, and Gabor C. Temes. Switched-capacitor


circuit design. Proceedings of the IEEE, 71(8):941–966, Aug 1983.

[27] Mikael Gustavsson, J. Jacob Wikner, and Nianxiong Nick Tan. CMOS Data
Converters for Communications. Kluwer, 2000.
BIBLIOGRAPHY 144

[28] Sarah K. Harriman. Custom Integrated Amplifier Chip for VLF Magnetic Re-
ceiver. Ph.D. Dissertation, Stanford University, Aug 2010.

[29] Alan Hastings. The Art of Analog Layout. Pearson, 2nd edition, 2006.

[30] Robert A. Helliwell. Whistlers and Related Ionospheric Phenomena. Stanford,


1965.

[31] Maxim Integrated. Dynamic testing of high-speed ADCs, part 2. Application


note, San Jose, CA, Jul 2002. APP 729.

[32] Max Klein. Autonomous Ultra-Low Power ELF/VLF Receiver Systems. Ph.D.
Dissertation, Stanford University, Nov 2009.

[33] Takayuki Konishi, Kenji Inazu, Jun Gyu Lee, Masanori Natsui, Shoichi Masui,
and Boris Murmann. Design optimization of high-speed and low-power opera-
tional transconductance amplifier using gm /ID lookup table methodology. IEICE
Transactions, E94-C(3):334–345, Mar 2011.

[34] Thomas H. Lee. The Design of CMOS Radio-Frequency Integrated Circuits.


Cambridge, 2nd edition, 2004.

[35] J. D. Li, M. Spasojevic, V. Harid, M. B. Cohen, M. Golkowski, and U. Inan. Anal-


ysis of magnetospheric ELF/VLF wave amplification from the siple transmitter
experiment. Journal of Geophysical Research: Space Physics, 119:1837–1850,
Mar 2014.

[36] Robert Andrew Marshall. Very Low Frequency Radio Signatures of Transient Lu-
minous Events Above Thunderstorms. Ph.D. Dissertation, Stanford University,
May 2009.

[37] Benjamin Joseph Mossawir. High-Fidelity, Broadband Amplification and Filter-


ing for Spaceborne Instrumentation. Ph.D. Dissertation, Stanford University,
December 2009.
BIBLIOGRAPHY 145

[38] M. Ortmanns and F. Gerfers. Continuous-Time Sigma-Delta A/D Conversion.


Springer, 2006.

[39] Evans W. Paschal. The design of broad-band VLF receivers with air-core loop
antennas. Technical report, Stanford University, Stanford, CA, May 1988.

[40] Donald O. Pederson and Kartikeya Mayaram. Analog Integrated Circuits for
Communication. Springer, 2nd edition, 2008.

[41] Marcel J. M. Pelgrom, AAD C. J. Duinmaijer, and Anton P. G. Welbers.


Matching properties of MOS transistors. IEEE Journal of Solid-State Circuits,
24(5):1433–1440, Oct 1989.

[42] Behzad Razavi. Design of Analog CMOS Integrated Circuits. McGraw-Hill, 2001.

[43] R. K. Said, U. S. Inan, and K. L. Cummins. Long-range lightning geolocation


using a VLF radio atmospheric waveform bank. Journal of Geophysical Research,
115:D23108, Dec 2010.

[44] Richard Schreier. Delta Sigma Toolbox for MATLAB. http://www.mathworks.


com/matlabcentral/fileexchange/, Jan 2000.

[45] Richard Schreier and Gabor C. Temes. Understanding Delta-Sigma Data Con-
verters. Wiley, 2005.

[46] David K. Su, Marc J. Loinaz, Shoichi Masui, and Bruce A. Wooley. Experimental
results and modeling techniques for substrate noise in mixed-signal integrated
circuits. IEEE Journal of Solid-State Circuits, 28(4):420–430, Apr 1993.

[47] Kevin E. Trenberth and Jerry G. Olson. Temperature trends at the South Pole
and McMurdo Sound. Journal of Climate, 2:1196–1206, Oct 1989.

[48] Martin A. Uman. The Lightning Discharge. Academic Press, 1987.

[49] Charles Chang-I Wang. High Fidelity Analog-to-Digital Conversion for Space-
borne Applications. Ph.D. Dissertation, Stanford University, Sep 2009.
BIBLIOGRAPHY 146

[50] Bo Xia, Shouli Yan, and Edgar Sánchez-Sinencio. An RC time constant auto-
tuning structure for high linearity continuous-time Σ∆ modulators and active
filters. IEEE Transactions on Circuits and Systems, 51(11):2179–2188, Nov 2004.

[51] Shouli Yan and Edgar Sánchez-Sinencio. Low voltage analog circuit design tech-
niques: A tutorial. IEICE Transactions on Analog Integrated Circuits and Sys-
tems, E00-A(2):1–17, Feb 2000.

[52] J. Yuan and C. Svensson. New single-clock CMOS latches and flipflops with im-
proved speed and power savings. IEEE Journal of Solid-State Circuits, 32(1):62–
69, Jan 1997.

[53] F. G. Zoghzoghy, M. B. Cohen, R. K. Said, and U. S. Inan. Statistical patterns in


the location of natural lightning. Journal of Geophysical Research: Atmospheres,
118:787–796, Jan 2013.

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy