micromachined inductors and transformers for

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 11-Sep-2021

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

Page 2: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

c© 2012 Christopher D. Meyer

2

Page 3: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

I dedicate this to my loving family.

3

Page 4: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

Page 5: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

Page 6: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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 ”m vs. 30 ”m . . . . . . . . . 1056.3.2.2 Inductor shape: square vs. circular spirals . . . . . . . . 106

6

Page 7: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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 ”m 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 ”m 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

Page 8: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

Page 9: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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 ”m thick layers . . . . 75

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

9

Page 10: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

Page 11: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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 ”m thick layers . . . . . . . . . . . 126

7-5 Impedance plots of 1 : 3.5 transformer with 10 ”m 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 ”m thick layers . . . . . . . . . . . . . 136

7-15 Impedance plots of 1 : 1 transformer with 30 ”m 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

Page 12: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

Page 13: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

Page 14: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

Page 15: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

Page 16: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

Page 17: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

Page 18: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

Page 19: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

Page 20: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

+−

+ 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

Page 21: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

Page 22: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

Page 23: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

Page 24: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

Page 25: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

Page 26: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

Page 27: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

Page 28: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

fabricated in standard 0.13 ”m 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 ”m toroidal inductor on a silicon

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

40 ”m-thick copper traces wound around a 30 ”m-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 ”r = 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 ”m × 370 ”m and was comprised of 4 turns of 2.8 ”m-thick sputter deposited

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

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

28

Page 29: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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 ”m × 3466 ”m in area and featured

50 ”m-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 ”m FeCoBN and 0.4 ”m 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 − 1000C to a final thickness

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

29

Page 30: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

windings were then electroplated 50 ”m-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 ”m. Characterization of the inductor

revealed an inductance of 1.4 ”H 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 ”m-thick

(Fe97.6Hf2.4)90N10 magnetic and 500 A-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

Page 31: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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 ”m thick

and patterned 100 ”m wide with 10 ”m 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 ”m-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

Page 32: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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 ”m-thick

electroplated Ni81Fe19 magnetic core. The copper coils were electroplated 43 ”m 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 ”H was measured to be constant

up to 5 MHz. Leakage inductance was determined at 0.4 ”H 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

Page 33: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

Page 34: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

Page 35: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

Page 36: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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 =ℑZℜ 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

Page 37: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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− V

I

)

/j2(

VI+ V

I

)

/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

Page 38: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

Page 39: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

Page 40: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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 =K1”0n

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 =”n2davgc12

[

ln

(

c2

ρ

)

+ c3ρ+ c4ρ2

]

, (3–10)

40

Page 41: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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 ”m [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

Page 42: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

Page 43: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

Page 44: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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 ”m, 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 ”m (Figure 3-3A) and D = 1000 ”m (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 ”m, 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 ”m, the trace width was fixed at w = 50 ”m, 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

Page 45: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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/R

dc (

nH/Ω

)

20 ”m30 ”m40 ”m50 ”m60 ”m70 ”m80 ”m

Width w

A Outer diameter D = 500 ”m

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/R

dc (

nH/Ω

)

20 ”m30 ”m40 ”m50 ”m60 ”m70 ”m80 ”m

Width w

B Outer diameter D = 1000 ”m

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

45

Page 46: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

0.5 1 1.5 2 2.520

40

60

80

100

120

140

Outer Diameter D (mm)

L dc/R

dc (

nH/Ω

)

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

10−1

100

101

102

103

80

100

120

140

160

Interlayer spacing (”m)

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

Page 47: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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 ”m. 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 ”m 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

Page 48: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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 =Ç«A

g, (3–18)

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

capacitive coupling are the permittivity Ç« 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 ”m 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

Page 49: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

Page 50: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

107

108

109

10−9

10−8

10−7

10−6

10−5

Frequency (Hz)

Indu

ctan

ce (

nH)

R

c=0 Ω

Rc=10 Ω

Rc=100 Ω

Rc=1000 Ω

A Effective Inductance

107

108

109

10−2

100

102

104

106

Frequency (Hz)

Res

ista

nce

(Ω)

R

c=0 Ω

Rc=10 Ω

Rc=100 Ω

Rc=1000 Ω

B Effective Resistance

107

108

109

0

20

40

60

80

100

120

Frequency (Hz)

Qua

lity

Fac

tor

R

c=0 Ω

Rc=10 Ω

Rc=100 Ω

Rc=1000 Ω

C Effective Quality Factor

Figure 3-8. Modeled effect of substrate resistance on overall inductor impedance.Impedance calculated from circuit model for inductor with dc inductanceLdc = 100 nH and dc resistance Rdc = 1 Ω. Substrate resistance Rc varied inseries with substrate capacitance Cs = 1 pF in circuit model.

50

Page 51: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

When the capacitance through the substrate is the dominant capacitance

contributing to resonance with the inductor, the substrate resistance causes a damping

of the resonant behavior. To illustrate the effect of this damping, the circuit model of

Figure 3-7 was simulated with values of inductance, resistance, and capacitance that

would be typical of microinductors fabricated according to the methods of this work.

The effective inductance, resistance, and quality factor looking into the lumped inductor

circuit were extracted from the modeled data and plotted in Figure 3-8 for different

values of substrate resistance. The plots show that, compared to the case of zero

substrate resistance, increasing substrate resistance results in a lower peak inductance

near resonance and a lower frequency point at which the effective resistance raises

above its dc value due to resonance. The overall effect is a smoothing of the resonant

peaks and a decrease in the quality factor of the inductor at higher frequencies.

Cs

Rdc

Ldc

Rc

CL

Figure 3-9. Circuit diagram of inductor model with dc inductance Ldc , series resistancethrough the inductor Rdc , shunt capacitance between windings of CL, shuntcapacitance to the substrate of Cs , and resistance along the capacitive paththrough the substrate of Rc .

At even higher values of substrate resistance, this simple model (including only

capacitance with the substrate) shows that the resonant behavior would become ever

increasingly damped. With more and more damping, the inductance would remain ever

flatter with frequency, and the rise in effective resistance would be pushed out to greater

frequencies. As this resonance is damped out with very high substrate resistance,

the quality factor would improve again as the inductor would behave more ideally (i.e.

without capacitance). In practice, however, as the substrate capacitance is damped

51

Page 52: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

107

108

109

10−9

10−8

10−7

10−6

10−5

Frequency (Hz)

Indu

ctan

ce (

nH)

CL=100 pF

CL=10 pF

CL=1 pF

CL=0.1 pF

A Effective Inductance

107

108

109

10−4

10−2

100

102

104

Frequency (Hz)

Res

ista

nce

(Ω)

CL=100 pF

CL=10 pF

CL=1 pF

CL=0.1 pF

B Effective Resistance

107

108

109

0

20

40

60

80

100

120

Frequency (Hz)

Qua

lity

Fac

tor

C

L=100 pF

CL=10 pF

CL=1 pF

CL=0.1 pF

C Effective Quality Factor

Figure 3-10. Modeled effect of competition between winding capacitance and substratecapacitance on overall inductor impedance. Impedance calculated fromcircuit model for inductor with dc inductance Ldc = 100 nH and dcresistance Rdc = 1 Ω. Winding capacitance CL varied with no seriesresistance while substrate capacitance Cs = 1 pF had series resistanceRc = 1000 Ω in circuit model.

52

Page 53: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

out, the intrinsic interwinding capacitance begins to dominate. To more accurately

model the behavior of the microinductors the circuit must include the interwinding

capacitance as drawn in Figure 3-9, with the interwinding capacitance CL placed in

parallel with the capacitance through the substrate. Unlike the capacitance through

the substrate, the interwinding capacitance is relatively lossless, assuming only air or

a low-loss dielectric between the traces. If the value of the interwinding capacitance

is of similar value or larger than the capacitance to the substrate, the interwinding

capacitance will dominate the overall resonant behavior of the inductor. The circuit of

Figure 3-9 with both interwinding and substrate capacitance was simulated to show the

competition of the two capacitances. The simulated impedance looking into an inductor

with typical parameters was plotted in Figure 3-10 for various values of interwinding

capacitance. As shown in the plots, the resonance showed sharper peaks in effective

inductance and resistance when the interwinding capacitance was greater than the

substrate capacitance, due to the absence of resistance in series with the interwinding

capacitance. The quality factor, however, is lowered even further by interwinding

capacitance compared to the case of assuming only substrate capacitance. This is

due to the resonant frequency of the inductor being lowered from the case in which

resonance is governed by the substrate capacitance.

3.6.2 Eddy Currents

Eddy currents disrupt the flow of current through current through inductor traces

and result from the interaction of a conductor with the time-varying magnetic field of high

frequency electrical currents. When an electric current is passed through any conductor,

a corresponding magnetic field is generated in the space surrounding the current.

When the direction of current flow alternates (ac excitation), so too does the polarity of

the associated magnetic field. According to Lenz’s law electrical currents are induced

in nearby conductors so as to oppose the incident alternating magnetic field. These

induced currents are called eddy currents, and these confine the current to flowing only

53

Page 54: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

through a reduced cross-section of the inductor traces. The effects of eddy currents

have commonly been separated into the effects of an ac current to its own distribution

(skin effect) and the effects to other nearby current paths (proximity effect).

The skin effect refers to the tendency of a high-frequency ac current to confine itself

only along the surface of a conductor rather than flowing evenly throughout the cross

section of the conductor. This effect is the result of eddy currents generated within the

conductor of the ac current itself. An oft-cited parameter related to this effect is the skin

depth, which refers to the distance from the surface of a conductor at which the current

density is reduced to 1/e ≈ 0.37 of its nominal value at the surface. For good conductors

(σ/ωǫ ≫ 1) the skin depth is approximately given by

ή ≈√

2

ω”σ, (3–19)

where σ is the conductivity and ” is the magnetic permeability of the conductor material

and ω is the angular frequency of the ac current.

However, the approximation above is only applicable in certain cases, such as that

of an electromagnetic wave incident on an infinite slab of conductor or of current through

a conductor with circular cross section. However, the traces of the microinductors in this

and most contemporary works are of rectangular cross section. To illustrate the effect

of shape on the skin effect, COMSOL simulations were performed to plot the current

density and magnetic fields across the cross sections of two copper windings with the

same area but different aspect ratios as in Figure 3-11. The trace with closer to 1 : 1

aspect ratio exhibited a more uniform distribution of current density around its perimeter,

while the flatter trace exhibited greater current crowding along its shorter edges. For the

same level voltage excitation, the maximum current density in the 50 ”m× 10 ”m trace is

15% greater than that of the 25 ”m × 20 ”m trace. The minimum current density is also

15% less in the thinner trace compared to the thicker one.

54

Page 55: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

A Copper trace: 50 ”m× 10 ”m

B Copper trace: 25 ”m× 20 ”m

Figure 3-11. COMSOL simulations at 100 MHz of current density and magnetic field incross sectional view of copper windings having the same area but differentaspect ratios.

55

Page 56: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

Regarding the general question as to which surfaces of a conductor would carry

high frequency current, Wheeler [54] answered with a simple rule:

“The rule is, that the current follows the path of least impedance.”

This simple rule aids in the understanding of the segregation of current along the short

sides of a rectangular conductor. At dc, impedance is purely resistive and current flows

uniformly through a straight conductor. With increasing ac frequency, a conductor’s

impedance is increasingly dominated by an inductive component. As Wheeler [54]

observed, the answer as to determining the distribution of high frequency current then

becomes one of finding the path of least inductance. In a flat rectangular conductor, the

regions of greatest current density at the short ends (see Figure 3-11A) can be thought

of as separate parallel wires with current flowing in the same direction. Inductance

is then minimized when the parallel current paths are farthest separated and their

magnetic fields cancel in the interior of the conductor.

The relationship between inductance and resistance at high frequencies is evident

in how these values change in an inductor as functions of frequency. Wheeler [54]

discussed an “incremental-inductance rule” by which the effective resistance of

conductors could be calculated as equal to the change in reactance resulting from eddy

currents in certain cases. Although the premise of this rule is invalid for the traces with

rectangularly-shaped cross sections [55], a similar result was found in the measurement

of the microinductors. After measuring the frequency-dependent resistance R (f )

and reactance ωL (f ) of an example inductor, the deviations in resistance (∆R) and

reactance (∆X ) from their values at lower frequencies (Rdc and ωLdc , respectively)

were plotted in Figure 3-12. While the large increases in both resistance and reactance

in the plot past 200 MHz are due to self-resonance, the deviations in resistance and

reactance from their lower-frequency values are roughly equal and opposite from about

20–200 MHz as indicated by the plot of ∆R + ∆X remaining relatively flat.

56

Page 57: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

107

108

109

−10

−5

0

5

10

Frequency (Hz)

Impe

danc

e (Ω

)

∆X=ωL(f)−ωLdc

∆R=R(f)−Rdc

∆R+∆X

Figure 3-12. Measured effect of eddy currents on inductor impedance. Plot shows themeasured deviations in resistance (∆R) and reactance (∆X ) from theirvalues at lower frequencies (Rdc and ωLdc , respectively).

3.7 Summary of Inductor Design

The preceding chapter outlined a methodology for designing air-core inductors for

switched mode power converters.

‱ Quality factor was shown to be a metric of how much energy is stored in aninductor per cycle compared to how much is dissipated by it.

‱ The three complementary effects of inductance, resistance, and parasiticcapacitance were discussed for their roles in affecting inductor quality factor,which motivated a stacked planar spiral design that balanced the three effects.

‱ An analytical model was presented with the goal of predicting inductor performanceat low frequencies.

‱ The model was analyzed to uncover trends affecting the inductance to resistanceratio when varying geometric parameters. An optimal geometry was found with atrace width of 50 ”m and a packing density of 40%.

‱ A circuit model for capacitive coupling between the inductor and substrate revealedthat increases in substrate resistance would result in a dampening of the resonantbehavior of the inductor and a reduction in the peak quality factor.

57

Page 58: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

‱ Eddy currents were discussed as leading to skin and proximity effects thatwould increase the resistance of inductor windings at higher frequencies as thecurrent would seek to flow through a path of minimum impedance (i.e. minimuminductance at high frequencies).

58

Page 59: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

CHAPTER 4TRANSFORMER DESIGN

This chapter discusses the considerations that shape the design of transformers.

The similarities and differences between the designs of inductors and transformers

are first established. Maximum transducer efficiency is introduced as a metric for

transformer optimizing transformer performance. In an effort to utilize lessons learned

from the inductor design, a high-level, abstracted look at energy flowing through a

transformer shows how the quality factors of individual inductors can provide insights

into the maximum efficiency of a transformer. From these insights, a winding layout

scheme for the transformers is selected that allows both high performance and

step-up/down opportunities. A circuit model for the transformers is then presented

to explain the frequency behavior of the devices. Finally, network analysis is used to

derive expressions for the load-dependence of the transformer efficiency and voltage

gain.

4.1 Overview and Goals

The transformers presented in this work consist essentially of a pair of coiled

inductors so positioned that their magnetic fluxes are linked and energy can be

transferred from one to another. While it is possible to create transformers incorporating

more than two coils for more complex power distribution, the analyses in this chapter

are limited to two-coil devices. The two coils are referred to as primary and secondary;

the power source is connected to the primary coil and power is transferred to the load

connected to the secondary coil.

Just as the quality factor is important in establishing the efficiency of magnetic

energy storage in an inductor, so too does the magnetic energy transfer in a transformer

rely on currents flowing through its windings, ideally with as little resistive loss as

possible for high efficiency. However, due to imperfect magnetic flux coupling between

the coils in a transformer, some energy is stored in the individual windings and not

59

Page 60: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

transferred. At high frequencies a significant proportion of energy is also capacitively

stored in the electric field between coils. The interactions between these effects is

further complicated by the load impedance, which can alter the phase at different

parts of the transformer to help or harm efficiency. The primary goal of the design is to

maximize the efficiency of power transfer through the transformer.

4.2 Maximum Efficiency

As two port devices, the transformers require a different metric for efficiency

than the quality factor used for inductors. Further, the efficiency of the transformer

will depend on the load that is attached to it. For the following analyses efficiency is

defined as the ratio of real power output to the load versus the real power input to the

transformer. One useful case is that of the maximum possible efficiency assuming

conjugate-impedance matched loading.

4.2.1 From Scattering Parameters

In the realm of radio frequency (RF) systems, efficiency is referred to as power

gain, the ratio of output to input power. Characterization of RF devices typically

entails determination of scattering parameters, which are obtained by measurement

of sinusoidal signals incident on, reflected from, and transmitted through the device to

be tested [56]. The maximum power gain (efficiency) that can be attained from a generic

two-port transducer (e.g. a transformer) is obtained for the case of conjugate-impedance

matched loading and can be calculated from the scattering parameters [57],

Gmax =|S21||S12|

(

K −√K 2 − 1

)

, (4–1)

where K is the Rollet Stability Condition, which is defined as

K =1− |S11|2 − |S22|2 + |S11S22 − S12S21|2

2 |S12S21|. (4–2)

However, this complicated expression for efficiency does not lend itself to being easily

interpreted to aid in the design of transformers.

60

Page 61: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

4.2.2 From Coil Quality Factors and Coupling Coefficient

A simplified analysis may be performed for the case of conjugate-impedance

matched loading, which enables the derivation of maximum transformer efficiency in

terms of targetable design variables, namely coil quality factor and coupling coefficient

between primary and secondary coils. Under matched conditions, no power is reflected

to the source. When a certain energy Ein is input to the primary transformer coil some

of that energy is stored in the magnetic field around the primary coil and some Ed1 is

dissipated by it. The magnetically stored energy is comprised of both that energy Em

which is mutually shared between the primary and secondary and that energy Es1 which

is not coupled with the secondary. Written algebraically, the sum of primary coil energies

is

Ein = Ed1 + Es1 + Em. (4–3)

Of the energy Em that is coupled to the secondary coil, some energy some Ed2 is

dissipated, while some Es2 is stored solely in the magnetic field of the secondary. Finally,

the remaining energy Eout is output to the load. The sum of secondary coil energies is

Em = Ed2 + Es2 + Eout . (4–4)

This energy flow is shown schematically in Figure 4-1.

Ed1 Ed2

Es1 Es2

Ein Eout

Primary Secondary

Em

Figure 4-1. Diagram of energy flowing into primary transformer coil. Some of that energyis stored or dissipated, and the rest is transferred to the secondary coil.Some of that energy is stored or dissipated, and the rest is output to theload.

61

Page 62: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

The coupling coefficient k specifies the fraction of the total magnetically-stored

energy that is transferred between the primary and secondary coils. In the primary coil,

the magnetically-stored energy is that which is input minus that which is dissipated by

the primary, so that the coupling coefficient can be written as

k =Em

Ein − Ed1=

Em

Es1 + Em. (4–5)

The same coupling coefficient can be equivalently written in terms of the secondary

coil energies. In the secondary coil, the total magnetically-stored energy is that which

is transferred from the primary plus that which is stored solely in the secondary. The

coupling coefficient can therefore be written as

k =Em

Es2 + Em. (4–6)

The quality factors of each coil are the other design variables of interest and

represent the total energy stored in the magnetic field of each coil—both coupled (Em)

and uncoupled (Es )—to that dissipated by it. For the primary coil the quality factor is

Q1 =Es1 + EmEd1

, (4–7)

and for the secondary the quality factor is similarly

Q2 =Es2 + EmEd2

. (4–8)

If Equations 4–7 and 4–8 are multiplied with Equations 4–5 and 4–6, the resulting

expressions are simply

kQ1 =Em

Ed1(4–9)

and

kQ2 =Em

Ed2. (4–10)

Finally, the above Equations 4–9 and 4–10 can be combined to derive a simple

expression for the maximum transformer efficiency. In terms of the energy variables,

62

Page 63: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

the overall efficiency η of the transformer is defined as the ratio of useful energy of the

secondary, that is, the output energy plus the secondary stored energy, over the useful

energy of the primary, that is, the input energy minus the primary stored energy,

η =Eout + Es2Ein − Es1

. (4–11)

From the sums at the primary (Equation 4–3) and secondary (Equation 4–4) energy

nodes, Equation 4–11 can be rewritten as

η =Em − Ed2Em + Ed1

=1− Ed2/Em1 + Ed1/Em

. (4–12)

Substituting the ratios of dissipated to transferred energy in the above equation with

expressions of coupling coefficient and quality factor of Equations 4–9 and 4–10,

Equation 4–12 becomes

η =1− 1

kQ2

1 + 1

kQ1

=k − 1/Q2k + 1/Q1

. (4–13)

This final expression calculates the maximum transformer efficiency given only

the individual coil quality factors and the coupling between coils. To validate this result,

Equation 4–13 and the classic expression for Gmax given by Equation 4–1 were both

used to calculate efficiency using the same measured data. Both calculations were

plotted as in Figure 4-2 and displayed excellent agreement at frequencies well below the

first resonant frequency of the transformer. Near the resonant frequency (> 200 MHz

in Figure 4-2 for example), the efficiency calculated as Gmax was much greater due to

its accounting for the capacitive energy storage and adjusting the load accordingly,

whereas the simple expression of Equation 4–13 was derived assuming no capacitive

storage.

Equation 4–13 reveals several insights that are useful for designing transformers for

optimal efficiency. The first insight is that the importance of coupling between primary

and secondary diminishes with increasing quality factor of each coil.

63

Page 64: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

107

108

0

20

40

60

80

100

Frequency (Hz)

Effi

cien

cy (

%)

G

max(S

11,S

12,S

21,S

22)

η(Q1,Q

2,k)

Figure 4-2. Transformer efficiency calculated by both Q factors and S parameters andmeasured data taken from an example inductor. Excellent agreement wasobtained up to 200 MHz, at which point the transformer approachedresonance and capacitive storage dominated.

The second insight is that, because the quality factors of Equation 4–13 are

dependent on frequency, the primary and secondary coils should attain their highest

quality factors at the same frequencies to obtain the highest overall transformer

efficiency. Isolation transformers with 1 : 1 turns ratios usually have nearly identical

primary and secondary coils and thus easily satisfy the requirement for matching

frequency behavior. Transformers with non-unity turns ratios, however, are comprised

by necessity of mismatching coils with unequal inductances. The coil with lesser

inductance is often physically smaller with less parasitic capacitance and higher self

resonant frequency than the coil with greater inductance. The smaller coil is therefore

likely to attain higher quality factor at higher frequency than the larger coil. This issue

limits the efficiency of high frequency transformers with large turns ratios.

64

Page 65: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

4.3 Layout

The layout for the transformers of this work employed a hybrid combination of two

winding techniques: interleaving and nesting. Utilizing the same analytical expressions

from the inductor design (Section 3.4), the primary coil was first laid out according to

the required specification, except with extra space provided between turns. Within this

space, the secondary coil was interleaved as an exact copy of the primary but rotated

180. The resulting layout would be that of a 1 : 1 transformer. In order to achieve

voltage/current gain, additional secondary turns were then nested within the space

cleared by the primary coil. An example layout of a microtransformer is depicted in

Figure 4-3.

A Lower winding layer B Upper winding layer

Primary Winding

Secondary Winding

Via

C Key

Figure 4-3. Diagrams illustrating transformer winding layout on lower and upper windinglayers. Key identifies windings belonging to primary and secondary coils andlocation of vias.

4.3.1 Turns Ratio

The turns ratio of a transformer is roughly a measure of its voltage or current gain.

This ratio has been traditionally useful for transformers with such high permeability

magnetic cores that the magnetic flux induced by wires wrapped around the core are

essentially fully contained within the core. For such transformers with nearly perfect

magnetic coupling, the magnetic flux induced by the primary coil is fully sensed by the

65

Page 66: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

secondary coil. The gain is then equal to the primary to secondary ratio of the number of

physical turns of wire around the core, as suggested by the term “turns ratio.”

However, microscale transformers with no or low permeability magnetic cores do

not exhibit perfect coupling. Furthermore, not all turns of the microtransformer coils have

equal contributions to the inductance due to the difference in areas enclosed by inner

and outer loops. The turns ratio of a 1 : n microtransformer is instead calculated by

n =

√

L2

L1, (4–14)

where L1 is the primary inductance that would be obtained if the secondary were

open-circuited and L2 is the secondary inductance that would be obtained if the primary

were open-circuited.

4.4 Performance Under Load

The Vector Network Analyzer (VNA) was used for characterizing microtransformers.

Because the VNA measurement is typically performed with 50 Ω loading, the results

must be re-interpreted to derive performance for other loading conditions.

Although Equation 4–1 can be used to quickly calculate the maximum attainable

transformer efficiency from measured scattering parameters, it has several shortcomings:

it assumes matched loading at every frequency point but does not actually indicate

the matched load impedance required, it does not provide information about how

big a performance hit is suffered if the frequency or load deviates from their optimal

intersections, and it does not provide the associated voltage gain. For this reason,

a set of expressions was derived that could provide more detailed information about

load-dependent performance.

4.4.1 Derivation of Efficiency and Voltage Gain for Arbitrar y Load

Modified transmission parameters Aâ€ČB â€ČC â€ČD â€Č were found that represented not true

transmission parameters but instead provided relationships between the input and

output voltages and currents. These modified parameters were calculated in terms of

66

Page 67: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

the original ABCD transmission parameters of the transformer and the load impedance

ZL. The ABCD parameters could be derived either from modeling or from measurement.

The voltage across the load and the current through it were represented by V â€Č

2

and I â€Č2, respectively. These quantities were found separately through network analyses

of two cascaded networks representing the same situation of a load attached to the

transformer. Relationships of each of these quantities to the input voltage V1 and input

current I1 were defined as

Aâ€Č = V1V â€Č

2

B â€Č = −V1I â€Č2

C â€Č = I1V â€Č

2

D â€Č = − I1I â€Č2

. (4–15)

Load-conditioned parameters Aâ€Č and C â€Č were found by cascading the transformer

network with a shunt load and leaving the output as an open circuit,

A BC D ZL

−I2 = 0

V â€Č

2

I1

V1

Figure 4-4. Circuit diagram of two-port transformer transmission (ABCD) networkcascaded with shunt load.

The resulting network with shunt load was

Aâ€Č B

C â€Č D

=

A B

C D

·

1 0

1/ZL 1

=

A+ B/ZL B

C + D/ZL D

. (4–16)

Load-conditioned parameters B â€Č and D â€Č were found by cascading the transformer

with a series network representing the load and then shorting the output, as shown in

the figure below.

The resulting network with parallel load network was

A B â€Č

C D â€Č

=

A B

C D

·

1 ZL

0 1

=

A AZL + B

C CZL +D

. (4–17)

67

Page 68: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

A BC D

ZL−I â€Č2

V2 = 0

I1

V1

Figure 4-5. Circuit diagram of two-port transformer transmission (ABCD) networkcascaded with series load.

The modified parameters taken from each of the two cascaded networks together

formed the following set of relationships,

Aâ€Č = V1V â€Č

2

= A+ BZLB â€Č = −V1

I â€Č2

= AZL + B

C â€Č = I1V â€Č

2

= C + DZLD â€Č = − I1

I â€Č2

= CZL +D. (4–18)

Efficiency was defined as the ratio of the real power delivered from the transformer

to a load versus the real power delivered from a source to the transformer, that is,

η =Pload

Pin=

ℜ

−V â€Č

2Iâ€Č

2

ℜ

V1I1 (4–19)

By properties of complex numbers, the expression was simplified as

η =

(

−V â€Č

2Iâ€Č

2 − V â€Č

2Iâ€Č

2

)

/2(

V1I1 + V1I1)

/2=

1 +V â€Č

2

I â€Č2

I â€Č2

V â€Č

2

−V1V â€Č

2

I1

I â€Č2

− V1

I â€Č2

I1V â€Č

2

=1 + ZL/ZL

Aâ€ČD â€Č + B â€ČC â€Č

. (4–20)

This expression could also be written in terms of the original transformer ABCD

parameters and the load impedance,

η =ZL + ZL

(AZL + B) (CZL + D) + (AZL + B) (CZL + D)=

ℜZLℜ

(AZL + B) (CZL + D) (4–21)

The efficiency was thus found to be a function of the load impedance and operating

frequency.

68

Page 69: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

The voltage gain provided to the transformer to varied loads was also derived from

this analysis, simply as the inverse of Aâ€Č,

Av =|V â€Č

2||V1|

=

∣

∣

∣

∣

1

Aâ€Č

∣

∣

∣

∣

=

∣

∣

∣

∣

ZL

AZL + B

∣

∣

∣

∣

. (4–22)

4.4.2 Conjugate Impedance Matched Loading

The maximum efficiency through the transformer occurs when the source and load

impedances attached to the transformer are conjugate matched to its input and output

impedances, respectively [57]. A generic two-port ABCD network was analyzed to

determine the ZL that would result in conjugate impedance matching for a given ABCD.

As labelled in Figure 4-6, the impedances looking into the primary and secondary

coils of the transformer were denoted Z1 and Z2, respectively, and the source and load

impedances were denoted ZS and ZL.

A BC D ZL

ZLZ2Z1ZS

Figure 4-6. Circuit diagram of two-port transformer transmission (ABCD) network withsource and load impedances ZS and ZL, respectively.

Conjugate impedance matching requires that ZS and Z1 are complex conjugate

pairs, that is,

ZS = Z1. (4–23)

The same condition is required of ZL and Z2,

Z2 = ZL. (4–24)

From network theory, the impedance looking into the transformer input port is given by

Z1 =AZL + B

CZL + D, (4–25)

69

Page 70: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

and the impedance looking into the output port is

Z2 =DZS + B

CZS + A. (4–26)

From Equation 4–23 and Equation 4–25, the matched source impedance is written,

ZS =

(

AZL + B

CZL +D

)

(4–27)

The above expression for matched source impedance is then replaced into the

expression for output port impedance of Equation 4–26,

Z2 =D(

AZL+B

CZL+D

)

+ B

C(

AZL+B

CZL+D

)

+ A

= ZL, (4–28)

which is related to ZL from Equation 4–24. Equation 4–28 can then be algebraically

manipulated into the form,

(

AC + AC)

ZL2+

(

BC − BC + AD − AD)

ZL −(

BD + BD)

= 0, (4–29)

which is identified as a quadratic equation.

Solving the quadratic equation for ZL yields the load impedance required for the

conjugate matched impedance condition for maximum power gain. The value of Gmax , as

calculated from Equation 4–1, is identically equal to the result obtained when the value

of ZL (Equation 4–29) is replaced into Equation 4–21. However, the latter calculation

reveals the required matched load impedance, which is not otherwise known.

4.5 Summary of Transformer Design

The preceding chapter outlined considerations affecting the design of transformers

intended for switched mode power converters.

‱ Efficiency was forwarded as a metric for maximizing the power delivered to theload while minimizing that dissipated in the transformer.

70

Page 71: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

‱ A two-layer hybrid combination of both interleaved and nested primary andsecondary coils was presented as a layout that would provide both strong couplingand opportunities for non-unity voltage gain.

‱ Three methods of determining efficiency were discussed based on the informationrequired to calculate each: measured scattering parameters, measured ordesigned quality factors and coupling coefficient, and measured or modeledABCD parameters.

‱ ABCD analysis was further used to calculate efficiency for any arbitrary loadimpedance and to determine the corresponding voltage gain.

71

Page 72: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

CHAPTER 5FABRICATION

This chapter describes a multilevel wafer-level microfabrication process that was

specially developed as a means to realizing three-dimensional (3D) electroformed

copper components. The process was tailored to deliver the fine dimensions and

complex routing needed for microinductors and transformers with high performance in

integrated high frequency power converters. In this chapter, an overview first outlines

the fundamental steps at the heart of this microfabrication process. Several variations

to enable extended capabilities of the core process are then discussed. Details of

the major steps in the process are then provided for a deeper understanding of the

considerations motivating the selection various sequences and parameters. Scanning

electron microscope (SEM) images are frequently used in this chapter to depict features

of the microfabricated devices.

The 3D copper microfabrication process was developed in response to deficiencies

that have so far prevented air-core microinductors from being integrated with power

converters. This process was required to simultaneously achieve three goals: thick

copper windings, multilayer stacking of windings, and low capacitance between

windings. While magnetic materials have been used in other works to increase the

inductance through a given length of conductor, the air core spirals presented here

required a longer electrical path to achieve the same inductance. The length of the

coils could lead to a high series resistance. Thick copper was necessary in order to

minimize the electrical resistance through the inductors. Because the planar spiral

design occupied a large area compared to its thickness, multilayer stacking provided

the best magnetic coupling to increase inductance densities. Multilayer stacking also

enabled the complex routing schemes needed for transformers with strong magnetic

coupling between primary and secondary coils. Removal of dielectric from between

72

Page 73: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

adjacent traces reduced the capacitance that limited the upper operating frequency.

These design decisions have been covered in greater detail in Chapters 3 and 4.

5.1 Process Overview

The microfabrication process consisted principally of two stages: an additive

stage in which copper was electroplated layer-by-layer through patterned photoresist

molds and a following subtractive stage in which the molds were removed leaving a

freestanding 3D copper structure.

During the additive stage, thick copper traces were formed via a through-mold

electroplating technique. A thin copper seed layer was first deposited across the entire

surface of the wafer to serve as a conductive base onto which thicker copper would

be electroplated. A photoresist mold was then patterned on top of the seed, and the

thick layer of copper was electroplated through the mold. These mold-filling steps were

repeated as illustrated in Figure 5-1 for each layer of the device so that structures with

three-dimensional features may be obtained.

After all desired layers were added, the fabrication process entered the subtractive

stage, during which the photoresist molds and copper seeds were removed in one of two

ways.

5.1.1 Sequential Layer Removal

Thin (10 ”m) suspended features with greater widths than thicknesses were prone

to snapping down during the wet removal process, a problem known as stiction [58]. For

wafers having such structures, the photoresist molds and copper seed layers had to be

sequentially removed in photoresist developer and copper etchant, respectively. The

sequential removal enabled the patterning of selective portions of the photoresist mold

into insulating structural elements to prevent stiction between floating traces. Figure

5-2 illustrates the progression of releasing a suspended inductor by sequential layer

removal.

73

Page 74: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

A Layer 1 mold over copper seed B Layer 1 plated through mold.

C Layer 2 plated through mold. D Layer 3 plated through mold.

E Layer 4 plated through mold.

Figure 5-1. Illustrations of additive process stage.

A microinductor that was released by sequential layer removal with 10 ”m thick

traces and an outer diameter of 500 ”m is depicted in the scanning electron micrograph

(SEM) image of Figure 5-3. Both upper and lower winding layers are visible in this image

along with the scaffolding and support posts that aid in anchoring and propping up the

windings.

5.1.2 Ultrasonic Agitation in Solvents

If the electroplated coppers were sufficiently robust, removal both photoresist

and copper seed was accomplished through ultrasonic agitation in a photoresist

74

Page 75: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

A Layer 4 mold and seed removed. B Layer 3 mold and seed removed.

C Layer 2 mold and seed removed. D Layer 1 mold and seed removed.

Figure 5-2. Illustrations of subtractive process stage.

Figure 5-3. SEM image of microfabricated inductor with 10 ”m thick copper windinglayers with photoresist support posts between winding layers.

75

Page 76: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

Figure 5-4. SEM image of microfabricated inductor with 30 ”m thick copper windinglayers with upper winding layer held in place only by vias.

stripper such as acetone or a solution containing n−methyl− 2− pyrrolidone (BAKER

PRS-3000).This simplified release method was appropriate for devices where any

suspended structures were at least 30 ”m thick and not considerably wider than thick.

Figure 5-4 shows a 600 ”m microinductor with 30 ”m thick layers, the molding layers

of which were removed by ultrasonic agitation in BAKER PRS-3000. No photoresist

support posts were required for this inductor, as the thicker traces provided ample

mechanical support to resist stiction.

5.2 Features and Variations on the Process

5.2.1 Planar Processing

The sequence of steps in this process was devised so that a flat, planar surface

would be maintained throughout the fabrication of the devices for compatibility with

planar microfabrication techniques. A consequence of the through-mold electroplating

technique was that the conductive copper seed in regions between electroplated

traces of each layer was covered by the mold. The copper seed electrically shorted

76

Page 77: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

all traces together and had to be removed, which in turn necessitated removal of the

mold to gain access. Removing the mold and seed immediately after electroplating

would have resulted in a non-flat surface topography with recessed regions where

the mold once existed between thicker electroplated regions. This topography would

create problems when building multilayer structures: most photoresists need to be spun

on top of a flat surface with topographical disturbances much less than the desired

thickness of the photoresist layer in order to obtain a uniform photoresist thickness.

A non-uniform photoresist thickness would not develop properly since thicker regions

would underdevelop, while thinner regions will overdevelop. Also, the topography of the

surface would not allow the regions of photoresist to be in contact with the mask in the

case of contact mask lithography or would lead to regions that are exposed out of focus

in the case of projection lithography. In either case the result would be poor resolution

with diffraction of light around the areas to be exposed.

Various works have presented options for creating a planar surface above

electroplated features. Some have utilized photosensitive benzocyclobutene (BCB)

applied over the electroplated features after removal of the mold [59, 60]. BCB exhibits

redistribution of material to fill gaps between underlying features to form a planar surface

while curing [61]. Another option that has been used extensively in industry for surface

planarization while processing the copper metal layers of microprocessors is chemical

mechanical polishing (CMP).

The fabrication method described in this chapter on the other hand maintained

a planar processing surface by leaving the copper seed layers in place throughout

the additive stage and finally removing each of the seed layers during the subtractive

stage. Additionally, because the electroplating steps were timed so that the deposited

copper filled to the height of the surrounding mold, a planar surface was maintained after

electroplating each layer without the need for any reflow or CMP planarizing steps.

77

Page 78: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

5.2.2 Photoresist as a Structural Element

One capability that differentiated this process from other multilevel metallization

processes was that portions of the photoresist mold could be left as structural

elements to provide mechanical support to the molded copper parts. The ability

to form photoresist structural elements was devised in response to finding that the

wet processing steps for the removal of the moldings caused bending and binding

of device elements to each other, a problem known as stiction [58]. In the case of

the microfabricated inductors and transformers this stiction led to electrical shorting

of windings that drastically lowered the performance of the devices. Because the

photoresist that formed the plating molds was a dielectric material with negligible

conductivity, it was usable to provide mechanical and electrical isolation between

windings.

The fabrication of the photoresist structural elements specifically required a

positive-tone photoresist, one in which exposure to ultraviolet light initiates a chemical

modification in the photoresist that makes it soluble (i.e. etchable) in a basic solution

(the developer). The capability also required use of the sequential removal method for

the subtractive stage. In this process each layer of photoresist was exposed twice. The

first exposure defined the mold. Regions of the photoresist were exposed and removed

in developer solution to form the desired mold. This photoresist mold was then exposed

a second time everywhere except the regions that would serve as structural elements.

However the photoresist was not immediately developed after the second exposure.

Instead, the process proceeded with the addition of more layers to the structure until all

parts of the device had been added. Then during the subtractive stage each mold was

removed in photoresist developer except those unexposed regions of photoresist that

became structural elements.

78

Page 79: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

5.2.3 Substrate Versatility

The multilevel copper microfabrication process was relatively insensitive to the

choice of substrate. Structures were formed on top of both silicon and Pyrex wafers

with minimal variation of the process steps required between the different substrates.

The conductive silicon wafers needed to be electrically isolated from the copper, which

accomplished by plasma-enhanced chemical vapor deposition (PECVD) of silicon

dioxide or silicon nitride dielectric layers over the surface of the silicon. In instances

where the inductors and transformers were intended to the silicon substrate throughout

testing, reactive ion etching (RIE) was used to selectively remove regions of the

dielectric layer to electrically ground the silicon substrate with the ground components of

the copper layers. In all cases, a thin layer of titanium was sputter-deposited on top of

the silicon dioxide, silicon nitride, or Pyrex to improve the adhesion of the copper parts to

these surfaces.

The dielectric layer was also used in other instances as a sacrificial material that

allowed detachment of multilevel copper parts from the fabrication wafer. As described

in greater detail in Section 8.2, integrated power converter modules were formed by

encapsulating microfabricated inductors and a multilevel routing and interconnect

framework in epoxy. The encapsulated copper modules were then released from a

silicon substrate by etching in concentrated 49% hydrofluoric acid the layer of silicon

dioxide that insulated the two from each other. An etching time of several hours was

required for the hydrofluoric acid to fully undercut the silicon dioxide from beneath the

3 mm× 3 mm modules.

5.3 Process Steps

The first steps of the microfabrication process concerned the adhesion and

insulation between the substrate and the multilevel copper structures to be fabricated.

For silicon substrates a dielectric layer of silicon dioxide or silicon nitride was first

deposited over the blank wafer by plasma-enhanced chemical vapor deposition

79

Page 80: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

(PECVD). The dielectric layer provided electrical insulation between the copper and

the conductive silicon substrate or in some cases was used as a sacrificial layer to

physically detach copper parts from the wafer after fabrication. The thickness of the

dielectric layer was varied from 200 nm up to 2 ”m depending on the intended purpose.

When the dielectric layer was used as a sacrificial material the thickest deposition was

used. When used only for electrical insulation, openings were formed in the dielectric

layer so that the the ground nodes of the copper were directly in contact with and would

ground the silicon substrate. To form the openings, resist was photolithographically

patterned on top of the dielectric and the exposed dielectric was removed by reactive ion

etching (RIE). The resist that remained after the RIE was stripped in oxygen plasma.

The wafer—either fresh Pyrex or dielectric coated silicon—then entered the sputter

tool for cleaning and deposition of the starting seed layer. Because the sputter tool had

multiple chambers, the surface of the wafer was able to be cleaned first with a short

bombardment etch of radio-frequency (RF) excited argon ions. While remaining under

vacuum to prevent any contamination or oxidation, the wafer was transferred to another

chamber where a 50 nm thin layer of titanium was sputter deposited across the full

surface of the wafer to provide improved adhesion between the Pyrex or dielectric layer

and the copper structures. The wafer was transferred to a third chamber where a 200 nm

thick copper seed layer was sputter deposited. On subsequent layers, only the RF etch

and copper deposition were used, as titanium was only utilized for the initial deposition

on the wafer.

Except for the first layer that additionally required titanium deposition before copper,

each layer of the multilevel copper structure was added by repetition of a basic set of

steps. The steps were:

1. In-situ argon sputter etch.2. Sputter deposition of 200 nm copper seed layer.3. Coating of AZ 9245 positive tone photoresist and spinning to targeted thickness.4. Heating wafer on hotplate to drive out solvent from photoresist.

80

Page 81: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

Table 5-1. Process parameters for passives fabrication at U.S. Army ResearchLaboratory

Step Tool Used DetailsDeposit dielectric LAM 790 SiO2 or SiN by timed deposition.Argon sputter etch Metron 3290 50 W, 40 s.Sputter Ti Metron 3290 300 W, 30 s.Sputter Cu Metron 3290 1.18 kW, 40 s.Spin photoresist SUSS MicroTec ACS200 AZ 9245 photoresist.Softbake Hotplate 95 C.Photoresist exposure SUSS MicroTec MA6 20 mW/cm2.Plasma descum Metroline M4L 200 sccm O2, 400 W, 30 s.Electroplate Cu Dynatronix DuPR 10-3-6 Timed, direct current.Drying bake Hotplate 95 C, 5 min.Skin removal Metroline M4L 200/20 sccm O2/CF4, 250 W, 5 min.

5. Alignment and contacting of mask to wafer and ultraviolet (UV) exposure ofphotoresist.

6. Development of photoresist.7. De-scum etch of residue out from trenches in photoresist mold.8. Optional second UV exposure of photoresist.9. Electrodeposition of copper.10. Heating wafer on hotplate to dry.

The second UV exposure was only used when a sequential layer removal was

required as described in Section 5.1. The tools and parameters used for each of these

steps are listed in Table 5-1. More processing parameters are discussed in Section

5.4 as some were dependent on the choice of layer thickness, which was tested in

thicknesses of 10 ”m and 30 ”m. Cross-section diagrams in Figure 5-5 illustrate how

each step of the additive process stage contributed to the building up of the multilevel

stack. After completion of the additive stage, the process entered the subtractive stage,

during which the molding was removed by either ultrasonic agitation in a solvent or

by the sequential layer removal process. Structures with thin copper layers required

sequential removal of the molding to enable parts of the photoresist mold to be used

as structural elements to prevent parts from snapping together. During the sequential

removal, the following steps were repeated until the entire molded copper structure was

released from the molding:

81

Page 82: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

1. Develop photoresist.2. Copper etch.3. Skin removal.

Cross-section diagrams in Figure 5-6 illustrate how each step of the sequential removal

contributed to releasing the copper structure from its mold.

5.4 Special Processing Considerations

5.4.1 Sputtering

Sputter deposition of copper was required in the process to form the conductive

seed layers onto which the thick copper traces of the device would be electroplated. The

only other metal used in this process was a thin layer of titanium deposited at the start

of the process to aid in adhesion of the copper device to the substrate, which was either

Pyrex or nitride- or oxide-coated silicon. The titanium was dc magnetron sputtered at

300 W for 30 s for a thickness of roughly 50 nm over the surface of the wafer. Copper

was then sputtered on top of the titanium adhesion layer to provide the seed for copper

electroplating.

Forming subsequent layers in the process required deposition of copper onto

a surface consisting of both photoresist and electroplated copper. Obtaining good

coverage of this surface by the copper seed presented several challenges: the sputtered

copper did not always adhere well to either the photoresist or the electroplated copper,

and the film often broke along the boundaries between the electroplated copper and the

photoresist.

Adhesion of the sputtered copper film on top of electroplated copper was improved

by argon sputter etching of the electroplated copper surface immediately prior to

sputter deposition (see Section 5.4.4 for discussion). Good adhesion of the sputtered

copper film to photoresist was achieved with high power dc magnetron sputtering at

1.18 kW. In Figure 5-7, a copper trace was peeled back and flipped over, revealing

underlying photoresist posts that remained attached to the trace. The fact that each

photoresist post remained attached to the trace above it—rather than to the trace below

82

Page 83: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

A Sputter deposit Ti and Cu. B Deposit, pattern, and descum photoresist.

C Flood expose photoresist. D Electroplate Cu and dry bake.

E Sputter etch surface and deposit Cu. F Deposit, pattern, and descum photoresist.

G Flood expose photoresist. H Electroplate Cu and dry bake.

I Sputter etch surface and deposit Cu. J Deposit, pattern, and descum photoresist.

K Expose photoresist with pattern. L Electroplate Cu and dry bake.

M Sputter etch surface and deposit Cu. N Deposit, pattern, and descum photoresist.

O Flood expose photoresist. P Electroplate Cu and dry bake.

Figure 5-5. Cross section diagrams of the additive process stage.

83

Page 84: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

A Develop photoresist. B Copper etch and O2/CF4 plasma etch.

C Develop photoresist. D Copper etch and O2/CF4 plasma etch.

E Develop photoresist. F Copper etch and O2/CF4 plasma etch.

G Develop photoresist. H Copper etch.

I Titanium etch.

Figure 5-6. Cross section diagrams of the subtractive process stage.

it—indicated that adhesion of the sputtered copper onto the photoresist was better than

that of the photoresist onto the copper.

Sputtering copper at lower powers resulted in delamination of the film from the

photoresist. This problem was evident when photoresist was cured on top of the

sputtered copper film. During solvent evaporation, tension from the resist would pull at

the copper film, detaching it from the photoresist and breaking the film at the boundaries

between the underlying photoresist and electroplated copper. Accurate mask alignment

was not possible due to the shifting film blocking the underlying features. During

84

Page 85: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

Figure 5-7. SEM of copper trace peeled back from inductor to reveal photoresist spacerblocks attached to copper. This indicated that adhesion of copper sputteredonto photoresist was better than that of photoresist spun onto copper.

development, the broken copper film would also allow photoresist developer to seep

down around and etch into lower photoresist layers, causing a widening of the mold and

electroplating of copper between traces as shown in Figure 5-8A.

At a sputtering power of 1.18 kW no breaking or shifting of the copper film was

observed and electroplating was well-confined to the molds, yielding good separation

of features such as those shown for comparison in Figure 5-8B. At sputtering powers

greater than 1.18 kW, excessive heating of the wafer caused bubbling of the photoresist.

5.4.2 Photolithography

AZ 9245 positive tone photoresist was used for its ability to form thick layers and

compatibility with the copper electroplating bath. Two separate sets of photolithography

parameters enabled the resist to be spun to a thickness of either 10 ”m or 30 ”m per

layer. In either case, the photolithography steps consisted of the following:

1. Photoresist was dispensed onto wafer and spun to obtain an even coat of uniformthickness. The time, acceleration, and speed of the spin were tailored yield the

85

Page 86: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

A Copper deposition between featuresdue to poor seed layer adhesion

B Good isolation of features achieved asa result of improved adhesion

Figure 5-8. SEM images of copper channels. Poor adhesion of copper seed layer tophotoresist resulted in cracks between features through which copper waselectroplated where not desired.

target thickness of 10 ”m or 30 ”m. Instead of applying the thickest layers inmultiple coats, a slow spin speed and short duration were used to obtain thethickest coat in a single spin. The single spin was found to result in the mostuniform coat of surface topography.

2. The wafer with a fresh coat of photoresist was placed on a hotplate for thesoft-bake step to drive the solvent out from and cure the photoresist. Due tothe multilayer nature of this process, which called for further photolithographyand processing on top of already-patterned photoresist layers, a low soft-baketemperature of 95 C was used to avoid deformation of the already-patternedlayers that would occur at temperatures beyond 100 C.

3. Edge-bead removal was performed by spinning the wafer and applying a steadystream of acetone along the edge. The purpose was to remove the thicker beadof photoresist that pooled up around the edge of the wafer during coating to allowflush contact with photomask and to prevent outgassing that would occur in theexcessively thick bead. This step also allowed copper to be plated around the rimof the wafer for improved deposit uniformity (Section 5.4.3).

4. An additional bake on the hotplate drove solvent out from the newly formed edgeduring edge-bead removal. The acetone used for edge-bead removal tended toabsorb into and soften the photoresist around the edge. The edge needed to becured again to prevent the wafer from sticking to the photomask.

5. Rehydration of the photoresist was necessary to restore the water content thathad been baked out of the photoresist during the curing process. The water wasneeded for the photo-reaction that occurred during ultraviolet exposure to make theresist developable. To reduce the time required for rehydration and the sensitivity

86

Page 87: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

to ambient humidity conditions, rehydration was accomplished by submerging thewafer in a dish of water.

6. The wafer was aligned to the photomask and contacted with vacuum to pull thewafer flush against the photomask. Ultraviolet (UV) exposure was then timed to theappropriate dose depending on photoresist thickness.

7. Development of photoresist was done by submersion in AZ400K potassiumhydroxide based developer. After development the wafer was rinsed in deionizedwater and blown dry with nitrogen.

8. A de-scum etch in oxygen plasma was required to remove residue out fromtrenches in photoresist mold. The scum was otherwise found to disrupt the initialelectrodeposition of copper in regions, leading to a nonuniform copper fill. Inparticular, most of the residue was found to accumulate at the ends of recessedchannels in the photoresist due to capillary wicking of the liquid developer as thewafer was dried [62].

9. An optional second exposure of the photoresist to UV enabled the layers to besequentially removed in photoresist developer during the subtractive stage ofthe process. This second exposure could optionally be masked so that someregions of the resist mold would remain as part of the final device structure. It wasimportant that the second exposure was performed following the plasma descumstep to avoid heating just-exposed resist. During exposure, diazonaphtoquinonephoto-active-compounds in the positive photoresist decomposed and releasednitrogen gas, which slowly diffused through the resist [63]. If the resist was heatedtoo soon after exposure, the excess nitrogen gas that had not yet diffused outwould expand and cause bubbling and cracking of the resist.

5.4.3 Electroplating

Copper electroplating was carried out in an acid copper sulfate bath. While

most commercially-available copper electroplating bath electrolytes have contained

proprietary concoctions of surfactants and other organic additives to improve various

deposit characteristics, such as surface smoothness and hardness, the addition agents

have also been known to co-deposit with the plated copper and cause embrittlement

or residual stress [64]. Because the inductors and transformers of this work called for

relatively long stretches of suspended copper traces, an electrolyte chemistry was

used without any additive agents in order to minimize the residual stress in the copper

deposit.

87

Page 88: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

Table 5-2. Recipe per 1 L acid copper sulfate electroplating bath. Adapted fromRothschild [65].

Step Ingredient QuantityStart Deionized water 500 mLStir in Copper sulfate pentahydrate crystals 60 gAdd Sulfuric acid 120 mLFill Deionized water Up to 1 L total solution

In place of the additive agents, other electroplating parameters were optimized

to yield a uniform deposit. Agitation of the bath electrolyte during plating yielded the

greatest improvement to uniformity when comparing the rates of copper plating in areas

of high versus low feature density. In a still bath without agitation, regions of a wafer

with large areas of exposed copper plated at a significantly slower rate than in regions

with little copper area (mostly masked). Bath agitation was accomplished by pumping

the fluid and attaching the wafer to a horizontally-oscillating holder. The bath electrolyte

itself was also formulated with a low concentration of dissolved copper and a high

concentration of sulfuric acid to increase its throwing power, the ability of the bath to

provide a uniform deposit thickness over an irregular shapes [64]. The recipe for the

copper plating bath was adapted from the work of Rothschild [65] and is listed in Table

5-2.

A 4 mm ring of photoresist was removed around the rim of the wafer to expose

the underlying copper seed. The function of this exclusion zone was two-fold. First, it

allowed for electrical connection from the frontside of the wafer to the negative terminal

of the power supply for electroplating. The second function of the exclusion rim was

as a current baffle to provide uniform electric field strength across the wafer surface.

Because it was also in contact with the electroplating bath, copper was electroplated

onto the seed layer exposed around the perimeter. Having such a large electroplating

area around the wafer edge was found to aid in the uniformity of the deposition rate

amongst features of different dimensions across the wafer surface.

88

Page 89: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

The anodes consisted of 0.5 in nuggets of copper with 0.04 − 0.06% phosphorus

content. The quantity of copper anode material was adjusted downward to prevent

copper ion concentration from increasing in the bath over time. Too high of a concentration

of copper ion in the bath presented the formation of wart-like nodules on the copper

surface.

Cantilever structures such as the one depicted in Figure 5-9 were co-fabricated

alongside the inductors and transformers. At up to 1 ”m long and 10 ”m thick, the

cantilevers did not exhibit any perceptible curvature after release, indicating that no

stress gradient was present throughout the thickness of the copper.

A Full 1mm cantilever B Cantilever tip

Figure 5-9. SEM images of 1 mm-long copper cantilever. Minimal residual stress ispresent in electroplated copper structures as evidenced by minimalcurvature of long copper cantilever.

5.4.4 Argon Sputter Etch

After each thick layer of copper was electrodeposited to fill the photoresist molds,

an argon sputter etch was required to clean the copper surface. This etch was required

in order to improve the adhesion between the electroplated copper and the copper seed

layer sputter deposited onto its surface. Figure 5-10 compares images of devices that

were fabricated with (left) and without (right) this argon sputter etch step.

Separation was evident between each of the layers of the devices fabricated

without the argon sputter etch as shown for example in Figure 5-10A. Although some

89

Page 90: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

A Without sputter etch B With sputter etch

Figure 5-10. SEM images of devices fabricated with and without argon sputter etchingbetween layer depositions. Separation between layers was evident withoutthe sputter etch and was eliminated using the sputter etch.

layers remained in place (such as the one shown in the figure), many devices exhibited

catastrophic delamination during the final wet processing steps of fabrication, and the

yield was consequently low. In contrast, the devices that received the argon sputter etch

steps exhibited strong adhesion between copper layers. As shown in Figure 5-10B, the

sputter etch was effective in eliminating the separation between layers.

Argon sputter etching was performed in situ immediately prior to sputter deposition

of copper. In this manner, the etch removed the oxidized surface of the electroplated

copper, and the cleaned surface remained under vacuum until after the coated was

completed.

5.4.5 Photoresist Skin Removal

An unintended byproduct of the previously described argon sputter etch was the

alteration of the photoresist surface. The sputtering resulted in the formation of a thin,

impermeable skin on the exposed surface of the photoresist that consequently blocked

the underlying photoresist from being removed in the photoresist developer during

sequential layer removal (Section 5.1.1). In Figure 5-11A, a portion of the blocking skin

removed by physically scratching it away and the rest of the photoresist was etched in

90

Page 91: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

acetone. The unbroken skins could be seen spanning trenches between copper traces

at the interfaces between layers.

A Without CF4 plasma etch B With CF4 plasma etch

Figure 5-11. SEM images of copper trenches. Photoresist has been etched byKOH-based developer in each case, but thin surface layers remain betweentrenches without CF4 plasma etch

These photoresist skins could not be removed by O2 plasma ashing alone but were

successfully removed by a plasma etch consisting of both O2 and CF4. Figure 5-11B by

comparison shows how one such trench should appear when the photoresist skins were

removed at each layer by the CF4 plasma etch.

5.4.6 Copper Seed Etch

After all layers were added by electrodeposition through photoresist mods, removal

of the copper seeds at each layer was necessary to electrically isolate electroplated

traces. When it was used to remove the molds, ultrasonic agitation (Section 5.1.2)

in acetone or BAKER PRS-3000 was able to break up all of the seed layers that had

been deposited on top of photoresist. The bottommost copper seed layer, however,

and each of the copper seed layers when sequential layer removal was performed,

needed to be etched away with a copper etchant. Because the seed and the thicker

traces were both composed of copper, the seed etchant also etched the traces. As

shown by the comparison between electroplated copper features before and after

seed layer removal in Figure 5-12, this etching produced noticeable roughening of the

91

Page 92: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

copper sidewall profile, which was smoothly defined by the photoresist mold. Without

etch selectivity between seed and trace, the seed removal relied on timing since the

200 nm thick seed would be fully etched much quicker than the 10 ”m thick traces.

Ceric-ammonium-nitrate-based Cyantek CR9 chromium etchant was used to etch the

copper seeds due to its slow etch rate, which allowed easy timing of the etch to minimize

over-etching the electroplated features. Full etching of a seed layer took approximately

1 min with moderate agitation.

A Before copper etch B After copper etch

Figure 5-12. SEM images showing sidewall of copper features before and after copperetch. The sidewall is shown to be much smoother before the copper etchthan afterwards.

92

Page 93: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

CHAPTER 6INDUCTOR CHARACTERIZATION

This chapter discusses characterization of the microfabricated inductors under

radio-frequency excitation. An overview of the testing setup provides relevant background

regarding measurement with the vector network analyzer (VNA) tool. The section on

characterization methods details how scattering data measured from the VNA were

converted to impedance characteristics. Measurements are reported first for one-port

inductors fabricated on Pyrex substrates. These results provide a broad view of the

design space in terms of the electrical effect of varying geometry parameters such

as diameters, trace widths, and spacings. Subsequent sections explore the effect of

modifying the interlayer dielectric, switching the substrate to silicon, and changing the

shape from square to circle.

6.1 Equipment and Setup

Characterization of the microinductors at radio frequencies (RF) was accomplished

using a Vector Network Analyzer (VNA). The measurements were made with either

an Agilent E8361A with useable frequency range of 10 MHz − 30 GHz or a Rohde

& Schwarz ZVA/B with a useable frequency range of 300 kHz − 8 GHz. The general

working principle of the VNA was to excite the device under test with a single frequency

signal and to sample the amplitude and phase of the incident, reflected, and transmitted

waves. For the work presented here, the excitation frequency was swept from 10 MHz

up to at least 8 GHz to obtain the full frequency-dependent behavior of the devices up to

and beyond the first resonant frequency of each. The data were recorded as complex

scattering (S) parameters, defined as various ratios of the measured wave vectors [56].

Electrical connection to the inductors and transformers was made by radio-frequency

(RF) probes with ground-signal-ground (GSG) tip footprint configuration and 150 or

200 ”m tip pitch. Fixture compensation was performed with a calibration substrate

having open, short, through, and 50 Ω load standards [66].

93

Page 94: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

6.2 Inductor Characterization Methods

The microinductors were designed in both one-port and two-port configurations as

shown in Figure 6-1. In all cases the signal pad was 100 ”m × 100 ”m in area and was

flanked on both sides by ground pads of considerably larger area and spaced 50 ”m

apart.

A One-port B Two-port

Figure 6-1. SEM images depicting inductors with either one- port or two-portconnections.

6.2.1 One-Port Inductor Methods

The ends of an inductor in the one-port configuration were terminated at the signal

pad and at one of the adjacent ground pads. One-port characterization was convenient

in that the inductor impedance was directly reported on the screen of the vector network

analyzer (VNA), but for inductors fabricated on silicon the capacitance through the

substrate between signal and ground terminals was lumped into the measurement. Also,

because the microinductors of this work were asymmetric in that one half of the winding

was physically closer to the substrate than the top half, the one-port measurement was

furthermore sensitive to whether the top or bottom half of the winding was connected to

the signal pad. Because the substrate was held at a ground potential, the capacitance

was greater if the half of the winding closest to the signal pad was located on the

bottom.

94

Page 95: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

Inductor data obtained with one-port measurement yielded one complex scattering

value S per frequency point. Each scattering value was converted to an impedance

value, Z , via the relationship,

Z = Z01 + S

1− S , (6–1)

where Z0 was the characteristic impedance of the measurement line; in this case

Z0 = 50 Ω.

6.2.2 Two-Port Inductor Methods

The two-port configuration was specifically utilized to identify and separate out

the effects of capacitive coupling through the substrate. Inductors in this configuration

were connected with the ends terminating at the signal pad of each port. Ground pads

still flanked the signal pads and were connected between the two ports but were not

connected by copper to either of the signal pads. This two-port configuration provided

a richer data set that enabled separation of the impedance through the signal terminals

(i.e. the inductor impedance) from the shunt impedance between the signal and ground

terminals at each port (i.e. the substrate capacitance between the connection pads).

ZS1 ZS2

ZT

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

The two-port inductor impedances were modeled according to the circuit shown in

Figure 6-2, where ZT was the impedance through the inductor and ZS1 and ZS2 were

the shunt impedances to the substrate at each port. Conversion of the measured S

parameters to ABCD facilitated characterization of the two-port inductors and extraction

of the through and shunt impedances. The ABCD parameters for the circuit network

of Figure 6-2 were found using the property that the ABCD parameter matrices of

95

Page 96: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

cascaded network sections could be multiplied together to obtain the ABCD parameters

of the overall circuit. The overall ABCD parameter matrix for the circuit with shunt

impedance ZS1 followed by series impedance ZT followed by shunt impedance ZS2 was

obtained as

A B

C D

=

1 0

1

ZS10

1 ZT

1 0

1 0

1

ZS20

=

1 + ZTZS2

ZT

1

ZS1+ 1

ZS2+ ZTZS1ZS2

1 + ZTZS1

(6–2)

From the ABCD parameters, ZT could therefore be extracted simply as

ZT = B. (6–3)

The shunt impedances ZS2 and ZS1 were then extracted by rearranging the equivalencies

for the A and D parameters, respectively,

A = 1 +ZT

ZS2; ZS2 =

ZT

A− 1 =B

A− 1 (6–4)

and

D = 1 +ZT

ZS1; ZS1 =

ZT

D − 1 =B

D − 1. (6–5)

The two-port model also enabled simulation of the impedance that would have been

measured if the inductor were excited in a one-port configuration with the other port

shorted to ground. The input impedance looking into port 1 with port 2 shorted to ground

was solved as

Z1 =1

1

ZT+ 1

ZS1

=B

D. (6–6)

Similarly, the input impedance looking into port 2 if port 1 were shorted to ground was

obtained as

Z2 =1

1

ZT+ 1

ZS2

=B

A. (6–7)

96

Page 97: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

6.2.3 Inductor Characteristics Obtained from Impedance

Inductor impedance values were split into frequency-dependent resistances and

inductances. Resistances represented the real part of the complex impedances,

R = ℜZ . (6–8)

Inductance was derived from the imaginary part of the complex impedance divided by

the angular frequency,

L =ℑZ2πf

. (6–9)

As discussed in Section 3.1.2, quality factor was calculated as the ratio of the imaginary

to the real part of the impedance,

Q =ℑZℜ Z . (6–10)

The example plots of inductance, resistance, and quality factor in Figure 6-3

illustrates the frequency-dependent characteristics that were obtained by converting

measured scattering values to impedance for a typical inductor. As shown in Figure

6-3A, the inductance remained relatively flat for low frequencies at a value denoted by

Ldc . The inductance was found to decrease slightly from Ldc before rising sharply as the

frequency increased towards the self resonant frequency, SRF , of the inductor. At the

SRF the phase of the impedance wrapped and the ‘inductance’ values read negative as

the impedance became capacitive. The plot of resistance in Figure 6-3B similarly shows

that the resistance reached a peak value at the SRF of the inductor. Compared to the

inductance, however, the resistance deviated to a greater extent from its dc value Rdc

prior to the peak at SRF . While the large peaks in inductance and resistance at the SRF

resulted from the effect of shunt capacitance on the series interpretation of impedance

used here, the gradual decrease in inductance and increase in resistance observed in

the region between the peaks at SRF and the flat, low-frequency values were caused by

the generation of eddy currents within the traces leading to current crowding along the

trace edges (Section 3.6.2).

97

Page 98: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

6.3 One-Port Inductor Characterization

6.3.1 One-Port Inductors on Pyrex Substrates

One-port inductors were fabricated on Pyrex substrates with varied geometries.

Although photoresist was used to separate the upper and lower winding layers, it

was patterned into small posts to minimize its contribution to the capacitance of the

inductors. These devices were ideal for validation of the modeling and design concepts

that were presented in Chapter 3. The characteristics of several inductors of various

geometries were measured and summarized in Table 6-1. The areas, trace widths,

spacings, and numbers of turns were reported in the table based on the computer-aided

drawings (CAD) of the photomasks used in fabrication. The reported values for Ldc , Rdc ,

SRF , Qmax , and f @Qmax were averaged over the number of known-good inductors. The

yield was reported in the table as the number of known-good inductors out of the total

number of inductors of a given design that were tested. The smaller devices exhibited

greater yield—100% of the tested inductors up to 2.4 mm2 functioned properly. Some

larger inductors suffered from shorting between windings, and the characteristics of

those malfunctioning inductors was not recorded. The names listed in the table were

used to refer to all copies of inductors of a specific design.

Several interesting performance trends were revealed by comparing data amongst

different designs. Inductors I1 and I2 were identical in every way except for trace width

and spacing—both had the same 60 ”m trace pitch, but one had 40 ”m trace width and

20 ”m spacing, while the other had 50 ”m width and 10 ”m spacing. Although the dc

resistance was significantly lesser for the wider-traced inductor, the maximum quality

factor was significantly greater for the narrow-traced inductor. The frequency values

related to resonance (SRF and f @Qmax) were almost equal between the two inductors,

indicating that the capacitance in the device that lead to self-resonance was dominated

by the trace length rather than proximity of adjacent traces.

98

Page 99: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

107

108

109

−2

−1

0

1

2x 10

−7

Indu

ctan

ce

Frequency (Hz)

SRF

Ldc

A Inductance with Ldc and SRF

107

108

109

100

102

104

Res

ista

nce

Frequency (Hz)

SRF

Rdc

B Resistance with Rdc and SRF

107

108

109

−20

0

20

40

Qua

lity

Fac

tor

Frequency (Hz) f@Qmax

SRF

Qmax

C Quality factor with Qmax , f @Qmax , and SRF

Figure 6-3. Identification of inductor specifications from plots of inductance, resistance, and quality factor data obtainedfrom VNA.

Table 6-1. Comparison of measured inductor performance. Values were averaged over the number of devices tested.Area Width Spacing # turns # devices Ldc Rdc SRF f @Qmax

Name (mm2) (”m) (”m) per layer tested (nH) (Ω) (MHz) Qmax (MHz)I1 0.28 40 20 3 16/16 14.8 0.73 4290 33.0 1740I2 0.27 50 10 3 16/16 14.4 0.55 4260 27.5 1760I3 1.02 50 10 5 4/4 90.2 1.87 922 20.0 395I4 0.97 50 5 6 4/4 108 2.13 865 15.5 376I5 2.40 50 10 5 4/4 198 3.40 479 21.2 207I6 2.40 50 10 8 3/4 327 4.77 330 17.3 148I7 4.12 50 10 10 2/4 676 8.25 176 15.8 81I8 4.33 100 10 6 3/4 226 2.86 293 15.8 131I9 6.65 60 10 10 3/4 894 9.93 111 13.9 51

99

Page 100: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

Table 6-2. Comparison of model-predicted to measured inductor performance.Analytical FastHenry Measured Analytical FastHenry Measured

Name Ldc (nH) Ldc (nH) Ldc (nH) Rdc (Ω) Rdc (Ω) Rdc (Ω)I1 14.3 14.1 14.8 0.70 0.69 0.73I2 13.1 13.8 14.4 0.56 0.55 0.55I3 89.9 89.5 90.2 1.89 1.87 1.87I4 105 108 108 2.08 2.06 2.13I5 214 205 198 3.26 3.27 3.40I6 344 343 327 4.55 4.52 4.77I7 754 747 676 7.59 7.54 8.25I8 246 251 226 2.29 2.24 2.86I9 1056 1046 894 8.31 8.24 9.93

0 200 400 6000.5

0.6

0.7

0.8

0.9

1

Current (mA)

Res

istn

ace

(Ω)

A Inductor I2 (0.5 mm × 0.5 mm)

0 200 400 6002

3

4

5

Current (mA)

Res

istn

ace

(Ω)

B Inductor I3 (1.0 mm × 1.0 mm)

0 200 400 6002.5

3

3.5

4

4.5

Current (mA)

Res

istn

ace

(Ω)

C Inductor I5 (1.5 mm × 1.5 mm)

Figure 6-4. Measured resistance of different-sized inductors as a function of applied dc current. Inductors tested up toonset of thermal runaway.

100

Page 101: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

6.3.1.1 Comparison to model predictions

The low frequency inductances Ldc and resistances Rdc measured from the

inductors were compared to those values calculated from the analytical expressions

used in their design (Section 3.4) and were also compared to FastHenry simulations

(Section 3.5.2). DC resistance was calculated using a copper resistivity of 3.3 ”Ω · cm,

a value that was obtained by measuring resistance test structures that were co-fabricated

alongside the inductors. The measured, calculated, and simulated values are listed

in Table 6-2. The closest agreement between the values was found for the smaller

inductors up to about 200 nH. The larger inductors consistently measured lower

inductances and higher resistances than calculated or simulated.

6.3.1.2 Current rating

While saturation of a ferromagnetic core often limits the upper current capability of

an inductor, the air-core microinductors tested here were limited by resistive heating.

The current rating of the microinductors was tested by running successively greater

currents through the coils until the point at which excessive heating led to thermal

runaway.

Thermal runaway occurred when a stable operating current could not be maintained.

Because the rate of heat generation in the device was proportional to its electrical

resistance and the resistance of the copper also increased with temperature, this

positive feedback resulted in an unstable condition at high currents. At this point heat

generation and resistance increased without bounds until the device burnt up.

Three inductor designs of different sizes were tested for thermal runaway to

estimate their current rating: I2, I3, and I5. The devices were fabricated on a 100 mm

diameter, 500 ”m-thick Pyrex substrate and were tested while remaining attached to

the whole wafer. The wafer was placed on a large stainless steel chuck held at room

temperature, which was approximately 25 F. The resistance of each inductor at direct

current (dc) was plotted in Figure 6-4 as a function of the applied current. Inductor

101

Page 102: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

I2 experienced thermal runaway at 600 mA, I3 at 470 mA, and I5 at 450 mA. Prior to

runaway, the photoresist between traces was seen to melt, and the copper windings of

each inductor could be seen to change color as the copper oxidized. The dc current at

which the copper of each inductor noticeably changed color was 550 mA for I2, 350 mA

for I3, and 450 mA for I5. The preceding data however were particular to the specific

setup and the actual maximum current would depend on the heat transfer characteristics

of a given application.

6.3.1.3 Interwinding capacitance

To test the effect of an interlayer dielectric on interwinding capacitance, a batch of

inductors were fabricated with the same layouts as those listed in Table 6-1 but, instead

of patterning the resist into support posts, these were fabricated with a continuous

layer of photoresist between the upper and lower winding layers. The images in Figure

6-5 illustrate the physical difference between an inductor with patterned resist support

posts and one with an interlayer of continuous photoresist. Two designs, I1 and I6, were

selected to highlight the effect of the interlayer photoresist on the measured impedance

of a small and a large inductor, respectively. Comparing the measured impedances as

plotted in Figure 6-6 for two inductors of design I1, the patterning of the photoresist into

support posts increased the self resonant frequency (SRF) from 3.77 GHz to 4.20 GHz,

an improvement of 11%. For the impedances of the two larger inductors of design I6 as

plotted in Figure 6-7, patterning the photoresist into support posts increased the SRF

from 265 MHz to 326 MHz, an increase of 23%. As a result of the increased SRF, the

measured quality factor also increased from 16.0 to 17.3. Comparing the impedances of

other designs bore similar results, in which larger inductors with more turns exhibited a

greater improvement from minimizing the interlayer dielectric through patterning of the

photoresist support posts.

102

Page 103: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

A Patterned resist posts

B Continuous resist layer

Figure 6-5. Scanning electron micrograph (SEM) images of one inductor with patternedphotoresist support posts between upper and lower winding layers and onewith a continuous layer of photoresist between windings.

103

Page 104: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

107

108

109

1010

101

102

103

104

Indu

ctan

ce (

nH)

Frequency (Hz)

Unpatterned layerPatterned posts

A Inductance

107

108

109

1010

100

102

104

Res

ista

nce

(Ω)

Frequency (Hz)

Unpatterned layerPatterned posts

B Resistance

107

108

109

1010

0

10

20

30

40

Qua

lity

Fac

tor

Frequency (Hz)

Unpatterned layerPatterned posts

C Quality Factor

Figure 6-6. Comparison of interlayer dielectric effect on impedance for a small inductor (design I1, outer diameterD = 500 ”m).

107

108

109

1010

101

102

103

104

Indu

ctan

ce (

nH)

Frequency (Hz)

Unpatterned layerPatterned posts

A Inductance

107

108

109

1010

100

102

104

Res

ista

nce

(Ω)

Frequency (Hz)

Unpatterned layerPatterned posts

B Resistance

107

108

109

1010

0

10

20

30

40

Qua

lity

Fac

tor

Frequency (Hz)

Unpatterned layerPatterned posts

C Quality Factor

Figure 6-7. Comparison of interlayer dielectric effect on impedance for a large inductor (design I6, outer diameterD = 1.5 mm).

104

Page 105: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

6.3.2 One-Port Inductors on Silicon Substrates

Several iterations of inductor designs were also fabricated on silicon substrates.

As detailed in Section 3.6.1 the measured impedances of the inductors were heavily

impacted by capacitive coupling of the inductor through the substrate. Compared

to those on Pyrex substrates, all inductors fabricated on silicon exhibited lower

self-resonant frequencies (SRF ) and resonances that were highly damped by the

resistance of the substrate. The effect was dependent on the thickness of the insulating

dielectric layer between silicon and copper and on the resistivity of the substrate.

Because the capacitance through the substrate was lumped into the measured

impedance of the one-port inductors, the results that looked at substrate effects

were measured with two-port inductors in Section 6.4. Two sets of experiments were

conducted, however, with one-port inductors fabricated on silicon substrates to highlight

the effect of copper layer thickness and inductor shape on measured impedance.

6.3.2.1 Copper layer thickness: 10 ”m vs. 30 ”m

Two inductor designs (small and large) were each implemented in a version with

10 ”m thick layers of copper and in another version with 30 ”m thick layers of copper.

The smaller of the two inductor designs had an outer diameter D = 500 ”m, trace width

w = 30 ”m, spacing between traces s = 10 ”m, and n = 3 turns per layer on each

of the two winding layers. The larger had an outer diameter D = 1000 ”m, trace width

w = 30 ”m, spacing between traces s = 10 ”m, and n = 5 turns per layer on each of the

two winding layers. The vertical gap height between the upper and lower winding layers

of each inductor was equal to the thickness of each of the winding layers of that inductor.

The impedances were plotted for the smaller and larger designs in both thicknesses

in Figure 6-8 and 6-9, respectively. The measured performance parameters obtained

from each were listed in Table 6-3 for comparison. The thicker inductors of both designs

were shown to have greatly reduced low frequency resistances, resulting in greater

quality factors up to 100 MHz. The resistances of the thicker traces also showed

105

Page 106: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

Table 6-3. Measured performance parameters of two inductor designs, eachimplemented in versions with 10 ”m and 30 ”m thick copper layers.

Design Layer Nominal Nominal Maximum Frequency forsize thickness inductance resistance quality factor maxmimum quality

Small 10 ”m 21 nH 2.0 Ω 3.9 116 MHzSmall 30 ”m 17 nH 0.25 Ω 8.7 60 MHzLarge 10 ”m 125 nH 3.9 Ω 5.1 40 MHzLarge 30 ”m 110 nH 0.85 Ω 12 28 MHz

steeper increases with frequency due to increased eddy current losses. As a result,

the frequencies at which the maximum quality factors were measured for the thicker

inductors was lower than for the thinner versions. The thicker inductors yielded slightly

lower inductances as well from reduced mutual coupling between layers as a result of

the increased vertical gap height compared to the thinner versions. The reduction was

greater between those of the smaller design since the difference in vertical gap height

was greater in proportion to its diameter.

6.3.2.2 Inductor shape: square vs. circular spirals

Another set of inductor designs was implemented in small and large diameters with

square and circular spiral layouts. All of the inductors were constructed with 30 ”m thick

copper layers, but the outer diameters were varied between the square and circular

layouts so that the larger and smaller copies of each had roughly matching inductances.

The geometric parameters are listed in Table 6-4 with the circular layouts having larger

outer diameters to offset their smaller areas.

Measurement of the impedances of each inductor revealed only minor differences

resulting from the shape of both the smaller and the larger designs as plotted in

Figures 6-8 and 6-9. The performance characteristics were extracted from this data

and listed Table 6-5. Overall, the inductance-to-resistance ratios of the circular-shaped

inductors of both sizes were improved by approximately 10%. The improvement was

also associated with a 6–8% improvement in the maximum quality factors measured for

the circular-shaped inductors.

106

Page 107: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

Table 6-4. Geometric parameters of large and small inductors in square- andcircular-shaped spiral layouts. All were implemented with 30 ”m thick copperlayers.

Size Shape Outer diameter Trace width Trace spacing Turns per layerSmall Square 540 ”m 20 ”m 12 ”m 5Small Circular 585 ”m 20 ”m 12 ”m 5Large Square 960 ”m 32 ”m 16 ”m 6Large Circular 1015 ”m 32 ”m 16 ”m 6

Table 6-5. Measured performance parameters of small- and large-sized inductors withsquare- and circular-shaped spiral layouts.

Design Nominal Nominal Maximum Frequency forsize Shape inductance resistance quality factor maxmimum quality

Small Square 42 nH 0.60 Ω 10.8 88 MHzSmall Circular 45 nH 0.59 Ω 11.4 60 MHzLarge Square 117 nH 0.97 Ω 10.7 30 MHzLarge Circular 117 nH 0.88 Ω 11.6 30 MHz

107

Page 108: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

107

108

109

101

102

103

Indu

ctan

ce (

nH)

Frequency (Hz)

10 ”m thick layers30 ”m thick layers

A Inductance

107

108

109

10−1

100

101

102

103

Res

ista

nce

(Ω)

Frequency (Hz)

10 ”m thick layers30 ”m thick layers

B Resistance

107

108

109

0

5

10

15

Qua

lity

Fac

tor

Frequency (Hz)

10 ”m thick layers30 ”m thick layers

C Quality Factor

Figure 6-8. Comparison of layer thicknesses for small (outer diameter D = 500 ”m) one-port inductor on silicon substrate.

107

108

109

101

102

103

Indu

ctan

ce (

nH)

Frequency (Hz)

10 ”m thick layers30 ”m thick layers

A Inductance

107

108

109

10−1

100

101

102

103

Res

ista

nce

(Ω)

Frequency (Hz)

10 ”m thick layers30 ”m thick layers

B Resistance

107

108

109

0

5

10

15

Qua

lity

Fac

tor

Frequency (Hz)

10 ”m thick layers30 ”m thick layers

C Quality Factor

Figure 6-9. Comparison of layer thicknesses for large (outer diameter D = 1000 ”m) one-port inductor on silicon substrate.

108

Page 109: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

107

108

109

101

102

103

Indu

ctan

ce (

nH)

Frequency (Hz)

SquareCircle

A Inductance

107

108

109

10−1

100

101

102

103

Res

ista

nce

(Ω)

Frequency (Hz)

SquareCircle

B Resistance

107

108

109

0

5

10

15

Qua

lity

Fac

tor

Frequency (Hz)

SquareCircle

C Quality Factor

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

107

108

109

101

102

103

Indu

ctan

ce (

nH)

Frequency (Hz)

SquareCircle

A Inductance

107

108

109

10−1

100

101

102

103

Res

ista

nce

(Ω)

Frequency (Hz)

SquareCircle

B Resistance

107

108

109

0

5

10

15

Qua

lity

Fac

tor

Frequency (Hz)

SquareCircle

C Quality Factor

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

109

Page 110: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

6.4 Two-Port Inductor Characterization on Silicon Substra tes

6.4.1 Capacitive Coupling through the Substrate

Compared to inductors fabricated on insulating substrates such as Pyrex, the

measurement of inductors fabricated on silicon was severely affected by capacitive

coupling of windings through the substrate. As discussed in Section 3.6.1 the effects

of this included significant reductions in self-resonance accompanied by significant

increases in the effective resistances through the inductors. Capacitive coupling

through the substrate was strongest when points of maximum potential difference were

positioned in close proximity to each other and to the substrate. Because the maximum

potential difference occurs between the end terminals of a properly functioning inductor,

the observed inductor characteristics were highly dependent on the manner in which the

measuring probe connections were made to the end terminals.

For characterization at high frequencies (> 10 MHz) the ends of the microfabricated

inductors were terminated at pads that were designed specifically to correspond to

standard radio-frequency (RF) probe tips in ground-signal-ground (GSG) configurations.

With a thin dielectric layer of silicon dioxide providing electrical isolation between

the pads and the conductive silicon substrate, capacitors were inadvertently formed

between the pads and the conductive substrate as illustrated in Figure 6-12.

The resistance through the substrate between the tightly-spaced adjacent ground

and signal pads of the one port inductors could be estimated as the resistance between

two points on the surface of an infinite conductive slab, calculated as

Rs =ρ

2πsp, (6–11)

with ρ being the bulk resistivity of the slab and s the lateral separation between the

points. From the above equation the resistance between two pads separated by sp =

50 ”m was calculated at approximately Rs = 300 Ω on a ρ = 10 Ω · cm silicon wafer.

In GSG configuration, the two parallel paths between the signal pad and the two

110

Page 111: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

Cu Pad Cu Pad

Conductive Substrate

Probe Tip Probe Tip

Dielectric Layer

Figure 6-12. Diagram illustrating capacitive coupling through substrate between inductormeasurement pads.

ground pads to either side would reduce the overall substrate resistance in half to about

Rs = 150 Ω.

The inductors in this section were fabricated with a 2 ”m insulating layer of

plasma-enhanced chemical vapor deposited (PECVD) silicon dioxide between the

electroplated copper and the silicon wafer. Some had openings in the oxide for electrical

connection of the ground pads to the silicon substrate while others had a continuous

oxide layer. In either case, because the area of the ground pad was much larger,

capacitance between pads was largely determined by the size of the signal pad, which

was fixed in lateral area at 100 ”m × 100 ”m. The capacitance of the oxide layer

between the signal pad and the substrate was estimated assuming a standard parallel

plate capacitance

Cs =ǫrǫ0A

t, (6–12)

where Ç«r is the relative permittivity of the oxide layer, A is the lateral area of the pad,

and t is the thickness of the oxide layer. Assuming the PECVD silicon dioxide to have a

111

Page 112: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

relative permittivity of Ç«r = 3.5, the capacitance between the signal pad and the silicon

was estimated at 0.15 pF.

A two-port inductor was fabricated with 30 ”m thick copper layers on a ρ =

10 Ω · cm silicon wafer. The inductor consisted of two winding layers in a circular

spiral configuration with n = 6 turns per layer, D = 940 ”m outer diameter, w = 36 ”m

trace width, and s = 15 ”m spacing between traces. The structure of the inductor

was depicted in Figure 6-1B with each of the two ports located on opposite sides

of the inductor. One of the ports, referred to as Port 1, was connected to the upper

winding layer, while the other, Port 2, was connected to the lower winding layer. The

shunt impedances between each of the ports and ground were extracted from the

measurement data according to Equations 6–4 and 6–5.

Plots of the equivalent series resistances and capacitances as functions of

frequency revealed significant differences between the shunt impedances at each port

as shown in Figure 6-13. The equivalent series capacitance at Port 2, Cs2 was greater

at a value of about 1.4 pF due to the large area of the lower winding layer to which it

was directly connected. By comparison, the equivalent series capacitance at Port 1,

Cs1, was roughly 0.7 pF and more closely represented only the capacitance between the

signal pad and the silicon substrate. The measured value of Cs1 was greater than the

estimated value of 0.15 pF, which could have resulted from deviations in the thickness

of the silicon dioxide layer or the relative permittivity. The equivalent series resistance

(ESR) of the shunt impedance at Port 2 was less than that at Port 1, Rc2 = 50 Ω vs.

Rc1 = 115 Ω, again due to the larger area of the lower winding layer that was directly

connected to Port 2. While the measured value of ESR at Port 1 was slightly less than

the estimated value of 150 Ω, this result showed that Equation 6–11 provided a simple

method of obtaining a good rough estimate.

Equations 6–6 and 6–7 were also used to calculate (from the two-port measured

data) the impedance of the inductor that would be seen at each port if the opposite port

112

Page 113: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

had been shorted to ground. The frequency-dependent inductances, resistances, and

quality factors were plotted in Figure 6-17 for the two cases. In shorting the greater

shunt capacitance of Port 2 to ground, the one-port impedance looking into Port 1

exhibited an increased self-resonant frequency (SRF ) of 875 MHz compared to an

SRF of 554 MHz when looking into Port 2 with Port 1 shorted. The maximum quality

factor was also greater at a value of 14.6 when looking into Port 1 compared to 13.3

when looking into Port 2. These results highlighted the importance of considering all

connections to the inductors when measuring on a conductive substrate such as silicon.

6.4.2 Winding Losses

As shown in Section 6.3.2.1 by the comparison of identical inductor designs

implemented with either 10 ”m or 30 ”m thick copper layers, increases in the thicknesses

of the copper windings led to more pronounced increases in series resistance at lower

frequencies. With the increased volume of copper crossed with magnetic fields in the

inductors with 30 ”m thick traces, an attempt was made to mitigate losses due to eddy

current generation by splitting the windings into several parallel filamented traces.

The concept of filamented traces stemmed the use of Litz wires in high frequency

transformers, which feature many small parallel wires that have been bunched into one.

Filamented traces have been proposed to benefit the high frequency resistance of coils

by creating gaps across which eddy currents should not be able to “flow” [67]. However,

experiments in this work provided evidence that simple parallel filaments exhibit nearly

identical impedance at high frequencies (10 MHz− 10 GHz).

Two inductors were fabricated on a high resistivity (> 10000 Ω · cm) silicon wafer

with 30 ”m thick copper layers and identical geometries: two winding layers in circular

spiral configurations with n = 6 turns per layer, D = 940 ”m outer diameter, w = 36 ”m

trace width, and s = 15 ”m spacing between traces. However, as shown in Figure 6-15,

while one had solid traces like those of all other inductors presented in this work, the

other inductor featured traces that were filamented with slits. Shown in greater detail in

113

Page 114: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

Figure 6-16, the filamented traces contained sets of two parallel 3 ”m-wide slits running

along the trace length. Copper crossbars bridged the slits every 10 to improve the

structural integrity of the coil.

The measured impedances were found to be almost indistinguishable between the

inductor with filamented traces the standard one with solid traces. Figure 6-17 compares

plots of the frequency dependent inductances, resistances, and quality factors of the

impedances through each inductor, which were extracted according to Equation 6–3 to

minimize effects due to the substrate. Due to the loss of cross-sectional area through

the traces, the low-frequency resistance of the inductor with filamented traces was

about 10% greater at 0.90 Ω vs. 0.82 Ω with solid traces. Plotted in Figure 6-18 is the

change to the measured resistance in implementing filamented traces as a percentage

of the resistance with solid traces. At about 200 MHz the resistances between the two

inductors crossed over, beyond which point the filamented traces measured up to 10%

lower resistance around 1 GHz.

114

Page 115: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

107

108

109

1010

10−1

100

101

102

103

Cap

acita

nce

(pF

)

Frequency (Hz)

C

s2

Cs1

A Shunt Capacitance

107

108

109

1010

100

105

Res

ista

nce

(Ω)

Frequency (Hz)

R

c2

Rc1

B Equivalent Series Resistance of shuntcapacitances

Figure 6-13. Plots of shunt capacitances (Cs1 and Cs2) and equivalent series resistances (Rc1 and Rc2) of shuntcapacitances vs. frequency at Ports 1 and 2 of inductor.

107

108

109

1010

101

102

103

Indu

ctan

ce (

nH)

Frequency (Hz)

L

2

L1

A Inductance

107

108

109

1010

100

102

104

Res

ista

nce

(Ω)

Frequency (Hz)

R

2

R1

B Resistance

107

108

109

1010

0

5

10

15

Qua

lity

Fac

tor

Frequency (Hz)

Q

2

Q1

C Quality Factor

Figure 6-14. Plots of impedance vs. frequency for two-port inductor looking in to Port 1 (L1, R1, Q1), and looking in to Port2 (L2, R2, Q2). In each case the opposite port was shorted to ground.

115

Page 116: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

A Inductor with solid traces B Inductor with filamented traces

Figure 6-15. SEM images depicting inductors with solid and filamented traces.

A Solid traces B Filamented traces

Figure 6-16. SEM images zoomed closer in on solid and filamented traces.

116

Page 117: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

107

108

109

1010

101

102

103

Indu

ctan

ce (

nH)

Frequency (Hz)

SolidFilamented

A Inductance

107

108

109

1010

100

102

104

Res

ista

nce

(Ω)

Frequency (Hz)

SolidFilamented

B Resistance

107

108

109

1010

0

5

10

15

20

25

Qua

lity

Fac

tor

Frequency (Hz)

SolidFilamented

C Quality Factor

Figure 6-17. Plots of impedance vs. frequency for two-port inductors with filamented and with solid traces.

107

108

109

1010

−20

−10

0

10

20

Cha

nge

in r

esis

tanc

e (

% )

Frequency (Hz)

Figure 6-18. Difference between resistances through inductor with filamented vs. solid traces plotted as a percent of thesolid trace resistance.

117

Page 118: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

6.5 Summary of Inductor Characterization

The preceding chapter presented and compared the measured characteristics of a

variety of inductors to highlight the electrical effects resulting from design decisions.

‱ The inductances and resistances measured from the inductors at low frequenciesmatched well with the values calculated from the analytical expressions used intheir design.

‱ Several inductors with 10 ”m thick copper winding layers were subjected to highcurrent levels and were found to sustain up to about 500 mA before thermalrunaway caused the windings to burn up.

‱ The bulk removal of photoresist from between the upper and lower winding layerswas found to increase the self-resonant frequencies of inductors by about 10–20%.

‱ Increasing the thickness of each copper winding layer from 10 ”m to 30 ”m yieldedinductors and transformers that had significantly improved direct current (dc)resistances with only slightly decreased inductances. However, the benefit ofthe thicker layers at dc was lost at frequencies greater than about 100 MHz asthe resistances of the thicker windings increased more rapidly with increasingfrequency due to increased eddy current losses in the copper.

‱ Changing the shape of the inductors from square to circular provided minorimprovements of about 10% in inductance-to-resistance ratios.

‱ The characterization of inductors fabricated on silicon substrates showedthat capacitive coupling to the substrate had a strong effect on the measuredperformance. In particular, higher self-resonant frequencies and quality factorswere measured by grounding the terminal connected to the lower winding layerand applying the signal at the terminal connected to the upper winding layer.

‱ Splitting the inductor windings into parallel filamented windings had almost noeffect on the measured impedance of the inductor.

118

Page 119: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

CHAPTER 7TRANSFORMER CHARACTERIZATION

This chapter discusses the methods and results of characterizing the microfabricated

transformers under radio-frequency excitation. A description of the equipment and setup

outlines the vector network analyzer tool as it was used to characterize the two-port

microtransformers. Conversion of the measured scattering parameters to impedance

parameters is then discussed as a route for interpretation of the data in terms that

were relevant to power converters. The load-dependence of the transformer efficiency

and voltage gain is addressed with appropriate mathematical tools to quantify the

dependence. The results of characterizing three microfabricated transformers is then

presented. The first two transformers were implemented with 10 ”m thick copper layers

with square spiral layouts and turns ratios of 1 : 1 and 1 : 3.5. The last transformer

was an updated 1 : 1 transformer with circular spiral coils and an improved design

implemented in 30 ”m thick copper. The performance characteristics are compared

amongst the microtransformers to highlight the measurable electrical effects associated

with the various design options.

7.1 Equipment and Setup

An Agilent E8361A Vector Network Analyzer (VNA) was used for two-port

measurement of the transformers. The VNA excited the transformer under test with

a frequency signal on one port and sampled the amplitude and phase of the incident,

reflected, and transmitted waves on both ports. Both ports were alternately excited

to fully characterize the transformer in both directions. For the work presented

here, the excitation frequency was swept from 10 MHz–8 GHz to obtain the full

frequency-dependent behavior of the devices up to and beyond the first resonant

frequency of each. The data were recorded as complex scattering (S) parameters,

defined as various ratios of the measured wave vectors [56].

119

Page 120: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

Electrical connection to the inductors and transformers was made by radio-frequency

(RF) probes with ground-signal-ground (GSG) tip configuration and 150 or 200 ”m tip

pitch. Fixture compensation was performed with a calibration substrate having open,

short, thru, and 50 Ω load standards [66].

7.2 Impedance Parameters

Two-port measurement of the transformers with the Vector Network Analyzer (VNA)

yielded a 2× 2 matrix of complex scattering parameters for each frequency point,

S11 S12

S21 S22

. (7–1)

While scattering parameters have proven to be useful in a wide variety of radio-frequency

(RF) applications such as communications systems, impedance parameters were

more appropriate for comparison with other power transformers and for extracting

performance characteristics such as coupling coefficients and turns-ratios that were

affected by design decisions.

The scattering values were converted to impedance values via the following

relationships,

Z11 = Z0(1 + S11) (1− S22) + S12S21(1− S11) (1− S22)− S12S21

, (7–2)

Z12 = Z02S12

(1− S11) (1− S22)− S12S21, (7–3)

Z21 = Z02S21

(1− S11) (1− S22)− S12S21, (7–4)

Z22 = Z0(1− S11) (1 + S22) + S12S21(1− S11) (1− S22)− S12S21

. (7–5)

The impedance parameters corresponded to the elements of the circuit diagram

drawn in Figure 7-1. For the transformers measured in this work, Z11 represented the

120

Page 121: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

+-

+- Z21I1Z12I2

Z22Z11I1 I2

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

impedance of the primary coil, Z22 represented that of the secondary coil, and Z12 and

Z21 represented the coupling between primary and secondary coils.

The complex impedance values were then split into real and imaginary parts, and

the frequency-dependent resistances and inductances were extracted in the same

manner as for the inductors (Section 6.2.3). Frequency-dependent resistances were

reported as the real part of the complex impedances,

Rxx = ℜZxx , (7–6)

where the subscripts of the resistance followed from the corresponding impedance

parameters. Frequency-dependent inductance was calculated as the imaginary part of

the complex impedance divided by the angular frequency,

Lxx =ℑZxx2πf

. (7–7)

Extraction of the nominal inductance Lxx,dc , resistance Rxx,dc , and coupling

coefficients k of the transformers was based on an assumed low-frequency model

of a transformer with imperfectly coupled inductors and series resistances. The circuit

diagram for this low-frequency model was drawn in Figure 7-2. This model represented

the behavior of the transformers in the low frequency range for which the inductance and

resistance values remained constant. In all measured cases, the mutual inductances in

the forward and reverse directions, L12,dc and L21,dc , were almost identically equal—as

were R12,dc and R21,dc—indicating that the coupling between coils was the same

121

Page 122: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

regardless of whether the transfer was from primary to secondary or in reverse. For

this reason, the mutual inductances L12,dc and L21,dc were reported as a single value Lm.

The coupling coefficient k was calculated from the nominal inductances,

k =Lm

√

L11,dcL22,dc. (7–8)

The resistances, R12,dc and R21,dc , of the mutual impedances were omitted from the

model as depicted in Figure 7-2 as these values were negligibly small since the

transformers had no magnetic cores.

L11,dc

R11,dc Lm R22,dc

L22,dc

Figure 7-2. Low frequency transformer model consisting of imperfectly coupled inductorswith series resistance.

7.3 Load-Dependent Efficiency and Voltage Gain

Although the quality factors of the primary and secondary coil could be calculated

from the measured impedances and would provide some insight into the maximum

transformer efficiency (see Section 4.2.2), the efficiency of power transfer through the

transformer was highly dependent on the load impedance. Since the vector network

analyzer (VNA) was calibrated to report scattering data in relation to 50 Ω characteristic

impedance, the efficiency and voltage gain were calculated directly from the scattering

parameter data. The efficiency, i.e. the power delivered to the load as a percent of

the power delivered into the transformer, was calculated for the 50 Ω characteristic

impedance of the VNA by

ηZ0 =|S21|2

1− |S11|2, (7–9)

122

Page 123: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

and the voltage gain was calculated as

AZ0 =

∣

∣

∣

∣

S21

1 + S11

∣

∣

∣

∣

. (7–10)

Determination of the transformer efficiency and voltage gain under other loading

conditions was estimated via the method described in Section 4.4, in which conversion

of the scattering parameters to ABCD parameters enabled simulation of the transformer

performance given an arbitrary load. The values for efficiency and voltage gain were

calculated from the measured data and arbitrary load impedances ZL using Equations

4–21 and 4–22, respectively. Repeating the equations here for convenience, the

efficiency was calculated by

η =ℜZL

ℜ

(AZL + B) (CZL + D) , (7–11)

and voltage gain was calculated by

Av =

∣

∣

∣

∣

ZL

AZL + B

∣

∣

∣

∣

. (7–12)

By sweeping the complex load impedance ZL, surfaces were obtained that provided

visual clues about the realms of impedance that would result in high efficiency or

voltage gain at a given fixed frequency. Each frequency point had a different set of

efficiency/voltage gain surfaces. By contrast, the maximum efficiency introduced in

Section 4.2 provided the value of the peak efficiency at each given frequency regardless

of the load impedance. The complex load impedance at each frequency point that would

result in maximum efficiency was referred to as the matched load, which was calculated

by solving for ZL in the quadratic equation given in Equation 4–29,

(

AC + AC)

ZL2+

(

BC − BC + AD − AD)

ZL −(

BD + BD)

= 0, (7–13)

where the overline above parameters denoted complex conjugates. The matched load

impedance was then fed back into Equations 7–11 and 7–12 to obtain the maximum

123

Page 124: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

efficiency and corresponding voltage gain under matched load conditions, which was

identically equal to the maximum efficiency as calculated by Equation 4–1.

7.4 Characterization of Transformers with 10 ”m Thick Layers

Two transformers with 10 ”m thick copper layers were tested on Pyrex substrates: a

1 : 1 isolation transformer and a 1 : 3.5 step-up transformer, both of which had identical

50 nH primary coils. The secondary coil of the 1 : 1 transformer was designed as a

mirror image of the primary coil with the goal of providing direct current (dc) isolation

between primary and secondary with unity gain. The secondary coil of the 1 : 3.5

transformer was composed of nine additional turns on both upper and lower winding

layers that were nested within the area cleared by the primary coil. The scanning

electron micrograph (SEM) images in Figure 7-3 compare the structures of the two

transformers, showing similar layout but with more turns inside the 1 : 3.5 transformer, all

of which belonged to the secondary coil.

The layout of both transformers began with the 1 : 1 transformer, which was

designed with interleaved primary and secondary coils with 1.5 mm outer diameter,

30 ”m wide traces, 50 ”m space between traces of the same coil, 10 ”m space between

adjacent coils, 10 ”m vertical space between upper and lower winding layers, and 2

turns of each coil per layer. The secondary coil of the 1 : 3.5 transformer was extended

with an additional 9 turns per layer in the inner region of the transformer with 10 ”m

space between turns.

The frequency-dependent inductances and resistances were plotted in Figures 7-4

and 7-5 for the 1 : 1 and 1 : 3.5 transformers, respectively. For the 1 : 1 transformer,

the inductances (L11 and L22) and resistances (R11 and R22) were essentially equal

between the primary and secondary by design. The mutual inductadisnces (L12 and

L21) would have been equal to L11 and L22 if coupling were perfect between primary

and secondary but in practice were less due to imperfect coupling. For the 1 : 3.5

transformer, with its secondary winding containing many more turns, the secondary

124

Page 125: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

A 1 : 1 transformer B 1 : 3.5 transformer

Figure 7-3. SEM images of 1 : 1 and 1 : 3.5 microtransformers. The structures of eachare similar except that the 1 : 3.5 transformer contains nine extra turnsnested within its area.

inductance L22 and resistance R22 were both much greater than the primary inductance

L11 and resistance R11. In both cases, the coupled resistances R12 and R21 were trivially

small at low frequencies and fell within the noise of the measurement, indicating that

core loss was practically nonexistent as there was no magnetic core. The coupled

resistances appeared to rise with frequency only due to capacitive coupling between the

coils affecting the phase of the coupling impedance.

7.4.1 Extraction of Nominal Inductances and Resistances

The nominal, low-frequency values for inductances and resistances were extracted

from the measured data and were listed in Table 7-1 for comparison. Also included are

data that were extracted from an improved 1 : 1 transformer with 30 ”m layers that is

discussed in Section 7.5. The primary coils of the 1 : 1 and the 1 : 3.5 transformers

proved to be identical both physically and electrically. The measurements bore out this

fact as the primary inductances L11,dc and resistances R11,dc were essentially equal

across both designs. The inductances and resistances of the secondary coil of the

1 : 1 transformer also matched that of the primary by design. The 1 : 3.5 transformer,

with its nested secondary windings, exhibited greatly increased secondary inductance

125

Page 126: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

107

108

109

1010

101

102

103

Frequency (Hz)

Indu

ctan

ce (

nH)

107

108

109

1010

10−2

10−1

100

101

102

103

104

Frequency (Hz)

Res

ista

nce

(Ω)

L11

, L22

L12

, L21

R11

, R22

R12

, R21

Figure 7-4. Plots of frequency-dependent inductance and resistance of 1 : 1 transformerwith 10 ”m thick layers.

Table 7-1. Comparison of transformer circuit parameters.Turns Layer Primary Secondary Mutualratio thickness Area L11,dc R11,dc L22,dc R22,dc Lm k

1 : 1 10 ”m 2.4 mm2 47 nH 2.1 Ω 47 nH 2.1 Ω 41 nH 0.871 : 3.5 10 ”m 2.4 mm2 47 nH 2.1 Ω 496 nH 9.0 Ω 96 nH 0.631 : 1 30 ”m 1.08 mm2 45 nH 0.5 Ω 44 nH 0.5 Ω 41 nH 0.92

L22,dc compared to that of the 1 : 1 transformer at the expense of greater secondary dc

resistance R22,dc as well. Additionally, the coupling coefficient was significantly better

between the interleaved windings of the 1 : 1 transformer compared to that of the nested

windings of the 1 : 3.5.

126

Page 127: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

107

108

109

1010

101

102

103

104

Frequency (Hz)

Indu

ctan

ce (

nH)

107

108

109

1010

10−2

10−1

100

101

102

103

104

Frequency (Hz)

Res

ista

nce

(Ω)

L12

, L21

L22

L11

R11

R22

R12

, R21

Figure 7-5. Plots of frequency-dependent inductance and resistance of 1 : 3.5transformer with 10 ”m thick layers

7.4.2 Load-Dependent Performance of 1 : 1 Transformer

The efficiency of the 1 : 1 transformer was plotted in Figure 7-6, both for the

as-measured case with 50 Ω loading and for the case of maximum efficiency with

matched loads at each frequency point. The peak efficiency measured with 50 Ω loading

was 84% at 350 MHz, while the maximum estimated efficiency with a matched load was

found to peak at 930 MHz with 92%. By definition, the maximum efficiency as estimated

with matched loads was greater over all frequency points than as measured with a 50 Ω

load. However, the efficiencies of the two cases most closely matched in values around

127

Page 128: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

the peak values of both. This suggested that the value of the matched load at these

frequencies approached 50 Ω.

107

108

109

1010

0

20

40

60

80

100

Frequency (Hz)

Effi

cien

cy %

Matched Loads

50 Ω Load

Figure 7-6. Efficiency of 1 : 1 transformer for both 50 Ω and conjugate matched loads.

107

108

109

1010

0

0.2

0.4

0.6

0.8

1

Frequency (Hz)

Vol

tage

Gai

n (V

/V)

50 Ω Load

Matched Loads

Figure 7-7. Voltage gain of 1 : 1 transformer for both 50 Ω and conjugate matched loads.

The voltage gain corresponding to the same loading conditions at each frequency

was plotted in Figure 7-7. As measured with 50 Ω loading, the voltage gain was

approximately 0.8 V/V at frequencies up to 100 MHz, beyond which point the voltage

128

Page 129: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

gain decreased with increasing frequency. With matched loads, by comparison, the

voltage gain was lower at frequencies < 45 MHz than the 50 Ω case but increased to a

peak value of 0.94 V/V at 500 MHz. In either loading case, the peak voltage gain was

found to occur at frequencies similar to those at which the peak efficiencies occurred.

107

108

109

1010

100

101

102

103

Frequency (Hz)

Mag

nitu

de o

f Loa

d Im

peda

nce

|ZL| (

Ω)

107

108

109

1010−90

−45

0

45

90

Ang

le o

f Loa

d Im

peda

nce

∠Z

L (D

egre

es)

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

The calculated impedance of the matched load for each frequency point was plotted

in Figure 7-8 in terms of its magnitude and phase. As indicated by the phase remaining

at a negatively value of about −45 up to almost 1 GHz, the matched load impedance

had equal components of capacitance and resistance over most of the usable frequency

range. This could have been implemented as a load resistance placed in series or in

parallel with a capacitor, with values depending on the frequency to equate with the

matched load impedance.

