micromachined inductors and transformers...

174
MICROMACHINED INDUCTORS AND TRANSFORMERS FOR MINIATURIZED POWER CONVERTERS By CHRISTOPHER D. MEYER A DISSERTATION PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY UNIVERSITY OF FLORIDA 2012

Upload: others

Post on 20-Oct-2020

10 views

Category:

Documents


0 download

TRANSCRIPT

  • MICROMACHINED INDUCTORS AND TRANSFORMERS FOR MINIATURIZEDPOWER CONVERTERS

    By

    CHRISTOPHER D. MEYER

    A DISSERTATION PRESENTED TO THE GRADUATE SCHOOLOF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT

    OF THE REQUIREMENTS FOR THE DEGREE OFDOCTOR OF PHILOSOPHY

    UNIVERSITY OF FLORIDA

    2012

  • c© 2012 Christopher D. Meyer

    2

  • I dedicate this to my loving family.

    3

  • ACKNOWLEDGMENTS

    I would like to thank everyone who has contributed to the success of the work

    presented in my dissertation. I thank my adviser, Dr. David Arnold, who provided me

    with the opportunity to work on exciting topics in power magnetics and who introduced

    me to microfabrication at the University of Florida cleanroom. I thank Dr. Rizwan

    Bashirullah who served on my committee and who is developing the very high frequency

    power converter circuits that motivated my work. I thank Drs. Yong-Kyu Yoon and Peng

    Jiang for their valuable insights while also serving on my committee. I thank Xue Lin

    for testing my microinductor within his hybrid boost converter. I thank Christopher

    Dougherty for enlightening me on the considerations that affect high frequency converter

    designs. I thank Jessica Meloy for her help in wirebonding.

    I thank the U.S. Army Research Laboratory (ARL) for funding the project and

    my colleagues at ARL for their support. I thank Dr. Brian Morgan not only for leading

    the Power for Microsystems project from which my research derived, but also for the

    clarity he brought and for his mentoring me. I thank Dr. Sarah Bedair for countless

    discussions and for her sage advice contributing to my growth both technically and

    professionally. I thank Manrico Mirabelli for his microfabrication assistance and for

    sharing his photolithography expertise. I thank James Mulcahy of the cleanroom staff for

    maintaining and fixing the tools that were vital to this work. I thank William Benard for

    heading the cleanroom and keeping it running smoothly.

    I thank my grandfather, whose pride in me inspired me to complete my doctoral

    degree. I thank my wife, Jennifer, for her steadfast love. Finally, I would like to thank my

    parents for their continuous support and loving devotion.

    4

  • TABLE OF CONTENTS

    page

    ACKNOWLEDGMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

    LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

    LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

    ABSTRACT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

    CHAPTER

    1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

    1.1 The Case for Small . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171.1.1 Distributed On-Chip Power for Microprocessors . . . . . . . . . . . 171.1.2 Mobile Autonomous Microsystems . . . . . . . . . . . . . . . . . . 18

    1.2 Switched-Mode Power Converters . . . . . . . . . . . . . . . . . . . . . . 181.3 Impedance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211.4 High Frequency Benefits and Challenges . . . . . . . . . . . . . . . . . . 221.5 Survey of Existing Microfabricated Inductors and Transformers . . . . . . 231.6 Air-Core Passive Components for Microscale Power Converters . . . . . . 24

    2 BACKGROUND . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

    2.1 High Frequency Power Converters . . . . . . . . . . . . . . . . . . . . . . 272.2 Inductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 282.3 Transformers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

    3 INDUCTOR DESIGN . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

    3.1 Quality Factor Definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . 353.1.1 Quality Factor of Non-Ideal Reactive Components . . . . . . . . . 353.1.2 Quality Factor of Inductor . . . . . . . . . . . . . . . . . . . . . . . 37

    3.2 Performance Trilemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 383.3 Stacked Planar Spiral Layout . . . . . . . . . . . . . . . . . . . . . . . . . 393.4 Low Frequency Analytical Inductor Model . . . . . . . . . . . . . . . . . . 403.5 Trends and Optimization . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

    3.5.1 Analytical . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 433.5.2 FastHenry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

    3.6 Radio Frequency Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . 473.6.1 Capacitive Coupling . . . . . . . . . . . . . . . . . . . . . . . . . . 473.6.2 Eddy Currents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

    3.7 Summary of Inductor Design . . . . . . . . . . . . . . . . . . . . . . . . . 57

    5

  • 4 TRANSFORMER DESIGN . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

    4.1 Overview and Goals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 594.2 Maximum Efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

    4.2.1 From Scattering Parameters . . . . . . . . . . . . . . . . . . . . . . 604.2.2 From Coil Quality Factors and Coupling Coefficient . . . . . . . . . 61

    4.3 Layout . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 654.3.1 Turns Ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

    4.4 Performance Under Load . . . . . . . . . . . . . . . . . . . . . . . . . . . 664.4.1 Derivation of Efficiency and Voltage Gain for Arbitrary Load . . . . 664.4.2 Conjugate Impedance Matched Loading . . . . . . . . . . . . . . . 69

    4.5 Summary of Transformer Design . . . . . . . . . . . . . . . . . . . . . . . 70

    5 FABRICATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

    5.1 Process Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 735.1.1 Sequential Layer Removal . . . . . . . . . . . . . . . . . . . . . . . 735.1.2 Ultrasonic Agitation in Solvents . . . . . . . . . . . . . . . . . . . . 74

    5.2 Features and Variations on the Process . . . . . . . . . . . . . . . . . . . 765.2.1 Planar Processing . . . . . . . . . . . . . . . . . . . . . . . . . . . 765.2.2 Photoresist as a Structural Element . . . . . . . . . . . . . . . . . . 785.2.3 Substrate Versatility . . . . . . . . . . . . . . . . . . . . . . . . . . 79

    5.3 Process Steps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 795.4 Special Processing Considerations . . . . . . . . . . . . . . . . . . . . . . 82

    5.4.1 Sputtering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 825.4.2 Photolithography . . . . . . . . . . . . . . . . . . . . . . . . . . . . 855.4.3 Electroplating . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 875.4.4 Argon Sputter Etch . . . . . . . . . . . . . . . . . . . . . . . . . . . 895.4.5 Photoresist Skin Removal . . . . . . . . . . . . . . . . . . . . . . . 905.4.6 Copper Seed Etch . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

    6 INDUCTOR CHARACTERIZATION . . . . . . . . . . . . . . . . . . . . . . . . . 93

    6.1 Equipment and Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 936.2 Inductor Characterization Methods . . . . . . . . . . . . . . . . . . . . . . 94

    6.2.1 One-Port Inductor Methods . . . . . . . . . . . . . . . . . . . . . . 946.2.2 Two-Port Inductor Methods . . . . . . . . . . . . . . . . . . . . . . 956.2.3 Inductor Characteristics Obtained from Impedance . . . . . . . . . 97

    6.3 One-Port Inductor Characterization . . . . . . . . . . . . . . . . . . . . . . 986.3.1 One-Port Inductors on Pyrex Substrates . . . . . . . . . . . . . . . 98

    6.3.1.1 Comparison to model predictions . . . . . . . . . . . . . . 1016.3.1.2 Current rating . . . . . . . . . . . . . . . . . . . . . . . . 1016.3.1.3 Interwinding capacitance . . . . . . . . . . . . . . . . . . 102

    6.3.2 One-Port Inductors on Silicon Substrates . . . . . . . . . . . . . . 1056.3.2.1 Copper layer thickness: 10 µm vs. 30 µm . . . . . . . . . 1056.3.2.2 Inductor shape: square vs. circular spirals . . . . . . . . 106

    6

  • 6.4 Two-Port Inductor Characterization on Silicon Substrates . . . . . . . . . 1106.4.1 Capacitive Coupling through the Substrate . . . . . . . . . . . . . . 1106.4.2 Winding Losses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113

    6.5 Summary of Inductor Characterization . . . . . . . . . . . . . . . . . . . . 118

    7 TRANSFORMER CHARACTERIZATION . . . . . . . . . . . . . . . . . . . . . 119

    7.1 Equipment and Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1197.2 Impedance Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1207.3 Load-Dependent Efficiency and Voltage Gain . . . . . . . . . . . . . . . . 1227.4 Characterization of Transformers with 10 µm Thick Layers . . . . . . . . . 124

    7.4.1 Extraction of Nominal Inductances and Resistances . . . . . . . . 1257.4.2 Load-Dependent Performance of 1 : 1 Transformer . . . . . . . . . 1277.4.3 Load-Dependent Performance of 1 : 3.5 Transformer . . . . . . . . 131

    7.5 Characterization of Transformer with 30 µm Thick Layers . . . . . . . . . . 1357.6 Summary of Transformer Characterization . . . . . . . . . . . . . . . . . . 140

    8 PACKAGING AND TESTING WITH CIRCUITS . . . . . . . . . . . . . . . . . . 142

    8.1 Microinductor Wire Bonded to Very High Frequency Boost Converter . . . 1428.1.1 About the Microinductor . . . . . . . . . . . . . . . . . . . . . . . . 1428.1.2 About the Converter and Test Results . . . . . . . . . . . . . . . . 143

    8.2 Testing with Commercial Surface-Mount Converter . . . . . . . . . . . . . 1458.2.1 About the Texas Instruments TPS61240 Converter . . . . . . . . . 1468.2.2 Module Design and Processing . . . . . . . . . . . . . . . . . . . . 1468.2.3 Converter Module Testing . . . . . . . . . . . . . . . . . . . . . . . 149

    8.3 Summary of Inductor Packaging and Testing within Converter Circuits . . 154

    9 CONCLUSION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157

    9.1 Summary of Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1579.2 Lessons Learned . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1589.3 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160

    REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166

    BIOGRAPHICAL SKETCH . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174

    7

  • LIST OF TABLES

    Table page

    2-1 Literature survey of microinductors . . . . . . . . . . . . . . . . . . . . . . . . . 33

    2-2 Literature survey of microtransformers . . . . . . . . . . . . . . . . . . . . . . . 34

    3-1 Coefficients for modified Wheeler and current sheet expressions . . . . . . . . 41

    5-1 Process parameters for passives fabrication . . . . . . . . . . . . . . . . . . . . 81

    5-2 Recipe for acid copper sulfate electroplating bath . . . . . . . . . . . . . . . . . 88

    6-1 Comparison of measured inductor performance . . . . . . . . . . . . . . . . . . 99

    6-2 Comparison of model-predicted to measured inductor performance . . . . . . . 100

    6-3 Performance comparison of inductors with different layer thicknesses . . . . . . 106

    6-4 Geometric parameters of square and circular inductors . . . . . . . . . . . . . 107

    6-5 Performance comparison of square and circular inductors . . . . . . . . . . . . 107

    7-1 Comparison of transformer circuit parameters . . . . . . . . . . . . . . . . . . . 126

    8-1 Component sizes in functional converter module . . . . . . . . . . . . . . . . . 149

    8

  • LIST OF FIGURES

    Figure page

    1-1 Common converter circuits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

    1-2 Review of microinductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

    1-3 Review of microtransformers . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

    3-1 Circuit diagram of simple inductor model . . . . . . . . . . . . . . . . . . . . . . 39

    3-2 Diagram of spiral dimensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

    3-3 Trends of inductance to resistance ratio vs. packing density . . . . . . . . . . . 45

    3-4 Trends of inductance to resistance ratio vs. outer diameter . . . . . . . . . . . 46

    3-5 Trend of inductance vs. vertical gap between stack simulated in FastHenry . . 46

    3-6 Diagram of capacitive coupling of traces through substrate . . . . . . . . . . . 49

    3-7 Circuit model of inductor with substrate capacitance . . . . . . . . . . . . . . . 49

    3-8 Substrate resistance effect on inductor impedance . . . . . . . . . . . . . . . . 50

    3-9 Circuit model of inductor with winding and substrate capacitances . . . . . . . 51

    3-10 Substrate vs. winding capacitance effect on inductor impedance . . . . . . . . 52

    3-11 COMSOL simulations of skin effect in winding cross sections . . . . . . . . . . 55

    3-12 Measured effect of eddy currents on inductor impedance . . . . . . . . . . . . 57

    4-1 Transformer energy flow diagram . . . . . . . . . . . . . . . . . . . . . . . . . . 61

    4-2 Transformer efficiency calculated by quality factors and scattering parameters . 64

    4-3 Transformer layout winding diagram . . . . . . . . . . . . . . . . . . . . . . . . 65

    4-4 Circuit diagram of two-port network cascaded with shunt load . . . . . . . . . . 67

    4-5 Circuit diagram of two-port network cascaded with series load . . . . . . . . . . 68

    4-6 Circuit diagram of two-port network with source and load impedances . . . . . 69

    5-1 Illustrations of additive process stage . . . . . . . . . . . . . . . . . . . . . . . . 74

    5-2 Illustrations of subtractive process stage . . . . . . . . . . . . . . . . . . . . . . 75

    5-3 Scanning electron micrograph (SEM) of inductor with 10 µm thick layers . . . . 75

    5-4 SEM of inductor with 30 µm thick layers . . . . . . . . . . . . . . . . . . . . . . 76

    9

  • 5-5 Cross section diagrams of process additive stage . . . . . . . . . . . . . . . . . 83

    5-6 Cross section diagrams of process subtractive stage . . . . . . . . . . . . . . . 84

    5-7 Adhesion of copper to photoresist . . . . . . . . . . . . . . . . . . . . . . . . . 85

    5-8 Electroplating leakage between features . . . . . . . . . . . . . . . . . . . . . . 86

    5-9 Electroplated copper cantilever . . . . . . . . . . . . . . . . . . . . . . . . . . . 89

    5-10 Comparison images showing argon sputter etch effect on adhesion . . . . . . . 90

    5-11 Photoresist blocking layer formed by argon sputter etch . . . . . . . . . . . . . 91

    5-12 Sidewall roughening caused by copper etch . . . . . . . . . . . . . . . . . . . . 92

    6-1 SEM images of one-port and two-port inductors . . . . . . . . . . . . . . . . . . 94

    6-2 Two-port inductor impedance network . . . . . . . . . . . . . . . . . . . . . . . 95

    6-3 Identification of inductor specifications from plots . . . . . . . . . . . . . . . . . 99

    6-4 Current rating of inductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100

    6-5 Comparison images of interlayer photoresist . . . . . . . . . . . . . . . . . . . 103

    6-6 Comparison of interlayer dielectric effect on impedance of small inductor . . . . 104

    6-7 Comparison of interlayer dielectric effect on impedance of large inductor . . . . 104

    6-8 Comparison of layer thicknesses for small inductor . . . . . . . . . . . . . . . . 108

    6-9 Comparison of layer thicknesses for large inductor . . . . . . . . . . . . . . . . 108

    6-10 Comparison of shape of small inductor . . . . . . . . . . . . . . . . . . . . . . . 109

    6-11 Comparison of shape of large inductor . . . . . . . . . . . . . . . . . . . . . . . 109

    6-12 Pad capacitance diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

    6-13 Shunt capacitance at two ports of inductor . . . . . . . . . . . . . . . . . . . . . 115

    6-14 Impedance plots from two-port inductor . . . . . . . . . . . . . . . . . . . . . . 115

    6-15 SEM images of inductors with solid vs. filamented traces . . . . . . . . . . . . 116

    6-16 SEM images of solid vs. filamented traces . . . . . . . . . . . . . . . . . . . . . 116

    6-17 Impedance plots of filamented vs. solid traces . . . . . . . . . . . . . . . . . . 117

    6-18 Change in resistance due to filamented vs. solid traces . . . . . . . . . . . . . 117

    7-1 Circuit representation of two-port impedance parameters . . . . . . . . . . . . 121

    10

  • 7-2 Low frequency transformer model . . . . . . . . . . . . . . . . . . . . . . . . . 122

    7-3 SEM images of microfabricated transformers . . . . . . . . . . . . . . . . . . . 125

    7-4 Impedance plots of 1 : 1 transformer with 10 µm thick layers . . . . . . . . . . . 126

    7-5 Impedance plots of 1 : 3.5 transformer with 10 µm thick layers . . . . . . . . . . 127

    7-6 Efficiency of 1 : 1 transformer . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128

    7-7 Voltage gain of 1 : 1 transformer . . . . . . . . . . . . . . . . . . . . . . . . . . 128

    7-8 Magnitude and phase of matched load impedance for 1 : 1 transformer . . . . . 129

    7-9 Efficiency and voltage gain vs. load impedance for 1 : 1 transformer . . . . . . 130

    7-10 Efficiency of 1 : 3.5 transformer . . . . . . . . . . . . . . . . . . . . . . . . . . . 132

    7-11 Voltage gain of 1 : 3.5 transformer . . . . . . . . . . . . . . . . . . . . . . . . . 132

    7-12 Magnitude and phase of matched load impedance for 1 : 3.5 transformer . . . 133

    7-13 Efficiency and voltage gain vs. load impedance for 1 : 3.5 transformer . . . . . 134

    7-14 SEM image of microtransformer with 30 µm thick layers . . . . . . . . . . . . . 136

    7-15 Impedance plots of 1 : 1 transformer with 30 µm thick layers . . . . . . . . . . . 137

    7-16 Efficiency of thicker 1 : 1 transformer . . . . . . . . . . . . . . . . . . . . . . . . 138

    7-17 Voltage gain of thicker 1 : 1 transformer . . . . . . . . . . . . . . . . . . . . . . 138

    7-18 Magnitude and phase of matched load impedance for thicker 1 : 1 transformer 139

    7-19 Efficiency and voltage gain vs. load impedance for thicker 1 : 1 transformer . . 141

    8-1 Microinductor wire bonded to circuit for testing . . . . . . . . . . . . . . . . . . 143

    8-2 Impedance of wire bonded inductor . . . . . . . . . . . . . . . . . . . . . . . . 144

    8-3 Measured efficiencies of converter with wire bonded microinductor . . . . . . . 145

    8-4 Copper layout of converter module . . . . . . . . . . . . . . . . . . . . . . . . . 147

    8-5 Photograph of released copper framework . . . . . . . . . . . . . . . . . . . . . 148

    8-6 Photographs of functional converter module . . . . . . . . . . . . . . . . . . . . 149

    8-7 Impedance of inductor used in converter module . . . . . . . . . . . . . . . . . 151

    8-8 Measured efficiency vs. output current of converter module . . . . . . . . . . . 152

    8-9 Boost converter circuit diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . 152

    11

  • 8-10 Inductor voltage waveform for several input voltages . . . . . . . . . . . . . . . 153

    8-11 Waveforms of inductor voltage for different load currents . . . . . . . . . . . . . 155

    8-12 Waveforms of output voltage for different load currents . . . . . . . . . . . . . . 155

    9-1 Review of microinductors including new results . . . . . . . . . . . . . . . . . . 159

    9-2 Review of microtransformers including new results . . . . . . . . . . . . . . . . 159

    9-3 Illustrations of package assembly . . . . . . . . . . . . . . . . . . . . . . . . . . 161

    9-4 SEM images of copper sockets . . . . . . . . . . . . . . . . . . . . . . . . . . . 163

    9-5 SEM images of teeth contacting to surface-mount component . . . . . . . . . . 163

    9-6 Surface-mount resistor and capacitor alongside microinductors . . . . . . . . . 164

    9-7 Measured impedances of socketed resistor and capacitor . . . . . . . . . . . . 165

    12

  • Abstract of Dissertation Presented to the Graduate Schoolof the University of Florida in Partial Fulfillment of theRequirements for the Degree of Doctor of Philosophy

    MICROMACHINED INDUCTORS AND TRANSFORMERS FOR MINIATURIZEDPOWER CONVERTERS

    By

    Christopher D. Meyer

    May 2012

    Chair: David P. ArnoldMajor: Electrical and Computer Engineering

    Switched-mode dc-dc power converters are a ubiquitous part of modern, feature-rich

    portable electronic devices and are essential for efficiently transferring electrical energy

    out of battery sources and into various power-hungry loads, such as microprocessors,

    displays, sensors, and communications systems. These converters often comprise

    a significant portion of total system size/weight, and the largest offenders are often

    the associated power inductors and transformers. A significant reduction in the size

    the inductors and transformers would have a transformative effect in enabling new

    applications, such as mobile autonomous microsystems.

