A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

221
A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL SHEAR STRESS SENSOR By VIJAY CHANDRASEKHARAN 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 2009 1

Transcript of A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Page 1: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL SHEAR STRESSSENSOR

By

VIJAY CHANDRASEKHARAN

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

2009

1

Page 2: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

c© 2009 Vijay Chandrasekharan

2

Page 3: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

To my parents and grandparents, who have been my inspiration throughout

3

Page 4: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

ACKNOWLEDGMENTS

Financial support for this work was provided by NASA Langley Research Center and

the University of Florida. I thank my advisor Prof. Mark Sheplak for his role as a mentor,

both technically and otherwise. I would specially like to thank him for providing the right

kind of guidance and the intellectual freedom to realize my research interests. I am also

grateful to my committee members David Arnold, Lou Cattafesta, and Nam Ho Kim for

their technical inputs, helping me succeed in this project.

I am especially grateful to many of my former and present colleagues at the Inter-

disciplinary Microsystems Group (IMG). David Martin and Karthik Kadirvel helped me

understand interface circuits for capacitive sensors through insightful discussions, even when

they were pressed for time. I probably do not have better words to thank Jeremy Sells and

Jessica Meloy for their help with this project at the design implementation and testing stage.

The numerous hours of discussions and the efforts of Jeremy and Jessica really brought out

the best in us as a team and not just my efforts. I would like to thank Benjamin Griffin and

Brian Homeijer for all those hours of brainstorming and technical discussions. It was the

best time in graduate school in terms of the learning process that we started together back

in 2003. My special thanks to Matt Williams for his willingness to help numerous times,

may it be on mechanics or proof reading my dissertation. I am grateful to John Griffin for

his help with the experimental setup. I thank Brandon Bertolucci for his technical, as well

as his photographic assistance. I thank Sara Homeijer, Sheetal Shetye, Naigang Wang, and

Janhavi Agashe for their help with SEM images. I am also grateful to all of the students at

IMG. Honestly, I learnt a great deal from discussions with fellow graduate students at IMG

than from discussions with faculty. Learning process was never as easy and effective before

and perhaps may not be the same after IMG. I would like to thank the faculty at IMG for

facilitating such a research environment in the group.

I am thankful to Ken Reed at TMR engineering for his excellent and timely machining

work for sensor packages. Other technical assistance was provided at the University of

4

Page 5: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Florida by Prof. Ho-Bun Chan and his student Konstantinos Ninios with the supercritical

release and wire bonding. I would also like to thank Al Ogden for his willingness to help in

the cleanroom.

Last but foremost of all, I thank my parents, G.Chandrasekharan and Ranjani Chan-

drasekharan, and my sister, Sudha Chandrasekharan, for their support and encouragement

at every stage of graduate school. My parents are my role models and have taught me by

example, good work ethic, diligence, and practical thinking, which is directly reflected in the

successful outcome of this project.

5

Page 6: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

TABLE OF CONTENTS

page

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

LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

ABSTRACT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

CHAPTER

1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

1.1 Wall Shear Stress and Boundary Layers . . . . . . . . . . . . . . . . . . . . 181.1.1 Laminar Boundary Layer . . . . . . . . . . . . . . . . . . . . . . . . 211.1.2 Turbulent Boundary Layer . . . . . . . . . . . . . . . . . . . . . . . 221.1.3 Small Scales in Turbulence and Sensor Requirements . . . . . . . . 26

1.2 Research Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 331.3 Dissertation Organization . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

2 BACKGROUND . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

2.1 Shear Measurement Techniques . . . . . . . . . . . . . . . . . . . . . . . . 352.1.1 Indirect Measurement Techniques . . . . . . . . . . . . . . . . . . . 352.1.2 Direct Measurement Techniques . . . . . . . . . . . . . . . . . . . . 37

2.2 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

3 DEVICE MODELING . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

3.1 Quasi-Static Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 523.1.1 Small Deflection Analysis . . . . . . . . . . . . . . . . . . . . . . . . 543.1.2 Large Deflection Analysis . . . . . . . . . . . . . . . . . . . . . . . . 573.1.3 Electrostatic Behavior . . . . . . . . . . . . . . . . . . . . . . . . . . 58

3.2 Dynamic Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 823.2.1 Lumped Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . 833.2.2 Electromechanical Transduction . . . . . . . . . . . . . . . . . . . . 843.2.3 Equivalent Sensor Circuit . . . . . . . . . . . . . . . . . . . . . . . . 89

3.3 Higher Order Effects and Noise . . . . . . . . . . . . . . . . . . . . . . . . 953.3.1 Fringing Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 953.3.2 Parasitic Capacitance . . . . . . . . . . . . . . . . . . . . . . . . . . 973.3.3 Noise Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97

4 DESIGN OPTIMIZATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101

4.1 Sequential Quadratic Programming-SQP . . . . . . . . . . . . . . . . . . . 1014.2 Optimization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

4.2.1 Sensor Performance Requirements . . . . . . . . . . . . . . . . . . . 103

6

Page 7: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

4.3 Problem Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1044.3.1 Objective and Design Variables . . . . . . . . . . . . . . . . . . . . 1044.3.2 Constraints . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1074.3.3 Design Specifications . . . . . . . . . . . . . . . . . . . . . . . . . . 110

4.4 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1104.4.1 Sensitivity Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . 114

4.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119

5 DEVICE FABRICATION AND PACKAGING . . . . . . . . . . . . . . . . . . . 121

5.1 Fabrication Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1225.1.1 Floating Element Trench . . . . . . . . . . . . . . . . . . . . . . . . 1235.1.2 Seedless Electroplating of Nickel . . . . . . . . . . . . . . . . . . . . 1245.1.3 Backside Release . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1265.1.4 Die Separation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1265.1.5 The MEMS Shear Stress Sensor . . . . . . . . . . . . . . . . . . . . 128

5.2 Packaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129

6 EXPERIMENTAL CHARACTERIZATION . . . . . . . . . . . . . . . . . . . . 131

6.1 Experimental Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1316.1.1 Impedance Measurements . . . . . . . . . . . . . . . . . . . . . . . . 1316.1.2 Mean Shear Stress Measurement . . . . . . . . . . . . . . . . . . . . 1336.1.3 Dynamic Measurements . . . . . . . . . . . . . . . . . . . . . . . . . 1376.1.4 Noise Measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . 141

6.2 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1426.2.1 Impedance Measurements . . . . . . . . . . . . . . . . . . . . . . . . 1426.2.2 Mean Shear Stress Characteristics . . . . . . . . . . . . . . . . . . . 1446.2.3 Dynamic Characteristics . . . . . . . . . . . . . . . . . . . . . . . . 1486.2.4 Noise Characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . 1566.2.5 Experimental Uncertainty Estimation . . . . . . . . . . . . . . . . . 159

6.3 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160

7 CONCLUSIONS AND FUTURE WORK . . . . . . . . . . . . . . . . . . . . . . 162

7.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1627.2 Non-Idealities in Sensor Design and Characterization . . . . . . . . . . . . 164

7.2.1 Design Aspects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1647.2.2 Characterization Aspects . . . . . . . . . . . . . . . . . . . . . . . . 168

7.3 Recommendations for Future Sensor Designs . . . . . . . . . . . . . . . . . 175

APPENDIX

A MECHANICAL ANALYSIS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178

A.1 Beam Deflection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178A.1.1 Governing Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . 178A.1.2 Small Deflection Analysis . . . . . . . . . . . . . . . . . . . . . . . . 181

7

Page 8: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

A.1.3 Large Deflection Analysis . . . . . . . . . . . . . . . . . . . . . . . . 183A.2 Lumped Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192

A.2.1 Lumped Mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193A.2.2 Lumped Compliance . . . . . . . . . . . . . . . . . . . . . . . . . . 195

B TWO PORT ELEMENT MODELING . . . . . . . . . . . . . . . . . . . . . . . 198

B.1 Two Port Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198B.2 Linear Conservative Transducers . . . . . . . . . . . . . . . . . . . . . . . . 199

C SHEAR STRESS IN PWT WITH REFLECTIONS . . . . . . . . . . . . . . . . 205

D PROCESS TRAVELER AND PACKAGING DETAILS . . . . . . . . . . . . . . 209

REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214

BIOGRAPHICAL SKETCH . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221

8

Page 9: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

LIST OF TABLES

Table page

1-1 Laminar boundary layer parameters [8]. . . . . . . . . . . . . . . . . . . . . . . 22

3-1 Sensor geometry for effective tether length calculation. . . . . . . . . . . . . . . 66

4-1 Lower and upper bounds for design variables and flow specifications. . . . . . . 110

4-2 Constant parameters for optimization. . . . . . . . . . . . . . . . . . . . . . . . 110

4-3 Capacitive shear stress optimization results for Cp + Ci = 2.2 + 0.3 pF . . . . . . 112

4-4 Tolerance chart for variables and constants. . . . . . . . . . . . . . . . . . . . . 119

4-5 Design sensitivity based on Monte Carlo (MC) simulation. . . . . . . . . . . . . 119

6-1 Measurement settings for static calibration in the flow cell. . . . . . . . . . . . . 145

6-2 Sensitivity estimates using Monte Carlo technique on the mean shear stress mea-surement data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148

6-3 Measurement settings of B&K PULSE Multi-Analyzer System (Type 3109) fordynamic calibration in the PWT. . . . . . . . . . . . . . . . . . . . . . . . . . . 149

6-4 Spectrum analyzer settings for noise measurement in a double Faraday cage. . . 156

6-5 Integrated voltage noise floor of the sensor at different frequency ranges. . . . . 158

6-6 Comparison of predicted and measured sensor performance parameters. . . . . . 161

7-1 Comparison of measured sensor performance with previous work. . . . . . . . . 163

9

Page 10: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

LIST OF FIGURES

Figure page

1-1 Velocity profile of viscous fluid flow over a flat plate. . . . . . . . . . . . . . . . 19

1-2 Schematic of a 2D boundary layer for flow over a flat plate showing both thelaminar and turbulent regions (adapted from [8]). . . . . . . . . . . . . . . . . 22

1-3 Near wall, mean velocity in a turbulent boundary layer or law of the wall (adaptedfrom [11]). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

1-4 Schematic of energy distribution in turbulence as a function of wavenumber atsufficiently high Re (adapted from [13]). . . . . . . . . . . . . . . . . . . . . . . 28

1-5 Reynolds number for different flight regimes at sea level (adapted from [15]). . 29

1-6 Variation of Kolmogorov length and time scales with flow Reynolds number (x =1m,U∞ = 50m/s). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

1-7 Variation of shear stress and corresponding shear force on sensor area (Asensor =η2) with flow Reynolds number (x = 1m,U∞ = 50m/s). . . . . . . . . . . . . . . 31

2-1 Simplified 2-d schematic showing the plan view (top) and cross-section (bottom)of a typical floating element shear stress sensor structure (adapted from reviewby Naughton and Sheplak [6]). . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

2-2 Schematic of cross section of polyimide capacitive shear stress sensor with differ-ential capacitive readout scheme using integrated depletion mode P-MOS tran-sistors (adapted from Schmidt et al. [37]). . . . . . . . . . . . . . . . . . . . . . 40

2-3 Schematic of a) folded beam with lateral capacitive change and b) force feedbackcapacitive structure (adapted from Pan et al. [38]). . . . . . . . . . . . . . . . . 41

2-4 Top view of cantilever based floating element capacitive shear sensor (adaptedfrom Zhe et al. [44]). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

2-5 Cross sectional schematic of a)two diode and b)’split’ diode (3 diodes) opticalshutter based micromachined shear stress sensor (adapted from Padmanabhan’swork [21]). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

2-6 Conceptual schematic showing bottom and side view of fabry-perot floating ele-ment shear stress sensor (adapted from [52]). . . . . . . . . . . . . . . . . . . . . 46

2-7 Schematic of the optical Moire shear stress sensor. The shift of the amplifiedMoire pattern is imaged using a microscope attached to a CCD camera (withpermission from author [54]). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

10

Page 11: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

2-8 3D schematic of a piezoresistive sensor with top side tether implants oriented inthe direction of flow for high shear stress measurements (adapted from Ng et al.and Goldeberg et al. [56, 57]). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

2-9 3D schematic of a side implanted piezoresistive shear stress sensor. The implantsare such that two tethers are in tension while two are in compression for highersensitivity (with permission from author [60]). . . . . . . . . . . . . . . . . . . 49

3-1 Schematic of geometry of the differential, capacitive shear stress sensor. . . . . 53

3-2 Simplified mechanical model of the floating element structure. . . . . . . . . . . 55

3-3 Schematic of parallel plate capacitor analogous to overlapping comb fingers. . . 59

3-4 Simplified schematic of individual a)tether, b)comb finger, and c)floating elementcapacitances, respectively. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

3-5 Schematic of non-uniform tether capacitors with primary and secondary gaps. . 64

3-6 Differential capacitance sensing model for sensing capacitors with the asymmetricgaps. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

3-7 Simplified classification of capacitive interface electronic circuits. . . . . . . . . 69

3-8 Simplified sensor circuitry with a charge amplifier along with noise sources. . . 72

3-9 Simplified sensor circuitry with a voltage amplifier along with noise sources. . . 74

3-10 Simplified differential capacitance sense circuitry using synchronous modulationand demodulation technique with a voltage amplifier. . . . . . . . . . . . . . . 75

3-11 Schematic indicating spectra of the sensor output at each stage of the sensorcircuitry. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

3-12 Schematic of capacitive transduction scheme using a single sense capacitance. . . 86

3-13 Schematic of capacitive transduction scheme using differential capacitive sensingtechnique. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87

3-14 Schematic of equivalent circuit for the differential capacitive shear stress sensor. 89

3-15 Resonant modes of floating element structure. . . . . . . . . . . . . . . . . . . . 91

3-16 Asymmetric comb finger structure to estimate effect of fringing electric fields. . 96

3-17 Simplified schematic of sensor circuitry with noise sources. . . . . . . . . . . . . 98

4-1 Schematic of operating space of the sensor. . . . . . . . . . . . . . . . . . . . . . 103

4-2 Sensitivity of optimized MDS to design variables for Design 1. . . . . . . . . . . 115

11

Page 12: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

4-3 Sensitivity of optimized MDS and constraints due to change in optimized variablevalues of Design 1 (cont· · · ). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118

4-4 Sensitivity of optimized MDS, Sτ,overall, and fmin to tolerances in variables inconstants of Design 1 using Monte Carlo simulation. . . . . . . . . . . . . . . . 120

5-1 Schematic showing the plan view of a single die of the proposed shear stress sensorand its section view indicating various layers. . . . . . . . . . . . . . . . . . . . 121

5-2 Step by step fabrication process. . . . . . . . . . . . . . . . . . . . . . . . . . . . 122

5-3 SEM images of comb finger structure etched using DRIE, giving 3D perpective. 124

5-4 SEM image of a cleaved wafer sample depiciting a) vertical sidewalls and b)uniformly plated nickel. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127

5-5 Microscopic image of a released floating element sensor structure from Design 3(Table 4.4). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128

5-6 A 2 mm× 2 mm sense element on a 5 mm× 5 mm sensor die. . . . . . . . . . . 128

5-7 Schematic of sensor package for shear stress characterization. . . . . . . . . . . . 129

5-8 Photographs of sensor packaged on a 30 mm× 30 mm PCB. . . . . . . . . . . . 130

6-1 Schematic showing the die-level impedance measurement setup for the sensor. . 132

6-2 Schematic showing the mean shear stress/static calibration setup using Poiseuilleflow in a 2-D channel. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133

6-3 Schematic with optical image of sensor die (5 mm × 5 mm) indicating float-ing element, contact pads, and interface circuit (voltage buffer) for mean shearcharacterization. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134

6-4 Schematic of the biasing circuit scheme to control phase and amplitude of biasingsignals to null out output from a potential mismatch in sensor capacitances. . . 137

6-5 A schematic of the dynamic calibration setup for measuring shear sensitivity usingrigid termination with the sensor located at pressure node and velocity maximum.139

6-6 A schematic of the dynamic calibration setup for measuring pressure sensitivityusing normal incidence acoustic waves. . . . . . . . . . . . . . . . . . . . . . . . 141

6-7 A schematic of the dynamic calibration setup for shear stress measurement withplane progressive acoustic waves. . . . . . . . . . . . . . . . . . . . . . . . . . . 142

6-8 A schematic of the noise measurement setup for the packaged shear stress sensorusing a double Faraday cage. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143

6-9 Capacitance from impedance measurements on the pre-packaged sensor die. . . 144

12

Page 13: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

6-10 Linear sensor output voltage as a function of mean shear stress at bias voltageamplitude of 2.5 V at 10 kHz. . . . . . . . . . . . . . . . . . . . . . . . . . . . 146

6-11 Distribution of the measured shear stress sensitivity due to uncertainty in themeasured shear stress and sensor output voltage. . . . . . . . . . . . . . . . . . 147

6-12 Sensor output voltage at 1.128 kHz at a bias voltage of 10 V . Sensor is placedat a quarter wavelength (velocity maxima) from the rigid termination. . . . . . 150

6-13 Voltage output as a function of pressure at 4.2 kHz at different bias voltages. . 151

6-14 Linear sensor output voltage as a function of shear stress at 4.2kHz at 3 differentbias voltages. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152

6-15 Sensor output voltage normalized by bias voltage as a function of shear stress at4.2 kHz. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153

6-16 Schematic showing a set of overlapping comb fingers deflection due to both shear(δs) and due to pressure (δp). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154

6-17 Frequency response of sensor at Vb = 10 V using τin = 0.5 Pa as the referencesignal up to the testing limit of 6.7 kHz. . . . . . . . . . . . . . . . . . . . . . 155

6-18 Measured output referred noise floor of the packaged sensor in Vrms/√

Hz atdifferent bias voltages. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157

6-19 Zoomed in plot of the output referred noise floor of the packaged sensor near1 kHz at different bias voltages. . . . . . . . . . . . . . . . . . . . . . . . . . . 157

7-1 Schematic with the top view and section of a sensor bond pad showing the asso-ciated electrical impedances. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165

7-2 Schematic with the top view and section of a sensor bond pad showing the asso-ciated electrical impedances. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166

7-3 Effective sense capacitance variation with frequency as a function of substrateresistance. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166

7-4 Capacitance measurement drift indicated via the first and last measurement priorto ensemble averaging. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169

7-5 Drift in mean capacitance with subsequent dc bias sweeps. . . . . . . . . . . . . 170

7-6 Magnitude of sensor to microphone transfer function as a function of acousticpressure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171

7-7 Phase of sensor to microphone transfer function as a function of acoustic pressure.171

7-8 Transfer function between the sensor and the reference microphone for normalacoustic incidence. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172

13

Page 14: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

7-9 Comparison of combined (shear and pressure) and pressure transfer functionsmeasured between sensor and reference microphone. . . . . . . . . . . . . . . . 173

7-10 Comparison mean shear stress measurements at Vb = 2.5 V at different channelwidths, h, in the flow cell. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175

7-11 Mechanical LEM for pressure sensitivity of the sensor. . . . . . . . . . . . . . . 176

A-1 A deflected beam with arbitrary curvature. . . . . . . . . . . . . . . . . . . . . . 178

A-2 Simplified mechanical model of the floating element structure. . . . . . . . . . . 180

A-3 Schematic of one half of a clamped-clamped beam under large deflection. . . . . 189

B-1 General representation of an ideal two-port element. . . . . . . . . . . . . . . . 198

B-2 Circuit representation of the transformer and gyrator. . . . . . . . . . . . . . . 199

B-3 Impedance to impedance analogy representation of a two port element. . . . . 199

B-4 Equivalent circuit representation of a transducer using impedance analogy. . . . 202

B-5 Schematic of a parallel plate capacitive transducer. . . . . . . . . . . . . . . . . 202

B-6 Circuit representation of a capacitive transducer for constant charge biasing. . 204

C-1 Setup for shear stress in PWT for a general impedance termination. . . . . . . 205

D-1 PCB Layout and dimensions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212

D-2 Drawing of Lucite plug. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213

14

Page 15: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

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

A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL SHEAR STRESSSENSOR

By

Vijay Chandrasekharan

May 2009

Chair: Mark SheplakMajor: Mechanical Engineering

A shear stress sensor measures the frictional force due to fluid flow adjacent to a surface.

Shear stress sensors have a wide variety of applications which include the fundamental study

of turbulence, aerodynamic drag measurement devices, feedback sensors for flow control

applications, and non-intrusive industrial flow measurement devices.

Several research efforts in the past have been directed towards developing a shear shear

stress sensor for time resolved wall shear stress measurements. Conventional techniques

rely on velocity measurements and heat/mass transfer techniques to infer the shear stress.

While these measurement techniques are suitable to determine mean shear stress in well-

known two dimensional flows, they are unsuitable for complex three dimensional flows as

they significantly alter the flow or rely on semi-empirical theory for a measurement. Shear

stress sensors require flush mounted sensors on the surface to avoid flow disturbance. It

is also necessary that these sensors measure the shear force directly and be small enough

to have sufficient resolution both in space and time. These sensors when assembled in the

form of an array on a surface, may be used to understand the basic nature of turbulent flow

and flow separation. Preventing flow separation via flow control techniques helps to reduce

separation drag and improve energy efficiency of moving vehicles.

A microelectromechanical systems (MEMS) shear stress sensor may serve as an instru-

ment grade sensor for many applications. A lack of such a sensor on a laboratory or a

commercial level, despite several research undertakings indicates the challenging nature of

15

Page 16: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

this problem. Challenges and requirements for shear stress sensors are explained. The previ-

ous work on MEMS scale shear stress sensors are reviewed in detail with emphasis on scope

of improvement.

A differential capacitive shear stress sensor design is presented for application in bound-

ary layer flows. The sensor is the first in its kind to use an asymmetric comb finger structure.

This enables a simple two mask fabrication process using bulk micromachining, in which a

single etching process defines the sensor structure, comb fingers, and electrical contact pads.

A thorough analysis of the mechanical, electrical, electromechanical transduction, and inter-

face circuit models is presented. A system level optimization is performed, which includes

the sensor and the associated interface electronics.

The characterized sensor has the highest dynamic range (> 102 dB) and the lowest noise

floor/MDS (14.9 µPa) reported to date for a micromachined direct shear stress sensor. The

sensor has a bandwidth of 6.2 kHz, which is defined by its resonance. It is the first sensor to

demonstrate a pressure rejection of 64 dB for comparable forcing in shear and pressure. In

comparison to previous work on direct MEMS shear stress sensors, this sensor outperforms

its best predecessors by roughly two orders of magnitude in minimum detectable shear stress

and at least an order of magnitude in dynamic range.

16

Page 17: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

CHAPTER 1INTRODUCTION

Viscous or skin friction drag due to relative velocity between a body and the surrounding

fluid medium is a major source of energy dissipation in aircraft and automobiles. Drag has

an adverse effect on the fuel efficiency of the vehicle. Annual consumption of hydrocarbon

based fuels runs into billions of gallons [1]. Furthermore, depleting petroleum reserves in the

world emphasize the importance of drag reduction to improve fuel efficiency.

In 2000, the annual inflation-adjusted aviation fuel cost in the United States was about

$10 billion [2]. The cost per barrel of crude oil has since risen from an average value of

$27.53 in 2000 to $147 in July 2008 [1,3]. While it has subsequently dropped to $40.26, the

long term prospects indicate market volatility and project prices as high as $186 per barrel

in 2030 [1, 4]. Increasing world automobile and air traffic together with ever increasing fuel

prices thus necessitate effective techniques to reduce drag. During landing, take off and

cruise conditions, skin friction drag is approximately 50% of the total drag [2, 5]. A 20%

reduction would translate into approximately $1 billion in annual fuel savings worldwide [2].

In underwater vehicles almost 90% of the drag is associated with skin friction. A 20%

reduction in drag in such vehicles can help increase their speeds by as much as 6.8% [2].

Viscous drag also determines the characteristics of different wall bounded shear flows

and is also important for characterizing the state of turbulent boundary layers. This helps

in understanding different flows from a fundamental fluids perspective and aids in their

control [6]. Quantifying drag reduction requires accurate measurement of wall shear stress,

which is drag force per unit area.

An ideal shear stress sensor should be able to accurately capture both the static and

dynamic characteristics of the flow. To capture the turbulent flow physics, a shear stress

sensor requires sufficiently high bandwidth to provide acceptable temporal resolution and

small dimensions for good spatial resolution. A microscale sensor possesses the reduced

physical size and low inertance required to satisfy both spatial resolution and bandwidth

17

Page 18: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

requirements [6]. But, none exist that provide a reliable shear stress measurement due to

sensitivity drift and insufficient dynamic range, bandwidth, and/or shear stress resolution.

The goal of this research is the design, fabrication, and characterization of a microscale

sensor for direct measurement of time resolved wall shear stress on a surface adjacent to a

flow.

In the sections that follow, the mechanism of wall shear stress generation and its rele-

vance is briefly reviewed. Design requirements for the sensor are determined from relevant

length and time scales of the flow physics. The objectives and expected contributions of

this research are outlined, and at the conclusion of the chapter, the overall dissertation

organization is presented.

1.1 Wall Shear Stress and Boundary Layers

A fluid in motion transmits both tangential and normal stress between adjacent fluid

particles. Tangential stress is a result of viscosity, a fluid property that quantifies the inter-

molecular forces inherent to a moving fluid. In the vicinity of a solid surface, there exists a

velocity gradient between the fluid and the solid. At the fluid-solid interface in most situa-

tions, the viscous forces cause the fluid molecules to adhere to the surface, commonly known

as the no-slip boundary condition. This gives rise to a velocity gradient normal to the wall,

resulting in a shear stress [7].

For a Newtonian fluid in a 2D boundary layer, the shear stress is linearly related to the

velocity gradient via the viscosity of the fluid. Mathematically, the fundamental relation is

τwall = µdu

dy

∣∣∣∣y=0

, (1–1)

where τwall is the shear stress, µ is the dynamic viscosity, u is the streamwise velocity of the

fluid, and y is the coordinate normal to the wall. Figure 1-1 depicts the velocity gradient

due to flow over a flat plate. The velocity is zero at the stationary wall due to the no-slip

boundary condition, and reaches the free stream velocity, U∞, at a finite distance from the

wall. This viscous dominated region close to the wall is known as a boundary layer. Usually,

18

Page 19: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

x

y

Direction of Flow

u = 0

u = 0.99 U∞

0

w

y

du

dyτ µ

=

=

δ

Figure 1-1. Velocity profile of viscous fluid flow over a flat plate.

the boundary layer thickness, δ, is defined as the distance from the surface where the local

fluid velocity is 99% of the free stream velocity i.e., u = 0.99U∞ [8].

Skin friction measurement has been of interest since the early 1870s [9]. The reduction

of skin friction is one of the important reasons behind wall shear stress measurement. In

addition to skin friction, drag is also caused due to flow separation. Flow separation occurs

when the adverse pressure gradient overcomes the momentum of the flow. The region of

adverse pressure gradient is also known as the pressure recovery region. When the flow

separates, it effectively encounters a deformed body after the point of separation. This

results in a net integrated pressure force in the direction of flow, causing pressure drag.

From a flow control perspective, separation drag is detrimental for an aircraft as it reduces

lift and its fuel efficiency. Thus, wall shear stress measurement is extremely important for

flow control where flow separation is delayed using a variety of control techniques [8]. From

the discussion so far, it is evident that the measurement of wall shear stress has practical

implications and also gives insight into the flow physics. As stated by Haritonidis, “The mean

stress is indicative of the overall state of the flow over a given surface while the fluctuating

stress is a footprint of the individual processes that transfer momentum to the wall” [10].

In general, the viscous diffusion of flow momentum from the wall due to shearing effects

results in a boundary layer [7]. The non-dimensional Reynolds number, Re, which is the

ratio of inertial to viscous forces, is used to describe the extent of viscous diffusion, resulting

19

Page 20: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

in a boundary layer. Mathematically, the Reynolds number is expressed as,

Rel =ρUl

µ, (1–2)

where ρ is the density of the fluid, U is a typical velocity scale of the flow, and l is a typical

length scale of interest in the flow. The Reynolds number may also be interpreted as the

ratio of viscous diffusion to convective time scales in the flow [8]. Thus, high values of Re

result in thin boundary layers while low values of Re result in thicker boundary layers in

both laminar and turbulent flows. In an incompressible flow over a thin body, the boundary

layer flow is usually laminar for 1000 < Re < 106 and turbulent at higher Re [8]. Before

delving into different kinds of boundary layers, it is informative to introduce some important

parameters. First is the displacement thickness, δ∗, which is a measure of the mass deficit

in the boundary layer due to viscous retardation of the fluid. For an incompressible flow, it

is expressed as [8],

δ∗ =

∞∫

0

(1− u

U∞

)dy. (1–3)

Second is the momentum thickness, θ, which represents the flow momentum lost due to shear-

ing effects transmitted to the bounding surface. The momentum thickness is mathematically

expressed as [8],

θ =

∞∫

0

u

U∞

(1− u

U∞

)dy. (1–4)

The wall shear stress and momentum thickness are related to the friction and drag coeffi-

cients, Cf and CD, for flow over flat surfaces with zero pressure gradient through [8],

Cf =τw

12ρU∞

2 = 2dθ

dx(1–5)

and CD =D

12ρU∞

2L=

2θ(L)

L, (1–6)

20

Page 21: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

where D is the drag force per unit width for a flat plate given by

D = θ(ρU∞

2)

=

L∫

0

τw(x)dx. (1–7)

Equations 1–5 and 1–6 are general in nature and are valid for both laminar and turbulent

boundary layers for incompressible flow over flat plates [8].

Section 1.1.1 describes a laminar boundary layer, which is formed at relatively low

Reynolds numbers. General characteristics of a steady, two-dimensional boundary layer and

their relevance to wall shear stress measurements are provided. In Section 1.1.2, turbulent

boundary layers formed at high Reynolds numbers are described. Section 1.1.3 provides

a discussion on relevant length scales and time scales essential for the understanding of

turbulent flow physics. These length and time scales are translated later into sensor design

requirements.

1.1.1 Laminar Boundary Layer

In general, laminar boundary layers are less complex than turbulent boundary layers.

A laminar boundary layer is characterized by smooth, orderly flow and is generally observed

at lower Reynolds numbers. The vorticity generated at the wall due to tangential pressure

gradients is transported away through viscous diffusion. The extent of discernable viscous

diffusion away from the bounding surface determines the boundary layer thickness, δ. Usu-

ally, δ is defined as the distance from the surface where the local fluid velocity is 99% of the

free stream velocity [8].

Blasius’ solution is used for laminar boundary layers with zero pressure gradient (ZPG).

This method assumes large values of Re but not enough for transition into turbulence.

Table 1-1 provides the relevant parameter values for a laminar boundary layer using the

Blasius solution. These relations may be used to theoretically validate the measurement

accuracy of the shear stress sensor in a ZPG laminar boundary layer.

21

Page 22: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Table 1-1. Laminar boundary layer parameters [8].

Parameter Blasius Solution

θx

√Rex 0.664

δ∗x

√Rex 1.721

δ99%x

√Rex 5.0

Cf

√Rex 0.664

CD

√Rex 1.328

Stable

laminar

flow

x0

U

Recrit

T/S

waves

Spanwise

vorticity

Three

dimensional

vortex

breakdown

Turbulent

spots

Fully

turbulent

flow

Edge

contamination

LaminarTransition length

Turbulent

Retr

U

Top View

Side View

Figure 1-2. Schematic of a 2D boundary layer for flow over a flat plate showing both thelaminar and turbulent regions (adapted from [8]).

1.1.2 Turbulent Boundary Layer

Turbulent boundary layers are formed approximately at Rex > 106 for flow over a flat,

smooth, plate. Turbulent flows are generally characterized by random fluctuations in the

flow, diffusivity, large Re, three dimensional vorticity fluctuations, and dissipation [11]. A

wall-bounded turbulent flow results in a turbulent boundary layer. Figure 1-2 is a schematic

22

Page 23: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

of flow over a flat plate that shows the laminar, transition, and turbulent regions of the

boundary layer.

Unlike laminar boundary layers, turbulent momentum transport occurs via both viscous

diffusion and macroscopic mixing where the latter dominates. A non-dimensional length,

y+, known as the viscous wall unit, is used to express the extent of the influence of viscous

diffusion. The characteristic viscous length scale used for non-dimensionalization is

l+ =ν

u∗, (1–8)

where u∗ is the friction velocity and ν is the kinematic viscosity of the fluid. A wall unit is

therefore defined as,

y+ = y/l+ = u∗y/ν. (1–9)

The friction velocity is defined as

u∗ =

√|τw|ρ

. (1–10)

It noteworthy that the fundamental turbulent boundary layer non-dimensional parameters

depend on u∗, which is estimated from τw, further emphasizing the importance of wall shear

stress measurements. As opposed to a laminar boundary layer, the shear stress in turbulent

boundary layers has an additional component referred to as the Reynolds stress [11],

τtr = −ρu′v′, (1–11)

where u′ and v′ are the fluctuating components of the velocities. The bar in Equation 1–11

represents the time average at fixed locations, assuming a statistically steady but inhomo-

geneous flow field [11]. The velocities decomposed in terms of a time averaged mean and

fluctuating components, are represented as

u = u + u′ (1–12)

and v = v + v′. (1–13)

23

Page 24: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

The stress represented by Equation 1–11 is a term in the Reynolds-averaged momentum equa-

tion derived from the Navier Stokes equation by decomposing velocity, pressure and stresses

into their respective mean and fluctuating components. Physically, this term serves as the

mechanism for momentum transport due to macroscopic mixing in a turbulent boundary

layer, similar to viscous diffusion in a laminar boundary layer. Note that this is true for wall

units, y+ > 5, whereas for y+ ≤ 5, viscous diffusion continues to dominate.

In the past, von Karman’s integral approach has been used to obtain flow parameters

for a turbulent boundary layer just as it has been for laminar boundary layers. In this

dissertation, it will be used to provide rough scaling information. An assumed velocity

profile suggested by Prandtl based on pipe flow theory is [8]

u

U∞≈

(y

δ

)1/7

. (1–14)

Using Equation 1–5 and Equations 1–6 with the assumed velocity of Equation 1–14 results

in [8]

Reδ ≈ 0.16Re6/7x ,

δ

x≈ 0.16

Re1/7x

, and Cf ≈ 0.027

Re1/7x

. (1–15)

The expressions in Equation 1–15 agree well with published flat plate data [8] and serve as

rough estimates for validating mean shear stress measurements in turbulent flows.

A typical near wall non-dimensional mean velocity profile of a turbulent boundary layer

is shown in Figure 1-3. The near wall turbulent boundary layer consists of three different

regions: the viscous sublayer, the buffer layer, and the inertial sublayer. The portion of the

boundary layer for which y+ ≤ 5 is called the viscous sublayer. The mean flow in this layer

scales as [12, 13]

u

u+= f(y+) ∼ y+. (1–16)

Equation 1–16 indicates that the mean velocity scales linearly with y+. Similar to a laminar

boundary layer, the momentum transport in this region is dominated by viscous diffusion.

However, unlike the laminar boundary layer, the viscous sublayer still possesses fluctuations

24

Page 25: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

10−1

100

101

102

103

104

0

5

10

15

20

25

30

y+

u+

=u/u

Viscous Sublayer Inertial Sublayer

u+ = y+

u+ = 1/κ log(y+) + B

BufferLayer

Figure 1-3. Near wall, mean velocity in a turbulent boundary layer or law of the wall(adapted from [11]).

due to turbulence in the flow [11]. As previously mentioned, the wall shear stress mea-

sured in a turbulent boundary layer thus possesses both mean and fluctuating components.

The viscous sublayer is also referred to as the laminar sublayer due to the similarity of its

properties and those of laminar boundary layers [12]. At y+ > 50, it is found that [12]

u

u+= f(y+) ∼ 1

κlog(y+) + B, (1–17)

where κ is known as the von Karman constant and B is a universal constant. Both are

determined empirically. Typically, κ = 0.4 and B ≈ 4 − 5. In this region, known as the

inertial sublayer, the shear stress is dominated by the turbulent exchange with negligible

contribution from viscous effects. Equation 1–17 shows that in the inertial sublayer, the

mean velocity has a logarithmic variation with y+.

The transition region between the viscous and the inertial sublayers is known as the

buffer layer. The buffer layer usually exists between 5 ≤ y+ ≤ 30. In this region the shear

effects are transferred by a combination of both the viscous effects and turbulent exchanges.

25

Page 26: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

The velocity profile in this region is thus a smooth combination of the linear and logarithmic

variations with y+. Spalding and Chui [14] provide a composite formula valid from y+ = 0

to y+ > 100 expressed as

y+ = u+ + eκB

[eκu+ − 1− κu+ − (κu+)2

2− (κu+)3

6

]. (1–18)

Since this dissertation focusses on developing a wall shear stress sensor, the present discussion

on turbulence is limited to the near wall region and excludes wakes. Furthermore, the analysis

is based on 2-D turbulent flow assumption.

1.1.3 Small Scales in Turbulence and Sensor Requirements

As stated previously, viscous forces tend to be negligibly small relative to inertial forces

at large values of Re. Negligible viscous effects would imply that there is no conceivable

energy dissipation in the flow, which is contrary to physical intuition. The non-linear terms in

the Navier Stokes equations drive an energy cascade. This maintains a finite dissipation level

in the flow by generating small scale motions that can be influenced by viscous effects [11].

A sensor is therefore necessary to measure these small scale motions. The wall shear stress,

if measured accurately, can aid in extracting valuable information about these small scale

fluctuations.

A shear stress sensor has to be non-intrusive in nature to ensure that the flow remains

unaltered. The sensor is therefore ideally placed flush with the wall. If the sensor roughness

is within the viscous sublayer, it does not disturb the flow and is considered hydraulically

smooth. The sensor measures shear stress contributions from both the small scale fluctua-

tions and the mean flow. For a reliable measurement while avoiding spatial averaging effects,

the sensor dimensions should be smaller than the length scale of these small scale fluctua-

tions of interest. Similarly, the sensor should have sufficient bandwidth to resolve the small

time scales (high frequencies) associated with small scale fluctuations. Kolmogorov derived

various scales of motion using an elegant energy based theory called Kolmogorov’s universal

equilibrium theory [11]. The respective “Kolmogorov scales” for length, time, and velocity

26

Page 27: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

are

η ≡ (ν3/ε

)1/4,T ≡ (ν/ε)1/2 , and υ ≡ (νε)1/4 , (1–19)

where ε is the dissipation rate per unit mass. The small scale motions are statistically

independent of large scale motions and the mean flow. However, they rely on the energy

supply from the large scale motions and the viscous dissipation to sustain their motion [11].

According to Kolmogorov’s theory, near equilibrium the rate of energy supplied should be

approximately the same as the rate of energy dissipated. The dissipation rate is therefore

expressed in terms of the large scale motions as

ε ∼ u3/`, (1–20)

where u is the characteristic velocity scale of large eddies (u ≈ 0.01U∞) and ` is the length

scale related to the size of those eddies, also referred to as the integral scale [11]. In a

turbulent boundary layer the size of large eddies are approximately equal to the boundary

layer thickness δ. Therefore the inviscid estimate for ε in Equation 1–19, is used with δ as

the length scale instead of ` to relate the large scale motions to the Kolmogorov scales as

η

δ∼

(uδ

ν

)−3/4

= Re−3/4δ , (1–21)

Tu

δ∼

(uδ

ν

)−1/2

= Re−1/2δ , (1–22)

andυ

u∼

(uδ

ν

)−1/4

= Re−1/4δ . (1–23)

Combining Equations 1–15, 1–21, 1–22, and 1–23, the Kolmogorov scales in terms of the

distance from the leading edge, x, of a flat plate and the corresponding Reynolds number,

Rex, are estimated via the 1/7th power law as follows:

η ∼ 0.632xRe−11/14x ,

T ∼ 40x

U∞Re−4/7

x , (1–24)

and υ ∼ 0.016U∞Re−3/14x .

27

Page 28: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

log k

log E

k = O(δ-1) k = O(η-1)

k-5/3

Figure 1-4. Schematic of energy distribution in turbulence as a function of wavenumber atsufficiently high Re (adapted from [13]).

A typical 1-D distribution of turbulent energy, E(k), as a function of the wavenumber,

k, is shown in Figure 1-4. The turbulent wave number k is defined as

k = ω/v ≡ 2π

δ, (1–25)

where ω is the frequency, v is the turbulent eddy velocity, and δ is the corresponding eddy

length scale. In general, turbulent energy in a boundary layer is concentrated in the large

turbulent eddies, which dominate momentum transfer via macroscopic mixing. These eddies

have length scales on the order of the boundary layer thickness δ. Therefore, from Equa-

tion 1–25, the frequencies corresponding to these structures scale as, O(δ−1). This energy

gets redistributed from the large scale structures to the small scale turbulent structures on

the order of the Kolmogorov scales (η) with frequencies that scale as O(η−1). As Rex in-

creases, these scales get smaller, resulting in challenging sensor design even with modern

micromachining technology. Thus, shear stress sensor length scales and bandwidths are

application specific with stringent design requirements with increasing values of Re. The

different flight regimes as a function of Reynolds number are shown in Figure 1-5.

28

Page 29: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Figure 1-5. Reynolds number for different flight regimes at sea level (adapted from [15]).

The challenges in sensor design for resolving Kolmogorov scales may be further explained

via the following study. Consider flow over a flat plate with U∞ = 50m/s and a shear

stress sensor of sensing area, Asensor = η2, placed at x = 1 m from the leading edge. For

this example, any variation in Re is due to change in thermodynamic parameters. From

Equations 1–5 and 1–15,

τw =1

2ρU2

∞Cf =1

2ρU2

∞0.027

Re1/7x

. (1–26)

29

Page 30: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

106

107

108

109

10−2

10−1

100

101

102

η(µ

m)

Kolmogorov Scales

Reynolds Number Rex

106

107

108

10910

3

104

105

106

1/T

(Hz)

Figure 1-6. Variation of Kolmogorov length and time scales with flow Reynolds number(x = 1m, U∞ = 50m/s).

The corresponding shear force, Fs, on the sensor is

Fs = τwAsensor = τwη2. (1–27)

The variation of η and T at different flight Reynolds numbers for x = 1 m is shown in

Figure 1-6. Figure 1-7 shows the variation of τw and Fs with Rex. While the shear stress levels

remain at moderate values, the sensing area (η2) and the corresponding Fs drop dramatically.

For example, at Rex ∼ 1 × 107, η ∼ 2 µm and 1/T ∼ 12 kHz. Furthermore, for Asensor =

η2 ∼ 4 µm2, Fs = 16 pN and τw = 4 Pa, which is the maximum measurable shear stress. For

a given range of Rex the variation in η is O(102) higher than the variation in T. Similarly,

the variation of Fs is also much higher than the change in τw, making shear stress sensing

more challenging at higher Rex. Now consider a measurement where the turbulent shear

stress is roughly 40 dB re 1 Pa below the mean shear stress of 4 Pa with a preferred signal

to noise ratio of 20 dB. Thus, the sensor needs to have a minimum detectable signal (MDS)

of 4 mPa or a dynamic range of 60 dB. The shear force corresponding to this MDS is 16 fN .

30

Page 31: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

The analysis clearly elucidates the stringent requirements and the challenging nature of shear

stress sensor design to resolve Kolmogorov scales in both space and time at high Rex.

106

107

108

109

2

2.5

3

3.5

4

4.5

5

5.5

6

τ w(P

a)

Reynolds Number Rex

106

107

108

10910

−9

10−8

10−7

10−6

10−5

10−4

10−3

Fs

(µN

)

Figure 1-7. Variation of shear stress and corresponding shear force on sensor area (Asensor =η2) with flow Reynolds number (x = 1m, U∞ = 50m/s).

It is important to note that the Kolmogorov scales represent the smallest fluid structures

in turbulence, but not necessarily the smallest scales of interest from the sensing perspective.

For instance, in flow control applications, shear stress sensors may be used as feedback

sensors for drag reduction. Numerous research articles on flow control for turbulent drag

reduction sight turbulent streaks as a major cause of skin friction drag. Turbulent streaks

are alternating spanwise regions of low and high speed fluid oriented in the streamwise

direction and are quiescent most of the time [16]. At low Re values, the streaks are spaced

100l+ apart [16] and are roughly 40l+ wide [17], circumventing the need for the sensor

to always resolve the Kolmogorov scales. For example, for measuring shear stress from

turbulent streaks that are 40l+ wide, let us assume the sensor may be at most 20l+ in its

maximum sensing dimension. Using the example in the previous paragraph, for τw = 4 Pa,

20l+ ≈ 200 µm, which is two orders of magnitude greater than the length scales required

31

Page 32: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

to resolve Kolmogorov scales. Thus, similar to the previous case, for a dynamic range of

60 dB for shear stress, the corresponding maximum and minimum shear forces acting on the

sensor are 160 nN and 160 pN , respectively. These numbers indicate that the stringency

of design in considerably lower for applications such as flow control as against resolution of

Kolmogorov scales.

Micromachined sensors offer the potential to satisfy both the spatial and temporal res-

olution requirements for different applications. No direct/indirect shear stress sensor exists

that satisfies both spatial and temporal resolution requirements while providing accurate

shear stress measurements at the Kolmogorov scales. This will be shown in the review of

previous research presented in Chapter 2. Padmanabhan [18] used a floating element shear

stress sensor while both Lofdahl et al. [19] and Alfredsson et al. [20] use thermal shear

stress sensors. Their research efforts have sensor length scales of 4l+ [21], 5l+ [19], and

10 − 20l+ [20]. As stated earlier, to accurately measure the Kolmogorov fluctuations, the

sensor dimensions must be equal to or smaller than these scales. Sensor resolution at the

Kolmogorov scales essentially allows one to use the sensor for a wide variety of applications.

However, previous efforts indicate that the sensor dimensions are restricted by signal resolu-

tion issues and fabrication limitations. Numerous direct and indirect micromachined sensors

have been previously developed and will be reviewed in Chapter 2 with emphasis on their

strengths and limitations.

Floating element sensors used for direct measurement of wall shear stress may also

be sensitive to vibrations and pressure fluctuations. If these undesirable sensitivities are

not eliminated, the shear stress measurements may be erroneous. The magnitude of the

fluctuating pressure forces is approximately two orders of magnitude higher than for the

shear forces. Direct numerical simulations to study the wall pressure effects show that, based

on frequency, the wall pressure fluctuations are 7 − 20 dB higher than the streamwise wall

shear stress and 15 − 20 dB higher than the spanwise component [22]. The sensor design

should also ensure minimum errors due to misalignment, pressure gradients and gaps [9].

32

Page 33: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

The gaps must be less than a few viscous length scales (< 5l+) to ensure that there is no

flow disturbance (hydraulically smooth) [7]. With micromachined floating element sensors

gaps as small as 4 µm are achievable. Detailed discussions on floating element sensors and

implementation of these design requirements are provided in Chapter 2 and 3, respectively.

1.2 Research Objectives

This dissertation provides details on the development of a MEMS based capacitive

floating element sensor for direct, time resolved wall shear stress measurements. The sen-

sor employs a differential capacitance measurement scheme, which helps to eliminate the

undesirable cross axis sensitivity of the sensor. A capacitive sensor is also insensitive to tem-

perature variations. The sensor employs simple fabrication technology and is suitable for

stand alone measurements of wall shear stress. The aim is to design, optimize, and fabricate

a benchmark shear stress sensor to reliably measure turbulent variations within its spatial

and temporal resolution limitations. The goal is to measure up to 10 Pa of shear stress at

frequencies up to 10 kHz and achieve a spatial resolution of ≈ O(100l+) to demonstrate the

proof of concept of the proposed sensor.

1.3 Dissertation Organization

The dissertation is organized into seven chapters. Chapter 1 introduced wall shear

stress and motivation for its measurement. Scaling studies and corresponding sensor re-

quirements were presented. Chapter 2 is a review of previous research efforts to develop

microscale, direct wall shear stress sensors. Existing sensing techniques, including capacitive

transduction using comb fingers, are discussed. Chapter 3 presents the sensor’s mechanical

and electrostatic model, electromechanical transduction, interface circuitry, and the noise

model. Mathematical representations for noise and sensitivity are derived to use in a design

optimization strategy. Chapter 4 contains the formulation of the optimization problem and

an explanation of the optimization scheme to be implemented. Optimization results are

presented together with sensitivity analysis for the design. Chapter 5 provides details of the

fabrication steps and the packaging scheme for the sensor. Chapter 6 explains the sensor

33

Page 34: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

characterization procedure and discusses the results. Chapter 7 provides conclusions and

gives recommendations for the next generation of capacitive shear stress sensors.

34

Page 35: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

CHAPTER 2BACKGROUND

In this chapter, previous research efforts directed towards developing shear stress sensors

are described. Existing measurement techniques are briefly explained along with key figures

of merit. Transduction schemes commonly used in direct MEMS wall shear stress sensors

are described. Finally, a review of direct MEMS wall shear stress sensors is presented. The

chapter summarizes the various techniques for wall shear stress measurement drawn mainly

from previous review articles [6, 9, 10,23].

Different techniques exist for shear stress measurement. As stated in Chapter 1, shear

stress may be directly determined from the measurement of shear force on a known area.

However, other methodologies also exist and are used in practice to estimate the shear stress

from measured flow parameters such as velocity, joulean heating rate etc. Thus based on

the quantity measured, the techniques are broadly classified as direct (measures shear force)

or indirect (measures other quantitites) measurement techniques. Shear stress sensors may

also be categorized as conventional (macroscale) or microscale sensors. Several research

efforts in the past have been directed towards developing both direct and indirect microscale

sensors. Microscale direct sensors, due to their direct nature and favorable scaling to capture

turbulent flow physics (see Chapter 1), are better suited for quantitative time-resolved shear

stress measurement. The discussion in this chapter mostly concentrates on previous research

on direct micromachined sensors.

2.1 Shear Measurement Techniques

2.1.1 Indirect Measurement Techniques

Indirect techniques rely on a known correlation between the measured parameter and

shear stress to estimate the latter. Heat/mass transfer based devices, surface/flow obstacle

devices and velocity profile measurement techniques are used to indirectly estimate wall shear

stress. Reviews by Winter [9] and Haritonidis [10] describe the benefits and limitations of

these measurement techniques.

35

Page 36: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Heat/mass transfer based devices rely on the exchange of heat and mass flux between

the device and the flow. Heated films/wires are examples of devices based on heat transfer

techniques. The benefits of these techniques include high sensitivity, reasonable dynamic

response, small size, simple structure, and non-intrusive mounting. However, temperature

sensitivity, tedious and non-repeatable calibration due to heat loss to substrate/air limit the

use of these sensors for reliable shear stress measurement. Naughton and Sheplak [6], and

Sheplak et al. [23] discuss the developments and limitations of micromachined thermal shear

stress sensors.

Preston tubes, stanton tubes, razor blades, steps and fences are surface obstacle based

shear stress measurement devices. Since these devices obstruct the flow they are useful

in thick boundary layer flows. The measurement is sensitive to the size and geometry of

these devices. These devices are also limited to mean measurements and lack time resolved

shear stress measurement capability. The shear stress estimation is based on empirical

correlations relating 2-D turbulent boundary layer profiles to the measured property. Lastly,

the calibration is sensitive to the location of the probe in the boundary layer. Therefore,

these techniques are of little use in complex three-dimensional flows.

Velocity profile based techniques use invasive probes like pitot tubes or optical (non-

invasive) techniques like laser doppler velocitmetry (LDV) or particle image velocimetry

(PIV). Optical techniques are limited to thick boundary layers because it is very difficult to

seed the near-wall region of the flow. The velocity profile measurements are then matched

with the logarithmic law of the wall also referred to as the Clauser plot [24]. This method

relies on the validity of the Clauser plot and the ability to identify the logarithmic region of

the near wall boundary layer.

The reported micromachined indirect wall shear stress sensors include thermal sensors

[25–28], micro-fences [29], micro-pillars [30–33], and laser based sensors [34, 35]. Micro-

fences consist of a cantilever structure whose deflection due to the flow is measured using

piezoresistve transduction. Micro-pillars use an array of pillars on the wall within the viscous

36

Page 37: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

sublayer. The tip deflections of the pillars are optically measured and correlated to the shear

stress. Laser Doppler based sensors are based on the measurement of the doppler shift of

light scattered from the particles passing through a diverging fringe pattern in the viscous

sublayer of a turbulent boundary layer.

All these techniques rely on the invariance of the near-wall structure of a two dimensional

turbulent boundary layer and are thus not suited for three-dimensional flows.

2.1.2 Direct Measurement Techniques

Direct measurement sensors as stated previously are most suitable since they measure

the shear force directly and do not require assumptions about the flow field [10]. Direct shear

stress measurement techniques include thin-oil film interferometry, liquid crystal coatings and

floating element based sensors. The recent review by Naughton and Sheplak [6] discusses the

operating principle, developments, uncertainties and limitations of these techniques. Both oil

film and liquid crystal coatings need optical access which limit their widespread use. Both

techniques involve complex post processing to analyze the data and are limited to slowly

varying mean measurements [6]. These two techniques are therefore not discussed here.

Most of the direct measurement shear stress sensors employ a sensing element with a

known area attached to a compliant structure. The compliant structure provides a restoring

force when the sensing element is moved by shear force. A floating element sensor is one of

the most widely used direct measurement sensor structure. The sensing element is suspended

over an open cavity by means of tethers that also act as restoring springs. Figure 2-1 shows a

schematic of a floating element sensor. The motion of the floating element is transduced into

a proportional electrical/optical signal to measure the shear stress. These devices, however,

have issues of their own as first explained by Winter [9]. The following problems are generally

associated with conventional floating element sensors [9, 10]:

1. Minimum detectable shear stress (MDSS) increases when the sensor size is decreased

to improve spatial resolution. A small sensor size results in smaller integrated shear

force (Figure 1-7) and thus a lower stress sensitivity and correspondingly higher MDSS.

37

Page 38: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Le

Flow

X

Floating ElementTether

X

t

gk k

Velocity

Lt Lt

Wt

Le

We

Figure 2-1. Simplified 2-d schematic showing the plan view (top) and cross-section (bottom)of a typical floating element shear stress sensor structure (adapted from reviewby Naughton and Sheplak [6]).

2. The effect of essential gaps around the floating element

3. The effect of misalignment of the floating element with respect to the surrounding

surface

4. Forces due to pressure gradients

5. Effects of gravity and acceleration when placed on a moving object or on a non-level

surface

6. Massive devices result in poor temporal resolution

Naughton and Sheplak [6], based on previous research efforts, provide a detailed dis-

cussion showing the favorable scaling of micromachined shear stress sensors for quantitative

time resolved shear stress measurements. Their findings show that micromachined sensors

offer the potential for high temporal and spatial resolution. They theoretically offer at least

five-orders of magnitude improvement in the sensitivity bandwidth product and about two

38

Page 39: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

orders of magnitude improvement in spatial resolution [6]. In a more recent review, Sheplak

et al. summarize the benefits of micro-machined sensors as follows: [23]

1. MEMS sensors can measure shear forces of the order 0.1 nN . This corresponds to a

shear stress of order 10 mPa for a floating element size of 100 µm× 100 µm.

2. Sensors fabricated using standard micromachining technology do not need assembly of

sensor components, eliminating some misalignment issues although sensor packaging

and installation on the test setup is still prone to misalignment errors.

3. Micromachining allows gaps of O(1 µm), rendering the surface hydraulically smooth

except at very high Re [7].

4. Three orders of magnitude reduction in scale compared to conventional sensors greatly

reduces pressure gradient errors in micromachined sensors.

5. The cross-axis sensitivity of these sensors, with respect to acceleration, is three orders

of magnitude smaller than conventional sensors because of reduced sensor mass.

6. Thermal expansion errors are mitigated due to monolithic fabrication techniques but

careful packaging is also important to avoid these errors.

To highlight the benefits and challenges faced by MEMS based shear stress sensors,

this section will focus on previous research, specifically on micromachined direct shear stress

sensors. Based on the transduction scheme used, the sensors are categorized as capacitive,

optical, and piezoresistive, respectively.

2.1.2.1 Capacitive Technique

Schmidt et al. were the first to develop a micromachined shear stress sensor with a

capacitive transduction scheme for measuring shear stress [36,37]. This was the first proof of

concept device indicating feasibility of micromachined floating element shear stress sensors.

The sensor was surface micromachined with polyimide using aluminum as the sacrificial

layer. The sensor used a differential capacitance read out scheme with integrated depletion

mode P-MOS transistors to sense the change in capacitance. A schematic of this device is

shown in Figure 2-2. The floating element was 500 µm× 500 µm× 30 µm and the tethers

39

Page 40: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Floating-ElementEmbedded

Conducter

Cps1 Cdp Cps2

Silicon

Passivated

Electrodes

VDS

Csb2Csb1

On Chip

Off Chip

+

-Vd

Figure 2-2. Schematic of cross section of polyimide capacitive shear stress sensor with differ-ential capacitive readout scheme using integrated depletion mode P-MOS tran-sistors (adapted from Schmidt et al. [37]).

were 1000 µm × 5 µm × 30 µm. The sensor demonstrated a sensitivity of 52 µV/Pa [37].

It however, suffered from drift due to moisture on the order of several µV/min. Moisture

has two different effects, 1) Moisture induces hydroscopic in-plane stresses that affects the

mechanical sensitivity and 2) It changes the dielectric properties of the polyimide resulting

in drift [36]. In general, any charge accumulation at the air dielectric interface will result in

drift [6].

Pan et al. and Hyman et al. developed comb finger-based capacitive shear stress

sensors [38–40]. They developed two sensor designs, one based on differential capacitive

measurement, and the other based on force balancing of moving capacitive plates using

a feedback signal. Figure 2-3 shows a schematic of both sensors. In the first design a

40

Page 41: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Comb finger

Drive

Folded

Beam

V+

Tether

Floating

Element Anchor

Release

Holes

V-

Expanded View of Comb Finger

Structures

V-V+

Flow

C2

C1

b)

Release

Hole

Floating

Element/Electrode

Bottom

electrode

Flow

a)

Figure 2-3. Schematic of a) folded beam with lateral capacitive change and b) force feedbackcapacitive structure (adapted from Pan et al. [38]).

folded beam structure support a shear stress sensing element. The folded beams provided

the requisite restoring force. In this sensor, the output due to capacitance change was

undetectable due to high parasitic attenuation. Hyman et al. attribute the parasitics to

the direct electrical contact being made to the device [39]. Optically measured deflection in

a known flow resulted in a mechanical sensitivity of 9Pa/µm or 0.11 µm/Pa. The second

41

Page 42: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

generation sensors used a structure similar to a commercial accelerometer where a shear stress

sensing element is used instead of a proof mass. Using the reciprocal and conservative nature

of electrostatic transducers [41, 42], the mechanical force sensed using the sense electrodes

is countered by applying electrostatic force on the actuation electrodes. This force feedback

method maintains the sensor at its nominal position i.e., zero displacement. On chip circuitry

with sufficiently high loop gain also mitigate parasitic effects [40]. These sensors had a high

sensitivity of 1.03 V/Pa. At high shear stresses, the electrical actuation force was not enough

to maintain null deflection of the sensor element, potentially limiting its dynamic range. The

dynamic and noise characteristics of these sensors were not reported.

Zhe et al. developed a cantilever based shear stress sensor, consisting of a sensing

element attached to the end of a cantilever beam, that provided the restoring force [43, 44].

The sensing element is 200 µm × 500 µm × 50 µm and the beam is 3000 µm × 10 µm ×50 µm. Two capacitors are formed across the gaps between the floating element and the

surrounding substrate as shown in Figure 2-4. The differential capacitance change due to

element deflection is measured using an off the shelf circuit component MS3110 from Irvine

Sensors. Flow calibration of the sensor resulted in a high noise floor of 0.04 Pa and a

large sensitivity of 337 mV/Pa due to compliance of the long beam [44]. The authors

expressed concern over sensor misalignment in the channel leading to errors. There was 13%

uncertainty in the measurement, which was attributed to the potential misalignment errors

and channel height uncertainty. There were other sources of uncertainty and scatter in the

data, which are still being investigated [44]. The dynamic behavior of the sensor was not

reported. A charge amplification scheme is used in the interface circuitry. Using this scheme

renders the sensor sensitivity independent of the parasitics in the system but, the noise floor

is sensitive to parasitics (see Section 3.1.3.3). This explains the high noise floor of the sensor.

Furthermore, all capacitive sensors inherently are high impedance devices and are therefore

susceptible to electromagnetic interference (EMI).

42

Page 43: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Sense Electrode

Actuation Electrode

Pad

Flow

t, C2t, C1

nl

n2

El g

b

L

SS’

E’ E

x

y

z

n2

n1

h

b

Figure 2-4. Top view of cantilever based floating element capacitive shear sensor (adaptedfrom Zhe et al. [44]).

McCarthy et al. and Tiliakos et al. have attempted to develop a similar sensor using

differential capacitance measurement and force feedback for high shear stress applications

(10 Pa−10 kPa). They however did not perform an experimental characterization and their

work is therefore not discussed any further [45–47].

Similarly, Desai et al. designed and fabricated a novel MEMS structure for 2-D shear

stress measurement using the wafer thickness for the floating element [48]. The sensor

proposed to use a differential capacitive transduction. Though the sensor was fabricated, no

experimental characterization has been reported.

2.1.2.2 Optical Techniques

Padmanabhan et al. designed the first micro-scale optical shear stress sensor with photo

diodes as detection elements using an optical shutter technique [18, 21, 49, 50]. This sensor

also had integrated electronics on the sensor die for electronic signal read-out. Two differ-

ent generations of this device were built. The sensor was comprised of a floating element

(120 µm× 120 µm and 500 µm× 500 µm) acting as the optical shutter for two photo diodes

located under the element at the leading and trailing edges as shown in Figure 2-5. Four

43

Page 44: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Photodiodes

n+ n+

n-Type Si

Floating-element shutterFlow

Incident Light from a Laser Source

p-Type Si Substrate

Photodiodes

n+ n+

n-Type Si

Floating-element shutterFlow

Incident Light from a Laser Source

p-Type Si Substrate

(a)

(b)

Figure 2-5. Cross sectional schematic of a)two diode and b)’split’ diode (3 diodes) opticalshutter based micromachined shear stress sensor (adapted from Padmanabhan’swork [21]).

tethers 500 µm long and 10 µm wide suspended the floating element 1.2 µm over a silicon

substrate. The tethers and floating element were 7 µm thick. A coherent uniform light

source located above the sensor illuminates the exposed area of the two photodiodes, result-

ing in a differential photocurrent. The photocurrent is ideally zero when there is no motion

as the illuminated areas of the two photodiodes are the same by design. In the presence of

a force due to shear stress, the lateral sensor motion increases the illuminated area of one

photodiode while the area for the other decreases by the same amount, resulting in a dif-

ferential photocurrent. The differential photocurrent is proportional to both the magnitude

and direction of the shear stress. Intensity gradients across the floating element resulted in

44

Page 45: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

erroneous outputs with the first generation sensors. The second generation of this sensor

implemented a split diode configuration and resulted in 94% reduction in sensitivity to in-

tensity variations [21]. Static calibration results indicated maximum nonlinearity of 1% over

shear stress ranging from 1.4 mPa to 10Pa. The sensor was also characterized dynamically

via Stokes’ layer excitation. This is was the first sensor use this dynamic calibration tech-

nique [51]. The dynamic response exceeded 10 kHz and the sensitivity for 120 µm sensor

was shown to be 0.097%/Pa (percentage in photocurrent/shear stress) [50]. The sensor was

designed to allow off-chip read-out of the output signal. These sensors were also relatively in-

sensitive to EMI and stray capacitance, unlike capacitive sensors [21]. The detection scheme

was effectively designed to minimize common mode signals due to pressure fluctuations or

out of plane acceleration and vibration with promising results [50]. The shortcomings of

this sensor were front side electrical contacts and remote location of the light source. The

remote location of the light source with respect to the sensor may make the measurement

sensitive to mechanical motion of the light source, such as vibration and thermal expansion

of its mounting [23].

Tseng and Lin used an optical fiber based micro-Fabry-Perot interferometry technique

[52]. The floating element is built on a flexible membrane using two layers of SU-8 resist

as shown in Figure 2-6. The 1.5 mm× 1.5 mm× 20 µm membrane provides the requisite

restoring force. The floating element was 200 µm × 200 µm in length and width. The

reflecting portion of the floating element was 400 µm high. With no gaps, this sensor

circumvented flow disturbance issues, was truly flush mounted, and could also be used in

different fluids. Two optical fibers aligned orthogonal to each other allow 2-D shear stress

measurements, though this ability was not demonstrated experimentally. The detection

technique was also immune to EMI. The static calibration results indicate a sensitivity of

0.65 Pa/nm or 1.54 nm/Pa and temperature sensitivity of 3.4 nm/K [52]. The temperature

sensitivity was thus higher than the shear stress sensitivity. The sensor also had a high noise

floor of ≈ 2 ∼ 3 nm, which corresponds to a shear stress of 1.3 Pa. The dynamic response

45

Page 46: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Floating Element

Glue Groove

A A'

Input (Output) Fiber

MRTV1 Membrane Air or Liquid Flow

Reflection Mirror

(Floating Element)d

IR1R2

Figure 2-6. Conceptual schematic showing bottom and side view of fabry-perot floating ele-ment shear stress sensor (adapted from [52]).

of the sensor was not reported, but the large height of the floating element suggests that

the sensor would have a large mass and therefore a lower bandwidth. The dynamic range

of the sensor was also not reported. The sensor may also have been sensitive to pressure

fluctuations and vibrations of the membrane, supporting the floating element.

Horowitz et al. designed an optical floating element shear stress sensor that used an

aligned wafer bonding/thin back process [53, 54]. Figure 2-7 shows a schematic of the sen-

sor. The sensor has optical gratings on the backside of the floating element and on the

top side of a transparent pyrex wafer. The gratings form a geometric Moire pattern that

amplifies the mechanical motion of the floating element. The Moire pattern displacement

46

Page 47: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Reflected Moiré

FringeIncident

Incoherent Light

Floating ElementLaminar Flow Cell

Pyrex

Silicon

Aluminum Gratings

(Floating Element &

Base Gratings)

Tethers

Figure 2-7. Schematic of the optical Moire shear stress sensor. The shift of the amplifiedMoire pattern is imaged using a microscope attached to a CCD camera (withpermission from author [54]).

was measured using a linescan CCD camera and imaged using a microscope. By design, the

sensor is insensitive to out of plane motion due to pressure, vibration etc. since the motion

produces no change in the Moire fringe pattern. Like most optical sensors, this sensor is

also insensitive to EMI. Drawbacks of this sensor include a bulky optical setup and the as-

sociated packaging requirements, restricting its use to an optical bench. The sensor had a

static mechanical sensitivity of 0.26 µm/Pa with a linear response up to the testing limit of

1.3 Pa. The dynamic testing indicated a resonant frequency of 1.7 kHz, and a noise floor

of 6.2 mPa/√

Hz [54].

2.1.2.3 Piezoresistive Technique

Ng and Goldberg et al. were the first to use piezoresistive implants on a floating element

sensor to measure large shear stress (1 kPa − 100 kPa) for polymer extrusion processes

[55–57], using axial loading on the tethers. They used a wafer bonding and etch back

47

Page 48: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

technology with top side tether implants to form piezoresistors. The piezo resistors were

connected in a half bridge configuration to obtain an electrical output. A simplified schematic

of the sensor is shown in Figure 2-8. The tethers were aligned along the direction of the flow

to withstand the high shear stresses. The floating element was 120 µm × 140 µm × 5 µm

while the tethers were 30 µm × 10 µm × 5 µm, respectively. The sensor had back side

electrical contacts which represented an important landmark in terms of achieving truly

flush mounted non-optical sensors for shear stress measurement. Sheplak et al. noted that

use of anisotropic wet etching process for the backside contact resulted in large die sizes [23].

Flow

X

Y

180 m

120 m

120 m

10 m

Plate Thickness = 5 m

Figure 2-8. 3D schematic of a piezoresistive sensor with top side tether implants oriented inthe direction of flow for high shear stress measurements (adapted from Ng et al.and Goldeberg et al. [56, 57]).

Barlian et al. studied side tether implants and additional top-side implants to also

sense out of plane motion in a floating element shear stress sensor [58, 59]. They presented

mechanical characterization, temperature and doping effects, and the effects of annealing on

noise for a single implanted cantilever beam (6000 µm × 400 µm × 15 µm) [59]. However,

Barlian’s initial flow characterization results were preliminary and had significant scatter in

their measurements, warranting further investigation [58].

48

Page 49: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Al-Si(1%)

p++interconnect

Silicon tether

Sidewall

implanted

piezoresistor

Floating

element

Bond

pad

n-well

Reverse

bias contact

tether

centerline

Figure 2-9. 3D schematic of a side implanted piezoresistive shear stress sensor. The implantsare such that two tethers are in tension while two are in compression for highersensitivity (with permission from author [60]).

Li et al. used side implanted piezoresistors on the tethers of the floating element sensor.

A schematic of a side implanted floating element sensor is shown in Figure 2-9 [60]. They

presented preliminary shear stress characterization results for a 1 mm2 floating element with

tethers 1 mm long and 30 µm wide. The floating element and tethers were 50 µm thick.

The sensor was calibrated using a Stokes layer excitation to generate shear stress in a plane

wave tube [51]. The sensor was linear up to a measured shear stress of 2 Pa and possessed

a normalized sensitivity of 2.83 µV/V/Pa. The noise floor was 11.4 mPa at 1 kHz and

the temperature sensitivity was 0.42 mV/C [61]. The upper end of the dynamic range was

limited by the testing limit of 2 Pa [61]. The dynamic response of the sensor indicated

a flat frequency response up to the testing limit of 6.7 kHz. This sensor demonstrated

promising results and offers a robust sensor for shear stress measurements. High sensitivity

to in-plane motion due to sidewall implants, low sensitivity to out of plane motion both

due to geometry and a simple read out scheme, are the benefits of this sensor design. In a

49

Page 50: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

balanced Wheatstone bridge, out of plane motion acts as a common change in all resistors

resulting in no output. The sensor however had an unbalanced bridge due to poor uniformity

in doping for side wall piezoresistive implants, resulting in higher than expected out of plane

sensitivity. The drawbacks of this sensor include a high temperature coefficient of resistance,

limiting applied bias voltages and thus lowering sensitivity, and drift due convective heat

transfer to the flow similar to hot wires.

2.2 Conclusion

A review of various shear stress sensor techniques was provided. Basic principle, merits,

and issues with existing non-direct measurement techniques were presented based on pre-

vious reviews. The benefits of micromachined direct shear stress sensors were highlighted,

followed by a detailed review of previous work on these sensors organized according to their

transduction mechanisms. Each transduction mechanism has strengths and limitations that

are yet to be resolved, leaving ample room for further development of micromachined shear

stress sensors.

Optical transduction has increased levels of complexity when the whole measurement

system is scaled down. At small scales optical alignment and noise issues can be significant.

Floating element piezoresistive shear stress sensors may drift due to Joulean heating and

require active temperature compensation. Spencer’s research in the area of pressure sensors

indicates that the raw sensitivity of capacitive pressure sensors is 10-20 times the sensitivity

of piezoresistive sensors with similar geometric shape and dimensions [62]. This is amply

supported by the fact that a large number of silicon microphones are mostly capacitive in

nature [63]. Spencer’s observation may be valid if capacitive and piezoresistive transducers

are compared in general. Further, bulk micromachined capacitive sensors are simple to

fabricate, inherently noiseless, and are insensitive to temperature changes. Susceptibility to

EMI, effects of circuit noise and parasitics (see Chapter 3) remain a concern for capacitive

sensors. However, the several benefits of capacitive sensors offer opportunities to design

around their limitations.

50

Page 51: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

In this dissertation, attempts are made to overcome some limitations of capacitive

sensors. The drifts due to charge accumulation at the air-dielectric interface observed by

Schmidt et al. [36] is avoided using metal coated electrodes for the capacitors. An unpack-

aged (die form) low noise voltage amplifier is used with very low input capacitance and is

directly wire bonded to the sensor to mitigate the adverse effect of parasitics. An overall

system optimization is performed which includes the sensor and the interface circuitry to

obtain the best design performance from the sensor. The next chapter discusses the design

and modeling of a floating element based differential capacitive shear stress sensor.

51

Page 52: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

CHAPTER 3DEVICE MODELING

Design and modeling of a micromachined direct capacitive wall shear stress sensor is

presented in this chapter. An electromechanical model is developed using a differential ca-

pacitive transduction scheme. Detailed quasi-static models are developed and used to predict

sensitivity, minimum detectable signal and frequency response of the electromechanical sen-

sor. The proposed shear stress sensor is required to meet stringent requirements on spatial

and temporal resolution while accurately measuring both the mean and the fluctuating shear

stress. In order to maximize overall performance and meet sensor requirements, a design

optimization is required (Chapter 4). Mechanical and electrical models developed in this

chapter are essential components for the optimization study.

This chapter is organized into three major sections. In Section 3.1, an existing quasi-

static deflection model presented in [36,64] is extended to include the effect of comb fingers.

Expressions for mechanical deflections are presented using both linear and nonlinear theories.

A novel asymmetric differential capacitive sensing scheme is proposed. Electrostatic behavior

of the sensor is explained. A short comparison of capacitive interface circuits is given,

followed by integration of the chosen circuitry with the device model. In Section 3.2, the

dynamic characteristics of the shear stress sensor are analytically studied. Lumped element

modeling is employed to estimate the dynamic characteristics of the sensor. In Section 3.3,

higher order effects such as the fringing fields and parasitic capacitance effects are explained.

The noise of the system is also discussed.

3.1 Quasi-Static Model

The proposed shear stress sensor model developed in this chapter consists of a floating

element structure similar to the ones discussed in Section 2.1.2. A rectangular proof mass,

suspended over a small cavity by four compliant tethers, forms the floating element. A flow

across the floating element surface exerts a shear force, which produces a proportional lateral

deflection. The tethers act as springs that restore the floating element to its mean/reference

52

Page 53: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

position in the absence of the flow. One end of each tether is attached to the floating element

and the other end is attached to a fixed substrate. Interdigitated comb fingers between

the fixed substrate and the floating element serve as parallel plate capacitors and provide

electrical transduction. The capacitance due to the additional gaps between the stationary

substrate and both the floating element and tethers is also utilized. The schematic of the

sensor, shown in Figure 3-1, depicts the comb fingers, tethers, and the floating element.

Direction of

flow

y

x

Anchors

Tethers

Wt

Lt

Le

We

Comb

Fingers

We

Tt

z

y

LoWf

TtTt

g01

g02

Lt

Wf

TtTt

g01

g02 Wt

g01

Tt

A

A

Side view

Section A-ATop View

Floating

Element

Figure 3-1. Schematic of geometry of the differential, capacitive shear stress sensor.

A deflection changes the capacitance between the surrounding substrate and the tethers,

floating element and between comb fingers, resulting in a proportional change in voltage

(Section 3.1.3.1). The resulting change in voltage is detected using a voltage buffer (see

Section 3.1.3.3). For analysis purposes, the tethers are modeled as clamped beams and

53

Page 54: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

the floating element is a rigid mass. Nonlinear deflection of the floating element structure

introduces undesirable harmonic distortion in the spectral content of the sensor output.

The nonlinear deflection may be restricted to a small percentage of the linear deflection to

preserve spectral fidelity of the time resolved measurement. Hence, both small deflection

(linear) and large deflection (non-linear) analyses are needed in the design process in order

to predict the linear range of a particular geometry.

3.1.1 Small Deflection Analysis

In this section, the mechanical model of the differential capacitive shear stress sensor

for small deflection is presented. The following assumptions enable the derivation of the

mechanical model using beam theory [36,64].

• The length of the tether, Lt, is much larger than its width, Wt, and tether thickness,

Tt. The tethers may therefore be treated as beams.

• The floating element and the attached comb fingers move as a rigid mass under the

applied shear stress and is modeled for in-plane motions.

• The material is homogeneous, isotropic and linearly elastic.

• Euler Bernoulli beam theory is used for both small and large deflection analyses, re-

spectively [65].

In the simplified mechanical model as shown in Figure 3-2, each pair of tethers form

a clamped-clamped beam, with a point load P applied at center by the shear stress, τw,

on the floating element. A uniformly distributed load, Q, along the length of the tethers

accounts for the shear force on the tethers [37]. Two such beams share the load applied on

the floating element because of the geometric symmetry. Thus a single pair of tethers that

form the beam experience a force due to the shear stress given as

P =τwWeLe

2+

τw (NWfLf )

2[N ] (3–1)

and Q = τwWt, [N/m] (3–2)

54

Page 55: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Direction

of Flow

Le

Clamped

Boundary

MB

RA

P Q

2Lt

P Q

RA

V

Wt

Q

xx=0

Tt

xx=0

x=Lt

MA

MA Mx

RB

x

z

y

Rigid body motion

P

Figure 3-2. Simplified mechanical model of the floating element structure.

where We is the width of the floating element, Le is the length of the floating element, N

is the number of comb fingers on the floating element, Wf is the width of each comb finger,

and Lf is the length of each comb finger. The governing differential equation is given by

Mx

EI=

d2w (x)/dx2

(1 +

(dw(x)

dx

)2) 3

2

. (3–3)

where E is the Young’s modulus of elasticity, Mx is the moment along the length of the

beam, I is the moment of inertia about the y-axis, and w is the in-plane deflection in the

direction of the applied load. For both small and large deflections, the slope dw(x)dx

¿ 1 and

the simplified governing equation becomes

Mx

EI= d2w (x)

/dx2. (3–4)

Substituting for the the moment, Mx, in terms of the forces P and Q results in a third order

differential equation, expressed in terms of internal moments and reaction forces, requiring

55

Page 56: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

three boundary conditions. Since the beam is symmetric, the solution is obtained by solving

for half the length of the beam i.e., a single tether. Furthermore, the length of the floating

element is not accounted for the moment computations because it does not affect the bound-

ary condition at the tethers where they connect to the rigid floating element and hence the

beam deflection characteristics. The boundary conditions for a clamped-clamped beam are

mathematically stated as,

w(x = 0) = 0, (no deflection)

dw

dx

∣∣∣∣x=0

= 0, (zero slope)

anddw

dx

∣∣∣∣x=Lt

= 0. (symmetry) (3–5)

The Euler Bernoulli theory and small beam deflections in response to an applied force are

assumed while solving Equation 3–4. The theory hypothesizes that a straight line transverse

to the neutral axis remains straight, inextensible, and normal before and after deformation

[65]. Assuming pure bending and small deflections allows the nonlinear extensional strain

along the length of the beam to be neglected. The detailed derivation of the model is

presented in Appendix A. The linear deflection using this theory is,

w (x) = − τw

4ETtW 3t

(3 (LeWeLt + NWfLfLt) + 8WtL2t ) x2

− (2LeWe + 2NWfLf + 8WtLt) x3 + 2Wtx4

. (3–6)

The center (maximum) deflection represents the movement of the floating element. Since

most of the transduction takes place via comb fingers and the floating element, the center

(maximum) deflection significantly contributes to the transduction process. Furthermore, in

the lumped element model developed in Section 3.2, this center deflection is used to estimate

the lumped parameters. The maximum deflection at the center of the beam (x = Lt) is

δ = −w (Lt) =τwWeLe

4ETt

(1 +

NWfLf

WeLe

+ 2WtLt

WeLe

)(L3

t

W 3t

). (3–7)

56

Page 57: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

In the summation term in Equation 3–7, there are three terms contributing to the deflection:

the first term is due to the floating element, the second term is due to the comb fingers and

the final one is due to the tethers.

3.1.2 Large Deflection Analysis

Beam bending results in rotation of the transverse normal. For large beam deflections

(rotation of 10− 15), not all of the nonlinear terms in the Green-Lagrange strain displace-

ment relationships are negligible [65]. Solutions are obtained using two different solution

techniques; the Rayleigh-Ritz method (Section 3.1.2.1) and an analytical solution technique

(Section 3.1.2.2). The validity of both these solution techniques have been verified previously

using finite element analysis and compared with the linear quasi-static solution in [64].

3.1.2.1 Energy Method

The energy method solution utilizes the principle of virtual work to arrive at a solution

using the Rayleigh-Ritz method [65]. Appendix A provides the detailed derivation of the

solution using this method. The expression for the floating element deflection based on this

method is

δ

(1 +

3

4

Wt

)2)

=τwWeLe

4ETt

(1 +

NWfLf

WeLe

+ 2WtLt

WeLe

)(L3

t

W 3t

). (3–8)

Equation 3–8 is nonlinear and is solved using numerical solution techniques. Unlike the

previous analytical technique, which yielded a solution for deflection as a function of the

position along the beam, the energy method is used only to solve for the central deflection

with an assumed shape function or mode shape. The center deflection is the quantity of

interest because the permissible nonlinear deflection is defined with reference to the linear

central deflection in Equation 3–7 (see Section 4.3.2.2). The term,

(1 +

3

4

Wt

)2)

, (3–9)

is a consequence of beam stiffening due to deformation along the neutral axis. The solution

approaches the linear solution in Equation 3–7 for small deflections i.e., δ/Wt << 1.

57

Page 58: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

3.1.2.2 Analytical Solution

In this method, an axial force, Fa, responsible for the nonlinear strain along the length

of the beam is estimated. The axial force is obtained by integrating the stress along the

length of the beam. The expression for the central beam deflection (Appendix A) using this

solution technique is,

δ =

− P

2Faλsinh (λLt) +

Q

2Fa

L2t +

P

2Fa

Lt

−(cosh (λLt)− 1)

Faλ sinh (λLt)

(QLt +

P

2− P

2cosh (λLt)

)

,(3–10)

where

λ =

√12Fa

ETtW 3t

, (3–11)

and Fa =EWtTt

2Lt

Lt∫

0

(∂w

∂x

)2

dx. (3–12)

An iterative step-wise method for solving Equation 3–10 is provided in Appendix A.

3.1.3 Electrostatic Behavior

The previous section presented the expressions for the mechanical deflection, δ, of the

floating element sensor structure in response to an applied shear stress, τw. In this section,

the electrical response to a mechanical motion is analyzed. Fundamental relationships for

an electrostatic transducer are provided. A detailed analysis of the interdigitated capacitive

comb fingers and the interface circuitry implemented in the current sensor is developed.

3.1.3.1 Electrostatic Transducers

The capacitive shear stress sensor developed in this thesis is an example of an electro-

static transducer. An accelerometer with comb fingers is another common example of an

electrostatic transducer. An electrostatic transducer consists of at least two electrodes, sep-

arated by a finite distance by a dielectric medium to form one or more variable capacitors.

58

Page 59: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

One of the electrodes remains fixed while the other electrode is free to move in response to

a physical input signal such as pressure, shear stress, acceleration, etc. For a more detailed

discussion on electrostatic transducers the reader is referred to Rossi [41] and Hunt [66].

Tt

g0

L0

Wf

+VB

Ground (0 V)

+VB

C

Direction of

motion

g0

Tt

Figure 3-3. Schematic of parallel plate capacitor analogous to overlapping comb fingers.

Background. A model of a parallel plate electrostatic transducer with variable gap is

shown in Figure 3-3. The transducer consists of two conducting parallel plates separated by

a dielectric medium such as air. One plate is fixed and the other is free to move, such that

the gap between the plates may change. The capacitance between the plates is

C = εA/g, (3–13)

where ε is the permittivity of the medium, A is the area of overlap, and g is the plate

separation distance. For the current configuration, the movable plate is assumed to have

one dimensional motion such that only the gap between the plates changes while the area of

overlap stays constant. For illustration, consider a pair of overlapping comb fingers as the

parallel plates, the initial gap between the plates is g0 and the area is A = L0Tt, where L0 is

the overlapping length. Consequently, when the system is at rest, the capacitance between

the plates is

C0 =εL0Tt

g0

. (3–14)

The following assumptions must be true for Equations 3–13 and 3–14 to hold:

1. g ¿ √A, i.e., g ¿ L0 and g ¿ Tt.

59

Page 60: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

2. The electric field, EE is normal to the plates.

An applied force results in a change in gap δ(t); the time varying capacitance is then given

by

C(t) = C0

[1− δ(t)

g0

]−1

. (3–15)

The voltage between the capacitor plates is V (t) = Q(t)/C(t), where Q(t) is the charge on

the plates. Substituting Equation 3–15 into this expression gives the voltage,

V (t) =Q(t)

C0

[1− δ(t)

g0

]. (3–16)

The change in electrical potential energy and co-energy(*) stored in the capacitor are ex-

pressed as [42]

dEp = Fedδ + V dQ (charge variation), (3–17)

and dE∗p = −Fedδ + QdV (voltage variation), (3–18)

respectively. If no additional charge or voltage is applied to the capacitor electrodes, a change

in stored potential energy is effected by the motion, δ of the movable plate. Consequently,

the electrostatic force is given as,

Fe =dEp

∣∣∣∣Q

= −dE∗p

∣∣∣∣V

. (3–19)

The stored electrical potential energies in the capacitor are

Ep =

Q(t)∫

0

V (t)dQ =

Q(t)∫

0

Q(t)

C(t)dQ =

1

2

Q(t)2

C(t). (3–20)

and E∗p =

V (t)∫

0

Q(t)dV =

V (t)∫

0

C(t)V (t)dV =1

2C(t)V (t)2. (3–21)

Substituting Equation 3–21 and 3–20 into Equation 3–19 results in

Fe =d

(1

2

Q(t)2

C(t)

)= − d

(1

2C(t)V (t)2

). (3–22)

60

Page 61: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Substituting the expression for the time dependent capacitance from Equation 3–15 into the

above expression results in,

Fe = −Q(t)2

2C0g0

(3–23)

or Fe = − V (t)2C0

2g0

(1− δ

g0

)2 . (3–24)

Note that the force can be expressed either in terms of the voltage V (t) or in terms of charge

Q(t). No prior assumptions were made regarding the method of applying charge or voltage

to the capacitor plates. Equation 3–23 indicates that when there is a constant charge across

the capacitor, Fe is independent of the plate motion δ. This avoids the electrostatic pull-in

instability for the constant voltage case in Equation 3–24 as δ approaches g0 [66].

A restoring force due to the mechanical compliance of the structure opposes the deflec-

tion of the movable plate. The mechanical force Fm(t) in terms of the compliance, Cme, is

expressed as

Fm =δ(t)

Cme

. (3–25)

The electrical force and the mechanical force are opposite in direction. Using Equations 3–

16, 3–23, 3–24, and 3–25, the characteristic electrostatic equations for the system are written

as

V (t) =Q(t)

C0

[1− δ(t)

g0

], (3–26)

and F (t) =δ(t)

Cme

− Q(t)2

2C0g0︸ ︷︷ ︸Charge Control

=δ(t)

Cme

− V (t)2C0

2g0

(1− δ

g0

)2

︸ ︷︷ ︸V oltage Control

. (3–27)

Inspection of Equations 3–26 and 3–27 indicates that voltage, V (t), and force, F (t), are

coupled. In addition, the voltage and the force on the capacitor are nonlinearly related. The

system therefore needs to be linearized. The linearization of the system and the approach

for two port modeling is explained in Appendix B. In the next section, a two port model of

the entire sensor structure is illustrated based on the derivations in Appendix B.

61

Page 62: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

3.1.3.2 Electrostatics of the sensor structure

In this section, the electrostatic behavior of the developed shear stress sensor is dis-

cussed. A two port model based on the approach explained in Appendix B is derived to

illustrate the electrostatic transduction mechanism. Consider Figure 3-1 and Figure 3-4,

where there are pairs of capacitances formed by different gaps on either side of the sensor

structure as follows:

• C1a , C2a , between interdigitated comb fingers with primary gap g01

• C1b, C2b

, between interdigitated comb fingers with secondary gap g02

• C1c , C2c , between the floating element edges normal to the flow and the surrounding

substrate with primary gap g01

• C1d, C2d

, between the tether and the surrounding substrate with primary gap g01

• C1e , C2e , between the tether and the surrounding substrate with secondary gap g02

The comb finger capacitance formed by N interdigitated comb fingers results in N − 1

capacitors connected in parallel. Half the comb fingers form capacitors with the gap g01 and

the other half form capacitors with the gap g02. The nominal comb finger capacitances for

both the primary and secondary gaps are therefore written as

C1a = C2a =(N − 1)

2

εTtL0

g01

(3–28)

and C1b= C2b

=(N − 1)

2

εTtL0

g02

. (3–29)

The capacitances C1c and C2c are,

C1c = C2c =εTtLe

g01

. (3–30)

While the comb finger and floating element capacitances are easily modeled as parallel

plate capacitors, the tethers acting as cantilevers form a non-uniform gap requiring further

model simplification. Consider the tether capacitor with non-uniform gap as shown in Fig-

ure 3-5. Since the electrode surfaces are metalized, a uniform surface charge density, σ, is

62

Page 63: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

L0 Wf

TtTt

g01

g02

Interdigitated

comb finger(C1a,C2a)

(C1b,C2b)

a)

g01

Floating Element

g01 = Primary gap

g02 = Secondary gap

(C1c,C2c)

Surrounding silicon

substrate

b)

Lt

Wf

TtTt

g01

g02 Wt

Surrounding silicon

substrate

Tether(C1d,C2d)

(C1e,C2e)

Surrounding

silicon substrate

c)

Figure 3-4. Simplified schematic of individual a)tether, b)comb finger, and c)floating elementcapacitances, respectively.

assumed on the tethers and the surrounding substrate. By definition, the capacitance is [67],

−∫

+

EE · ds = Q/C, (3–31)

where EE is the electric field, ds is the differential vector distance/gap between the two

electrodes (positive [+] and negative [−]). For a parallel plate capacitor with a dielectric

63

Page 64: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

g20

g10

Tether

y

Lf

Surrounding

Substrate

Surrounding

Substrate

(C1d,C2d)

(C1e,C2e)

Figure 3-5. Schematic of non-uniform tether capacitors with primary and secondary gaps.

with constant permittivity, ε, EE is normal at both conductor surfaces, which simplifies

Equation 3–31 to

−∫

+

EE · ds =σ

ε

−∫

+

ds = Q/C. (3–32)

Assuming small deflection of the tethers, it is reasonable to assume that the electric field is

normal to the conducting surfaces, forming the tether capacitors. Now consider an elemental

area dA of the tether; using Equation 3–32, the capacitance formed by this element is

dC =dQ

σε

g(x)∫0

dg

. (3–33)

Expressing charge as dQ = σdA, Equation 3–33 is rewritten as

dC =εdA

g(x). (3–34)

Since the thickness stays constant over the length of the tether, dA = Ttdx. The gap

reduction/increase due to the deflection of the tethers is

g(x) = g0 ± w(x). (3–35)

64

Page 65: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

For decreasing gap, the capacitance is evaluated by using Equation 3–6 for w(x) and solving

the integral,

Cng =

Lt∫

0

εTt

g0 − τw

4ETtW 3t

(3 (LeWeLt + NWfLfLt) + 8WtL2t ) x2

− (2LeWe + 2NWfLf + 8WtLt) x3 + 2Wtx4

dx, (3–36)

where ng indicates non-uniform gap. Equation 3–36 is numerically evaluated in MAPLE.

The change in capacitance due to the non-uniform gap variation is computed using

∆Cng = Cng − Ctether, (3–37)

where Ctether =εTtLt

g0

. (3–38)

The capacitance change is also computed using a simple parallel plate model with tether

length, Lt and using the center deflection δ for the gap change. The capacitance change for

this case is

∆Cparallel = Ctetherδ

g0

. (3–39)

Equations 3–37 and 3–39 are numerically compared for a few geometries for small deflections

using MAPLE. Table 3-1 shows the geometry and parameters for this study. The results in

Table 3-1 indicate that for small deflections, ∆Cparallel is approximately twice the value of

∆Cng. For ease of modeling, this difference is accounted as an effective tether length, Lteff ,

allowing the tethers to be modeled as parallel plate capacitors. The effective tether length

is therefore,

Lteff ≈ Lt

2. (3–40)

65

Page 66: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Table 3-1. Sensor geometry for effective tether lengthcalculation.

Parameter Value

Shear stress τw 10 Pa

Tether Length Lt 1000 µm

Tether Width Wt 10 µm

Floating Element Length Le 1000 µm

Floating Element Width We 1000 µm

Tether Thickness Tt 45 µm

Number of comb fingers N 100

Comb Finger Width Wf 4 µm

Comb Finger length Lf 150 µm

Initial gap for tethercapacitance g0

3.5 µm

Dielectric Permittivity ε 8.85× 10−12 F

Floating Element Young’sModulus E †

160 MPa

∆Cparallel 6.637 fF

∆Cng 13.65 fF

† Average value in the 100 plane of silicon.

Using the parallel plate model, the tether capacitances for primary and secondary gaps are

C1d= C2d

=εTtLteff

g01

(3–41)

and C1e = C2e =εTtLteff

g02

, (3–42)

respectively.

All the capacitances with the same subscript (1 and 2) are in parallel by design. The

net capacitance is therefore the sum of the individual capacitances expressed as

C1 = C1a + C1b + C1c + C1d + C1e (3–43)

and C2 = C2a + C2b + C2c + C2d + C2e. (3–44)

66

Page 67: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Each component capacitor of C1 forms a differential capacitance with the corresponding

component of C2. Thus, C1 and C2 together form the differential capacitance pair for the

sensor. The nominal capacitances obtained by substituting Equations 3–28 - 3–30, 3–41,

and 3–42 in Equations 3–43 and 3–44 are

C0 = C10 = C20 =(N − 1)

2

εTtL0

g01

+(N − 1)

2

εTtL0

g02

+εTtLe

g01

+εTtLteff

g01

+εTtLteff

g02

or C0 = εTt

[3Lteff + Le + (N−1)

2Lf

g01

+Lteff + (N−1)

2Lf

g02

]. (3–45)

The differential capacitance model thus formed using the asymmetric gaps is graphically

represented using two capacitors for each of the sensing capacitors C1 and C2 (Figure 3-6).

C1

g01

g02

C2

g01

g02

Figure 3-6. Differential capacitance sensing model for sensing capacitors with the asymmetricgaps.

67

Page 68: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

For ease of representation an overall effective length is defined based on Equation 3–45 and

is given as

Leff =

(3Lteff + Le +

(N − 1)

2Lf

)+

(Lteff +

(N − 1)

2Lf

)g01

g02

. (3–46)

Substituting Equations 3–46 into 3–45, the nominal capacitance is simplified and rewritten

as

C10 = C20 =εTtLeff

g01

. (3–47)

In this section, the electrostatic equations were derived for a parallel plate transducer in

general and for a single pair of comb fingers in particular. Capacitances formed by different

geometric elements of the sensor structure, tethers, comb fingers, and floating element, were

described. An effective tether length was determined for modeling tether capacitance change

using the parallel plate assumption. Lastly, the nominal capacitances for the differential

capacitor scheme to be implemented for the proposed sensor were presented. The next

section describes details of the sensing circuitry.

3.1.3.3 Interface Circuits

In a capacitive sensor the change in capacitance usually manifests itself in terms of a

proportional charge or voltage change. Either the voltage is measured while holding the

charge constant or vice-versa. A sensing circuitry is always required to measure charge

or voltage change in capacitive sensors due to the large sensor impedance. This section

concentrates on the circuit requirements to measure the voltage/charge output obtained from

the capacitive floating element sensor with interdigitated comb fingers shown in Figure 3-

1. The measurement requirements for the sensor include, measuring dc and ac change in

capacitance due to shear stress. Typical sensor parameters that influence the interface circuit

design are C0 ≈ 0.4− 1.5 pF , ∆C ≈ 3− 25 fF . The sensor designs target a maximum shear

stress of 10 Pa. A high sensor resolution enables measurement of weak turbulent motion

and maximizes dynamic range. This requires the interface circuit to have a low noise floor.

68

Page 69: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Furthermore, parasitic capacitances in the system which effect sensor performance have to be

considered while designing the interface circuit. With these requirements in mind, existing

analog interface circuits for capacitive transducers are briefly reviewed. Next, equations

for the synchronous modulation and demodulation (MOD-DMOD) technique, used for the

capacitive shear stress sensor, are derived.

Interface circuits for capacitive sensors can be classified into two basic categories, open-

loop and closed-loop. Each method can be further sub-categorized into analog (no clock

signals) and digital (uses clock signals) based on the type of electronic components. Figure 3-

7 shows the block diagram illustrating the classification of interface circuitry. This work

focuses on open-loop techniques with analog electronic components.

Interface Electronic

Circuits

Open Loop Closed Loop

Continuous

Time

Discrete

Time

Switched

Capacitor

Methods

Charge

Amplifier

Voltage

Amplifier

Synchronous

MOD/

DEMOD

Oscillators

(LC and RC)

Figure 3-7. Simplified classification of capacitive interface electronic circuits.

Before describing the details of the proposed sensor circuitry, existing open-loop, analog

interface circuit configurations will be discussed. The most commonly used analog, open-loop

69

Page 70: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

interface circuits include charge amplifiers, voltage amplifiers, oscillators (RC and LC), and

synchronous modulation/demodulation (MOD-DMOD) techniques [68–71]. The principle of

operation of oscillators (RC and LC) formed using sense capacitances is different from the

other three techniques. Oscillators require complicated interface circuitry and are beyond the

scope of this work. The primary goal of the current work is to obtain reliable sensor output

voltage, while limiting the complexity of the sensor fabrication and interface electronics.

At a constant capacitance, charge and voltage were established in Equation 3–16 to be

directly proportional. Any change in capacitance with a constant charge or voltage applied

to the capacitor produces a proportional voltage or charge output, respectively. An interface

circuit can therefore be designed to sense either the charge or the voltage across the capacitor.

Constant voltage biasing with a charge amplifier and constant charge scheme in conjunction

with a voltage amplifier are the two techniques generally used in capacitive sensors, resulting

in an output voltage.

Comparative Study. The floating element sensor has two sets of capacitors C1 and

C2 as discussed previously in Section 3.1.3.1 and also shown in Figure 3-1. This sensor geom-

etry enables measurement of a differential capacitance change. Both single and differential

capacitance sensing schemes have been implemented for sensors previously. Martin [72] pro-

vides a comparison of single and differential capacitance sensing schemes using both charge

(constant voltage bias) and voltage (constant charge bias) amplifiers. An appropriate bias-

ing scheme for an electrostatic transducer is important from the interface circuit and overall

performance perspective. It is therefore relevant to perform a comparative study of charge vs

voltage amplifiers and single vs differential capacitance measurement schemes. Kadirvel [73]

and Martin [72] provide insight into the aspects of different biasing techniques. In the past

several other researchers have also studied the interface circuit schemes using charge and

voltage amplifiers [69–71]. In this section, a comparative study of the constant charge and

constant voltage biasing schemes is presented.

70

Page 71: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

A common input signal to the capacitances, such as hydrostatic pressure fluctuations,

ideally results in no net differential capacitance change. Thus, output due to any common

mode signal is suppressed by the differential measurement scheme. An added benefit of using

the differential capacitance measurement scheme, in conjunction with a charge amplifier, is

the doubling of the sensitivity compared to the single capacitance measurement [72]. For

a variable gap capacitive sensor, such as a microphone, a constant voltage biasing with a

voltage amplifier makes the sensor electrodes susceptible to electrostatic pull-in [42,66]. Pull-

in occurs when the mechanical restoring force is unable to balance the electrostatic force,

causing instability. In this biasing scheme, the reduction of the gap between the electrodes

increases the electrostatic force (refer Equation 3–24), eventually leading to pull in. The use

of constant charge biasing in combination with a voltage amplifier avoids this problem. The

constant charge biasing creates a constant electric field across the gap independent of the

plate motion (see Equation 3–23), preventing electrostatic pull-in [66].

The study conducted by Kadirvel [73] reveals several details about sensitivity and noise

in charge and voltage amplifiers. The sensitivity of charge amplifiers is independent of the

parasitic capacitance, while parasitics have a pronounced effect on the sensitivity of voltage

amplifiers. The adverse effect of parasitics in voltage amplifiers increases significantly with

a drop in sensor capacitance. However, with an increase in parasitic capacitance, the noise

increases in charge amplifiers. The study also revealed that noise performance in both charge

and voltage amplifiers is better when they are optimized for low current noise.

Charge Amplifiers. A charge amplifier circuit typically has an integrating capacitor,

Cf , in the feedback path that is inversely proportional to sensitivity, which is expressed as

Schg−amp = −2Vb

g0

C0

Cf

δ

τw

. (3–48)

Figure 3-8 shows a typical charge amplifier circuit with noise sources. While decreasing Cf

increases the sensitivity, it also increases the cut-on frequency of the circuit. The cut-on

71

Page 72: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

B

B

1

2

f

outin

f

iRf

Va

ia

p i

Figure 3-8. Simplified sensor circuitry with a charge amplifier along with noise sources.

frequency is

fcut−on =1

2πRfCf

, (3–49)

where Rf is the feedback resistor that sets the dc operating point. Lowering Cf also lowers

the bandwidth and increases the noise in charge amplifiers [73]. This can be seen from

the expression for the total noise power spectral density (PSD) at the output of a charge

amplifier, given as

Sv0chg−amp= Sva

(1 +

Ctot

Cf

)2

+

∣∣∣∣1

jωCf

∣∣∣∣2 [

Sia + SiRf

], (3–50)

where Ctot = C1 +C2 +Cp +Ci. Cp and Ci represent the parasitic and the input capacitances

in the system, respectively. SiRf= 4kT/Rf is the noise PSD of Rf represented as a current

source and Sva and Sia are the amplifier voltage and current noise PSDs, respectively. The

noise due to Rf is 1/f 2 shaped and increases significantly at low frequencies [73]. A high

value of Rf can help mitigate this effect by lowering the current noise, 4kT/Rf . This also

lowers the cut-on frequency, thereby increasing the bandwidth. The highest value of Rf that

72

Page 73: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

may be used is limited by the availability of off-the-shelf components (1 GΩ) for a board

level circuit implementation. The limiting values of Rf for a CMOS based interface circuit

implementation can be much smaller (O(100 KΩ)). A study of the specific circuit under

consideration is needed to truly understand the noise performance tradeoffs.

Voltage Amplifiers. A voltage amplifier typically has a bias resistor, Rb, that sets

the dc operating point for the amplifier. Figure 3-9 shows a typical voltage amplifier circuit

including dominant noise sources. Any capacitance at the input of the amplifier, except

the sensor capacitance, forms a potential divider reducing the voltage at the input of the

amplifier. The expressions for sensitivity and noise for the voltage amplifier circuit are

Svamp =C1 − C2

C1 + C2 + Cp + Ci

(Vb

g0

)(δ

τw

)(3–51)

and Sv0v−amp= Sva + Sia

(RbZi

Rb + Zi

)2

+ Svb

(Zi

Rb + Zi

)2

, (3–52)

where Zi = 1/(jωCtot) and Svb = 4kTRb is the voltage noise PSD of Rb. Higher values of

parasitics in comparison to the sensor capacitance lower the sensitivity (see Section 3.2.3). A

larger sensor capacitance should therefore help mitigate the ill effects of parasitics. However,

the ratio, ∆CC0

, should not drop significantly because it results in the loss of sensitivity.

This will be illustrated mathematically while explaining the expression for sensitivity of

the proposed sensor in Section 3.2.3. The combination of the bias resistor Rb and the total

capacitance Ctot (including parasitics), form a high pass filter. The cut-on frequency, 12πRbCtot

can be lowered using a large bias resistor. However, this will increase the noise contribution

at the output. The simplicity of the voltage amplifier circuit makes it an attractive option

for capacitive sensors. The reader is referred to Kadirvel’s work [73] on interface circuits for

a detailed understanding of the tradeoffs involved in using charge and voltage amplifiers.

Martin [74] reported better results using a voltage amplifier compared to a charge ampli-

fier with off the shelf components for a dual back plate capacitive microphone with nominal

capacitances similar to the sensor developed in this work. The same voltage amplifier will

be used for the proposed sensor. Desirable traits of shear stress sensors include their ability

73

Page 74: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

-

Rb

Cp Ci

Svb=

4kTRb

Sva

Sia

Vout

VB

-VB

C1

C2

Figure 3-9. Simplified sensor circuitry with a voltage amplifier along with noise sources.

to measure both the mean and fluctuating shear stress components. In the proposed sensor,

mean and fluctuating shear stress components would correspond to mean and fluctuating

changes in capacitance respectively. As stated previously, both charge and voltage amplifiers

possess a characteristic cut-on frequency, 1/2πRC, making them unsuitable for measuring

dc (mean) capacitance changes.

The synchronous MOD-DMOD scheme enables the measurement of static capacitance

changes or making mean measurements. This technique has been used widely in the past for

capacitive inertial sensors [69, 70]. The following section explains the MOD-DMOD scheme

for differential capacitance measurement using a voltage amplifier. The interface circuit

is explained and derivations for the output voltage of the system are provided. For the

derivations presented, each capacitor is assumed to be a parallel plate capacitor.

MOD-DMOD with Voltage Amplifier. A MOD-DMOD circuit for differential

capacitance measurement using a voltage amplifier is as shown in Figure 3-10. Each sense

capacitor is represented as a parallel combination of two capacitors which change in the

opposite sense. The capacitors in parallel correspond to the gaps, g01 and g02 that change in

a opposite sense with the sensor motion. A unity gain voltage follower is used to represent

the voltage amplifier. The non-inverting input is connected to the common electrode of the

capacitor (the floating element in this case). The other two electrodes of the capacitors

74

Page 75: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

+

-

vc=vacsin(ct)

-vc=-vacsin(ct)

C1

C2 vout1

RbCp Ci

vc=vrefsin(ct)

vout

g02

g02

g01

g01

Multiplier/

Demodulator

low-pass

filter

Figure 3-10. Simplified differential capacitance sense circuitry using synchronous modulationand demodulation technique with a voltage amplifier.

are biased directly using sinusoidal voltage signals, +vc and −vc respectively. For similar

implementations for inertial sensors, square waves are generally used [75]. However, square

waves inherently have harmonic content which may interfere with the measurement signal.

Square waves also possess less power in the fundamental frequency compared to sine waves

at a given amplitude, lowering the sensitivity (Section 3.2.3). Cp is the parasitic capacitance

due to on board connection lines and wire bonds. Ci is the input capacitance of the amplifier.

The bias resistor Rb sets the dc operating point of the amplifier. The combination of the

total capacitance at the input of the amplifier Ctot and Rb form a high-pass filter with

a cut-on frequency, 1/2πRbCtot. The charge at the common node remains approximately

constant provided the ac bias voltage frequency is greater than the cut-on frequency, i.e.,

fcut−on > 1/2πRbCtot [66]. The ac bias voltages on the two capacitors are equal in amplitude

and frequency, but opposite in phase. Thus, in the absence of a physical excitation, i.e., no

change in capacitance, the voltage at the common node is ideally zero (C1 ≈ C2). In case

of mismatched capacitors, the nominal output voltage will have an amplitude proportional

to the mismatch at the biasing frequency, restricting the dynamic range of the sensor. An

appropriate phase adjustment circuitry may be used to null the output resulting from an

initial mismatch in nominal capacitances (see Chapter 6). A shear stress input causes a

change in capacitance resulting in a voltage change at the middle node. The amplifier

75

Page 76: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

provides a voltage output corresponding to the capacitance change. Note, the mean wall

shear stress acting on the sensor produces a dc change in capacitance. In the absence of

fluctuating bias voltages, an output due to a dc change in capacitance is high-pass filtered.

The electromechanical coupling is not considered in the present analysis. The expressions

derived in this section are purely based on a change in capacitance due to a finite change in

gap between capacitive sensor elements i.e., free electrical impedance (refer Appendix B).

For ease of understanding, Figure 3-11 shows a simple schematic showing the expected

sensor output spectrum at every stage of the circuitry presented. The biasing results in a

modulated output at the amplifier output, where the mean change in capacitance is at the

carrier frequency and the dynamic change is observed at the sidebands of the modulated

output. There is an additional noise component from the amplifier. The modulated output

is then demodulated and low pass filtered to extract the original modulating/input signal.

There will be additional dc offset voltages and noise contribution from the demodulator

and the low pass filter, which are not shown in the figure. Especially, the net 1/f noise

component at the output, at low frequencies, needs consideration to ensure sufficient signal

resolution. The derivations that follow provide a better understanding of the benefits of the

modulation technique to measure static capacitance change and hence the mean shear stress.

The instantaneous capacitances for the sensor are

C1 = εTt

[3Lteff + Le + (N−1)

2Lf

g11

+Lteff + (N−1)

2Lf

g12

](3–53)

and C2 = εTt

[3Lteff + Le + (N−1)

2Lf

g21

+Lteff + (N−1)

2Lf

g22

], , (3–54)

where g11 and g12 are instantaneous gaps for C1 and g21 and g22 are for C2. The instantaneous

charge on the capacitors C1 and C2 are

q1 = vcC1 (3–55)

and q2 = −vcC2. (3–56)

76

Page 77: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

dc change

ac change ()

frequency

PS

D

Amplifier Noise PSD

frequency

~ dc change x

carrier amplitudeamplifier

PSD

frequency dc change multiplier/

demodulator

low-pass filter

PS

D

frequency dc changeFigure 3-11. Schematic indicating spectra of the sensor output at each stage of the sensor

circuitry.

where, ±vc = ±Vac sin (ωct) are the sinusoidal bias voltages that act as the carrier signals

for the modulation process.

The bias voltage frequency is chosen to be higher than the bandwidth of the sensor while

ensuring that no charge dissipates through Rb. Consequently, the bias resistor is essentially

treated as an open circuit for the voltage analysis. The combination of sense and parasitic

capacitances form a potential divider, with bias voltage sources connected to one electrode

of each of the sense capacitors. The analysis assumes one sense capacitor is biased at a time,

followed by superposition to obtain the net output voltage.

77

Page 78: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Consider the case when capacitor C1 is biased. The total charge, Qin1 , is the same as

the charge across the sense capacitance, C1. Therefore,

Qin1 = Q1 (3–57)

and Ctot1vc = C1(vc − vinC1), (3–58)

where Ctot1 = C1(C2+Cp+C1)

C1+C2+Cp+Ci. Substituting for Ctot1 and rearranging the equation results in,

vinC1= vc

C1

C1 + C2 + Cp + Ci

. (3–59)

Similarly when the capacitor C2 is considered,

Qin2 = Q2 (3–60)

and Ctot2(−vc) = C2

(−vc − vinC2

), (3–61)

where Ctot2 = C2(C1+Cp+C1)

C1+C2+Cp+Ci. Substituting for Ctot2 and rearranging the equation results in

vinC2= −vc

C2

C1 + C2 + Cp + Ci

. (3–62)

The total input voltage is due to the bias voltages applied to both C1 and C2, obtained by

adding Equations 3–59 and 3–62. The result is,

vin = vinC1+ vinC1

and vin = vcC1 − C2

C1 + C2 + Cp + Ci

. (3–63)

The output voltage of the amplifier is

vout1 = Gvin,

and vout1 = GvcC1 − C2

C1 + C2 + Cp + Ci

, (3–64)

where G is the closed loop gain of the amplifier. The voltage follower shown in Figure 3-10

has a unity gain; i.e., G = 1. The equations however, are derived for an amplifier with

arbitrary gain. Equations 3–63 and 3–64 verify that the input, and hence the output of the

78

Page 79: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

amplifier, is proportional to the differential capacitance. Substituting for the capacitances

from Equations 3–53 and 3–54, Equation 3–64 can be simplified and rewritten as,

vout1 = Gvc

(3Lteff + Le + (N−1)

2Lf

) [g21−g11

g11g21

]+

(Lteff + (N−1)

2Lf

) [g22−g12

g12g22

](3Lteff + Le + (N−1)

2Lf

) [g21+g11

g11g21

]+

(Lteff + (N−1)

2Lf

) [g22+g12

g12g22

]+ Cp+Ci

εTt

.

(3–65)

By design, the nominal capacitances C10 and C20 are assumed to be equal as shown in

Equation 3–45. Thus, the changes in capacitances are also equal in magnitude but opposite

in direction. Consider a floating element motion such that C1 increases and C2 decreases,

resulting in

g11 = g01 − δ, g12 = g02 + δ, g21 = g01 + δ, and g22 = g02 − δ. (3–66)

Note here the deflection δ is only used to analyze the change in capacitance. The effect of

electromechanical coupling is not included and will be accounted for later (see Section 3.2.3).

Substituting Equations 3–66 and 3–45 into Equation 3–65 gives

vout1 = Gvc

1−

(Lteff +

(N − 1)

2Lf

)

(3Lteff + Le + (N−1)

2Lf

)[g201

g202

]

1 +

(Lteff + (N−1)

2Lf

)(3Lteff + Le + (N−1)

2Lf

)[g01

g02

]

(2C0

2C0 + Cp + Ci

g01

. (3–67)

The term,

Hgap =

1−

(Lteff + (N−1)

2Lf

)(3Lteff + Le + (N−1)

2Lf

)[g201

g202

]

1 +

(Lteff + (N−1)

2Lf

)(3Lteff + Le + (N−1)

2Lf

)[g01

g02

] , (3–68)

in Equation 3–67 represents an attenuation term due to the presence of the secondary gap

g02 which partially cancels out the capacitance change due to the primary gap g01. This is

79

Page 80: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

represented graphically using two capacitors for C1 and C2 in Figure 3-6 and Figure 3-10.

As g01/g02 → 0, Hgap → 1 i.e., no attenuation. On the contrary, as g01/g02 → 1, Hgap → 0

i.e., 100% attenuation. If g01/g02 > 0, the differential scheme still holds, but the two gaps

switch functionalities, where g01 acts as the secondary gap and g02 acts as the primary gap.

The term,

Hc1 =2C0

2C0 + Cp + Ci

, (3–69)

in Equation 3–67 represents the attenuation due to parasitics [63]. However, the effect of

electromechanical transduction is not accounted yet and will be incorporated in the next

section. The effects of attenuation and the design considerations that are required to mini-

mize them are also discussed later. Substituting Equations 3–68 and 3–69 in Equation 3–67,

and setting bias voltage, vc = Vac sin (ωct), gives

vout1 = GHgapHc1

g01

)Vac sin (ωct) . (3–70)

Equation 3–70 represents the sensor output voltage amplitude that is modulated by the

carrier signal. The modulated output signal is subsequently demodulated using a multiplier

(demodulator) to retrieve the original signal.

During the demodulation process, the output from the voltage amplifier is multiplied

with a reference signal having the same frequency, ωc, as the sinusoidal bias voltages. As-

suming a reference voltage, vref = Vref sin (ωct), the output of the multiplier is

vout =1

U(vout1vref ) , (3–71)

where U is the scaling factor of the multiplier with units in volts (V). Combining Equations 3–

70 and 3–71 and substituting for vref results in

vout = HgapHc1

(δ(t)

g01

)(G

U

)VacVref sin2 (ωct)

or vout = HgapHc1

(δ(t)

g01

)(G

2U

)VacVref (1− cos (2ωct)) . (3–72)

80

Page 81: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

In a turbulent flow, the shear stress exerted by the flow has two components; shear stress due

to the mean flow and the fluctuating shear stress due to turbulent structures. A sinusoidal

fluctuation is assumed as a simple case for analysis purposes. The resulting deflection is

written as

δ(t) = δm + δt sin (ωt + φ) , (3–73)

where δm is the deflection due to the mean flow, and δt sin (ωt + φ) is the deflection due to

turbulent flow. Combining Equation 3–73 for deflection with Equation 3–72 results in

vout = VacVrefHgapHc1

(G

2U

)(1

g01

)(δm + δt sin (ωt + φ)) (1− cos (2ωct)) (3–74)

or vout = VacVrefHgapHc1

(G

2U

)(1

g01

)

δm + δt sin(ωt + φ)

−δm cos(2ωct)

−δt sin(ωt + φ) cos(2ωct)

. (3–75)

Equation 3–75 gives the net demodulated voltage output from the multiplier. This

signal has several frequency components, a dc component, δm, a low frequency component

from the flow, δt sin(ωt + φ), and high frequency components due to the carrier signal,

δm cos(2ωct) − δt sin(ωt + φ) cos(2ωct). The components of interest are the dc and the low

frequency signals, corresponding to the mean and fluctuating parts of the flow, respectively.

The higher frequency components are therefore low-pass filtered after the demodulation

process. The low-pass filtered output voltage is then given as

vout = VacVrefHgapHc1

(G

2U

)(1

g01

)(δm + δt sin(ωt + φ)) . (3–76)

In this section, the expression for output voltage using the proposed interface circuitry

using capacitors C1 and C2 were presented. Voltage output was derived purely based on

capacitance change resulting from electrode motion, disregarding the effect of electrome-

chanical coupling. An attenuation factor, Hgap due to the secondary gap g02 was identified.

So far, a mechanical model and an electrical model have been developed for the proposed

shear stress sensor. In the next section, simplified passive elements (lumped elements) for

81

Page 82: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

the sensor are obtained from the mechanical model. The mechanical and electrical domain

are then related using an electromechanical transduction scheme.

3.2 Dynamic Modeling

The dynamic model of the sensor is developed using the principle of lumped element

modeling. A brief description of the concept of lumped element modeling is provided here

before delving into the particular details for the sensor. In a lumped element model (LEM),

the distributed properties of the system such as mass, compliance and dissipation are lumped

about a single point of interest. This simplifies the mathematical modeling of the sensor

structure, which is necessary for predicting the sensor response to a given input. However,

the validity of LEM requires that the wavelength of physical input signal be much larger

than the length scale of the device. This allows for the decoupling of spatial and temporal

variations in the governing system equations and application of the quasi-static solution to

solve for the dynamic system response [41].

Lumping about a single point allows distributed system representation using discrete cir-

cuit/mechanical elements (inductor/intertance, capacitor/compliance and resistor/damper).

There are three kinds of passive circuit elements: the generalized resistor, the generalized

inductor and the generalized capacitor. The generalized resistor represents the dissipation,

the generalized capacitor represents the storage of potential energy, and the generalized in-

ductor represents the storage of kinetic energy in the system for the respective domains. In

the mechanical domain the deflected spring stores the potential energy, the moving mass

possesses kinetic energy, and the damper acts as the dissipator.

To track power transfer during electromechanical transduction, the conjugate power

variables of effort and flow are generally used. The product of the conjugate power variables

in each energy domain results in power. For example, in the mechanical domain, force is

the effort variable and velocity is the flow variable. The dynamic mechanical system equa-

tions are later studied using equivalent Kirchoff’s laws for velocity (analogous to current),

and shear force (analogous to voltage). An equivalent circuit is formed based on common

82

Page 83: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

flow or common effort shared by individual lumped elements. The analysis of the circuit

yields information such as bandwidth and sensitivity of the sensor. A noise analysis of the

equivalent circuit helps to estimate the minimum detectable signal (MDS).

3.2.1 Lumped Parameters

In this section, the lumped mechanical elements for the floating element structure are

presented. Since, the transduction is effected by the mechanical shear force acting on the rec-

tilinearly translating floating element structure, it is intuitive to use the mechanical lumped

elements to model the sensor. First, the lumped mass, Mme is estimated followed by the

lumped compliance, Cme. The derivations for Mme and Cme are presented in Appendix A.

When a deflection occurs due to shear force on the sensor, the tethers act as the springs

(compliance), storing potential energy. The distributed mass of the tethers and the rigid

mass of the floating element with comb fingers, lumped to the center, possess kinetic energy.

Since, majority of the transduction occurs at the floating element (x = Lt), the center de-

flection, δ = w(Lt), seems to be the natural choice for lumped parameter estimation. There

are two different sources of dissipation: elastic damping due to internal friction and fluidic

damping due to viscous effects under the suspended floating element structure [42].

A linear system is required to ensure spectral fidelity of the sensor. The energy and the

co-energy of a linear system are equal [41]. Therefore, the linear mechanical (Equation 3–6)

is used to compute the lumped mass and compliance of the floating element structure instead

of the nonlinear model.

3.2.1.1 Lumped Mass

To compute the lumped mass, Mme, the total kinetic co-energy due to the motion of

the beam is equated to the equivalent kinetic co-energy of the lumped system. For a linear

system,

W ∗KE =

∫dW ∗

KE =

∫ f0

0

pdf =1

2Mmef0

2, (3–77)

where p represents momentum and f0 represents flow. The total lumped mass is the sum

of the effective mass of the tethers lumped about the center, and the mass of the floating

83

Page 84: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

element. As shown in Appendix A, the lumped mechanical mass of the sensor is

Mme = ρTtLeWe

1 + 19235

(WtLt

LeWe

)+ 3

(NWf Lf

LeWe

)+ 384

35

(WtLt

LeWe

)(NWf Lf

LeWe

)

+1072105

(WtLt

LeWe

)2

+ 3(

NWf Lf

LeWe

)2

+ 1072105

(NWf Lf

LeWe

)(WtLt

LeWe

)2

+3(

NWf Lf

LeWe

)2 (WtLt

LeWe

)+ 2048

315

(WtLt

LeWe

)3

+(

NWf Lf

LeWe

)3

(1 +

NWf Lf

LeWe+ 2 WtLt

LeWe

)2 . (3–78)

3.2.1.2 Lumped Compliance

The total potential energy is equated to the equivalent potential energy of the lumped

system to estimate the lumped compliance. For a linear system,

W ∗PE =

∫dW ∗

PE =

∫ q0

0

edq =1

2

δ2

Cme

, (3–79)

where e and q are effort and displacement respectively. As shown in Appendix A, the

compliance of the tethers is given as

Cme =1

4ETt

(Lt

Wt

)3

(1 +

NWf Lf

WeLe+ 2

(WtLt

WeLe

))2

1 + 2(

NWf Lf

WeLe

)+ 4

(WtLt

WeLe

)+ 4

(NWf Lf WtLt

(WeLe)2

)

+6415

(WtLt

WeLe

)2

+(

NWf Lf

WeLe

)2

. (3–80)

The packaging and the flow around the sensor decide the damping in the system. An

accurate model for the damping is not available at the moment. The damping in the system

is therefore neglected for modeling purposes and will be experimentally deduced.

3.2.2 Electromechanical Transduction

This section presents the approach used to model energy transduction from the me-

chanical domain to the electrical domain. This energy transfer is modeled using idealized

two-port elements which permit energy transduction sans losses. A general approach for

modeling linear conservative transducers and an example of a capacitive transducer is pre-

sented in Appendix B. The two port model, transduction factor, and the electromechanical

84

Page 85: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

coupling factor for the shear stress sensor are provided. First, the transduction process us-

ing a single capacitor transducer is explained followed by one for the differential capacitance

scheme.

An impedance analogy is used for modeling the transduction process. The approach

enables transduction of force to voltage (both effort variables). For the capacitive shear

stress sensor, a transformer is used to model the transduction of the mechanical energy into

electrical energy. The electromechanical transduction factor, ϕ′, also known as the turns

ratio in a transformer, accounts for the coupling between two different energy domains.

The actual change in capacitance due to the effective shear force, fe, that is transduced, is

included via the fictitious transduction voltage, v0, to maintain a constant charge. Using

the equations developed in Appendix B, the sensor two-port model using a single sensor

capacitance, C10, is expressed as,

Vout

Fshear

=

1

jωC10

− Vb

jωgeff

− Vb

jωgeff

1

jωCme

I

U

, (3–81)

where Vb is the bias voltage (ac or dc) across the electrodes. In Equation 3–81, CMO and

CEB in Equation B–30 have been replaced with the Cme and C10, respectively. Therefore,

CMO = Cme (open mechanical compliance) (3–82)

and CEB = C10 = C20. (blocked electrical capacitance) (3–83)

Similarly, the electromechanical transduction factor is

ϕ′ =Vb

geff

Cme, (3–84)

where

geff =g01

Hgap

. (3–85)

85

Page 86: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

me fe o

10inC1

p i

φ’

Figure 3-12. Schematic of capacitive transduction scheme using a single sense capacitance.

Equation 3–85 represents the effective gap, which accounts for the loss in sensitivity due

to gap attenuation discussed earlier using Equation 3–68. The electromechanical coupling

factor is

κ2 =V 2

b

g2eff

C10Cme. (3–86)

For illustration purposes, Figure 3-12 shows the transduction mechanism for a capacitive

transducer in general, and the proposed shear stress sensor in particular with a single ended

capacitance (C10). The free electrical capacitance C ′10 is defined as

C ′10 =

C10

(1− κ2). (3–87)

In Equation 3–87, the effect of the electromechanical coupling has been incorporated. The

voltage, v0 is related to the effective shear force, fe, acting on the floating element sensor

through the turns ratio as follows,

v0 = ϕ′fe. (3–88)

Substituting Equation 3–84 in the expression above, the voltage, vin at the input of the

interface circuitry is

vin =C ′

10

C ′10 + Cp + Ci

v0. (3–89)

86

Page 87: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

me fe

φ’

φ’

in1

in2

10

20

in

fe

fe

Figure 3-13. Schematic of capacitive transduction scheme using differential capacitive sens-ing technique.

Substituting Equations 3–88 and 3–84 into the equation above gives

vin =

(C ′

10

C ′10 + Cp + Ci

)Vb

geff

Cmefe. (3–90)

Equation 3–90 represents the voltage available at the input of the interface circuitry i.e.,

amplifier (ideal amplifier).

So far, a single sense capacitance was used to illustrate the transduction process and an

expression for the amplifier input voltage was obtained. Next, the amplifier input voltage

using both the sense capacitors for the differential capacitance sensing scheme is described.

Figure 3-13 shows the differential transduction scheme for the floating element sensor.

The same principle used in deriving Equation 3–90 is used to obtain the net output

voltage for the differential capacitance scheme. One sense capacitor is considered at a time

while the other acts as parasitic. Subsequently, superposition is used to obtain the total

87

Page 88: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

amplifier input voltage. Using Equation 3–90, the individual amplifier input voltages are

vinC1=

(C ′

10

C ′10 + C ′

10 + Cp + Ci

)Vb

geff

Cmefe (3–91)

and vinC2=

(C ′

20

C ′10 + C ′

10 + Cp + Ci

)Vb

geff

Cmefe. (3–92)

The net voltage available to the interface circuit is given by the sum of Equations 3–91 and

3–92 as

vin =

(C ′

10 + C ′20

C ′10 + C ′

10 + Cp + Ci

)Vb

geff

Cmefe. (3–93)

Substituting for the electromechanical transduction factor of Equation 3–87 and the effective

gap of Equation 3–85 into Equation 3–93, gives

vin =

(C10 + C20

C10 + C10 + (Cp + Ci)(1− κ2)

)Hgap

Vb

g01

Cmefe. (3–94)

The first term in Equation 3–93 includes the effect of electromechanical coupling in the

parasitic attenuation term (see Equation 3–69). Recognizing that the nominal capacitances

are the same from Equation 3–45 the new parasitic attenuation term is

Hc =

(2C0

2C0 + (Cp + Ci)(1− κ2)

). (3–95)

Thus, expression for the voltage at the interface circuit input is

vin = HgapHcVb

g01

Cmefe. (3–96)

A typical electromechanical coupling factor for the sensor structure is about 3.6%. As cou-

pling gets weaker the attenuation Hc ≈ Hc1 in Equation 3–69 i.e., the free electrical capaci-

tance (C ′10) approaches the blocked electrical capacitance (C10). In the next section, an

equivalent circuit is provided using lumped elements from Section 3.2.1 and the transduc-

tion scheme developed in this section. The derived transduced voltage is compared with the

expression in Equation 3–67 to further explain the effect of electromechanical transduction

via the electromechanical coupling factor, κ2 (see Equation 3–86).

88

Page 89: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

fshear

e

me

me

me fe

φ’

φ’

in1

in2

10

20

in

fe

fe

To interface

circuit

Figure 3-14. Schematic of equivalent circuit for the differential capacitive shear stress sensor.

3.2.3 Equivalent Sensor Circuit

The electromechanical transduction model derived in Section 3.2.2 links the mechanical

and the electrical system of the sensor. In this section, the complete equivalent circuit of the

sensor is presented as shown in Figure 3-14. The resonant frequency (bandwidth) and the

static sensitivity of the sensor is estimated from the frequency response of this equivalent

circuit.

As discussed in Section 3.1.3.3, the output voltage is buffered using a high impedance

voltage follower. Ideally, the infinite input impedance of the voltage follower allows to assume

an open circuit condition after the transduction. Furthermore, weak electromechanical cou-

pling (κ2 ¿ 1) strengthens this assumption. Thus, loading of the system in Figure 3-14 due

to the interface circuitry and parasitics, is neglected. The frequency response is therefore de-

termined using the open circuit configuration and is dominated by the mechanical response

of the system. Although, the loading due to parasitics is neglected for the transduction,

subsequent parasitic attenuation (Hc) of the transduced signal is important to determine

the overall sensor sensitivity. It is noteworthy that the open circuit assumption leads to a

89

Page 90: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

stiffer/less sensitive device, resulting in a conservative estimate of the sensor performance.

However, this also slightly overpredicts the bandwidth estimated from the resonance, which

is usually between the open circuit and short circuit resonance frequencies. The mechanical

system in Figure 3-14 is a simple second order system (spring mass damper or LCR) and is

governed by the differential equation,

fshear(t) = md2δ

dt2+ R

dt+ kδ, (3–97)

where fshear represents the shear force, Mme represents the lumped mass, k represents the

effective stiffness of the tethers, and R represents the effective damping in the system. The

stiffness, k is

k =1

Cme

. (3–98)

The differential equation in 3–97 is solved using Fourier transform, resulting in the following

frequency response function:

H(ω) =δ(jω)

fshear(jω)=

Cme

MmeCme(jω)2 + RCme(jω) + 1. (3–99)

The undamped natural frequency (open circuit resonant frequency), of the system is

fr =1

√1

CmeMme

. (3–100)

For modeling purposes, the damping is assumed to be negligibly small. In that case, the

damped natural frequency is approximately equal to the undamped natural frequency of the

system. The validity of this assumption will be verified during experimental characterization

of the sensor.

The accuracy of the resonant frequency prediction is verified via finite element analy-

sis (FEA) in COMSOL using modal analysis. The simulation involves the moving floating

element (1 µm × 1000 µm × 45µm), tethers (1000µm × 15µm × 45µm), and comb fingers

(170µm× 4µm× 45µm) with a clamped boundary condition at the free end of the tethers.

90

Page 91: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

The structure has a total of 102 comb fingers on either side of the floating element. An eigen

frequency analysis is performed with Lagrange quadratic elements, allowing large deforma-

tions. For the first mode or the shear/in-plane mode, the LEM predicts a resonant frequency

of 5.23 kHz compared to 5.18 kHz from FEA, which are in agreement to within 1.2%. The

first six mode shapes for the simulated sensor geometry and the corresponding frequencies

are shown in Figure 3-15.

(a) Mode 1 (fr = 5.177 kHz)

(b) Mode 2 (fr = 14.39 kHz)

(c) Mode 3 (fr = 25.08 kHz)

(d) Mode 4 (fr = 46.91 kHz)

(e) Mode 5 (fr = 140.9 kHz)

(f) Mode 6 (fr = 141.0 kHz)

Figure 3-15. Resonant modes of floating element structure.

Next, the expression for the output voltage and static sensitivity of the sensor is derived.