129

Page 130: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

0 200 400 600 800 1000 −1000−500

0500

10000

20

40

60

80

100

XL (Ω)

RL (Ω)

Effi

cien

cy (

%)

A Efficiency

0 200 400 600 800 1000 −1000−500

0500

10000.6

0.8

1

1.2

1.4

1.6

XL (Ω)

RL (Ω)

Vol

tage

Gai

n (V

/V)

B Voltage Gain

Figure 7-9. Efficiency and voltage gain of 1 : 1 transformer (10 ”m thick winding layers)plotted as functions of a complex load at 100 MHz fixed frequency.

130

Page 131: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

By sweeping the complex load impedance for a given fixed frequency, surfaces were

obtained for the efficiency and voltage gain of the 1 : 1 transformer such as plotted in

Figure 7-9 for 100 MHz. The surface plot in Figure 7-9A showed that peak efficiency of

the 1 : 1 transformer at 100 MHz would be obtained for impedances with real resistances

less than 100 Ω and a slight capacitive component. The surface plot in Figure 7-9B

showed that the voltage gain at 100 MHz was nearly flat at 0.88 V/V over most values

of impedance, except in the region of nearly zero resistance, around which very slightly

inductive load impedances resulted in drastically reduced voltage gain and very slight

capacitive load impedances resulted in increased voltage gains.

7.4.3 Load-Dependent Performance of 1 : 3.5 Transformer

The efficiency of the 1 : 3.5 transformer was plotted in Figure 7-10, both for

the as-measured case with 50 Ω loading and for the case of maximum efficiency

with matched loads at each frequency point. The peak efficiency of 35% at 55 MHz

measured with 50 Ω loading was much lower than for the 1 : 1 transformer. However

the estimated maximum efficiency with a matched load was found to peak up to 78%

at 150 MHz. In contrast to the 1 : 1 transformer, the measured efficiency of the 1 : 3.5

transformer with 50 Ω deviated to greater extents from the maximum efficiency with

matched loads as the frequency increased away from 10 MHz. This suggested that the

value of the matched load was close to 50 Ω at 10 MHz but progressively deviated from

this value.

The voltage gain corresponding to the same loading conditions at each frequency

was plotted in Figure 7-11. A similar result was found for the voltage gain, where the

voltage gain calculated for matched loads deviated to a larger extent from that measured

with 50 Ω loading in contrast to the result with the 1 : 1 transformer. As measured with

50 Ω loading, the voltage gain was approximately 1.3 V/V only up to 15 MHz, beyond

which point the voltage gain quickly decreased with increasing frequency. With matched

loads, however, the voltage gain started low at low frequencies and increased to 3.1 V/V

131

Page 132: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

107

108

109

1010

0

20

40

60

80

100

Frequency (Hz)

Effi

cien

cy %

Matched Loads

50 Ω Load

Figure 7-10. Efficiency of 1 : 3.5 transformer for both 50 Ω and conjugate matched loads.

107

108

109

1010

0

1

2

3

4

Frequency (Hz)

Vol

tage

Gai

n (V

/V)

Matched Loads

50 Ω Load

Figure 7-11. Voltage gain of 1 : 3.5 transformer for both 50 Ω and conjugate matchedloads.

132

Page 133: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

at 100 MHz. In the matched load case, the peak voltage gain was found to occur at

frequencies similar to those at which the peak efficiencies occurred, whereas the peak

efficiency and voltage gain did not occur at the same frequencies for the 50 Ω load.

107

108

109

1010

101

102

103

104

Frequency (Hz)

Mag

nitu

de o

f Loa

d Im

peda

nce

|ZL| (

Ω)

107

108

109

1010−90

−45

0

45

90

Ang

le o

f Loa

d Im

peda

nce

∠Z

L (D

egre

es)

Figure 7-12. Magnitude and phase of matched load impedance for 1 : 3.5 transformerwith 10 ”m thick layers.

The calculated impedance of the matched load for each frequency point was plotted

in Figure 7-12 in terms of its magnitude and phase. As indicated by the phase remaining

at a negatively value of nearly −65 up to 100 MHz, the matched load impedance was

more capacitive than for the 1 : 1 transformer. The results for the 1 : 3.5 transformer

showed in summary that a strong value capacitance in the load would be an essential

component for reaching high efficiencies and voltage gains.

By sweeping the complex load impedance for a given fixed frequency, surfaces were

obtained for the efficiency and voltage gain of the 1 : 3.5 transformer such as plotted in

133

Page 134: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

0 200 400 600 800 1000 −1000−500

0500

10000

20

40

60

80

100

XL (Ω)

RL (Ω)

Effi

cien

cy (

%)

A Efficiency

0 200 400 600 800 1000 −1000−500

0500

10000

2

4

6

8

10

12

14

XL (Ω)

RL (Ω)

Vol

tage

Gai

n (V

/V)

B Voltage Gain

Figure 7-13. Efficiency and voltage gain of 1 : 3.5 transformer (10 ”m thick windinglayers) plotted as functions of a complex load at 100 MHz fixed frequency.

134

Page 135: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

Figure 7-13 for 100 MHz. The surface plot in Figure 7-13A showed peak efficiencies at

100 MHz would be obtained from the 1 : 3.5 transformer for a wide range of impedances

with real resistances of 100–400 Ω and a strong capacitive reactance of 100–500 Ω. The

surface plot in Figure 7-13B showed that a voltage gain of 2 V/V could be obtained for

a wide range of impedances at 100 MHz. In the region of nearly zero resistance, loads

with capacitance of about 7.5 pF were shown to result in voltage gains of almost 14 V/V.

7.5 Characterization of Transformer with 30 ”m Thick Layers

In response to the relatively high series resistance through the coils of the

transformers presented in the previous section, an improved 1 : 1 transformer with

30 ”m thick copper windings was fabricated on a Pyrex substrate. As seen in the

scanning electron micrograph image of the transformer in Figure 7-14, the layout also

featured interleaved circular shaped spiral coils for improved inductance to resistance

ratio. Because the purpose of this design was to maximize performance of the 1 : 1

transformer without comparison to a higher turns-ratio design, a greater fill fraction was

used than for the transformer with 10 ”m thick windings, which led to higher inductance

density and better coupling. Both primary and secondary coils were laid out with the

same geometry: circular-shaped spirals with 1.1 mm outer diameter, 32 ”m trace width,

64 ”m spacing between traces of the same coil, 16 ”m spacing between adjacent coils,

30 ”m vertical gap between the upper and lower winding layers, and 4 turns of each coil

per layer.

The frequency-dependent inductances and resistances were plotted in Figure 7-15.

The agreement between the inductances L11 and L22 and between the resistances R11

and R22 verified the designed goal of the primary and secondary coils to be in all ways

equal to each other. The mutual inductances L12 and L21 dropped slightly below L11

and L22 due to imperfect coupling. The coupled resistances R12 and R21 were very low

at low frequencies, indicating the absence of core loss as there was no magnetic core,

135

Page 136: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

Figure 7-14. SEM image of 1 : 1 microtransformer featuring 30 ”m thick copper layersand circular spiral layout. The primary and secondary coils are interleaved.

and rose with frequency only due to capacitive coupling between the coils affecting the

phase of the coupling impedance.

Nominal low-frequency parameters were extracted from the impedance data and

are reported in Table 7-1 for comparison with the transformers of Section 7.4, which

were implemented with 10 ”m thick layers. The primary and secondary inductances

(L11,dc and L22,dc ) were very similar to those of the previous 1 : 1 transformer with 10 ”m

thick layers. This thicker 1 : 1 transformer with 30 ”m layers, however, featured a 4×

improvement in resistances of the coils down to 0.5 Ω. The redesigned transformer

also utilized an area of 1.08 mm2, less than half that of the previous result owing to the

increased turn count (4 vs. 2). The increased utilization of the area also resulted in an

improved coupling coefficient of k = 0.92, up from the previous k = 0.87.

The efficiency of the thicker 1 : 1 transformer was plotted in Figure 7-16, both

for the as-measured case with 50 Ω loading and for the case of maximum efficiency

with matched loads at each frequency point. The peak efficiency of 90% at 333 MHz

136

Page 137: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

107

108

109

1010

101

102

Frequency (Hz)

Indu

ctan

ce (

nH)

107

108

109

1010

10−2

10−1

100

101

102

103

Frequency (Hz)

Res

ista

nce

(Ω)

L12

, L21

R11

, R22

R12

, R21

L11

, L22

Figure 7-15. Plots of frequency-dependent inductance and resistance of 1 : 1transformer with 30 ”m thick layers.

measured with 50 Ω loading was very close to the peak efficiency of 91% with matched

loads. Compared to Figure 7-6 for the previous 1 : 1 transformer with 10 ”m thick

layers, the measured efficiencies with a 50 Ω load of the thicker transformer more

closely approached the matched load efficiency at peak values but deviated to a greater

extent at lower frequencies. The matched load efficiency of the redesigned transformer

was much improved at these lower frequencies, reaching 70% efficiency at 10 MHz as

compared to 20% in the previous generation.

The voltage gain corresponding to the same loading conditions at each frequency

was plotted in Figure 7-16. As shown in the plot, there was a high degree of overlap

137

Page 138: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

107

108

109

1010

0

20

40

60

80

100

Frequency (Hz)

Effi

cien

cy %

Matched Loads

50 Ω Load

Figure 7-16. Efficiency of 1 : 1 transformer (30 ”m thick winding layers) as a function offrequency for both 50 Ω and conjugate matched loads.

107

108

109

1010

0

0.2

0.4

0.6

0.8

1

Frequency (Hz)

Vol

tage

Gai

n (V

/V)

Matched Loads

50 Ω Load

Figure 7-17. Voltage gain of 1 : 1 transformer (30 ”m thick winding layers) as a functionof frequency for both 50 Ω and conjugate matched loads.

138

Page 139: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

between the voltage gains for the 50 Ω load and for the matched load. As measured

with 50 Ω loading, the voltage gain was approximately 0.90 V/V up to 180 MHz. With

matched loads, however, the voltage gain increased up to a value of 0.94 V/V at

160 MHz. Both for the 50 Ω load and for the matched loads, the voltage gain and

efficiency were each maintained at values near their peaks throughout the frequency

range from 100–400 MHz.

107

108

109

1010

100

101

102

103

Frequency (Hz)

Mag

nitu

de o

f Loa

d Im

peda

nce

|ZL| (

Ω)

107

108

109

1010−90

−45

0

45

90

Ang

le o

f Loa

d Im

peda

nce

∠Z

L (D

egre

es)

Figure 7-18. Magnitude and phase of matched load impedance for 1 : 1 transformer with30 ”m thick layers.

The calculated impedance of the matched load for each frequency point was plotted

in Figure 7-18 in terms of its magnitude and phase. Like the 1 : 1 transformer with 10 ”m

thick layers, the phase remained at a value of −45 up to 1 GHz, similarly indicating that

the matched load impedance had equal components of capacitance and resistance over

most of the usable frequency range. Throughout the frequency range up to 450 MHz the

139

Page 140: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

magnitude of the matched load impedance for the thicker redesigned transformer was

roughly half that of the transformer with 10 ”m thick layers.

By sweeping the complex load impedance for a given fixed frequency, surfaces

were obtained for the efficiency and voltage gain of the 1 : 1 transformer with 30 ”m

thick layers such as plotted in Figure 7-19 for 100 MHz. The surface plot in Figure 7-19A

showed that at 100 MHz peak efficiency would be obtained for impedances with real

resistances less than 100 Ω and a slight capacitive component. The surface plot in

Figure 7-9B showed that the voltage gain at 100 MHz was nearly flat at 0.93 V/V over

most values of impedance, except in the region of nearly zero resistance, around which

very slightly inductive load impedances resulted in drastically reduced voltage gain and

very slight capacitive load impedances resulted in increased voltage gains.

7.6 Summary of Transformer Characterization

The preceding chapter presented and compared the measured characteristics of

two inductors fabricated with 10 ”m thick copper layers with turns ratios of 1 : 1 and

1 : 3.5 and one inductor fabricated with 30 ”m thick copper layers with a turns ratio of

1 : 1.

‱ The two transformers with 10 ”m both had the identical primary coils, but thetransformer with 1 : 3.5 turns ratio had extra turns of the secondary coil nestedinto the inner area cleared by the transformer. Mutual magnetic coupling wasconsequently less for the 1 : 3.5 transformer at k = 0.63 compared to k = 0.87 forthe 1 : 1 transformer.

‱ The efficiencies and voltage gains of the transformers were shown to be highlydependent on the load. Greater efficiencies and voltage gains were found to beachievable when the loads had capacitive components. The degree of desiredcapacitance was greater for the 1 : 3.5 transformer than for the 1 : 1 transformers.As a result, the 1 : 1 transformers exhibited better efficiencies and voltage gainsthan the 1 : 3.5 transformer with 50 Ω loading as measured with the vector networkanalyzer.

140

Page 141: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

0 200 400 600 800 1000 −1000−500

0500

10000

20

40

60

80

100

XL (Ω)

RL (Ω)

Effi

cien

cy (

%)

A Efficiency

0 200 400 600 800 1000 −1000−500

0500

1000

0.8

1

1.2

1.4

1.6

XL (Ω)

RL (Ω)

Vol

tage

Gai

n (V

/V)

B Voltage Gain

Figure 7-19. Efficiency and voltage gain of 1 : 1 transformer (30 ”m thick winding layers)plotted as functions of a complex load at 100 MHz fixed frequency.

141

Page 142: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

CHAPTER 8PACKAGING AND TESTING WITH CIRCUITS

This chapter discusses the packaging of the microfabricated components and

their testing within power converter circuits. Two distinct tests are described. The

first utilized wire bonds to connect a microfabricated inductor onto a printed circuit

board (PCB) to be tested with a prototype 100 MHz hybrid boost converter circuit

implemented in Complimentary Metal Oxide Semiconductor (CMOS). The second

test involved microfabrication of an inductor alongside a routing circuit for connection

to a commercially-available ball grid array (BGA) boost converter and surface-mount

capacitors. In this second test, the substrate on which the inductor and copper

framework were fabricated was completely removed, resulting in a converter module

with minimal packaging overhead.

8.1 Microinductor Wire Bonded to Very High Frequency Boost C onverter

For preliminary testing with a very high frequency (VHF) power converter, a

processed Pyrex wafer was diced to form a separate chip containing a microfabricated

inductor. The chip was then affixed to a printed circuit board (PCB) with tape and the

terminal ends of the microinductor were electrically connected to the pads on the PCB

with gold wire bonds. The attached chip, depicted in Figure 8-1, was then encapsulated

in epoxy for mechanical protection.

8.1.1 About the Microinductor

The inductor utilized for this test had an outer diameter D = 520 ”m, winding

trace width w = 40 ”m, spacing between traces s = 20 ”m, two winding layers, each

with thickness t = 10 ”m, and n = 3 winding turns per layer. The impedance of

this inductor was measured with a vector network analyzer (VNA) in its as-fabricated

state before wire bonding and was then measured with an impedance analyzer after

attachment to the PCB and after wire bonding. The measurements with the VNA

were taken by probing directly to the copper pads at the terminals of the inductor and

142

Page 143: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

represented the impedance through only the microfabricated copper inductor. A nominal

inductance of 14nH and resistance of 0.8 Ω were so measured with the VNA. The

probes of the impedance analyzer however were landed on the tin PCB pads, and so

the measurement included the impedances through the wire bonds in series with the

inductor. As plotted in Figure 8-2, the wire bonds added approximately 7 nH to the

inductance through the device as measured at the PCB pads and also contributed an

additional 0.2 Ω to the resistance.

Figure 8-1. Microinductor wire bonded to a PCB for testing with 100 MHz CMOS hybridboost converter.

8.1.2 About the Converter and Test Results

The hybrid boost converter consisted of a single switched-inductor boost stage

followed by two switched capacitor stages. It was fabricated in a 130 nm 1.2 V CMOS

process and operated at 100 MHz. Details of the converter implementation were

presented by Xue et al. [68]. With the fabricated microinductor connected in the

switched-inductor stage, the converter achieved a conversion ratio of 6 from a 1.2 V

source with up to 37% efficiency [69]. The measured converter efficiency was plotted

in Figure 8-3 as a function of load current for conversion ratios of 6 (Vout = 7 V) and

143

Page 144: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

104

106

108

100

101

102

Frequency (Hz)

Indu

ctan

ce (

nH)

Impedance Analyzer

Network Analyzer

A Inductance

104

106

108

10−1

100

101

Frequency (Hz)

Res

ista

nce

(Ω)

Impedance Analyzer Network Analyzer

B Resistance

104

106

108

10−2

100

102

Frequency (Hz)

Qua

lity

Fac

tor

Impedance Analyzer

Network Analyzer

C Quality Factor

Figure 8-2. Plots of impedance vs. frequency for microfabricated inductor used with100 MHz hybrid boost converter. Data measured with network analyzerbefore wire bonding to PCB and measured with impedance analyzer afterwire bonding to PCB.

144

Page 145: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

8 (Vout = 10 V) [69]. For comparison, the measured efficiency was also plotted for

the converter when using a commercially-available 43 nH surface-mount inductor.

Although the values of inductance between the two did not match, the overall efficiency

was very similar at load currents less than 1 mA between the converter using the

microinductor and that using the surface-mount inductor. At load currents greater than

1 mA, the efficiencies between the two began to diverge, with the microfabricated

inductor resulting in greater loss, due likely to its greater series resistance.

0%

5%

10%

15%

20%

25%

30%

35%

40%

0 0.5 1 1.5 2 2.5 3

fsw=50MHz, Vout=7V

fsw=50MHz, Vout=10V

fsw=45MHz, Vout=7V

fsw=45MHz, Vout=10V

Microfabricated

inductor L=25nH

Surface-mount

inductor L=43nH

Load Current (mA)

Eff

icie

nc

y

Figure 8-3. Measured efficiencies of converter as a function of load current with 1.2 Vinput source and 7 V and 10 V output voltage. Plots compare efficiencies ofconverter using microfabricated inductor to that using a surface-mountinductor. Adapted from [69].

8.2 Testing with Commercial Surface-Mount Converter

Due to the experimental nature of the very high frequency (VHF) converters,

subsequent testing of the microinductors was done using a commercially-available

converter: the Texas Instruments TPS61240 step-up converter with fixed 5 V output. In

addition to the inductor, the converter chip also required two surface-mount capacitors,

one each at the input and output nodes for filtering. To connect the microinductor,

converter chip, and capacitors a packaging method was devised to utilize the same

145

Page 146: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

multilevel copper process to form a routing and interconnect framework surrounding the

inductors under fabrication. The multilevel copper, including both inductors and routing,

was then encapsulated and detached from the substrate, resulting in a flexible film with

intricate, embedded copper traces. The converter and capacitors were then attached to

the copper to form the completed test module.

8.2.1 About the Texas Instruments TPS61240 Converter

Featuring a switching frequency of up to 4.5 MHz, the TPS61240 was selected

because it featured the fastest switching frequency amongst boost converters that were

commercially-available at the time of this publication. As a result of the high switching

frequency, the datasheet for the converter called for an external inductor with a value of

only 1 ”H, which, while high for microinductors, was considerably less than that required

for many other converters. The switching frequency of the TPS61240 varied at run time

depending on the load current. At light loads, the converter operated in Pulse Frequency

Modulation (PFM) mode. In PFM mode, the converter switched the inductor with a train

of several pulses only as necessary to maintain an output greater than an internally-set

threshold voltage. When the output current was too great to be supported by PFM

mode, the converter automatically switched to Pulse Width Modulation (PWM) mode.

The converter chip was obtained in a die-sized ball grid array (DSBGA) package with six

solder bumps arranged on its underside with 0.4 mm pitch.

8.2.2 Module Design and Processing

The test platform for the converter consisted of a microinductor and routing that

were fabricated on a silicon wafer using the multilevel copper process detailed in

Chapter 5 with three layers of copper, each 30 ”m thick. The layout illustrated in Figure

8-4A shows the outline of the module formed in copper. Pads were also formed in

the copper for connection to the six solder bumps of the converter chip and to the

ends of the two surface-mount capacitors. The microindutor consisted of two stacked

square-shaped spiral windings with 992 ”m outer diameter, 32 ”m trace width, 20 ”m

146

Page 147: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

spacing between traces, and 6 turns per layer. The scanning electron micrograph (SEM)

image in Figure 8-4B depicts the copper structure of the module after the molding

process. Although the manufacturer’s datasheet for the TPS61240 converter advised to

use an inductor with a minimum inductance of 1 ”H [70] the microinductor was designed

for a value of 130 nH out of concerns over the high series resistance through greater

valued microinductors.

A CAD drawing B SEM of fabricated copper

Figure 8-4. Copper layout of converter module as drawn in Computer-Aided Design(CAD) software and as fabricated on a silicon wafer.

The electroplated copper structures on the wafer surface were then coated with

Brewer Science A2-22 resist. This resist was selected for its chemical resistance to

hydrofluoric acid, which was used to separate the copper framework from the silicon

fabrication substrate. Prior to electroplating copper, the silicon wafer was first coated

by plasma-enhanced chemical vapor deposition (PECVD) of a 2 ”m thick layer of

silicon dioxide. Then a layer of titanium, which had served the purpose of improving

the adhesion of copper to the silicon dioxide, was sputter-deposited to an increased

thickness of 300nm. The silicon dioxide and titanium constituted a sacrificial layer that

was etched away in 49% hydrofluoric acid to detach the encapsulated copper from the

147

Page 148: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

silicon substrate. Figure 8-5 depicts an example of an embedded copper framework that

was detached from a wafer.