    Increasing the switching frequency of the power converters offers to reduce the

    values of the required passives. However, the expected switching frequencies of

    next generation power converters fall into a gap between magnetic film inductors

    and transformers operable at < 10 MHz and microwave air-core devices with high

    performance at > 1 GHz. In answer, a new class of air-core microinductors and

    microtransformers is presented in this document that leveraged microfabrication-enabled

    advancements to attain high performance in the desirable very high frequency (VHF)

    switching range and to enable fully integrated power management systems in the

    smallest possible packages.

    In order to design these devices, models were analyzed to uncover the ideal

    characteristics for operating in the VHF range. Compared to traditional air-core

    13

  • components, these new ones featured thicker windings and had more intricate windings

    for lower loss and higher density. A multilevel microfabrication process was developed

    for molding three-dimensional (3D) copper parts with the necessary characteristics of

    thickness, minimum feature size, and out-of-plane stacking.

    The 3D copper process enabled the microfabrication of inductors with measured

    inductance densities up to 170 nH/mm2 and quality factors as great as 33. Transformers

    were measured with even greater inductance densities: up to 325 nH/mm2 was

    obtained in a configuration for voltage gain of 3.5 with up to 78% efficiency. Performance

    figures for both inductors and transformers were shown to outstrip a number of other

    microfabricated examples found in the literature, particularly in the frequency range of

    10 MHz–1 GHz.

    Microfabricated inductors were tested within the circuits of both a prototype

    100 MHz switched-mode hybrid boost converter and a commercially-available surface-mount

    converter with up to 4 MHz switching frequency. With up to 37% efficiency at a

    conversion ratio of 6, the performance of the prototype 100 MHz converter when using

    a 14 nH microfabricated inductor largely matched that obtained when a larger 43 nH

    surface-mount inductor was used in the same converter at up to 1 mA load current.

    A packaging solution was devised for testing with the surface-mount converter. An

    embedded multilevel copper module consisting of both an inductor and interconnects

    was detached from its silicon fabrication substrate and served as a platform to which a

    surface-mount converter and capacitors were soldered.

    14

  • CHAPTER 1INTRODUCTION

    Switched-mode dc-dc power converters are a ubiquitous part of modern, feature-rich

    portable electronic devices. These power converters are essential for efficiently

    transferring electrical energy out of battery sources and into various power-hungry

    loads, such as microprocessors, displays, sensors, and communications systems.

    The need for power converters arises from the fact that electricity is utilized in

    many different forms even within a single system. Often each subsystem has a different

    expectation for electrical current (the “rate”), voltage (the “force”), and duty cycle (the

    on/off times). Loads may operate erratically or not at all if the source is incapable of

    delivering enough electrical current, at a given voltage level, and for a certain amount of

    time. Batteries are designed to provide current at a fixed voltage, which may not match

    the needs of the loads. The battery voltage also often decreases with higher current

    draws or with time as its energy storage is depleted. Power converters provide the

    handshaking that is necessary for the sources and loads to interoperate with each other.

    One basic role of the dc-dc switched-mode power converter is to accept electrical

    power that is input to it at one voltage level and output that power at a different voltage

    level. Intelligent control mechanisms within the converter can respond to fluctuations in

    source and load conditions to help smooth the delivery of power and prevent levels from

    falling out of specification.

    Although the term “dc” implies that the input and output voltages of the dc-dc

    converter are ideally constant to the outside world, inside the converter are switches

    that dynamically reconfigure the electrical current paths of the converter circuit many

    thousands to millions of times per second. Power conversion utilizes the reactions of

    inductors and transformers to the switch-induced changes in electrical current within

    the converter to raise or lower the voltage to desired levels. Such inductive components

    possess the characteristic of inducing a voltage difference to resist changes in the level

    15

  • of current passing through them. The fundamental equation describing this behavior is

    vL(t) = Ldi

    dt, (1–1)

    where vL(t) is the voltage induced across the inductor, di/dt is the change in current

    through it, and the ratio L is defined as the inductance, measured in henries (H). For an

    electrical current initially at 0 and rising to a level I , the energy E stored in the inductor is

    E =1

    2LI 2. (1–2)

    Inductance is generally proportional to the area enclosed by a coiled conductor.

    Because physical volume and mass are placed at a premium in portable electronics

    applications, small inductors are desired but have correspondingly small inductances.

    The inductors and transformers must, for the sake of efficiency however, store enough

    energy per switching cycle to overshadow the power lost during conversion. As a result,

    the inductive components can comprise a major portion of the entire converter system

    size and mass, especially when the rest of the converter circuit is integrated onto a

    single, tiny semiconductor chip with nm-scale transistors.

    The large inductive components are, due to their size, generally added as discrete

    components connected outside the converter package. External connections further

    add to the bulk of the system as each component requires its own packaging, pads,

    and solder joints. A significant size and weight savings would be obtained if the

    inductors and transformers could be integrated with the rest of the converter circuit

    within the same package. Bulky external connections could be replaced by much leaner

    wire bonds, embedded interconnects, or flip-chip bumps via a System-in-Package

    (SiP) approach. Further size savings would be obtained through the System-on-Chip

    (SoC) approach of monolithically fabricating inductors and transformers directly on the

    integrated circuits chip.

    16

  • To realize these SiP or SoC concepts for power converters, inductive components

    must first decrease in size to the point where integration is not cost-prohibitive. One

    route to decreasing the inductor/transformer size is to increase the switching frequency

    of the converter, so that the current variations, di/dt, are greater and occur more often.

    The other means of reducing size is to increase the inductance density of the inductors

    and transformers.

    The inductors and transformers presented in this document leverage microfabrication

    technologies for increased density and are designed to operate at increased frequencies

    that have been emerged from innovative integrated-circuit converters. The goal of this

    work is to enable fully-integrated high-frequency switched-mode dc-dc power converters

    with ultra-miniaturized, high-density inductors and transformers.

    1.1 The Case for Small

    From the advent of the integrated circuit in the 1950s up to the present day,

    electronic systems have been continually packed into rapidly shrinking devices with

    ever-greater processing power. Contemporary consumer electronics are marked

    by examples of portable computers, mobile phones, and media players in svelte

    forms with increasingly convergent functionality. The need for fully integrated power

    converters is reaching a critical point, however, as the scaling of power components

    has struggled to keep pace with that of data processing and storage. But beyond just

    the consumer-driven aesthetic of small for the sake of small, a significant reduction in

    the size of power subsystems could also have a transformative effect in enabling new

    applications, like mobile autonomous microsystems, and in improving the distribution of

    power, such as for microprocessors.

    1.1.1 Distributed On-Chip Power for Microprocessors

    Modern microprocessors are highly parallel in operation. Facing the upper limits

    of using higher clock frequencies to process data quicker, designers have leveraged

    the benefits of a continually-shrinking transistor size and are integrating multiple

    17

  • microprocessor copies, or cores, on a single chip. The cores are able to process

    data in parallel, but have a tendency to be under-utilized in situations where tasks

    require serial processing. Although todays microprocessors receive power that is

    supplied by a converter located outside the microprocessor package, the ability to

    integrate many power converters on the chip itself could enable individual portions of

    the microprocessor circuit to be rapidly turned on and off as needed, reducing power

    consumption and increasing the thermal budget of the active portions. On-die power

    converters would additionally reduce the complexity of utilizing independent voltage

    levels for portions of the microprocessor operating at different frequencies for further

    reductions in power consumption [1, 2].

    1.1.2 Mobile Autonomous Microsystems

    An emerging research effort is focused at developing mobile autonomous

    microsystems, tiny robotic devices that can navigate through their environment by

    flying, crawling, or hopping. The number and complexity of subsystems envisioned for

    these mobile microsystems is staggering. In addition to the actuators for locomotion,

    these manmade bugs are expected to contain sensors for situational awareness, logic

    blocks for data processing, communications systems for relaying information, and

    possibly generators for harvesting energy from the environment. Interoperation of these

    subsystems is likely to be a challenge as each is likely to require operation at unique

    voltages, currents, and duty cycles for best performance, and will require an advanced

    power management system that must furthermore be vanishingly small and light so as

    to not interfere with the mobility of the bug [3].

    1.2 Switched-Mode Power Converters

    A multitude of circuit topologies exist to achieve switched-mode power conversion.

    Some provide a step-up from lower input voltage to a higher output voltage, and some

    provide a step-down. Some are capable of providing either step-up or step-down

    on-the-fly, while others have input and output voltages that are equal to one another but

    18

  • provide isolation to protect the output from high voltage spikes that may occur on the

    input side of the circuit. Common switched-mode converter topologies include the boost,

    buck, buck-boost, and flyback circuits.