Consider the effective force, fe used in the the transduction process. This force is same as

the last term in Equation 3–97. This is also evident from a comparison of Equation 3–25,

91

Page 92: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

3–99, and the transfer function or gain factor, G(ω), between fe and fshear,

G(ω) =fe

fshear

=1

MmeCme(jω)2 + RCme(jω) + 1. (3–101)

Therefore,

fe = kδ =δ

Cme

, (3–102)

Thus, the input voltage in Equation 3–96 given to the interface circuitry is obtained by

substituting Equation 3–102 to give

vin = HgapHcδ

g01

Vb. (3–103)

Considering a general case, a voltage amplifier with a gain G, an input vin, and a sinusoidal

bias voltage, Vac sin(ωct) is considered for the interface circuitry. The output voltage of this

amplifier is

vout1 = Gvin = GHgapHc

g01

)Vac sin(ωct). (3–104)

The modulated sensor output was previously presented in Equation 3–70, which did

not account for the electromechanical coupling between electrical and mechanical domains.

Comparison with Equation 3–70 shows that all terms are the same except the term Hc in

Equation 3–104, which accounts for the electromechanical coupling in the system. Similarly,

the effect of electromechanical coupling is incorporated in the demodulated output signal in

Equation 3–76, rewritten with Hc instead of Hc1 as

vout = VacVrefHgapHc

(G

2U

)(1

g01

)(δm + δt sin(ωt + φ)) . (3–105)

A steady laminar flow results only in a mean wall shear stress with no fluctuations (ω = 0).

A steady laminar flow is therefore used to estimate the static sensitivity of the sensor to wall

92

Page 93: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

shear stress. The static sensitivity is defined as,

Sτ, static =Output voltage

(vout

∣∣ω=0

)

Mean wall shear stress (τw). (3–106)

Using the above definition, the expression for the overall static sensitivity, including atten-

uation, is

Sτ, overall = VacVref︸ ︷︷ ︸1

HgapHc︸ ︷︷ ︸2

(G

2U

)

︸ ︷︷ ︸3

(δm/τw

g01

)

︸ ︷︷ ︸4

. (3–107)

The static sensitivity of the sensor alone, neglecting the attenuation terms, is

Sτ, sensor = Vac

(δm/τw

g01

). (3–108)

It is instructive to study the effect of each term in Equation 3–107 on the sensitivity. Pro-

ceeding term by term, term 1 indicates that the static sensitivity is directly proportional to

the amplitude product of the bias voltage (Vac) and the reference voltage (Vref ) supplied to

the multiplier. The upper limit of Vac is determined from the pull-in limit or the dielectric

breakdown limit of air, with a built in factor of safety. Usually the pull-in constraint domi-

nates (see Chapter 4). Vref is limited by the maximum input voltage that can be applied to

the multiplier. While the sensitivity may be improved by increasing Vac and Vref , the source

noise contribution may also increase and should be accounted for while choosing the values

of Vac and Vref .

Term 2 captures two attenuating effects on the sensitivity: the effect of the secondary

gap g02 via Hgap and the sensitivity scaling in relation to the nominal sensor capacitance, C0,

and parasitic capacitances, Cp + Ci via Hc. Equation 3–68 indicates that Hgap < 1, lowering

the sensitivity. Hgap increases as g02 increases in comparison to g01. It is evident that the

condition g01 ¿ g02 needs to be satisfied to minimize attenuation from this term. Higher g02,

however, lowers the number of comb fingers and thus the total capacitance, which is needed

to overcome parasitics. Thus, there is a trade-off between the number of comb fingers and

g02.

93

Page 94: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

The condition Hc ¿ 1 or ((2C0)/(2C0 + Cp + Ci)) ¿ 1 or 2C0 À Cp + Ci is desirable

to make the sensitivity independent of the adverse effects of parasitics. However, it becomes

increasingly difficult to satisfy this condition as the sensor size is scaled down. Guarding,

a shielding technique in which the terminals of the parasitic capacitance are tied to the

common-mode potential of the sensor system, is often used to minimize the attenuation due

to parasitics [42, 68]. However, the finite input capacitance of the amplifier still remains,

which cannot be guarded. Using an amplifier with very low or negligible input capacitance

may therefore improve sensitivity. Term 2 is always less than unity if the conditions g01 ¿ g02

and ((2C0)/(2C0 + Cp + Ci)) ¿ 1 are not satisfied. A good estimate of parasitic capacitances

and an appropriate choice of g02 is essential during the design process to ensure this term is

as close to unity as possible i.e., minimum attenuation.

Term 3 in Equation 3–107 is the ratio of the closed loop gain, G, of the amplifier to

twice the scaling factor, U , of the multiplier. Thus, to increase sensitivity, a high gain

amplifying stage and a multiplier with a low scaling factor are favorable. Furthermore, a

high gain amplifying stage helps to boost the signal to noise ratio (SNR), minimizing the

effect of noise added in the subsequent circuitry. The value of gain G is again limited by the

maximum allowable voltage input to the multiplier.

Term 4 in Equation 3–107 represents the contribution from the actual sensitivity of the

sensor via the transduction itself (see Equation 3–108). The numerator (δm/τw) represents

the mechanical sensitivity of the sensor. The ratio δm/g01 represents a simplified form of

the change in capacitance to the original capacitance or the change in gap to the original

gap. The combined ratio thus represents the simplest form of sensitivity, which is percentage

change in capacitance per unit shear stress. Ideally, this value should be as large as possible

for the sensor design; however, bandwidth and linearity requirements constrain the upper

limit of this ratio.

In summary, the dynamic modeling section discussed lumped element modeling and

this concept was used to compute lumped elements from the distributed mechanical system.

94

Page 95: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Electromechanical transduction for the sensor system was mathematically illustrated. An

equivalent circuit representation of the sensor was provided. Expressions for the frequency

response and resonant frequency were given. Expressions for the demodulated sensor output

voltage and static sensitivity were given. Lastly, the relevance and contribution of different

terms in the sensitivity expression were explained.

3.3 Higher Order Effects and Noise

In this section, some of the foreseeable higher order effects like fringing fields and par-

asitic capacitances are discussed. The fringing fields are edge effects that add additional

parasitic capacitance, increasing attenuation. A reasonable model for fringing fields aids in

better design and prediction of sensor performance. A model to estimate the noise from the

interface circuitry is also presented. This noise model is used to estimate the minimum de-

tectable signal of the sensor, which is used as the objective function for design optimization

in Chapter 4.

3.3.1 Fringing Fields

The capacitance equations discussed up to this point assume an ideal parallel plate

capacitor. In an ideal parallel plate capacitor, the electric field lines are normal to the parallel

planes of the capacitor. However, due to the finite dimensions of the plate, some curved

electric field lines exist at their edges. These electric fields are referred to as fringing electric

fields. These fringing electric fields manifest themselves as a fringing field capacitance,

Cfr. Any capacitance which is unaccounted for when using the parallel plate assumption

is treated as fringing field capacitance. It is noteworthy that any additional capacitance

that does not change with the physical input to the sensor is parasitic in nature. Thus, the

fringing field capacitance also adds to the parasitic capacitance of the system. The fringing

field capacitance is assumed to be invariant with the change in nominal capacitance for small

input signals i.e., small change in gaps. This is a reasonable assumption because a small

change in gap does not change the fringing field capacitances. This is also verified via FEA

simulations, which are performed for both an initial gap size and that of a typical full scale

95

Page 96: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

deflection of the sensor. Figure 3-16 shows the electric field lines in the gaps g01 and g02 for

a single comb finger simulated in 3D, using FEA in COMSOL.

Lf

g02

Wf

g01

g02

Insulating

Boundary

1 V

0 V (ground)

Figure 3-16. Asymmetric comb finger structure to estimate effect of fringing electric fields.

The comb fingers are simulated because their contribution dominates the sensor nominal

capacitance in most designs (Table 4.4) and due to their asymmetric gaps. Comb fingers

made of perfect electrical conductors are assumed for this simulation and the dielectric

medium is air (εr = 1). The various geometric dimensions for this analysis are, g01 = 3.5 µm,

g02 = 20 µm, Wf = 4 µm, and Lf = 170 µm. The color gradient of the field lines indicates the

variation of electric potential in the gaps. Typically for the comb finger, the sensor fringing

field capacitance is ≈ 10% of the nominal capacitance i.e.,Cf r

C0≈ 0.1. The ratio of change in

fringing field capacitance to the total change in capacitance is ≈ 13% i.e.,∆Cf r

∆C0≈ 0.13 and

hence neglected. Such an approximation only leads to a conservative estimate of the sensor

performance while maintaining modeling simplicity.

96

Page 97: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

3.3.2 Parasitic Capacitance

As stated earlier in Section 3.1.3.3, parasitic capacitance has a detrimental effect on

the sensitivity of the sensor. An accurate estimate of the parasitic capacitances is thus

essential to accurately predict device performance characteristics. Theoretical models are

not presently available to obtain such an estimate.

Some sources of parasitics in a typical MEMS sensor are polysilicon/metal lines, wire

bonds, PCB traces and BNC cables. The contribution of these sources may vary from

sensor to sensor, based on length of metal lines, quality of wire bonds etc. Martin [72]

compares sensitivities of a dual backplate microphone, using both a charge amplifier and

a voltage amplifier. The sensitivity measured using a charge amplifier is independent of

parasitic capacitances, but is influenced by parasitics when measured with a voltage amplifier.

Thus, a comparison of the sensitivities measured using the two biasing schemes helps in

isolating the associated parasitic capacitance of the sensor. Similar to Martin’s work, the

sensor developed in this dissertation also uses the SiSonicTM microphone amplifier , courtesy

Knowles Acoustics [76]. It uses the same differential capacitance measurement scheme as

used in Martin’s sensor. Martin’s work shows experimental value of parasitic capacitances

vary from 0.92 pF to 2.23 pF from tests conducted on seven different microphones. The

proposed sensor has similar interface circuitry/sensor packaging and the same amplifier as

Martin’s microphone. The higher value of parasitic capacitance, 2.23 pF , should therefore

provide a conservative estimate of parasitics for design purposes.

3.3.3 Noise Model

In this section the noise model of the sensor is described. The ratio of the noise floor

[V ] to the sensitivity [V/Pa] of the sensor determines the minimum detectable shear stress

(MDS). The MDS is also the lower end of the dynamic range of the sensor and is discussed

in greater detail in Chapter 4. For a given bandwidth, the MDS is mathematically defined

97

Page 98: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

as

MDS =Noise[V ]

Sensitivity[V/Pa]. (3–109)

As a first order estimate, the noise from the mechanical and fluidic system is considered

negligible compared to the noise contribution from the interface circuitry. Thus, a noise

model for the interface circuitry is developed to obtain the noise floor for the sensor. A

modulation-demodulation scheme is used for the sensor interface circuitry as previously

shown in Figure 3-10. A voltage amplifier amplifies the modulated signal, the output of which

is supplied to the input of the multiplier (demodulator). If an amplifier with a sufficient closed

loop gain is used, the noise contribution from the later stages may be neglected provided

they have a comparable or lower noise floor. Therefore, in the noise model the input referred

noise of the amplifier is analyzed without the contribution of the subsequent stages.

-

Rb

Cp Ci

Svb=

4kTRb

Sva

SiaVout

C1

C2

Figure 3-17. Simplified schematic of sensor circuitry with noise sources.

In Figure 3-17 three noise sources are considered: thermal noise of the bias resistor Rb,

the input referred voltage noise Sva, and the current noise Sia of the amplifier. The noise

sources are assumed to be uncorrelated, allowing superposition to compute the total noise

at the output. The noise sources are presented in terms of their spectral density and must

be integrated over the noise bandwidth to obtain the total rms noise voltage at the amplifier

output. A single noise source is considered at a time and the corresponding output noise is

98

Page 99: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

determined. The total impedance at the amplifier input is

Zi =1

jω(C1 + C2 + Cp + Ci)=

1

jωCT

, (3–110)

where CT = C1 + C2 + Cp + Ci. (3–111)

First, consider the noise due to the source Sva. No current flows into the impedance Zi and

all the voltage is amplified and appears at the output. Thus,

S1 = G2Sva. (3–112)

Now, consider output noise due to the current source Sia. The impedance for this current

source is (Rb||Zi). The noise contribution to the output is therefore,

S2 = G2Sia

(RbZi

Rb + Zi

)2

. (3–113)

Finally, the noise contribution due to the thermal noise source Svb is considered. The noise

contribution from the resistor is

S3 = G2Svb

(Zi

Rb + Zi

)2

. (3–114)

The total input referred noise in the circuit is obtained by summing the output noise con-

tribution from each individual source and dividing it by the square of the closed loop gain.

The total input noise is given as,

Snoise,in =S1 + S2 + S3

G2,

and =1

G2

[Sva + Sia

(RbZi

Rb + Zi

)2

+ Svb

(Zi

Rb + Zi

)2]

. (3–115)

Combining Equation 3–107 and Equation 3–115, the MDS for the sensor is

MDS =Snoise,in

Sτ,overall

. (3–116)

99

Page 100: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

The thermal noise of the resistor is usually much smaller than the input referred noise

of the amplifier. The amplifier noise usually dominates the noise floor. Equation 3–115

and Equations 3–111 show that the contribution of the current noise increases as the total

capacitance, CT decreases due to small device capacitance. Hence it is desirable to have

minimum contribution from this term for microscale electrostatic sensors. Moreover, no

theoretical/empirical models exist for the accurate estimation of parasitics that contribute

to CT . Thus to have a predictable noise estimate, it is desirable to use an amplifier designed

for a very low current noise.

100

Page 101: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

CHAPTER 4DESIGN OPTIMIZATION

Optimization is a widely used design tool both in industry and academia. It is a method-

ical approach to making design choices that lead to the best possible outcome. Optimization

helps to explore and exploit the performance of a given sensor structure as a function of

geometry, materials, etc. It also gives physical insight into the design problem itself and

helps understanding of design tradeoffs.

This chapter explains the methodology used for the optimized design of the capaci-

tive shear stress sensor described in the previous chapter. An appropriate cost function

representing sensor performance is identified and optimized. Constraint functions are also

formulated based on performance requirements of the sensor and limitations of the fabrica-

tion process. The expressions and relations developed in Chapter 3 are used to formulate

mathematical functions and constraints of the resulting nonlinear constrained optimization

problem. Sequential quadratic programming (SQP), which is one of the several available

optimization techniques, is used for optimization. The constrained optimization problem

is implemented using the optimization toolbox in MATLAB. Specifically, the SQP func-

tion fmincon in MATLAB uses a gradient descent based method to optimize nonlinearly

constrained problems.

Section 4.1 gives a brief overview of the optimization technique used, its benefits, and

its shortcomings. Section 4.2 explains the design requirements, identifies the design vari-

ables, objective function, and the constraints. Section 4.4 provides optimization results and

sensitivity analysis.

4.1 Sequential Quadratic Programming-SQP

Sequential quadratic programming uses a direction seeking algorithm to arrive at the

minimum/maximum of a given function. SQP implemented in the current optimization

problem uses a function known as the Lagrangian formed using both the objective and

101

Page 102: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

constraint functions [77]. The Lagrangian is represented as,

L(x, λ) = f −n∑

i=1

λiGi. (4–1)

where f is the objective function, Gi represents the ith inequality constraint, gi, transformed

into an equality constraint, and λi is a parameter called the Lagrangian multiplier. The

Lagrangian in Equation 4–1 accounts for the constraints in a given optimization problem

and λi represents the relation between the objective and constraint functions when the

constraints are active. For instance, if no constraints are present or are inactive (λi = 0),

the Lagrangian simply reduces to the objective function. Optimality is ensured by solving

for λi such that the Kuhn-Tucker conditions are satisfied [77]. The Kuhn-Tucker conditions

are necessary conditions for optimality and are stated as

∇L (x, t, λ) = ∇f −n∑

i=1

λi · ∇Gi = 0,

λi ·Gi = 0 for i = 1, · · · , n

λi ≥ 0 for i = 1, · · · , n.

(4–2)

The first condition in Equation 4–2 shows that the optimization algorithm relies on the

reducing gradients of both the objective and constraint functions and solves for λi such

that ∇L = 0. Consequently, based on the initial values of design variables, the solver may

arrive at a local minimum instead of the global minimum of the objective in the design

space. Repeating the optimization for several different initial conditions provides greater

possibility of arriving at the global minimum [77]. The second and third conditions are for

the constraints, which indicate if the constraints are satisfied while being active or inactive

for optimality i.e., Gi = 0 and λi 6= 0 or Gi < 0 and λi = 0.

The benefits of this optimization technique include faster convergence, easy implementa-

tion, and high accuracy. Drawbacks of SQP are that it requires smooth analytical functions

to compute gradients and is not a global optimization algorithm.

102

Page 103: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

4.2 Optimization

4.2.1 Sensor Performance Requirements

Sensors in general need to satisfy several target specifications like large dynamic range

and bandwidth, in addition to high sensitivity and low noise floor. These terms and their

desirable values are briefly explained in this section. The operating space for a sensor may be

represented by a plot as shown in Figure 4-1. The operating space is bound by the limits of

the sensor performance. Ideally, a good design is aimed at maximizing the sensor operating

space. The extents of the operating space are bound by the dynamic range and bandwidth

of the sensor.

Frequency (Hz)

Shear Stress (Pa)

Bandwidth (Hz)

Dynamic Range (Pa)

Non-linearity

(Mechanical

structure, Pull-in)

±3dB point

(flat band

response)

Minimum detectable signal

(Sensor structure, electronics)

Figure 4-1. Schematic of operating space of the sensor.

The dynamic range of a sensor is defined as the maximum range of the physical input

that the sensor can reliably measure. The upper end of the dynamic range is determined

103

Page 104: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

by the maximum input signal (shear stress) the sensor can measure without significant non-

linearity/harmonic distortion. The lower end of the dynamic range is limited by the noise

floor/minimum detectable signal (MDS) of the sensor. For an input within its dynamic

range, the sensor bandwidth is defined by the range of frequencies over which the sensor has

a constant response (sensitivity).

In an optimization problem, the lower and upper end of the bandwidth are set via

constraints based on the frequency range of interest. The lower end of the bandwidth is at

dc since it is required to measure mean shear stress. The cut-on frequency of the interface

circuit is considered negligible in this case, allowing dc measurements. The upper end of the

dynamic range is constrained by the maximum permissible non-linearity in the output that

is ideally set by the target physical input to be measured. Thus, with the three sides bound

via constraints, the operating space is maximized by minimizing the MDS (fourth side) of

the sensor. This is consistent with Spencer’s study, which indicates that MDS is useful as

true figure of merit for studying sensor performance [62]. Minimizing MDS in Equation 3–

109 thus translates into maximizing the overall sensitivity while minimizing the noise floor

of the sensor.

4.3 Problem Formulation

4.3.1 Objective and Design Variables

The goal of this optimization is to minimize the minimum detectable signal (shear stress)

(MDS) for the differential capacitive shear stress sensor. In the expression for the output

given by Equation 3–105, the termVref

2U, represents a contribution from the demodulation

circuitry. In the current circuit implementation, this term is not variable since off-the-shelf

components are used for demodulation. A unity gain voltage follower is used as the amplifier

i.e., G = 1. Thus, the simplified expression for the overall sensitivity is

Sτ,overall = VacHgapHc

(δm/τw

g01

). (4–3)

104

Page 105: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

As discussed in Section 3.3.3, the mechanical thermal damping in the system is assumed

to be negligible. As a result, the electronics dominate the overall noise floor. The SiSonic

amplifier used in this implementation has a measured input referred voltage noise power

spectral density of 4×10−16 V 2

Hzfor thermal noise with a corner frequency, fce = 395 Hz [72].

The current noise power spectral density of this amplifier was negligibly small, 5×10−31 ( A2

Hz)

and it has an internal bias resistor. If an amplifier with a higher current noise is used, a change

in the nominal sensor capacitance alters the input impedance, Zi, affecting the overall noise

floor (Equation 3–115). Also, it may need an external bias resistor, whose noise contribution

varies with the sensor capacitance. With the present noise specifications for the amplifier,

the noise floor reduces to its voltage noise and is represented using a curve fit as

So = 4× 10−16 (1 + 395/f) . (4–4)

The optimization is performed for a overall static sensitivity, Sτ,overall, at 1 kHz, with

1 Hz bin-width. Since the amplifier has a fixed noise floor at 1 kHz, the objective function

only maximizes Sτ,overall, assuming fixed parasitic capacitance. However, to maintain gen-

erality, the MDS is retained as the objective function. The objective function is therefore

given as

F (−→X ) = MDS (Pa) =

√S0

∣∣f=1 kHz

Soverall

or F (−→X ) =

√S0

∣∣f=1 kHz

VacHgapHc

(δm/τw

g01

) , (4–5)

where−→X is the design variable vector as shown in Equation 4–6. Note, the overall sensitivity

includes the non-idealities and attenuation factors (Hc and Hgap) due to parasitics and

asymmetric gaps. The LEM developed in Chapter 3 is used to estimate the frequency

response of the coupled electromechanical system. The floating element, tethers, and comb

fingers have the same thickness, Tt, which is decided by the material layer thickness during

105

Page 106: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

fabrication. A constant charge biasing scheme is used in conjunction with a voltage buffer.

The voltage change corresponding to a change in capacitance is sensed using the voltage

buffer. Assuming silicon as the sensor material with known properties, the variables of

interest for this optimization problem are as follows:

1. Bias voltage amplitude Vac

2. Tether length Lt

3. Tether width Wt

4. Element length Le

5. Element width We

6. Tether thickness Tt

7. Comb finger overlap L0

8. Comb finger width Wf

9. Primary gap g01

10. Secondary gap g02

This results in 10 design variables represented as a vector,

−→X = (Vac, Lt,Wt, Le,We, Tt, L0,Wf , g01, g02) . (4–6)

The comb finger length, Lf , is longer than the overlap, L0, to accommodate the gap at the

end of each comb finger, given as

Lf = L0 + 20 µm. (4–7)

The number of comb fingers is computed using the the design variables as

N =2(We − 2Wt)

(2Wf + g01 + g02). (4–8)

Using this expression, the maximum number of comb fingers that can be accommodated

along the width of the floating element are obtained. The larger the number of comb fingers,

the higher the nominal capacitance, which is essential to overcome the parasitic attenuation,

106

Page 107: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Hc. This may marginally increase the mass of the sensor and correspondingly lower the

resonant frequency. However, the optimizer has more than one variable to vary to offset this

increase in mass.