Figure 8-5. Photograph of microfabricated copper framework encapsulated in epoxy andreleased from substrate by sacrificial layer etch.

After detachment, the bottommost layer of the multilevel copper was directly

accessible on the underside of the module, enabling electrical connection of the external

components to the copper. For stability while connecting to components and while

probing, the embedded copper film was temporarily affixed bottom-side-up with a drop of

silicone to an aluminum nitride chip. Attachment of the components was accomplished

by depositing solder paste onto the pads of the copper framework where the converter

chip and two filtering capacitors were manually positioned. The solder in the paste was

then reflowed on a hotplate set to 210 C to fix the components in place.

Figure 8-6 depicts the final converter module as it was tested with the TPS61240

converter chip and two 4.7 ”F surface mount capacitors soldered onto pads of the

embedded copper framework. As seen in the photographs in Figure 8-6, the surface

148

Page 149: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

A Top view B Side view

Figure 8-6. Photographs of functional converter module with controller and capacitorsmounted onto multilayered copper framework.

Table 8-1. Component sizes in functional converter module.Component Lateral area Thickness VolumeTPS61240 converter 1.26 × 0.86 = 1.08 mm2 625 ”m 0.67 mm3

Filtering capacitors (each) 1 × 0.5 = 0.5 mm2 500 ”m 0.25 mm3

Microinductor 1.04 × 0.99 = 1.03 mm2 90 ”m 0.09 mm3

Total 3 × 3 = 9 mm2 700 ”m 1.98 mm3

mount components were several times thicker than the copper framework. The footprint

of the module dominated by the three pads implemented in the copper framework that

were connected to the input, output, and ground nodes of the converter. While the

inductor also had a relatively large footprint at roughly 1 mm2, an inventory of the sizes

of the various components (listed in Table 8-1) revealed that, owing to its thinness, the

inductor comprised only 4.5% of the total volume of the converter module.

8.2.3 Converter Module Testing

Before the converter chip and capacitors were connected to the module, the

microinductor (embedded in ProTek A2-22 resist and detached from the silicon

substrate) was first characterized using a Rohde & Schwarz ZVA/B vector network

analyzer (VNA) to extract its impedance with frequency swept from 1 MHz–1 GHz. The

149

Page 150: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

inductance, resistance, and quality factor of the inductor were plotted as functions of

frequency in Figure 8-7. The measurements showed an inductance of 124 nH and a

resistance of 0.88 Ω at low frequencies (< 2.5 MHz), and peak quality factor of 12.4 at

177 MHz. A quality factor of 3.2 was measured at 4 MHz.

With the TPS61240 converter chip and the two 4.7 ”F capacitors soldered onto the

module, needle probes were used to make electrical connection to the input, output, and

ground pads implemented in the copper framework. Two Keithley 2400 SourceMeters

were used in the test setup: one at the input acted as the source to regulate the input

voltage Vin and to measure the input currents Iin, and one at the output acted as the

load to regulate the output current at set values Iout and to measure the resulting output

voltages Vout . Input and output powers were calculated from the set and measured

voltages and currents reported by each of the SourceMeters. Efficiency ηconv was

calculated as the ratio of the output power Pout to the input power Pin,

ηconv =Pout

Pin=Vout Iout

VinIin. (8–1)

An additional high impedance probe connected to a LeCroy oscilloscope was used to

measure the voltage waveforms across various nodes in the circuit.

The measured converter efficiency ηconv was plotted in Figure 8-8 as a function of

output current for several values of input voltage. In every case the output voltage was

consistently regulated at a value near Vout = 5.05 V. The peak measured efficiency of

62.4% was obtained for an input voltage Vin = 4 V and an output current Iout = 5 mA.

The maximum measured power output was 152 mW at 58.2% efficiency with an input

voltage Vin = 4 V and an output current Iout = 30 mA. As seen in Figure 8-8, the

overall converter efficiency decreased considerably with decreasing input voltage (i.e.

increasing conversion ratio) but remained comparatively consistent across the range of

output currents. At higher output currents than those plotted in Figure 8-8 the converter

150

Page 151: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

106

107

108

109

101

102

103

Indu

ctan

ce (

nH)

Frequency (Hz)

A Inductance

106

107

108

109

100

102

104

Res

ista

nce

(Ω)

Frequency (Hz)

B Resistance

106

107

108

109

0

5

10

15

Qua

lity

Fac

tor

Frequency (Hz)

C Quality Factor

Figure 8-7. Plots of impedance vs. frequency for microfabricated inductor used inconverter module with TPS61240 converter. Data measured with VNA afterinductor was embedded in ProTek A2-22 resist and detached from siliconsubstrate.

151

Page 152: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

0 10 20 3030

40

50

60

70

Output current (mA)

Effi

cien

cy (

%)

Vin

=3.5V

Vin

=4.0V

Vin

=3.0V

Figure 8-8. Measured efficiency vs. output current of converter module with outputvoltage regulated at 5.05 V.

exhibited severe ripple in the output voltage, which indicated control loop instability as

the converter transitioned from the PFM mode of operation to PWM.

+−Vin

L

Cout

+ VL −

Cin

+

Vout

−

Figure 8-9. Boost converter circuit diagram with marked voltages corresponding toreported measured waveforms.

Measurement of the voltage waveform across the inductor, VL as marked on

the circuit diagram in Figure 8-9, provided insight into the trends in efficiency across

operating points. As in Figure 8-10, a comparison of the inductor voltage waveforms

for different input voltages revealed a similar series of pulses in each case. The duty

cycle, the fraction of each period during which voltage was applied across the inductor,

increased with increasing conversion ratio. A pulse width of 67 ns was measured with

152

Page 153: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

0 0.5 1 1.5−6

−4

−2

0

2

4

6

8

Time (”s)

Indu

ctor

Vol

tage

(V

)

A Vin = 5 V

0 0.5 1 1.5−6

−4

−2

0

2

4

6

8

Time (”s)

Indu

ctor

Vol

tage

(V

)

B Vin = 4 V

0 0.5 1 1.5−6

−4

−2

0

2

4

6

8

Time (”s)

Indu

ctor

Vol

tage

(V

)

C Vin = 3 V

Figure 8-10. Voltage waveforms measured across the inductor in the converter modulefor several different values of input voltages.

153

Page 154: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

Vin = 4 V, while a width of 124 ns was measured with Vin = 3 V. The longer duration

of applied voltage across the inductor resulted in reduced efficiency as there was more

time for the current through the inductor to reach its peak value, at which point no further

energy would have been stored magnetically but rather would have been dissipated

through the resistance of the coil. The current iL through a semi-ideal inductor with

inductance L and series resistance R in response to a step voltage input vin can be

represented as a function of time t by the expression,

iL (t) =vin

R

(

1− e−RL t)

. (8–2)

From the above expression, the time taken for current through an inductor to reach its

peak value is determined by the ratio R/L, with greater R and lower L contributing to

a shorter time taken to reach peak current. This helps to explain why the drop-off in

efficiency at higher conversion ratios was more drastic with the greater R/L ratio of the

microinductor than was portrayed in the datasheet with a 1 ”H surface-mount inductor

with series resistance 80 mΩ [70].

Whereas increasing the conversion ratio increased the pulse duty cycle, increasing

the output current increased only the frequency at which the series of pulses were

issued. As can be seen in comparing Figures 8-11A and 8-11C, the pulse series were

nearly identical regardless of output current. However, sampling over a longer period of

time as in Figures 8-11B and 8-11D showed that the series of pulses were issued with

greater frequency at greater output currents. The pulse frequency was also seen in the

ripple of the output voltage as in Figure 8-12, with the ripple remaining relatively equal in

magnitude across various load currents but increasing in frequency with increasing load

current.

8.3 Summary of Inductor Packaging and Testing within Conver ter Circuits

The two tests presented in this chapter demonstrated the viability of microfabricated

inductors to be used in next-generation very-high-frequency switched-mode power

154

Page 155: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

0 0.5 1 1.5−4

−2

0

2

4

6

8

Time (”s)

Indu

ctor

Vol

tage

VL (

V)

A Load current Iout = 10 mA

0 5 10 15 20−4

−2

0

2

4

6

8

Time (”s)

Indu

ctor

Vol

tage

VL (

V)

B Load current Iout = 10 mA

0 0.5 1 1.5−4

−2

0

2

4

6

8

Time (”s)

Indu

ctor

Vol

tage

VL (

V)

C Load current Iout = 30 mA

0 5 10 15 20−4

−2

0

2

4

6

8

Time (”s)

Indu

ctor

Vol

tage

VL (

V)

D Load current Iout = 30 mA

Figure 8-11. Measured voltage waveforms across the inductor for an input voltageVin = 4.0 V and output currents of Iout = 10 mA and Iout = 30 mA.Waveforms at right show single series of pulses for each load current.Longer time sampled on plots at right show multiple series of pulses.

0 20 40 60 80 1004.9

4.95

5

5.05

5.1

5.15

5.2

Time (”s)

Out

put V

olta

ge (

V)

A Load current Iout = 1 mA

0 20 40 60 80 1004.9

4.95

5

5.05

5.1

5.15

5.2

Time (”s)

Out

put V

olta

ge (

V)

B Load current Iout = 5 mA

Figure 8-12. Measured voltage waveforms at output for different load currents with inputvoltage Vin = 3.0 V.

155

Page 156: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

converters and to enable new technologies for integrating all components with minimal

packaging overhead:

‱ A 14 nH microinductor was attached to a printed circuit board (PCB) usingwire bonds and tested within a 100 MHz boost converter implemented inComplementary Metal Oxide Semiconductor (CMOS). A maximum efficiencyof 37% was obtained with a conversion ratio of 6.

‱ A microinductor was fabricated with 3× thicker copper layers to deliver 124 nH withseries resistance similar to that of the 14 nH inductor in the previous test — an 8×improvement in the inductance-to-resistance ratio. The thicker copper also enableda packaging solution whereby the microinductor and a copper routing frameworkwere embedded in resist and the fabrication substrate was removed. The inductorwas tested with a surface-mount converter and capacitors soldered onto theembedded copper. The surface-mount components together accounted for 13×greater volume than that of the microinductor. A maximum efficiency of 62% wasobtained with a conversion ratio of 1.26. Accounting for the total volume of allconverter components, a maximum power density of 76 mW/mm3 was recordedfor the overall system.

156

Page 157: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

CHAPTER 9CONCLUSION

This dissertation documented the realization of microscale air-core inductors and

transformers with high inductance densities and high efficiency for next-generation

miniaturized power converters with switching frequencies on the order of 100 MHz.

Fabrication of the components was enabled by an advanced microfabrication process

that was devised specifically to address limitations that had so far prevented air-core

microinductors from being integrated with power converters.

This chapter highlights the accomplishment of the effort to date and then outlines

several ways in which future work can build on this foundation to enable new capabilities.

9.1 Summary of Work

The inductors and transformers of this work filled the void in high quality microscale

inductive components for power converters in the very high frequency (VHF) range.

While microscale components based on magnetic films had already provided high

quality at frequencies < 10 MHz and air-core inductive components have been used

in communications applications at frequencies > 1 GHz, the multilayered thick-film

microfabricated inductors and transformers demonstrated excellent performance in

the frequencies spanning the gap between these pre-existing groups. Comparing

results obtained from the new inductors to those of prior magnetic film and air-core

examples as in Figure 9-1, the newly developed components have been shown to exhibit

excellent inductance densities and quality factors in the VHF frequency range that had

not previously been addressed by microscale inductors. The results from the multilayer

thick-film microtransformers have been similarly outstanding when compared to previous

examples as in Figure 9-2.

The research work has resulted in several major accomplishments:

‱ A methodology was established for designing microinductors and microtransformerswith an emphasis on high density and high efficiency in power conversionapplications.

157

Page 158: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

‱ The design methods were shown to accurately predict the values of inductanceand resistance of a variety of components with stacked windings in sizes from0.25–2.5 mm2 and inductances from 15–350 nH.

‱ A new microfabrication process was devised that enabled the multilayer constructionof three-dimensional, high-density, freestanding copper structures with layerthicknesses from 10 ”m up to 30 ”m.

‱ Air-core inductors were fabricated in the multilayer process with measuredinductance densities up to 170 nH/mm2 and quality factors as high as 33.

‱ Air-core transformers were fabricated with inductance densities up to 325 nH/mm2

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

‱ Microfabricated inductors were tested within functional power converter circuits:a prototype step-up converter with a switching frequency of 100 MHz yielded aconversion ratio of 6 at up to 37% efficiency using a 14 nH microinductor; and acommercial surface-mount step-up converter with a maximum switching frequencyof 4 MHz yielded a conversion ratio of 1.26 at up to 62% efficiency using a 124 nHmicroinductor.

‱ A substrate detachment process was devised to form a miniaturized powerconverter module with an embeded microfabricated inductor and interconnectstructure to which a commercial surface-mount converter and capacitors wereattached and tested. The power converter module produced a maximum outputpower of 152 mW from a package occupying in total less than < 2 mm3.

9.2 Lessons Learned

In addition to the major accomplishments listed above, this research also yielded

important findings that should shape any future efforts to improve microscale air-core

power inductors and transformers:

‱ Increasing the thickness of each copper winding layer from 10 ”m to 30 ”m yieldedinductors and transformers that had significantly improved direct current (dc)resistances with only slightly decreased inductances. However, the benefit ofthe thicker layers at dc was lost at frequencies greater than about 100 MHz asthe resistances of the thicker windings increased more rapidly with increasingfrequency due to increased eddy current losses in the copper. Measurements werecompared between inductors of different thicknesses in Section 6.3.2.1.

‱ The bulk removal of dielectric material (e.g. photoresist) from between theupper and lower winding layers of stacked inductors increased the self-resonantfrequencies by about 10–20%.

158

Page 159: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

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

I2 I1

I3

I4

I5 I6 I7

I9

1

10

100

1 10 100 1000 10000

Pe

ak

Qu

ality

Fa

cto

r

Frequency for Peak Quality Factor (MHz)

Figure 9-1. Measured results of microfabricated inductors presented in this work(orange) plotted in terms of peak quality factor and the frequency at whichthe peak quality factor was obtained. Results are compared to reviewedmagnetic-film (blue) and air-core (green) inductors. Bubble size isproportional to inductance density.[8–22]

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

Meyer, 1:1

Meyer, 1:3.5

0%

20%

40%

60%

80%

100%

1 10 100 1000 10000

Eff

icie

nc

y

Frequency for Maximum Efficiency

Figure 9-2. Measured results of microfabricated transformers presented in this work(orange) plotted in terms of maximum efficiency and the frequency at whichmaximum efficiency was obtained. Results are compared to reviewedmagnetic-film (blue) and air-core (green) transformers. Bubble size isproportional to voltage gain.[23–40]

159

Page 160: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

‱ Changing the shape of the microinductors from square to circular resulted in onlyminor improvements in inductance-to-resistance ratios on the order of 10%.

‱ Segmenting the thick traces of an inductor horizontally into several paralleltraces of thinner widths was found to yield an almost immeasurable change tothe impedance of the inductor. The concept of filamented traces, motivated bythe stranded Litz wires found in larger scale high-frequency transformers, wasineffective in reducing the eddy currents in the windings. In order to reduce theeddy currents induced by the proximity effect between adjacent conductors, thewindings would have to be crossed in manners that increase the orthogonality ofadjacent current paths, as is achieved in Litz wires by twisting or braiding individualstrands to reduce bundle-level effects.

‱ The characterization of inductors fabricated on silicon substrates demonstratedthe strong influence of capacitive coupling on the measured performancecharacteristics. Increased substrate resistance was found to dampen the resonantbehavior of the inductors and to decrease the peak quality factor.

9.3 Future Work

Building off the success presented in Chapter 8.2 in embedding and detaching

microfabricated copper inductors and interconnects from silicon wafers, the multilevel

copper process could enable a new platform for embedding the three-dimensional (3D)

inductors and transformers side-by-side with heterogeneous components all within a

high-density package.

The envisioned packaging process would consist of the following steps, which are

illustrated in Figure 9-3:

1. Fabrication of thick copper inductor, transformer, and routing framework on top ofan oxide-coated silicon substrate.

2. Population of the framework with active circuits and surface-mount componentssnapped into copper sockets.

3. Filling and coating of the framework and components with an epoxy-based pottingcompound.

4. Release of the filled package from the silicon substrate by etching the oxide outfrom between the composite and the substrate.

In this way the thick 3D inductors and transformers would placed side-by-side with

the active circuits, rather than on top as has been the case with monolithic integration or

160

Page 161: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

A Copper framework on silicon.

B Components placed into sockets.

C Package filled with potting compound.

D Package released by sacrificial etch. E Underside showing input/output pads.

Figure 9-3. Illustrations of package assembly process.

161

Page 162: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

chip stacking (3D integration). This configuration is advantageous for avoiding capacitive

and inductive coupling of the passives with the conductive substrates of the active

circuits.

One unique feature of this packaging technology would be the ability to form the

copper into sockets that would allow tiny components to be snap-fit into place using

with deformable copper to provide electrical conncection. This ability was motivated by

the fact that the state-of-the-art surface-mount component size as of this writing, metric

0402 (imperial 01005), with a nominal footprint of 0.4 mm× 0.2 mm, had dimensions of a

similar order as the thickness of the electroplated copper stack. These tiny components

have been difficult to use in industry due to misalignment and the tomb-stoning that

can occur from imbalances between the wetting characteristics at the terminals of the

component [71–74]. While the film-type surface-mount inductors of this size still have

too great of resistance (e.g., up to 8 Ω for a 68 nH inductor with a 0.6 mm × 0.3 mm

footprint [75]), the more-energy-dense surface-mount capacitors would be well-suited for

high-frequency power converters.

As a test of this concept, a power package framework was fabricated in three layers

of copper, each 30 ”m thick, that consisted of the high-density multilayer inductors

and transformers that have been discussed throughout this work along with sockets to

accept metric 0402-sized resistors and capacitors. Two such sockets are depicted in

the scanning electron micrograph (SEM) image of Figure 9-4A. In the topmost copper

layer, 20 ”m-long 10 ”m-wide teeth protruded into the socket area as depicted in Figure

9-4B. The teeth served several purposes. The first was to account for deviations in

the exact sizes of the components resulting from imperfect manufacturing tolerances.

The second purpose was to physically secure and electrically contact the component

after placement. By securing the component in place, a solder reflow could then be

performed to ensure durable contact without the possibility of the component shifting or

tomb-stoning during the reflow.

162

Page 163: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

A Two 0402 sockets B Close-up of teeth

Figure 9-4. Scanning electron micrograph (SEM) images of metric 0402-sized coppersockets and a close-up of the copper teeth that protrude into the socket areato contact the surface-mount component.

By sizing the socket larger than the largest size expected of a component and sizing

the teeth longer than the deviation in expected component sizes, the teeth could deform

to contact a component over the entire range of sizes tolerated in its specification.

Figure 9-7 shows how the copper teeth deformed and buried into the tin contact of a

surface-mount component that was pressed to fit. The concept of deformable copper

teeth could be extended to form contacts with vertically-oriented chips.

A Horizontal deformation B Vertical deformation

Figure 9-5. Scanning electron micrograph (SEM) images of the deformable copper teethcontacting the tin pad of a surface-mount component.

163

Page 164: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

Figure 9-6. SEM image depicting a surface-mount resistor and capacitor mounted intosockets alongside microfabricated high-density inductors.

Finally, Figure 9-6 shows a surface-mount resistor and capacitor placed into sockets

alongside a bank of four copper microinductors. To demonstrate that electrical contact

was made to these components, the impedance was measured with an Agilent HP

4294a impedance analyzer by landing probes on the copper sockets to either side of the

components. The resistance of a 47 Ω resistor and the measured capacitance of a 47 pF

capacitor were measured after each had been socketed. The measurements, plotted in

Figure 9-7 matched the expected values of these components. The sharp deviations at

the higher frequencies were the result of resonance in the measurement cables.

Forming sockets to embed surface-mount components within high-density packages

is only one example of how the versatility of the multilayer thick film copper process in

producing fine-featured 3D conductive parts could enable new capabilities across a wide

range of applications.

164

Page 165: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

104

106

45

46

47

48

49

50

Frequency (Hz)

Res

ista

nce

(Ω)

A Resistor

104

106

42

43

44

45

46

47

Frequency (Hz)

Cap

acita

nce

(pF

)

B Capacitor

Figure 9-7. Measured resistance of 47 Ω resistor and capacitance of 47 pF capacitorafter each had been socketed.

165

Page 166: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

REFERENCES

[1] S. K. Mathew, M. A. Anders, B. Bloechel, T. Nguyen, R. K. Krishnamurthy,and S. Borkar, “A 4-GHz 300-mW 64-bit integer execution ALUwith dual supply voltages in 90-nm CMOS,” IEEE Journal of Solid-StateCircuits, vol. 40, no. 1, pp. 44–51, Jan. 2005. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=1374989

[2] P. Hazucha, G. Schrom, J. Hahn, B. A. Bloechel, P. Hack, G. E. Dermer, S. Narendra,D. Gardner, T. Karnik, V. De, and S. Borkar, “A 233-MHz 80%-87% efficientfour-phase DC-DC converter utilizing air-core inductors on package,” IEEE Journalof Solid-State Circuits, vol. 40, no. 4, pp. 838–845, Apr. 2005. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=1424213

[3] B. Morgan, S. Bedair, J. S. Pulskamp, R. G. Polcawich, C. Meyer, C. Dougherty,X. Lin, D. Arnold, R. Bashirullah, R. Miller, and M. Roosz, “Power considerationsfor mast platforms,” in Micro- and Nanotechnology Sensors, Systems, andApplications II, vol. 7679, no. 1. SPIE, 2010, p. 76790V. [Online]. Available:http://link.aip.org/link/?PSI/7679/76790V/1

[4] R. W. Erickson and D. Maksimovic, Fundamentals of Power Electronics, 2nd ed.Springer, 2001.