    The basic boost converter is drawn in Figure 1-1A. Current flows through the

    inductor L when switch Q is closed, and energy is stored in the magnetic field of the

    inductor. When Q is opened, a voltage is induced across the inductor to oppose any

    sudden reduction in current, pushing current through diode D, onto the output capacitor

    C , and out to the load R. The voltage induced across the inductor in this last step is

    negative with respect to the reference for vL(t) labelled on Figure 1-1A, meaning that the

    voltage across the load is greater than the input voltage Vin. The role of the capacitor C

    is to store charge between switching cycles and ensure that the output voltage remains

    at a relatively steady value. The conversion ratio M for the boost converter is controlled

    by the duty cycle of the switch in its closed position, as plotted in Figure 1-1B. When the

    switch is closed for a longer portion of the cycle than it is closed, the conversion ratio of

    the converter is larger.

    The buck (Figure 1-1C) and buck-boost (Figure 1-1E) circuits operate similarly in

    that transient current through charged inductors induce voltages across the inductors

    that are utilized to create voltage differences with respect to the input. The duty cycle of

    the switch-closed time again provides modulation of the output voltage. As per Figure

    1-1D the buck circuit is able to provide an output voltage that is less than the input, while

    the buck-boost provides an inverted voltage that can range in magnitude from levels that

    are both greater and lesser than the input (Figure 1-1F).

    The flyback converter of Figure 1-1G derives from the buck-boost converter, except

    that the inductor of the buck-boost is replaced by an isolating transformer. The switch

    in Figure 1-1G is positioned for non-inverting output, so the conversion ratio plotted in

    Figure 1-1H for a 1 : 1 transformer is the same as that of the buck-boost but opposite in

    polarity. A step-up transformer with non-unity gain may be utilized in this circuit to realize

    19

  • +−

    + vL(t) -

    L

    Q

    D

    C R

    +

    Vout-

    Vin

    A Boost converter circuit

    0 0.25 0.5 0.75 10

    1

    2

    3

    4

    5

    Con

    vers

    ion

    Rat

    io M

    (V

    /V)

    Duty Cycle D

    B Boost conversion ratio

    +−

    LQ

    D C R

    +

    Vout-

    Vin

    C Buck converter circuit

    0 0.25 0.5 0.75 10

    0.5

    1

    Con

    vers

    ion

    Rat

    io M

    (V

    /V)

    Duty Cycle D

    D Buck conversion ratio

    +− L

    Q D

    C R

    +

    Vout-

    Vin

    E Buck-boost converter circuit

    0 0.25 0.5 0.75 1−5

    −4

    −3

    −2

    −1

    0

    Con

    vers

    ion

    Rat

    io M

    (V

    /V)

    Duty Cycle D

    F Buck-boost conversion ratio

    +−

    Q

    D

    C R

    +

    Vout-Vin

    1 : n

    G Flyback converter circuit

    0 0.25 0.5 0.75 10

    1

    2

    3

    4

    5

    Con

    vers

    ion

    Rat

    io M

    (V

    /V)

    Duty Cycle D

    H Flyback conversion ratio

    Figure 1-1. Common converter circuits and their idealized conversion ratios M asfunctions of switching duty cycle D. Figures adapted from Erickson andMaksimovic [4].

    20

  • more drastic conversion ratios than would be had from the buck-boost. The transformer

    additionally provides isolation protection between input and output.

    1.3 Impedance

    While Equation 1–1 describes the characteristic transient behavior of inductive

    components in inducing voltages in opposition to changes in the current passing through

    it, an impedance analysis is useful for characterizing the behavior of an inductor when

    the changes are sinusoidal or periodic. The impedance Z of an inductor relates the

    voltage V across the inductor to the current I passing through it

    Z =V

    I, (1–3)

    where both voltage and current are sinusoidally varying. Other non-sinusoidal periodic

    excitations can be considered using Fourier analysis to decompose the signal into a

    summation of sinusoidal signals.

    When determining impedance, the voltage and current waveforms are represented

    by phasors, each being a vector with magnitude equal to the amplitude of the waveform

    and with angle equal to the phase difference between the waveform and some common

    reference. In complex form, the real part of the impedance represents the in-phase

    energy-dissipative (resistive) component and the imaginary part represents the

    out-of-phase energy-storage component. The impedance of an ideal inductor with

    inductance L at an angular frequency of ω is,

    ZL = jωL. (1–4)

    While the current through an ideal inductor lags the voltage across it by 90◦ (a quarter

    wavelength), resistive and capacitive effects cause frequency-dependent magnitude and

    phase relationships.

    21

  • 1.4 High Frequency Benefits and Challenges

    Because the impedance of an ideal inductor scales with frequency, power

    converters with higher switching frequencies can utilize lower-valued and physically

    smaller inductors. However, increasing to very high frequency (VHF) switching

    (> 10 MHz) also introduces several challenges that could severely limit performance of

    these converters if not addressed.

    Many magnetic core materials, which are used to increase magnetic induction in

    inductors/transformers, are unable to physically switch their magnetization fast enough

    in response to a VHF applied field. The time lag between changes in the applied field

    and the responding change in magnetic induction in the material leads to power losses

    in the core. Designers often utilize magnetic material anisotropy (or sometimes a dc bias

    magnetic field) perpendicular to the applied magnetic field in order to improve the high

    frequency response time of magnetic materials at the expense of lower permeability [5].

    Eddy current generation within electrically conductive materials results in the skin

    effect, the confinement of electric and magnetic fields to the materials’ surface at high

    frequency excitation. The skin effect limits the effective cross sectional area of both

    the electric winding and the magnetic core, leading to greater resistance and lesser

    inductance.

    In the VHF switching range, the converter circuit design demands components with

    inductance and capacitance values that are on the order of the unintended parasitic

    inductances and capacitances that inherently occur between components and in the

    interconnections between them [6]. The design of the inductors and transformers must

    consider the large parasitic capacitance experienced as energy is stored in the electric

    field between adjacent conductor traces. This parasitic capacitance limits the maximum

    operating frequency and efficiency of the inductor/transformer.

    22

  • 1.5 Survey of Existing Microfabricated Inductors and Trans formers

    Gardner et al. [7] published in late 2009 a comprehensive review of contemporary

    on-chip inductors with magnetic films and evaluated their application to integrated

    power converters. The review reached some interesting conclusions. A majority of

    the works featured inductors with inductance densities of less than 100 nH/mm2,

    calling into question whether the magnetic films are providing sufficient inductance

    improvement to warrant their inclusion. Additionally, few results were applicable to high

    frequency switching (> 100 MHz). The review identified quality factor as performance

    parameter of interest for efficient power conversion, with quality factor of 1 as the

    minimum below which inductors acted more like resistors than as intended as energy

    storing components [7]. Air-core microinductors, on the other hand, have mostly been

    designed for GHz radio frequency (RF) applications. Such devices can attain high quality

    factors when suspended above conductive substrates but typically have low inductances

    on the order of only several nH.

    Ahn, NiFe

    Yamaguchi, FeAlO

    Sato, FeCoBN

    Song, FeZrBAg

    Fukuda, NiZn

    Wang, NiFe Viala, FeHfN

    Flynn, NiFe

    Orlando, NiFe

    Lee, CoTaZr

    Park, Air

    Young, Air

    Choi, Air

    Weon, Air Yoon, Air

    1

    10

    100

    1 10 100 1000 10000

    Pe

    ak

    Qu

    ali

    ty F

    ac

    tor

    Frequency for Peak Quality Factor (MHz)

    Figure 1-2. Review of both magnetic-film (shaded in blue) and air-core (shaded in green)microinductors with each plotted in terms of peak quality factor and thefrequency at which the peak quality factor was obtained. Bubble size isproportional to inductance density.[8–22]

    23

  • A survey of existing microinductors, including both magnetic film and air cores,

    revealed that there was a significant gap where few microfabricated inductors were

    designed for frequencies ranging from tens of MHz up to 1 GHz. The gap was evident

    in Figure 1-2, where a number of microinductors were plotted against their peak quality

    factor and the frequency at which the peak quality factor was attained. A typical

    magnetic film microinductor had an inductance density of about 55 nH/mm2, almost

    twice that of the typical air core counterpart, which had about 30 nH/mm2. The situation

    was reversed for the peak quality factor where the median air core inductor had a peak

    quality factor of 50, far greater than the median magnetic film inductor at a quality factor

    of 9.

    A similar frequency gap was found amongst the results gathered from works

    reporting existing microtransformers as can be seen on the plot in Figure 1-3. Both

    magnetic film and air core microfabricated inductors were plotted against maximum

    efficiency of power transfer through the transformer and the frequency at which

    the maximum efficiency occurred. Most of the works were found to focus only on

    transformers with 1 : 1 turns ratios with near-unity voltage/current gain.

    1.6 Air-Core Passive Components for Microscale Power Conve rters

    Between magnetic-film-core devices operable at < 10 MHz and microwave RF

    air-core devices with high performance at > 1 GHz lies a large frequency gap amongst

    the reported microinductors and microtransformers. Coincidentally, the expected

    switching frequencies of next generation power converters fall right into this gap for

    which no inductor/transformer technology can yet claim championship. However,

    the development is presented here for a new class of air-core microinductors and

    microtransformers that leverage microfabrication-enabled advancements to attain high

    performance in this desirable switching frequency range and to enable fully integrated

    power management systems in the smallest possible packages.

    24

  • Yamaguchi, Air

    Laney, Air

    Zolfaghari, Air

    Ng, Air Aly, Air

    Mino, CoZrRe

    Kurata, CoFeSiB

    Yamaguchi, CoNbZr

    Mino, CoZrRe

    Xu, NiFe

    Sullivan, NiFe

    Sullivan, NiFe

    Brunet, NiFe

    Park, NiFe

    Rassel, NiFe

    Yun, NiFe

    Wang, NiFe

    0%

    20%

    40%

    60%

    80%

    100%

    1 10 100 1000 10000

    Eff

    icie

    nc

    y

    Frequency for Maximum Efficiency

    Figure 1-3. Review of both magnetic-film (shaded in blue) and air-core (shaded in green)microtransformers with each plotted in terms of maximum efficiency and thefrequency at which the maximum efficiency was obtained. Bubble size isproportional to voltage gain. [23–40]

    In order to design these devices, models are analyzed to uncover the ideal

    characteristics for operating in the VHF range. Compared to traditional air-core

    components, the devices here feature thicker metal traces arranged into intricate

    three-dimensional windings for lower loss and higher density. A multilevel microfabrication

    process is developed for molding copper parts with the necessary characteristics of

    thickness, minimum feature size, and out-of-plane stacking.

    This dissertation has been organized as follows. In Chapter 2 information gathered

    from a survey of existing microfabricated inductors and transformers is presented.

    Chapter 3 highlights the goals and considerations that motivated the design of

    the microinductors. Similarly, the design of the microtransformers is discussed in

    Chapter 4 alongside an introduction to the math and methods used to characterize the

    microtransformer performance. Presented in Chapter 5 is the fabrication process that

    was developed to response to the aforementioned design needs and the shortcomings

    of existing processes in meeting these needs. Characterization of the microfabricated

    25

  • inductors and transformers at radio frequencies is covered separately in Chapters 6 and

    7 for the inductors and transformers, respectively. Chapter 8 presents the packaging

    and testing of microfabricated inductors within a prototype VHF hybrid boost converter

    circuit and with a commercial converter chip. Chapter 9 concludes the dissertation with

    a summary of the advancements led by this work in filling the gap for microscale power

    inductors and transformers at VHF and enabling fully integrated power converters.

    26

  • CHAPTER 2BACKGROUND

    This chapter provides background information on existing works that have

    contributed to the state of the art of microfabricated inductors and transformers.

    Highlighted first are several significant demonstrations of very high frequency switched

    mode power converters that have created the possibility for full integration of all

    converter components in a single package. Power inductors and transformers from prior

    works are then surveyed with attention focused on the challenges and accomplishments

    met by each. Quantitative results from the surveyed inductors and transformers are

    outlined in tabular form at the end of the chapter along with results from selected

    GHz RF air core components for comparison. The works have been selected for their

    inclusion of detailed performance characteristics relevant toward enabling integrated

    switched mode power converters.

    2.1 High Frequency Power Converters

    Examples of very high frequency switching power converters are summarized

    here to demonstrate the viability of this new breed of converters in providing high

    performance with nH-level inductive components. The results from these works provided

    an idea of what switching frequencies would be used in next-generation converters and

    what size inductors would be required.

    Hazucha et al. [2] reported results from a four-phase dc-dc buck converter

    implemented in 90nm CMOS and designed to operate at switching frequencies ranging

    from 100− 600 MHz. The optimal switching frequency was determined by the size of the

    inductors. Four 6.8 nH discrete inductors were soldered onto the package for a switching

    frequency of 233 MHz. The authors quoted quality factors for the inductors of Q = 20 at

    100 MHz and Q = 30 at 300 MHz. The chip area of the converter was 1.26 mm2. The

    converter delivered 0.3 A at 0.9 V from a 1.2 V input with 83% efficiency.

    27

  • Li et al. [41] reported two discontinuous conduction mode dc-dc boost converters

    fabricated in standard 0.13 µm CMOS, both utilizing off-chip discrete inductors. One

    was a 100 MHz 4-phase boost converter that delivered 240 mW from a 1.2 V supply with

    output ranging from 3 − 5 V and peak efficiency of 64%. This converter used four 22 nH

    inductors, one per phase, and the CMOS area alone comprised 0.55 mm2. The other

    reported converter was a 45 MHz hybrid boost converter delivering 20 mW at 6 − 10 V

    also from a 1.2 V supply, with peak efficiency of 37%. The hybrid converter used a single

    43 nH inductor, while the CMOS area was 0.17 mm2. The authors stated that both high

    switching frequency and discontinuous conduction mode were utilized to reduce the size

    of the required off-chip components.

    2.2 Inductors

    Ahn et al. [22] constructed a 4 mm × 1 mm × 130 µm toroidal inductor on a silicon

    wafer via a multilevel metallization process. The inductor consisted of 33 turns of

    40 µm-thick copper traces wound around a 30 µm-thick electroplated Ni81Fe19 magnetic

    core. This composition of NiFe was cited as being chosen for achieving maximum

    permeability, minimum coercivity, minimum anisotropy, and maximum mechanical

    hardness. Permeability of the magnetic core was determined at approximately µr = 800

    both by vibrating sample magnetometry and magnetic circuit evaluation with a core

    of known dimensions. The measured inductance was 400 nH at 10 kHz, but this value

    decreased with frequency to a value of approximately 50 nH at 1 MHz. Such applications

    for the inductor were listed as sensors, actuators, and power converters.

    Yamaguchi et al. [42] demonstrated a 7.6 nH thin-film inductor with quality factor

    of 7.4 at 1 GHz intended for use in impedance matching at the front-end receiver

    of a 1 GHz mobile communication handset. The square-spiral inductor measured

    370 µm × 370 µm and was comprised of 4 turns of 2.8 µm-thick sputter deposited

    AlSi windings. After encapsulating the windings with a 3.5 µm-thick insulating layer of

    polyimide, a 0.1 µm-thick Fe61Al13O26 magnetic film was sputter deposited over the coils

    28

  • and was patterned by ion milling. The authors acknowledged that simply covering the

    inductor with the magnetic film could only double the inductance at best but predicted

    the improvement would prove sufficient for commercial use. An inverse relationship was

    found in the FeAlO films between the magnetic film resistivity and its resonant frequency,

    although high values of each were desired to avoid excess losses at the GHz range. Slits

    were created in the magnetic film to inhibit eddy current generation, resulting in a 31%

    reduction in the ac resistance at 1 GHz compared to the case of the film without slits.

    Sato et al. [11] developed a rectangular spiral inductor for 5 MHz switching

    dc-dc converters. The inductors measured 6310 µm × 3466 µm in area and featured

    50 µm-thick electroplated copper windings capped with a FeCoBN magnetic thin film.

    The magnetic film was deposited by dc magnetron sputtering with four alternating layers

    of 1.5 µm FeCoBN and 0.4 µm AlNx to suppress eddy currents. Film permeability was

    estimated at 900 up to 300 MHz. The multilayer film was etched in a single wet step

    with mixture of phosphoric, acetic, and nitric acids used to dissolve both constitutive

    materials at once. Inductance was measured at 370 nH with a peak quality factor

    of 15 at 7 MHz. The inductor was tested in 5 MHz switched mode power converters

    constructed of discrete components in both boost and buck configurations. The buck

    converter produced an output of 3 V from a 5 V input with a peak efficiency of about

    82% at an output current up to 500 mA. The boost converter operated with the same

    conversion ratio in reverse; an output of 5 V was obtained from a 3 V input. A similar

    peak efficiency of about 82% was achieved from the boost converter at 150 mA output

    current.

    Fukuda et al. [9] fabricated a 6 mm × 6 mm planar square spiral inductor that

    was fully encapsulated between two NiZn ferrite magnetic layers. Ferrite composition

    was NiO/CuO/ZnO/Fe2O3 in ratios of 16/12/23/49. The lower ferrite layer was first

    screen-printed over a silicon wafer and sintered at 900 − 1000◦C to a final thickness

    of 40 µm. Relative permeability of the lower ferrite layer was measured at 120. Copper

    29

  • windings were then electroplated 50 µm-thick on top of the lower ferrite layer. The

    upper ferrite layer was screen printed on top of the windings and was hardened but not

    sintered. Due to the lower relative permeability—measured at 25—the final upper ferrite

    layer was deposited more than twice as thick at 100 µm. Characterization of the inductor

    revealed an inductance of 1.4 µH with a peak quality factor of 40 at 5 MHz. Magnetic

    field analysis by the finite element method indicated that the inclusion of magnetic ferrite

    in the spaces between adjacent turns of the coil were beneficial in confining magnetic

    flux to the core, minimizing eddy current loss in the copper coil.

    Viala et al. [14] reported a square spiral inductor with a density around 90 nH/mm2

    and peak quality factor of about 10 at 1.5 GHz with sputtered FeHfN films over spiral

    inductors, but only noted a modest increase in inductance of 35% over the air-core case.

    The magnetic films were laminated and consisted of ten alternations of 0.1 µm-thick

    (Fe97.6Hf2.4)90N10 magnetic and 500 Å-thick SiO2 insulating layers. The authors noted

    difficulty in using magnetic films with spirals due to their having both in-plane and

    out-of-plane magnetic field components.

    Characteristics of the above mentioned inductors were summarized in Table 2-1

    alongside those from other significant works.

    2.3 Transformers

    Mino et al. [23] presented a 3 mm × 4 mm transformer fabricated by a completely

    dry process on a silicon substrate and consisting of copper coils wrapped around a

    magnetic layer of CoZrRe. The magnetic film was deposited by ion beam sputtering and

    was quoted as having a relative permeability > 3000. The copper coils were wrapped

    around the core in a primary:secondary ratio of 12 : 3 to obtain associated primary and

    secondary inductances of 350 nH and 40 nH, respectively. The microtransformer was

    mounted in a ceramic package and tested within a forward converter circuit operating at

    32 MHz. Output from the converter was 0.6 V to a 10 Ω load with 10 V source input. The

    30

  • efficiency of the converter was not given but was reportedly “low” due to the low primary

    inductance of the transformer.

    Yamaguchi et al. [33] fabricated a 2.4 × 3.1 mm2 microtransformer with stacked

    primary and secondary spiral copper coils sandwiched between multilayered CoNbZr/SiO2

    magnetic films on a glass substrate. RF sputtering was used to deposit both the copper

    windings and the magnetic films, which were each patterned by a photoresist lift-off

    method. Annealing of the magnetic core was performed at 250 ◦C for 1 hour under

    vacuum with a rotating magnetic field. The copper traces were deposited 7.5 µm thick

    and patterned 100 µm wide with 10 µm spacing. The turns ratio of primary:secondary

    coils was 8 : 7.3. A 10 Ω load was attached to the secondary winding for measurement

    of the transformer efficiency with 1 V sinusoidal input to the primary in the frequency

    range of 1 − 20 MHz. A maximum efficiency of 67% was obtained at 10 MHz, beyond

    which point efficiency was said to decrease due to core loss.

    Sullivan and Sanders [27] measured the performance of microfabricated power

    conversion transformers with areas on the order of 10 mm2 and primary:secondary

    turns ratios of 8 : 4. Primary and secondary windings were interleaved in an elongated

    spiral (racetrack) and consisted of 20 µm-thick electroplated copper. A multilayer

    laminated NiFe/SiO2 material acted as the magnetic core with a relative permeability of

    2000. Two designs were fabricated: one had a sandwich configuration with the copper

    windings embedded between separate layers of magnetic material, while the other

    design featured a closed core that fully enclosed the windings. The sandwich design

    was said to not only decrease inductance by a factor of 5 compared to the closed core,

    but also produced an “unfavorable field distribution” that further increased losses. A

    half-bridge forward converter served as a test bed for the sandwich transformer and

    measured 43.4% efficiency with 3.74 W input at 30 V and 1.625 W output at 4.21 V. The

    same converter circuit was tested again with a litz wire transformer having the same

    inductance as the sandwich transformer but assumed to have no loss. By comparing

    31

  • the difference in efficiency with the litz wire versus the sandwich transformer, the

    sandwich transformer efficiency was estimated at 61%. The closed core transformer was

    tested with a network analyzer and projected to have an efficiency of 70%. Higher than

    expected losses were attributed to hysteresis losses and shorting between layers in the

    core.

    Brunet et al. [28] presented a 30 mm2 microfabricated transformer consisting of

    interleaved, racetrack-shaped primary and secondary coils encapsulated in 4 µm-thick

    electroplated Ni81Fe19 magnetic core. The copper coils were electroplated 43 µm thick

    and were arranged in a turns ratio of 4 : 2. Electrical characteristics were obtained using

    an impedance analyzer. A primary inductance of 0.9 µH was measured to be constant

    up to 5 MHz. Leakage inductance was determined at 0.4 µH measuring the primary

    inductance while shorting the secondary coil. The transformer was tested in a full-bridge

    dc-dc converter at 2 MHz. Converter efficiency was measured at 40% for input voltages

    > 2 V. The core was found to saturate at an input voltage of 4.5 V for a maximum output

    power of 0.4 W.

    Characteristics of the above mentioned inductors were summarized in Table 2-2

    alongside those from other significant works.

    32

  • Table 2-1. Literature survey of microinductors.Inductance Peak Q DC

    Reference Layout Core Area Inductance density Peak frequency resistancematerial (mm2) (nH) (nH/mm2) Q (MHz) (Ω)

    Ahn et al. [22] Toroid Ni81Fe19 4 400 100 1.5 1 0.3Yamaguchi et al. [42] Spiral Fe61Al13O26 0.137 7.6 56 7.4 1000 6.5Sato et al. [11] Racetrack FeCoBN 21.9 370 16.9 15 7Song et al. [13] Racetrack FeZrBAg 146 1000 6.84 25 10 3.95Fukuda et al. [9] Spiral NiZn 36 1400 38.9 40 5 0.67Wang et al. [10] Racetrack NiFe 5.69 160 28.1 6 4 0.261Viala et al. [14] Spiral FeHfN 0.09 10 111 10 1500Flynn et al. [20] Toroid NiFe 10 1940 194 2 2Orlando et al. [21] Toroid NiFe 31.4 500 15.9 20 2 0.095Lee et al. [19] Solenoid CoTaZr 0.88 70.2 79.7 6.5 25 0.67Park and Allen [8] Spiral Air 1.69 37.8 22.4 44 1200 2.76Young et al. [15] Solenoid Air 0.25 14 56 18 820Choi et al. [16] Spiral Air 0.144 4.6 31.8 50 3500Weon et al. [17] Solenoid Air 0.05 2.1 42 78 4000 0.342Yoon et al. [18] Solenoid Air 0.06 1.17 19.5 84 2600

    33

  • Table 2-2. Literature survey of microtransformers.Primary Secondary

    Reference Area inductance inductance Coupling Voltage Efficiency Frequency Core(mm2) (nH) (nH) coefficient gain (MHz) material

    Mino et al. [23] 12.0 350 40 0.5 0.3∗ 3% 32 CoZrReYamaguchi et al. [33] 7.44 500 450 0.7 67% 10 CoNbZrKurata et al. [24] 1.38 50 50 0.92 1∗ 54% 100−250 CoFeSiBMino et al. [25] 25 820 820 0.93 1∗ 58%∗ 25 CoZrReXu et al. [26] 4 800 800 0.9 0.63 77%∗ 10 Ni80Fe20Sullivan and Sanders [27] 8.42 1380∗ 345 0.5∗ 61% 8 NiFeSullivan and Sanders [27] 11.85 3176∗ 794 0.5∗ 70% 10 NiFeBrunet et al. [28] 29.92 900 225∗ 0.58∗ 40% 2 Ni81Fe19Park and Bu [29] 5.7 440 440 0.85 1∗ 32% 25 Ni80Fe20Rassel et al. [30] 4.95 100 80 0.9 0.9∗ 1% 0.5 NiFeYun et al. [31] 78.4 830 830∗ 0.91 0.9 84%∗ 5 Ni81Fe19Wang et al. [32] 23.7 400 400 0.93 0.89 72% 5 NiFeYamaguchi et al. [33] 7.44 70 65 0.4 30% 10 AirCheung et al. [34] 0.16 8∗ 8∗ 0.75∗ 1∗ 1000 AirCheung et al. [34] 0.25 0.5∗ 12∗ 0.75∗ 5∗ 1000 AirLaney et al. [35] 0.16∗ 1.65 1.65 0.55 1∗ 56% 2500 AirLong [36] 0.16 8.5 8.5 0.84 1∗ 2000∗ AirRibas et al. [37] 0.09 8.6 8.6 0.79 1∗ 32%∗ 10000∗ AirZolfaghari et al. [38] 0.06∗ 11∗ 180∗ 3 29%∗ 1500 AirNg et al. [39] 0.09 2 3 0.8 1.2∗ 60% 8000 AirAly and Elsharawy [40] 0.32 4.79 4.79 0.88 1∗ 56%∗ 2000∗ Air∗Asterisked values were estimated based on the other information gathered from the respective references.

    34

  • CHAPTER 3INDUCTOR DESIGN

    This chapter outlines a methodology for designing microinductors for use in

    high-frequency switched-mode power supplies. The quality factor is investigated as

    a metric for the efficiency of the inductor in storing energy. Three inductor attributes

    affecting peak quality factor—inductance, resistance, and maximum operating

    frequency—are discussed in terms of the trade-offs in attempting to maximize any

    one of these quantities. The stacked planar spiral layout is chosen for the microinductors

    with the goal of reaching high quality factor by maximizing inductive coupling while

    minimizing electrical resistance. A modeling strategy is presented for optimization

    of the design of such inductors and prediction of their performance. The models

    provide a method for determining the optimal geometric proportions based on certain

    combinations of desired criteria: inductance, size, operating frequency, and maximum

    quality factor.

    3.1 Quality Factor Definition

    The quality factor, Q, of a circuit is a dimensionless quantity that generally provides

    a metric of how much energy is stored in a circuit versus how much energy is dissipated

    by it. However, the generality of this concept has led to confusion of the definition of

    Q amongst researchers since there are many application-specific interpretations and

    methods of extraction of Q [38, 43–46]. For example, one traditional use of Q is in

    quantifying the selectivity of a resonant filter circuit [47]. In such filtering applications, Q

    is defined as the ratio of the circuit resonant frequency to its half-power bandwidth.

    3.1.1 Quality Factor of Non-Ideal Reactive Components

    In contrast to the single-valued quality factor of resonant circuits, the quality factor

    Q of an energy storing circuit component (e.g. inductor or capacitor) quantifies how

    much energy is stored in the component versus how much is dissipated by it at each

    35

  • frequency. However, even this alternate definition can take on different meanings to

    different communities when the device is operated near its resonant frequency.