[5] J. Salvia, J. A. Bain, and C. P. Yue, “Tunable on-chip inductors up to 5 GHz usingpatterned permalloy laminations,” IEEE InternationalElectron Devices Meeting,2005. IEDM Technical Digest., vol. 00, no. 1, pp. 943–946, 2005. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=1609516

[6] J. M. Rivas, D. Jackson, O. Leitermann, A. D. Sagneri,Y. Han, and D. J. Perreault, “Design Considerations for VeryHigh Frequency dc-dc Converters,” in 37th IEEE Power ElectronicsSpecialists Conference. IEEE, 2006, pp. 1–11. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=1712114

[7] D. S. Gardner, G. Schrom, F. Paillet, B. Jamieson, T. Karnik, and S. Borkar,“Review of On-Chip Inductor Structures With Magnetic Films,” IEEE Transactionson Magnetics, vol. 45, no. 10, pp. 4760–4766, Oct. 2009. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=5257291

[8] J. Y. Park and M. G. Allen, “Packaging-compatible high Q microinductorsand microfilters for wireless applications,” IEEE Transactions on AdvancedPackaging, vol. 22, no. 2, pp. 207–213, May 1999. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=763193

[9] Y. Fukuda, T. Inoue, T. Mizoguchi, S. Yatabe, and Y. Tachi, “Planarinductor with ferrite layers for DC-DC converter,” IEEE Transactions onMagnetics, vol. 39, no. 4, pp. 2057–2061, Jul. 2003. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=1211182

166

Page 167: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

[10] N. Wang, T. O’Donnell, S. Roy, M. Brunet, P. McCloskey, and S. C. O Mathuna,“High-frequency micro-machined power inductors,” Journal of Magnetism andMagnetic Materials, vol. 290-291, pp. 1347–1350, Apr. 2005. [Online]. Available:http://linkinghub.elsevier.com/retrieve/pii/S0304885304016956

[11] T. Sato, K. Yamasawa, H. Tomita, T. Inoue, and T. Mizoguchi, “FeCoBN MagneticThin Film Inductor for MHz Switching Micro DC-DC Converters,” Transactions of theInstitute of Electrical Engineers of Japan, vol. 121-D, no. 1, pp. 84–89, 2001.

[12] M. Yamaguchi, K. Suezawa, Y. Takahashi, K. I. Arai, S. Kikuchi, Y. Shimada,S. Tanabe, and K. Ito, “Magnetic thin-film inductors for RF-integrated circuits,” Journalof Magnetism and Magnetic Materials, vol. 215-216, no. 1, pp. 807–810, Jun. 2000.[Online]. Available: http://linkinghub.elsevier.com/retrieve/pii/S0304885300002936

[13] J. Song, B. Min, D. Jeong, H. Kim, and J. Heo, “Characteristics of filmtype inductors using a FeZrBAg magnetic material,” Journal of Magnetism andMagnetic Materials, vol. 234, no. 3, pp. 494–498, Sep. 2001. [Online]. Available:http://linkinghub.elsevier.com/retrieve/pii/S030488530100261X

[14] B. Viala, S. Couderc, a.S. Royet, P. Ancey, and G. Bouche, “Bidirectionalferromagnetic spiral inductors using single deposition,” IEEE Transactions onMagnetics, vol. 41, no. 10, pp. 3544–3549, Oct. 2005. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=1519365

[15] D. J. Young, V. Malba, J.-J. Ou, A. F. Bernhardt, and B. E. Boser,“Monolithic high-performance three-dimensional coil inductors for wirelesscommunication applications,” in International Electron Devices Meeting.IEDM Technical Digest. IEEE, 1997, pp. 67–70. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=649460

[16] Y.-s. Choi, E. Yoon, and J.-B. Yoon, “Encapsulation of the micromachinedair-suspended inductors,” in IEEE MTT-S International Microwave SymposiumDigest, no. i. IEEE, 2003, pp. 1637–1640. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=1210452

[17] D.-H. Weon, J.-H. Jeon, and S. Mohammadi, “High-Q micromachinedthree-dimensional integrated inductors for high-frequency applications,”Journal of Vacuum Science & Technology B: Microelectronics and Nanome-ter Structures, vol. 25, no. 1, p. 264, 2007. [Online]. Available:http://link.aip.org/link/JVTBD9/v25/i1/p264/s1\&Agg=doi

[18] Y.-K. Yoon, J.-W. Park, and M. G. Allen, “Polymer-core conductorapproaches for RF MEMS,” Journal of Microelectromechanical Sys-tems, vol. 14, no. 5, pp. 886–894, Oct. 2005. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=1516170

[19] D. W. Lee, K.-P. Hwang, and S. X. Wang, “Design and fabrication of integratedsolenoid inductors with magnetic cores,” in 58th Electronic Components and

167

Page 168: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

Technology Conference. IEEE, May 2008, pp. 701–705. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=4550049

[20] D. Flynn, N. S. Sudan, a. Toon, and M. P. Y. Desmulliez, “Fabricationprocess of a micro-inductor utilising a magnetic thin film core,” MicrosystemTechnologies, vol. 12, no. 10-11, pp. 923–933, Apr. 2006. [Online]. Available:http://www.springerlink.com/index/10.1007/s00542-006-0138-7

[21] B. Orlando, R. Hida, R. Cuchet, M. Audoin, B. Viala, D. Pellissier-Tanon,X. Gagnard, and P. Ancey, “Low-Resistance Integrated ToroidalInductor for Power Management,” IEEE Transactions on Magnetics,vol. 42, no. 10, pp. 3374–3376, Oct. 2006. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=1704631

[22] C. H. Ahn, Y. J. Kim, and M. G. Allen, “A fully integrated planartoroidal inductor with a micromachined nickel-iron magnetic bar,”IEEE Transactions on Components, Packaging, and Manufacturing Technology:Part A, vol. 17, no. 3, pp. 463–469, 1994. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=311757

[23] M. Mino, T. Yachi, A. Tago, K. Yanagisawa, and K. Sakakibara, “A newplanar microtransformer for use in micro-switching converters,” IEEE Transactionson Magnetics, vol. 28, no. 4, pp. 1969–1973, Jul. 1992. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=144755

[24] H. Kurata, K. Shirakawa, O. Nakazima, and K. Murakami, “Solenoid-Type Thin-FilmMicro-Transformer,” IEEE Translation Journal on Magnetics in Japan, vol. 9, no. 3,pp. 90–94, May 1994. [Online]. Available: http://ieeexplore.ieee.org/xpls/absall.jsp?arnumber=4565864

[25] M. Mino, T. Yachi, a. Tago, K. Yanagisawa, and K. Sakakibara,“Planar microtransformer with monolithically-integrated rectifierdiodes for micro-switching converters,” IEEE Transactions on Magnet-ics, vol. 32, no. 2, pp. 291–296, Mar. 1996. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=486336

[26] M. Xu, T. M. Liakopoulos, and C. H. Ahn, “A microfabricated transformerfor high-frequency power or signal conversion,” IEEE Transactions onMagnetics, vol. 34, no. 4, pp. 1369–1371, Jul. 1998. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=706551

[27] C. R. Sullivan and S. R. Sanders, “Measured performance of ahigh-power-density microfabricated transformer in a DC-DC converter,”in PESC Record. 27th Annual IEEE Power Electronics Specialists Confer-ence. Baveno, Italy: Ieee, 2000, pp. 287–294. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=548594

168

Page 169: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

[28] M. Brunet, T. O’Donnell, L. Baud, N. Wang, J. O’Brien, P. McCloskey,and S. C. O’Mathuna, “Electrical performance of microtransformersfor DC-DC converter applications,” IEEE Transactions on Magnetics,vol. 38, no. 5, pp. 3174–3176, Sep. 2002. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=1042488

[29] J. Y. Park and J. U. Bu, “Packaging compatible microtransformerson a silicon substrate,” IEEE Transactions on Advanced Packaging,vol. 26, no. 2, pp. 160–164, May 2003. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=1233575

[30] R. J. Rassel, C. F. Hiatt, J. DeCramer, and S. A. Campbell, “Fabricationand characterization of a solenoid-type microtransformer,” IEEE Transactionson Magnetics, vol. 39, no. 1, pp. 553–558, Jan. 2003. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=1179917

[31] E.-J. Yun, M. Jung, C. I. Cheon, and H. G. Nam, “Microfabrication and Characteristicsof Low-Power High-Performance Magnetic Thin-Film Transformers,” IEEETransactions on Magnetics, vol. 40, no. 1, pp. 65–70, Jan. 2004. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=1264124

[32] N. Wang, T. O’Donnell, S. Roy, S. Kulkarni, P. Mccloskey, and C. O’Mathuna, “ThinFilm Microtransformer Integrated on Silicon for Signal Isolation,” IEEE Transactionson Magnetics, vol. 43, no. 6, pp. 2719–2721, Jun. 2007. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=4202864

[33] K. Yamaguchi, E. Sugawara, O. Nakajima, H. Matsuki, and K. Murakami,“Load characteristics of a spiral coil type thin film microtransformer,” IEEETransactions on Magnetics, vol. 29, no. 6, pp. 3207–3209, 1993. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=281138

[34] D. T. Cheung, J. R. Long, R. Hadaway, and D. Harame, “Monolithictransformers for silicon RF IC design,” in Proceedings of the1998 Bipolar/BiCMOS Circuits and Technology Meeting (Cat. No.98CH36198).Minneapolis, MN: Ieee, 1998, pp. 105–108. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=741890

[35] D. Laney, L. Larson, P. Chan, J. Malinowski, D. Harame, S. Subbanna, R. Volant,and M. Case, “Lateral microwave transformers and inductors implemented in aSi/SiGe HBT process,” in 1999 IEEE MTT-S International Microwave SymposiumDigest (Cat. No.99CH36282). Anaheim, CA: Ieee, 1999, pp. 855–858. [Online].Available: http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=779521

[36] J. R. Long, “Monolithic transformers for silicon RF IC design,” IEEE Journal ofSolid-State Circuits, vol. 35, no. 9, pp. 1368–1382, 2000. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=868049

169

Page 170: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

[37] R. P. Ribas, J. Lescot, J.-L. Leclercq, J. M. Karam, andF. Ndagijimana, “Micromachined microwave planar spiral inductorsand transformers,” IEEE Transactions on Microwave Theory and Tech-niques, vol. 48, no. 8, pp. 1326–1335, 2000. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=859477

[38] A. Zolfaghari, A. Chan, and B. Razavi, “Stacked inductors andtransformers in CMOS technology,” IEEE Journal of Solid-State Cir-cuits, vol. 36, no. 4, pp. 620–628, Apr. 2001. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=913740

[39] K. T. Ng, B. Rejaei, and J. N. Burghartz, “Substrate effects in monolithicRF transformers on silicon,” IEEE Transactions on Microwave Theory andTechniques, vol. 50, no. 1, pp. 377–383, 2002. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=981289

[40] H. A. Aly and B. Elsharawy, “Modeling and measurements of novelhigh k monolithic transformers,” in IEEE MTT-S International MicrowaveSymposium Digest, 2003. Ieee, 2003, pp. 1247–1250. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=1212595

[41] P. Li, L. Xue, D. Bhatia, and R. Bashirullah, “Digitally assisteddiscontinuous conduction mode 5V/100MHz and 10V/45MHz DC-DC boostconverters with integrated Schottky diodes in standard 0.13m CMOS,”in 2010 IEEE International Solid-State Circuits Conference Digest of TechnicalPapers (ISSCC), vol. 26, no. 7, 2010, pp. 206–207. [Online]. Available:http://ieeexplore.ieee.org/xpls/abs all.jsp?arnumber=5433983

[42] M. Yamaguchi, K. Suezawa, K. I. Arai, Y. Takahashi, S. Kikuchi, Y. Shimada,W. D. Li, S. Tanabe, and K. Ito, “Microfabrication and characteristicsof magnetic thin-film inductors in the ultrahigh frequency region,” Journalof Applied Physics, vol. 85, no. 11, p. 7919, 1999. [Online]. Available:http://link.aip.org/link/JAPIAU/v85/i11/p7919/s1\&Agg=doi

[43] C. P. Yue and S. S. Wong, “On-chip spiral inductors with patternedground shields for Si-based RF ICs,” IEEE Journal of Solid-State Cir-cuits, vol. 33, no. 5, pp. 743–752, May 1998. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=668989

[44] K. O, “Estimation methods for quality factors of inductors fabricated insilicon integrated circuit process technologies,” IEEE Journal of Solid-StateCircuits, vol. 33, no. 8, pp. 1249–1252, 1998. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=705364

[45] H. Lakdawala, X. Zhu, H. Luo, S. Santhanam, L. R. Carley, andG. K. Fedder, “Micromachined High-Q Inductors in a 0.18-m CopperInterconnect Low-K Dielectric CMOS Process,” IEEE Journal of Solid-State

170

Page 171: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

Circuits, vol. 37, no. 3, pp. 394–403, Mar. 2002. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=987092

[46] A. M. Niknejad and R. G. Meyer, “Analysis, design, and optimization ofspiral inductors and transformers for Si RF ICs,” IEEE Journal of Solid-StateCircuits, vol. 33, no. 10, pp. 1470–1481, 1998. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=720393

[47] C. K. Alexander and M. N. O. Sadiku, Fundamentals of Electric Circuits, 2nd ed.New York: McGraw-Hill, 2004.

[48] H. A. Wheeler, “Simple Inductance Formulas for Radio Coils,” Proceedings ofthe Institute of Radio Engineers, vol. 16, no. 10, pp. 1398–1400, 1928. [Online].Available: http://ieeexplore.ieee.org/xpls/abs all.jsp?arnumber=1669896

[49] S. S. Mohan, M. del Mar Hershenson, S. P. Boyd, and T. H. Lee,“Simple accurate expressions for planar spiral inductances,” IEEE Journal ofSolid-State Circuits, vol. 34, no. 10, pp. 1419–1424, 1999. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=792620

[50] E. B. Rosa, “Calculation of the self-inductances of single-layer coils,” Bulletin ofBureau Standards, vol. 2, no. 2, pp. 161–187, 1906.

[51] M. Kamon, M. J. Tsuk, and J. K. White, “FASTHENRY: a multipole-accelerated3-D inductance extraction program,” IEEE Transactions on Microwave Theoryand Techniques, vol. 42, no. 9, pp. 1750–1758, 1994. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=310584

[52] R. H. Jansen, L. Wiemer, H. Finlay, J. Suffolk, B. Roberts, and R. Pengelly,“Theoretical and Experimental Broadband Characterisation of Multiturn SquareSpiral Inductors in Sandwich Type GaAs MMICs,” in 15th European MicrowaveConference, 1985. IEEE, Oct. 1985, pp. 946–952. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=4132288

[53] C. Yue and S. Wong, “Physical modeling of spiral inductors on silicon,” IEEETransactions on Electron Devices, vol. 47, no. 3, pp. 560–568, Mar. 2000. [Online].Available: http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=824729

[54] H. Wheeler, “Formulas for the Skin Effect,” Proceedings of theIRE, vol. 30, no. 9, pp. 412–424, Sep. 1942. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=1694542

[55] G. Antonini, a. Orlandi, and C. Paul, “Internal impedance of conductorsof rectangular cross section,” IEEE Transactions on Microwave Theory andTechniques, vol. 47, no. 7, pp. 979–985, Jul. 1999. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=775429

171

Page 172: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

[56] “Understanding the fundamental principles of vector network analysis,” AgilentTechnologies, Application Note 1287-1, 2000.

[57] D. M. Pozar, Microwave Engineering, 3rd ed. Wiley, 2005.

[58] C. H. Mastrangelo and C. H. Hsu, “Mechanical stability and adhesionof microstructures under capillary forces. I. Basic theory,” Journal ofMicroelectromechanical Systems, vol. 2, no. 1, pp. 33–43, Mar. 1993. [Online].Available: http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=232593

[59] P. B. Chinoy and J. Tajadod, “Processing and microwavecharacterization of multilevel interconnects using benzocyclobutenedielectric,” IEEE Transactions on Components, Hybrids, and ManufacturingTechnology, vol. 16, no. 7, pp. 714–719, 1993. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=257862

[60] M. Ohnmacht, V. Seidemann, and S. Buttgenbach, “Microcoils andmicrorelays an optimized multilayer fabrication process,” Sensors and ActuatorsA: Physical, vol. 83, no. 1-3, pp. 124–129, May 2000. [Online]. Available:http://linkinghub.elsevier.com/retrieve/pii/S0924424700002867

[61] M. E. Mills, P. Townsend, D. Castillo, S. Martin, and A. Achen, “Benzocyclobutene(DVS-BCB) polymer as an interlayer dielectric (ILD) material,” MicroelectronicEngineering, vol. 33, no. 1-4, pp. 327–334, Jan. 1997. [Online]. Available:http://linkinghub.elsevier.com/retrieve/pii/S0167931796000615

[62] S. S. Bedair and G. K. Fedder, “Polymer Mass Loading ofCMOS/MEMS Microslot Cantilever for Gravimetric Sensing,” in 2007IEEE Sensors. Ieee, 2007, pp. 1164–1167. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=4388614

[63] “Thick resist processing,” MicroChemicals, Tech.Rep., November 2009. [Online]. Available:http://www.microchemicals.com/technical%5Finformation/thick%5Fresist%5Fprocessing.pdf

[64] M. Schlesinger and M. Paunovic, Eds., Modern Electroplating. New York: JohnWiley & Sons, Inc., 2000.

[65] B. F. Rothschild, “Factors Involved in the Development of Plating Solutions with HighThrowing Power,” Plating and Surface Finishing, vol. 66, p. 70, 1979.

[66] “Applying error correction to network analyzer measurements,” Agilent Technologies,Application Note 1287-3, 2002.

[67] J. Craninckx and M. Steyaert, “A 1.8-GHz low-phase-noise CMOSVCO using optimized hollow spiral inductors,” IEEE Journal of Solid-StateCircuits, vol. 32, no. 5, pp. 736–744, May 1997. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=568844

172

Page 173: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

[68] L. Xue, C. M. Dougherty, and R. Bashirullah, “50100 MHz, 8xstep-up DC-DC converters in 130nm 1.2V digital CMOS,” in 2011Twenty-Sixth Annual IEEE Applied Power Electronics Conference and Expo-sition (APEC). IEEE, Mar. 2011, pp. 892–896. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=5744700

[69] L. Xue, C. Meyer, C. M. Dougherty, S. Bedair, B. Morgan,D. P. Arnold, and R. Bashirullah, “Towards miniature step-uppower converters for mobile microsystems,” in 2011 Twenty-Sixth Annual IEEE Applied Power Electronics Conference and Exposition(APEC). IEEE, Mar. 2011, pp. 1451–1455. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=5744783

[70] TPS61240, TPS61241 3.5-MHz High Efficiency Step-Up Converter, TexasInstruments, P.O. Box 655303, Dallas, Texas 75265, May 2009.

[71] H. T. Ming, T. K. Ming, and L. Khor, “Tombstone reduction by reflowprofile optimization, SMT stencil design and pad design,” in 201034th IEEE/CPMT International Electronic Manufacturing Technology Symposium(IEMT), no. M. IEEE, Nov. 2010, pp. 1–5. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=5746716

[72] “0201 and 01005 adoption in industry,” DfR Solutions, Tech. Rep., 2011.

[73] R. Jarvina, S. Greiner, and R. Warren, “01005 SMT component assemblyfor wireless SIP modules,” Proceedings Electronic Components and Technology,2005. ECTC ’05., vol. 2, pp. 1500–1505, 2005. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=1441986

[74] J. Hoo, A. Lingley, R. Baskaran, X. Xiong, and K. F. Bohringer, “Parallelassembly of 01005 surface mount technology components with 100%yield,” in 2010 IEEE 23rd International Conference on Micro Electro MechanicalSystems (MEMS). IEEE, Jan. 2010, pp. 532–535. [Online]. Available:http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=5442449

[75] Chip Inductors, Murata Manufacturing Co., Ltd., 1-10-1, Higashi Kotari,Nagaokakyo-shi, Kyoto 617-8555, Japan, October 2011.

173

Page 174: MICROMACHINED INDUCTORS AND TRANSFORMERS FOR

BIOGRAPHICAL SKETCH

Christopher David Meyer was born October 23, 1983 in Ft. Lauderdale, Florida

to Patricia and Donald Meyer. He grew up in Coral Springs, Florida, not far from his

birthplace, and graduated from Coral Springs High School in 2002. He attended the

University of Florida (UF), where he earned the Bachelor of Science degree cum laude

in electrical engineering with a minor in German in 2006 and the Master of Science

degree in electrical engineering with a minor in mechanical engineering in 2009.

Inspired by a graduate course on Micro-Electromechanical Systems (MEMS),

Chris began his doctoral research with contributions to the development of thin-film

thermoelectric power generators. While continuing his research on power components,

Chris spent several years on internship at the U.S. Army Research Laboratory (ARL)

in Adelphi, MD. At ARL he developed the microfabrication process that enabled the

power components presented in this dissertation. Upon graduation with the Doctor of

Philosophy degree, Chris will join ARL full-time as an electronics engineer.

In November 2011 Chris married Jennifer nee Thompson, whom he had met as an

undergraduate student at UF.

174