    The discrepancy in definition arises from the fact that although the impedance of a

    device under alternating current (ac), sinusoidal excitation at its resonant frequency is

    purely resistive, energy is being stored and transferred within the electrical and magnetic

    fields within the device. By definition, the reactive part of the impedance falls to zero at

    resonance, and none of the energy stored in the device is available to the external circuit

    attached to it. Passive components in power conversion applications, however, need to

    store energy from the circuit and then provide that energy back to the circuit. For this

    reason, a separate definition of Q is most appropriate for power conversion application

    with the property that Q falls to zero at resonance. The simplified definition of Q used for

    quantification of power-passive performance is the ratio of the imaginary to the real part

    of the complex impedance looking into the device,

    Q =ℑ{Z}ℜ {Z} . (3–1)

    It can be shown that the above definition of Q for any one-port, energy-storing

    component correctly provides a measure of the storage efficiency as seen by the

    external circuit. The expected measure for the component under sinusoidal excitation is

    the rate energy is stored in and retrieved from the device to the rate energy is dissipated

    in it,

    Q =Reactive power transfer

    Real power transfer. (3–2)

    Assuming root mean square (RMS) quantities for voltage and current, the complex

    power transfer to/from a component is the product of the ac voltage V across the

    component and the complex conjugate of the current I through it, so that Equation 3–2

    is rewritten in terms of voltages and currents as

    Q =ℑ{

    V I}

    ℜ{

    V I} . (3–3)

    36

  • Applying properties of complex conjugates, the real and imaginary part operators can be

    expanded into equivalent algebraic expressions as

    Q =

    (

    V I − V I)

    /j2(

    V I + V I)

    /2, (3–4)

    which simplifies through manipulation to,

    Q =

    (

    VI− VI

    )

    /j2(

    VI+ VI

    )

    /2. (3–5)

    The voltage and current terms in Equation 3–5 are arranged so that the impedance

    equivalent is readily identified as,

    Q =

    (

    Z − Z)

    /j2(

    Z + Z)

    /2. (3–6)

    By properties of complex conjugates, Equation 3–6 is identical to the original formulation

    of Q in Equation 3–1 as the ratio of the imaginary to the real part of the complex

    impedance of component.

    3.1.2 Quality Factor of Inductor

    As discussed in Section 3.1.1, the quality factor Q provides a metric of the ac

    energy storage efficiency of actual, non-ideal reactive circuit components, such as

    microinductors. The formulation of Q as the ratio of imaginary to the real part of the

    inductor impedance (as in Equation 3–1) provides a figure of merit that quantifies the

    degree to which an inductor acts like an inductor to an attached circuit.

    For example, when used with direct current (dc) there is no electrical characteristic

    that differentiates an inductor from a trace of wire with some resistance. Although

    energy is stored in the magnetic field induced by the current flowing through the inductor

    even at dc, in the absence of any variation in current over time, that energy is never put

    back into the circuit.

    When there are ac fluctuations in the current flowing through the inductor, energy

    is stored as the current increases in magnitude to its peak level and is then retrieved

    37

  • from the magnetic field back into the circuit as the current decreases in magnitude.

    Some energy is dissipated as heat due to the resistance of the electrical path through

    the inductor. At low frequencies, the rate of energy storage/retrieval is less than the

    power dissipated by the inductor, and the inductor has consequently low quality. As

    the frequency of the current oscillation increases, however, so too does rate of energy

    storage/retrieval while the power dissipated remains relatively constant, and the inductor

    therefore attains a higher Q. If the inductor is modeled as the serial combination of an

    ideal resistor R and an ideal inductor L, the expression for quality factor at an angular

    frequency of ω as calculated from Equation 3–1 is

    QRL =ωL

    R. (3–7)

    This simplified expression ignores the changes in resistance that occur at very high

    frequencies and also ignores capacitive energy storage in the electric field that invariably

    exists in the inductor.

    Because of parasitic capacitance, Q diminishes near the self-resonant frequency of

    the inductor as more energy is stored in the electric field between traces. By definition,

    Q = 0 at resonance as equal amounts of energy are traded between electric and

    magnetic fields, and the spiral again appears as a resistor to the circuit.

    3.2 Performance Trilemma

    From Equation 3–7 the quality factor of a quasi-ideal inductor (ignoring effects

    of capacitance) is dependent on its inductance, resistance, and operating frequency.

    Ideally, the quality factor would be maximized if the inductance and operating frequency

    were maximized and the resistance minimized. In practice, however, all of these

    quantities are linked, so that improvements to any one of these three attributes is

    often done at the detriment of the other two.

    Consider the simple model of the inductor shown in Figure 3-1 with inductance

    L, series resistance R, and shunt capacitance C . Self-resonance limits the maximum

    38

  • operating frequency of the inductor and for the case of low resistance is approximately

    equal to the natural frequency,

    ω0 =

    1

    LC. (3–8)

    The above equation clearly shows that increasing inductance directly results in

    decreasing resonant frequency. However, attaining higher inductance often entails

    increasing the trace length of the inductor winding, resulting in higher capacitance,

    which in turn further decreases the resonant frequency. The increased trace length

    also increases the resistance of the inductor. Designing inductors must balance these

    competing goals to deliver a device that is tailored to the application.

    C

    R

    L

    Figure 3-1. Circuit diagram of simple inductor model with inductance L, series resistanceR, and shunt capacitance C .

    3.3 Stacked Planar Spiral Layout

    In response to the previously mentioned concerns for maximizing inductance while

    minimizing resistance and capacitance, the stacked planar spiral layout was selected

    for the inductors of this work. The planar spiral layout features conductive traces that

    are concentrically wound into a flat spiral as depicted in Figure 3-2. This layout is the

    most popular amongst all integrated inductors because it offers high density through

    tight spiral packing and it is easy to fabricate via conventional planar microfabrication

    steps. When all traces are constrained to only a single plane, however, performance is

    limited by poor magnetic coupling between outer and inner windings. As the number of

    spiral turns is increased, the separation between inner and outer windings can become

    so great that the inner turns contribute more towards increasing the resistance of the

    inductor than towards its inductance.

    39

  • To overcome the planar limitation, vertical stacking of planar spirals is used to

    increase both the inductance density and the quality factors of the inductors. Because

    the planar spiral is by nature wider in diameter than it is thick (given typical conductor

    thicknesses), stacking spirals provides excellent magnetic field coupling in the vertical

    direction. Assuming perfect coupling, the inductance of a two-layer stacked-spiral

    inductor can reach up to four times that of a single layer while the resistance is only

    doubled. In this simplified example, the inductance to resistance ratio of the two-layer

    device is improved to twice that of a single-layer device.

    3.4 Low Frequency Analytical Inductor Model

    The low-frequency model of the inductor includes only the electrical resistance

    along the length of the trace winding and the magnetic field generated when an

    electric current passes through the winding. The current is assumed to flow uniformly

    through the cross section of each trace, ignoring current crowding due to interactions

    between moving charge carriers. Inductance and resistance are first calculated for a

    single-layer winding of uniform trace width and thickness. The values are then extended

    as appropriate when two layers are vertically stacked.

    By the year 1928 Wheeler [48] had derived by empirical data an expression to

    predict the inductance of discrete radio coils. More than 70 years later Mohan et al. [49]

    modified the existing expression only slightly to be valid also for microinductors,

    Lmw =K1µ0n

    2davg

    1 + K2p. (3–9)

    In the above expression, K1 and K2 are empirically derived values that are specific to

    the shape of the spiral (i.e. square, hexagonal, octagonal) and are listed in Table 3-1.

    An additional expression was presented in Mohan et al. [49] for calculating inductance

    based on a current sheet approximation [50],

    Lgmd =µn2davgc12

    [

    ln

    (

    c2

    ρ

    )

    + c3ρ+ c4ρ2

    ]

    , (3–10)

    40

  • Table 3-1. Coefficients for modified Wheeler (Equation 3–9) and current sheet (Equation3–10) expressions [49].

    Shape K1 K2 c1 c2 c3 c4Square 2.34 2.75 1.27 2.07 0.18 0.13Hexagonal 2.33 3.82 1.09 2.23 0.00 0.17Octagonal 2.25 3.55 1.07 2.29 0.00 0.19Circle - - 1.00 2.46 0.00 0.20

    for which expression the shaped-dependent coefficients (c1, c2, c3, and c4) are provided

    not only for square, hexagonal, and octagonal shapes but also for circular. These

    coefficients are also listed in Table 3-1. The choice between using the two previously

    listed expressions depends on the situation. If a circular layout is desired, the current

    sheet expression in Equation 3–10 provides the best accuracy. If rearranging the

    expression to solve for a different variable, the modified Wheeler expression in Equation

    3–9 is simpler to solve.

    The authors of these expressions noted that each had been validated only for

    inductors < 100 nH with outer diameters ranging from 100 − 480 µm [49]. As part of

    this dissertation work, the inductance values calculated from Equation 3–9 were verified

    against magnetoquasistatic simulations with less than 5% error for inductors up to

    1050 nH and outer diameters up to 2.5 mm (see Section 6.3.1.1).

    All of the other variables in both Equation 3–9 and Equation 3–10 — i.e. the number

    of turns n, the packing density p, and the average diameter davg — are obtained from the

    geometry of the spiral. The geometry of the spiral can be uniquely specified in terms of

    the winding trace width w , the spacing between adjacent winding traces s, the number

    of winding turns n, and the outer diameter D. These dimensions are marked on the

    diagram of an example spiral in Figure 3-2. Inner (d) and outer (D) diameters were

    measured from the centerlines of the innermost and outermost traces, respectively.

    The inner diameter d represents the space contained within the spiral that is clear of

    41

  • w s

    d

    D

    Figure 3-2. Diagram of planar spiral layout with n = 3 turns and all other dimensionslabelled.

    windings and is calculated as a function of the other dimensions,

    d = D − 2 [wn + s (n − 1)] . (3–11)

    The average diameter is then simply calculated as

    davg =D + d

    2. (3–12)

    The packing density p represents the fraction of the inductor area that is filled with

    windings and is defined as

    p =D − dD + d

    . (3–13)

    Extending the aforementioned inductance and resistance calculations for a single

    layer spiral, the total inductance for an inductor with two identical spirals stacked

    vertically is calculated in proportion to L0, the inductance of a single-layer spiral from

    either Equation 3–9 or 3–10,

    Ldc = 2 (1 + k)L0, (3–14)

    42

  • where k is the coupling coefficient representing the portion of shared magnetic flux

    linking the top and bottom spirals. The value of k can vary between −1 and 1. If the

    spirals are positioned so that no magnetic flux is shared between spirals, k = 0 and the

    total inductance is twice that of the single-layer coil. When all flux is shared between

    coils, k = 1 and the total inductance is four times that of the single-layer coil. If the

    coils are stacked so that the magnetic fluxes of each coil are in opposition, k can have

    a negative value as the opposing magnetic fields nullify and reduce the total amount of

    flux linking the coils.

    The dc resistance of the inductor can be calculated by the familiar expression for

    resistance,

    Rdc =ρl

    wt, (3–15)

    where ρ is the electrical resistivity of the trace material, l is the total electrical path length

    of the inductor, and t is the thickness of the trace. The total trace length of the stacked

    spiral windings can be calculated from the geometry design variables. For the two-layer

    stacked square spiral, the total electrical trace length is evaluated as

    l = 2[

    4nD − (2n − 1)2 (w + s)]

    . (3–16)

    The trace length for the two-layer stacked circular spiral case is calculated as

    l = 2π [nD − n (n − 1) (w + s)] . (3–17)

    3.5 Trends and Optimization

    3.5.1 Analytical

    The expressions listed in Section 3.4 for estimating the low-frequency inductance

    and resistance of spirals were used to explore performance trends associated with

    sweeping certain design variables. A square spiral shape was assumed for ease in

    rapidly iterating layout and modeling. Perfect coupling (k = 1) was assumed for all

    cases. The target metric for this simplified analysis was the inductance to resistance

    43

  • ratio L/R, which is proportional to the quality factor at low frequencies (see Equation

    3–7).

    The first case was to determine the optimal packing density. The spacing between

    turns was fixed at s = 10 µm, while the width of the traces was varied. The number of

    turns was swept from the minimum number of turns (n = 1) turn up to the maximum

    number of turns that could physically be packed within the allotted area. Separate

    runs were completed for each different outer diameter D. The results were plotted for

    D = 500 µm (Figure 3-3A) and D = 1000 µm (Figure 3-3B). For all outer diameters and

    widths, the Ldc/Rdc ratios increased drastically as the number of turns was increased

    from 1 but then reached their maximal values at a packing density of approximately

    p = 0.4, which is equivalent to the points at which inner diameters were barely greater

    than 40% of the outer diameters. Also from these plots, the maximum Ldc/Rdc ratio

    increased with increasing trace width up to about w = 50 µm, past which no significant

    further increases were recorded.

    The plots of Figures 3-3A and 3-3B also suggested that the Ldc/Rdc might also

    increase with outer diameter. To test this idea, outer diameters from D = 0.5 mm up

    to D = 2.5 mm were evaluated using the optimal widths and packing densities already

    learned. For each trial with different outer diameter, the trace spacing was fixed at

    s = 10 µm, the trace width was fixed at w = 50 µm, and the number of turns n was

    calculated such that the packing density would be approximately p = 0.4. In this setup,

    the Ldc/Rdc computed for each outer diameter should represent approximately the peak

    values of the curves seen in Figure 3-3. The results from sweeping the outer diameter

    are plotted in Figure 3-4 and indicate a linear relationship between the Ldc/Rdc ratio and

    the outer diameter of the inductor.

    3.5.2 FastHenry

    FastHenry is a software program that can solve the magnetoquasistatic inductance

    and resistance of a three-dimensional structure using integral equation-based mesh

    44

  • 0 0.2 0.4 0.6 0.8 16

    8

    10

    12

    14

    16

    18

    20

    22

    24

    26

    Packing Density p

    L dc/

    Rdc

    (nH

    /Ω)

    20 µm30 µm40 µm50 µm60 µm70 µm80 µm

    Width w

    A Outer diameter D = 500 µm

    0 0.2 0.4 0.6 0.8 15

    10

    15

    20

    25

    30

    35

    40

    45

    50

    55

    Packing Density p

    L dc/

    Rdc

    (nH

    /Ω)

    20 µm30 µm40 µm50 µm60 µm70 µm80 µm

    Width w

    B Outer diameter D = 1000 µm

    Figure 3-3. Trends of inductance to resistance ratio vs. packing density for various tracewidths and outer diameters.

    45

  • 0.5 1 1.5 2 2.520

    40

    60

    80

    100

    120

    140

    Outer Diameter D (mm)

    L dc/

    Rdc

    (nH

    /Ω)

    Figure 3-4. Trends of inductance to resistance ratio vs. outer diameter using w = 50 µm,s = 10 µm, and n such that p ≈ 0.4

    10−1

    100

    101

    102

    103

    80

    100

    120

    140

    160

    Interlayer spacing (µm)

    Indu

    ctan

    ce (

    nH)

    4L0

    2L0

    FastHenry simulation results

    k=0

    k=1

    Figure 3-5. Trend of inductance vs. vertical gap between stack in 1 mm× 1 mm assimulated in FastHenry.

    46

  • analysis combined with a multipole-accelerated iterative solution algorithm [51]. A

    variety of stacked inductor designs were simulated in FastHenry, first to validate the

    analytical model results and then also to determine the effect of vertical separation

    between the two stacked layers on the mutual inductance between them. The same

    inductor design was simulated several times in FastHenry but with increasing vertical

    layer separation in each simulation trial. Plotted in Figure 3-5 is the low frequency

    inductance obtained for a 1 mm × 1 mm with interlayer separation varied from 0.1 −

    1000 µm. Drawn on the plot are lines indicating 2× and 4× the inductance L0 that would

    be obtained for a single winding layer. For the simulation with two winding layers the

    inductance asymptotically approaches 4L0 (k = 1) and 2L0 at the extremes of short

    and long separations, respectively. FastHenry simulations indicated that there would be

    minimal improvement to the inductance with separations less than 1% of the inductor

    diameter. A vertical separation of 10 µm was used for the microfabricated devices as

    shorter separations would serve only to detrimentally increase the parasitic capacitance

    between layers.

    3.6 Radio Frequency Effects

    Although direct current (dc) assumptions (e.g. uniform current distribution) enabled

    a simplified optimization of the inductor layout (such as diameter and number of

    turns), inductors designed for microscale power systems need to operate at such

    high frequencies (> 10MHz) that complex electromagnetic behavior alters the

    apparent inductances and resistances from the expected dc values. The dominant

    electromagnetic effects can be classified as due to capacitive coupling or due to eddy

    current generation.

    3.6.1 Capacitive Coupling

    Capacitive coupling results from variations in voltage potential that exist within

    different parts of the inductor. It is especially prominent between the terminal ends of the

    inductors where the difference in potential is the greatest. Because the terminal ends

    47

  • are where the inductor is connected to the rest of a circuit or where large landing pads

    provide electrical connection to probe tips for measurement and characterization, the

    effect of capacitive coupling can be highly dependent on factors that are external to the

    design of the inductor.

    Intrinsic to the design of the inductor, however, is the capacitive coupling that

    occurs between adjacent windings of an inductor. Microinductors intended for GHz RF

    applications typically consist of a single winding layer and a metal underpass providing

    electrical connection to the innermost turn. The interwinding capacitance of these

    single-layer inductors has long been known to be dominated by the areas where the

    windings and the underpass overlap [52, 53]. The stacking of windings for greater

    inductance density, as in the inductors of this work, would result in even greater levels of

    interwinding capacitance due to the significantly increased area of overlap. The general

    expression for capacitance between parallel plate electrodes,

    C =ǫA

    g, (3–18)

    shows that in addition to the overlap area between plates A, the other aspects affecting

    capacitive coupling are the permittivity ǫ of the material between the plates and the

    distance g of the gap between them. The multilevel thick-film fabrication technology

    presented in Chapter 5 minimizes the capacitive coupling between upper and lower

    winding layers by separating the layers by up to 30 µm and removing all dielectric

    material from between layers.

    If the inductors are fabricated on a conductive substrate such as silicon, the

    substrate creates an additional path for capacitive coupling. To electrically isolate

    the inductor from the substrate a thin dielectric layer such as silicon dioxide is often

    deposited over the substrate. Considering a scenario where two traces of an inductor at

    different voltage potentials sit atop the dielectric layer in close proximity, capacitors are

    48

  • formed with the dielectric layer between each trace and the substrate with the substrate

    providing an electrical connection between the two traces as illustrated in Figure 3-6.

    Copper Trace Copper Trace

    Conductive Substrate

    Dielectric Layer

    Figure 3-6. Diagram illustrating capacitive coupling through substrate between coppertraces of inductor winding.

    Like the interwinding capacitance, the shunt capacitance through the substrate

    contributes to resonant behavior as energy oscillates between inductive and capacitive

    storage elements. However, the finite resistance of the capacitive link through the

    substrate can have a profound effect on the perceived inductor behavior near the

    resonance. The substrate resistance can be modeled as a resistor Rc in series with the

    capacitance to the substrate Cs , as drawn in the circuit diagram in Figure 3-7.

    Cs

    Rdc

    Ldc

    Rc

    Figure 3-7. Circuit diagram of inductor model with dc inductance Ldc , series resistancethrough the inductor Rdc , shunt capacitance to the substrate Cs , andresistance along the capacitive path through the substrate Rc .

    49

  • 107

    108

    109

    10−9

    10−8

    10−7

    10−6

    10−5

    Frequency (Hz)

    Indu

    ctan

    ce (

    n