4.3.2 Constraints

Constraints for the optimization are set based on performance criteria, design specifi-

cations, and fabrication limitations. The bounds on the design variables are set based on

application and fabrication limitations. As stated in Chapter 1 the goal is to design a sensor

to measure τmax ≤ 10 Pa. The characteristic viscous length scale, l+, for this shear stress

is approximately 5.3 µm. The spatial resolution is set by the maximum sensor size which

is restricted to ≤ 400l+ via upper bounds (UB) for a conservative design to demonstrate

proof of concept. Three different scales are considered 100l+, 200l+, and 400l+, respectively.

The lower bounds (LB) are set by the sensing area needed to have a reasonable signal to

noise ratio sans attenuation. Nominal capacitance, fabrication and hydraulic smoothness

requirements are used to determine the bounds on the gaps around the sensor. The largest

gap is roughly 3.8l+, thus ensuring that hydraulic smoothness is maintained. The upper

(UB) and lower (LB) bounds on the design variables are (Table 4-1)

LB ≤ (Vac, Lt, Wt, Le,We, Tt, L0,Wf , g01, g02) ≤ UB. (4–9)

4.3.2.1 Bandwidth

This constraint ensures that the sensor possesses the bandwidth specified by the design

requirements. The bandwidth constraint is given as

fhigh − dc ≥ fmin, (4–10)

fmin is the minimum acceptable bandwidth set by the design specification. A ±3 dB fre-

quency is chosen which is defined as the ±3 dB change in the value of the frequency response

function from its flat band value (sensitivity). The lower end of the bandwidth is at dc since

the sensor needs to measure the mean flow. There is no mechanical cut-on frequency for the

107

Page 108: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

sensor (second order system) and the electrical cut-on frequency is assumed to be negligible.

fhigh is ±3 dB frequency at high frequencies, set via the damping in the system. As stated

in Section 3.2, the mechanical thermal damping in the system is neglected. Therefore, fhigh

is defined as the undamped natural frequency, fres of the system for estimation purposes.

Thus, the bandwidth constraint may be simply written as

fmin ≤ fres. (4–11)

4.3.2.2 Mechanical Nonlinearity

Any non-linearity in the system results in harmonics at the output. To maintain spec-

tral fidelity of time resolved measurements, the nonlinearities in the system have to be con-

strained. The linearity of the estimated MDP under maximum incident shear stress τmax, is

ensured using the nonlinearity constraints mathematically stated as follows,

δNL − δ

δ≤ 3%, (4–12)

where δNL is the nonlinear deflection computed using Equation 3–8 and δ is the linear

deflection obtained from Equation 3–7.

4.3.2.3 Electrostatic Pull-in

Ideally the capacitances that form the differential scheme remain at their mean position

under increasing bias voltages due to equal and opposite electric fields balancing each other.

However, when the pull-in voltage is reached, even a small perturbation causes the diaphragm

to jump to one of the stationary electrodes. Pull-in occurs when the mechanical force is no

longer sufficient to counter the electrostatic force causing the movable electrode to collapse to

the stationary electrode in an electrostatic transducer. A discussion on the pull-in instability

can be found in [42, 66]. For a differential capacitance scheme, the pull-in voltage is given

as [78]

VPI =

√g30

2CmeεA. (4–13)

108

Page 109: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Assuming g02 >> g01 i.e., pull-in in dominated by the primary gap, and using the small-

est possible electrode separation at maximum forcing, the conservative estimate for pull-in

voltage for the proposed sensor is written as

VPI =

√(g01 − δ)3

2Cmeε(3Lteff + Le + (N − 1)L0). (4–14)

The amplitude of the bias voltage amplitude Vac is constrained to 85% of the pull-in voltage

given as

Vac ≤ 0.85VPI . (4–15)

There is an additional nonlinearity due to nonlinear output voltages at large deflections

and is illustrated here mathematically. Assuming g02 À g01, the nonlinear output voltage,

VNL, for a differential capacitive sensing scheme with matched capacitors is

VNL = Vacδ

g01

C0

C0 + (Cp+Ci)

2

(1− δ2

g201

) . (4–16)

Typically, the full scale deflection δ ≈ 0.1 µm and g01 ≈ 3 µm. Thus, 1 − δ2/g201 ≈ 0.999,

indicating that a nonlinear output voltage due to large defections, i.e., δ ∼ g01 is unlikely. A

constraint to limit VNL is therefore not formulated to simplify the optimization problem.

Mathematically, the problem is to minimize the MDS subject to constraints or

Minimize

F (−→X ) = MDS,

subject to

g1 = fmin/fres − 1 ≤ 0,

g2 =δL − δNL

0.03δL

− 1 ≤ 0,

g3 =Vac

0.85VPI

− 1 ≤ 0,

109

Page 110: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

and gi which are to be determined from

LB ≤ (Vac, Lt, Wt, Le,We, Tt, L0,Wf , g01, g02) ≤ UB.

4.3.3 Design Specifications

This section provides the target design specifications and the constants used in the

optimization. The maximum input shear stress and the required bandwidth of the sen-

sor determine the target designs. Upper and lower bounds are based on spatial resolution

requirements and micromachining limitations. The target design specifications and the vari-

able bounds are shown in Table 4-1. The constants and material properties used for the

optimization are shown in Table 4-2.

Table 4-1. Lower and upper bounds for design variables and flow specifications.

Vac Lt Wt Le We Tt L0 Wf g01 g02 fmin τmax

(V ) (µm) (µm) (µm) (µm) (µm) (µm) (µm) (µm) (µm) (kHz) (Pa)

5−25 100−1000

10−40

100−2000

100−2000

45−50

5−150

4−20 3.5−15

3.5−20

10 5&10

Table 4-2. Constant parameters for optimization.

Parameter Value

Electrical Permittivity of Free Space ε0 8.854× 10−12 F/m

Relative Permittivity of Air εr 1.005

Floating Element Young’s Modulus E † 168 GPa

Density of Floating Element ρ 2330 kg/m3

Amplifier Input Capacitance Ci 0.3 pF

Parasitic Capacitance Cp 2.2 pF

† Average value in the 100 plane of silicon.

4.4 Results and Discussion

Optimization results for five different designs are presented in this section. Two target

shear stresses are used, τmax = 5 Pa and τmax = 10 Pa. The bandwidth target is fmin =

5 kHz and 10 kHz. The upper bounds on the floating element length and width are also

110

Page 111: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

varied to study the effect of different sensing areas. The results of the optimization are

summarized in Table 4-3.

The bias voltage, Vac is restricted either by its upper bound or by the active pull-in

constraint for all the designs. The tether length, Lt always hits its UB tending to maximize

the sensitivity by increasing the tether compliance. However, the tether width is not always

at its LB because of the bandwidth and the active pull-in constraint. Thus, the bandwidth

is maximized and the pull-in instability is avoided, but, at the cost of sensitivity. This is

better understood from the relations

Sensitivity ∼ Cme, (4–17)

, (4–18)

and VPI ∼ 1√Cme

. (4–19)

As stated already, the active pull-in constraint also restricts Vac which in turn lowers the

sensitivity (refer Equation 3–107). Thus, the detrimental effect of the pull-in voltage is

two-fold, it limits Vac and Cme, lowering the sensitivity.

The optimizer always maximizes the sensor sensing area via Le and We to increase the

available shear force, improving sensitivity. The tether thickness is always pushed to its

lower bound. Equation 3–80 indicates,

Cme ∝ 1

Tt

, (4–20)

which means a lower tether thickness increases Cme, improving the sensitivity. Ideally, a thin

floating element structure would greatly improve sensitivity but it would lower the resonance

frequency and hence the sensor bandwidth. A thinner element also reduces the overlapping

area (L0Tt) of the capacitors, lowering nominal capacitance. A high nominal capacitance is

essential to minimize attenuation due to parasitics. Also, smaller Tt makes the mechanical

structure more sensitive to undesirable pressure fluctuations.

111

Page 112: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

The initial finger overlap, L0 and the primary gap, g01 are always at their UB and LB

respectively. Both are essential to achieve a high nominal capacitance to minimize parasitic

attenuation. The secondary comb separation g02 is also maximized because it has a change

opposite to g01. As a result of g02, the overall capacitance change,∆C is lower than the

achievable value with only g01. A large value of g02 is therefore desirable to maximize ∆C

and hence the sensitivity. This is also explained in Chapter 3 via the gap attenuation factor,

Hgap in Equation 3–68. Large g02 also decreases the number of comb fingers N , lowering

nominal capacitance. Thus, a tradeoff exists between nominal capacitance needed to mitigate

parasitic attenuation, and larger g02 required to lower Hgap.

In some cases, the comb finger width, Wf is also maximized. Increased width con-

tributes to increase in sensing area and decrease in number of comb fingers N . Therefore,

thicker fingers result in lower nominal capacitance. Similar to the discussion in the previous

paragraph, there is tradeoff between improved sensitivity and lower C0 i.e., higher parasitic

attenuation, resulting from wider fingers.

Table 4-3. Capacitive shear stress optimization results for Cp + Ci = 2.2 + 0.3 pF .

DesignParameter

Design 1 Design 2 Design 3 Design 4 Design 5

τmax = 5 Pa τmax = 5 Pa τmax = 5 Pa τmax = 5 Pa τmax = 10 Pa

fmin = 5 kHz fmin = 5 kHz fmin = 5 kHz fmin = 10 kHz fmin = 5 kHz

UB Le = 2000 µm Le = 1000 µm Le = 500 µm Le = 1000 µm Le = 1000 µm

UB We = 2000 µm We = 1000 µm We = 500 µm We = 1000 µm We = 1000 µm

Bias VoltageVac (V )

25.00 22.78 16.46 25.00 22.21

TetherLengthLt (µm)

1000 1000 1000 1000 1000

Tether WidthWt (µm)

23 15 10 23 15

ElementLengthLe (µm)

2000 1000 500 1000 1000

ElementWidthWe (µm)

2000 1000 500 1000 1000

112

Page 113: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Table 4-3. Continued

DesignParameter

Design 1 Design 2 Design 3 Design 4 Design 5

τmax = 5 Pa τmax = 5 Pa τmax = 5 Pa τmax = 5 Pa τmax = 10 Pa

fmin = 5 kHz fmin = 5 kHz fmin = 5 kHz fmin = 10 kHz fmin = 5 kHz

UB Le = 2000 µm Le = 1000 µm Le = 500 µm Le = 1000 µm Le = 1000 µm

UB We = 2000 µm We = 1000 µm We = 500 µm We = 1000 µm We = 1000 µm

Tether/ElementThicknessTt (µm)

45 45 45 45 45

Comb FingerOverlapL0 (µm)

150 150 150 150 150

Comb FingerLengthLf (µm)

170 170 170 170 170

Comb FingerWidthWf (µm)

5.1 20 20 4.0 20

PrimaryCombSeparationg01 (µm)

3.5 3.5 3.5 3.5 3.5

SecondaryCombSeparationg02 (µm)

20 20 20 15.4 20

Number ofComb FingersN

116 31 15 71 31

ResonantFrequency/Bandwidthf (kHz)

5 5 5.2 10 5

Design ShearStressτmax (Pa)

5 5 5 5 10

SensitivitySoverall

(dB re 1 V/Pa)

−28.98 −34.33 −40.57 −43.05 −34.50

SensitivitySoverall

(mV/Pa)

35.60 19.20 9.360 7.040 18.80

MDS at1 kHz (µPa)

0.562 1.040 2.140 2.840 1.060

Attenuationdue to Cp +Ci, Hc (dB)

−5.10 −9.86 −12.65 −6.90 −9.86

113

Page 114: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Table 4-3. Continued

DesignParameter

Design 1 Design 2 Design 3 Design 4 Design 5

τmax = 5 Pa τmax = 5 Pa τmax = 5 Pa τmax = 5 Pa τmax = 10 Pa

fmin = 5 kHz fmin = 5 kHz fmin = 5 kHz fmin = 10 kHz fmin = 5 kHz

UB Le = 2000 µm Le = 1000 µm Le = 500 µm Le = 1000 µm Le = 1000 µm

UB We = 2000 µm We = 1000 µm We = 500 µm We = 1000 µm We = 1000 µm

TetherCapacitanceCtet (pF )

0.18 0.18 0.18 0.18 0.18

FloatingElementCapacitanceCE (pF )

0.23 0.11 0.06 0.11 0.11

Comb FingerCapacitanceCf (pF )

1.16 0.30 0.14 0.73 0.30

CapacitanceC0 (pF )

1.56 0.59 0.38 1.03 0.59

Full ScaleDeflectionδmax (nm)

60 60 50 10 120

CapacitanceChange∆C (fF )

22.6 8.73 5.22 3.55 17.5

PercentageChange(∆C/C0) %

1.45 1.48 1.37 0.35 2.95

Pull inVoltageVPI (V )

29.5 26.9 19.4 39.1 24.2

4.4.1 Sensitivity Analysis

The sensitivity of the optimized MDS to variation of the design variables is studied in

this section. The sensitivity analysis of Design 1 in Table 4.4 is presented. The MDS and

the variables are both normalized by their respective optimum values. The variables are

perturbed 20% on either side of their optimum value. Figure 4-2 shows the sensitivity of

normalized MDS to the ten normalized design variables. However, some of these solutions

may be invalid because the constraints may be violated and are studied separately.

The results indicate the MDS is highly sensitive to variation in tether dimensions, Lt

and Wt. The design has minimum sensitivity to the secondary gap, g02 and the comb finger

114

Page 115: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

0.8 0.9 1 1.1 1.20.5

1

1.5

2

X/Xopt

f obj(X

)/f ob

j(Xop

t)

Vac

Lt

Wt

Le

We

Tt

L0

Wf

g01

g02

Figure 4-2. Sensitivity of optimized MDS to design variables for Design 1.

width Wf . The MDS variation with each individual variables and constraints is also studied.

Only one design variable is perturbed at a time to understand the effect on the MDS. The

region of the plot that is not shaded or hatched indicates the feasible region. The shaded

portion indicates the region where the constraint is violated whereas the hatched portion

indicates the infeasible region set by the bounds. Figure 4-3 shows the results for each design

variable.

Consider the result in Figure 4-3(b) for Lt, for which the optimal MDS has the highest

sensitivity. The optimal MDS is limited by the upper bound of Lt. Any further increase

would have lowered the MDS but again restricted simultaneously by the pull-in and resonant

frequency constraints. The presented analysis provides insight into the expected variations

in the objective and constraint functions resulting from change in the optimized variables.

However, this study is limited to change in one variable at a time and as stated earlier may

result in violation of constraints for other variables.

The expected change in sensor performance (Design 1) due to simultaneous variations

in the design variables and parameters of interest is also studied statistically, using a Monte

115

Page 116: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

(a) Vac

(b) Lt

(c) Wt

(d) Le

Figure 4-3. Sensitivity of optimized MDS and constraints due to change in optimized variablevalues of Design 1.

116

Page 117: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

(e) We

(f) Tt

(g) L0

(h) Wf

Figure 4-3. Sensitivity of optimized MDS and constraints due to change in optimized variablevalues of Design 1 (cont· · · ).

117

Page 118: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

(i) g01 (j) g02

Figure 4-3. Sensitivity of optimized MDS and constraints due to change in optimized variablevalues of Design 1 (cont· · · ).

Carlo (MC) simulation. In this simulation, a Gaussian distribution is assumed for each

parameter of interest. The approximate tolerances and uncertainties, that may be expected

in the variables and constants, are used to build the Gaussian profile. Table 4-4 shows the

tolerance chart for the various parameters used in this study.

The performance parameters of interest for this analysis are MDS, overall sensitivity

(Sτ,overall), and bandwidth (fmin). The simulation uses 50000 samples for each input/design

parameter of interest. The variations in MDS, Sτ,overall, and fmin are represented in the form

of histograms after outlier rejection in Figure 4-4. The distribution is slightly skewed to the

nonlinearity in the system and due to correlation between the variables. For instance, when

gap g01 increases g02 proportionally decreases and vice versa. However, for this analysis

all variables were assumed to be uncorrelated. For small perturbations these effects are not

pronounced, resulting in a gaussian distribution. The vertical red line in the figures indicates

the optimized value from Table 4.4. The mean and standard deviation for each of the design

parameters are summarized in Table 4-5. The table shows that despite finite tolerances

118

Page 119: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

and uncertainties, the predicted performance of the sensor is likely to be within the 95%

confidence intervals for MDS, Sτ,overall, and fmin.

Table 4-4. Tolerance chart for variables and constants.

Parameter Tolerances (σ)

Bias Voltage Vac ±0.2 V

Tether Length Lt ±5 µm

Tether Width Wt ±0.5 µm

Floating Element Length Le ±10 µm

Floating Element Width We ±10 µm

Tether Thickness Tt ±1 µm

Initial Finger Overlap L0 ±5%

Comb Finger Width Wf ±5%

Primary Comb Separation g01 ±15%

Secondary Comb Separation g02 ±1 µm

Dielectric Permittivity ε ±10%

Density of Floating Element (Si) ρ † ±10%

Floating Element Young’s Modulus E ±10%

Amplifier Input Capacitance Ci ±10%

Parasitic Capacitance Cp ±10%† Average value in the 100 plane of silicon.

Table 4-5. Design sensitivity based on Monte Carlo (MC) sim-ulation.

DesignParameter

OptimizedValue

Mean Value(µMC)

95% ConfidenceInterval (2σ)

MDS (µPa) 0.562 0.528 0.292− 0.765

Sτ,overall

(mV/Pa)35.60 34.92 20.70− 49.13

fmin (kHz) 5.00 4.96 4.28− 5.65

4.5 Conclusion

This chapter provided details of the optimization for the proposed differential capacitive

shear stress sensor. The optimization principle was briefly discussed followed by the problem

119

Page 120: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

0.2 0.3 0.4 0.5 0.6 0.70

200

400

600

800

1000

MDS (µPa)

(a) MDS

15 20 25 30 35 40 450

200

400

600

800

Sτ,overall (mV/Pa)

(b) Sτ,overall

4 4.5 5 5.50

200

400

600

800

1000

1200

Resonant Frequency (kHz)

(c) Bandwidth (fmin)

Figure 4-4. Sensitivity of optimized MDS, Sτ,overall, and fmin to tolerances in variables inconstants of Design 1 using Monte Carlo simulation.

formulation i.e., design variables, objective function, constraints, and variable bounds. The

optimized results for five different designs were presented. Lastly, a sensitivity analysis was

performed to study the effect of variation of the optimized design variables on the MDS and

constraints for Design 1. A Monte Carlo simulation was performed to investigate the effect

on sensor performance due to simultaneous variation in design variables and constants. The

optimization provided better insight into the design physics. Dominant design variables that

influence the design most are identified easily via the sensitivity analysis.

120

Page 121: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

CHAPTER 5DEVICE FABRICATION AND PACKAGING

This chapter describes a detailed fabrication process for the metal coated capacitive

floating element shear stress sensor. The fabrication process uses a 2-mask Si bulk microma-

chining and metal plating processes to form a released, metal passivated, floating element

sensor. The fabrication process starts with a 100 mm (100) p-type silicon on insulator (SOI)

wafer with a 45 µm highly doped silicon layer above a 2.0 µm buried silicon dioxide (BOX)

layer. The highly doped device layer has a low resistivity of 0.001 − 0.005 Ω − cm. The

top silicon (device layer) and BOX are backed by a 500 µm thick, float zone, bulk silicon

substrate needed to provide mechanical support. The wafers are also double-side polished to

allow patterning on both sides of the wafer. Deep reactive ion etching (DRIE), an anisotropic

silicon etch process, defines the floating element structure on the device layer. The fabrica-

tion process is described followed by the sensor packaging. Figure 5-1 shows the schematic

of a single die of the shear stress sensor.

A A

A-A

p++ Si (45 )SiO2 (2.0

)Float Zone Si

(500 )v+ v-

Electroplated

Ni (~0.2 )Plan View

Section

View

Figure 5-1. Schematic showing the plan view of a single die of the proposed shear stresssensor and its section view indicating various layers.

121

Page 122: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

5.1 Fabrication Process

The sensor fabrication is a simple process involving two lithography steps, three etching

steps, and an electroplating step shown in Figure 5-2. First, the floating element structure is

defined using DRIE. Next, a thin layer of nickel is electroplated on the device layer. A front

to back alignment step defines the mask for the backside etch. This is followed by backside

DRIE with the BOX layer serving as an etch stop. Lastly, the floating element structure is

released by wet etching the BOX layer.

p++

p++

p++

Highly doped

device layer

Float Zone

bulk substrate

SOI Wafer

Photoresist

AZ1512

Pattern using

EVG 620

Mask Aligner

DRIE on Si

Device layer

(STS)

Pattern Back

Side with AZ

9260

DRIE on back

side.

BOX etch

stop

EVG 620

Mask Aligner

BOE dip to

remove BOX

Directly

electroplate

Ni onto highly

doped Si

Device layer

(0.2 )

Photoresist

SiO2 (BOX)

Silicon

Nickel

(a)

(b)

(c)

(d)

(d)

(e)

(f)

(g)

(h)

Figure 5-2. Step by step fabrication process.

The motivation for electroplating on the silicon is briefly explained here before illustrat-

ing the fabrication process. The highly doped micromachined device layer enables seedless

electroplating on the entire exposed sensor silicon surface. Seedless electroplating eliminates

several steps associated with seeded electroplating, significantly simplifying the fabrication

process. Electroplating eliminates unwanted charge accumulation on the dielectric native

122

Page 123: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

oxide formed on the device surface by providing a conductive metal coating. This allevi-

ates drifts due to trapped charge at dielectric interfaces. Electroplating also eliminates an

additional alignment step for metallization to make electrical contacts to the device silicon.

The details of the fabrication process are described in the sections that follow. A detailed

process flow describing the process parameters, equipment and the labs used for fabrication

are provided in Appendix D.

5.1.1 Floating Element Trench

An RCA clean is performed to clean the wafer surface before beginning the sensor

fabrication process. Photoresist is patterned on the device surface of the SOI wafer to define

the floating element. Standard anisotropic DRIE technique is used to form the floating

element, tethers, and comb fingers on the device layer with the BOX layer as an etch stop.

In addition to having an etch stop for DRIE, using an the SOI also enables precise device

layer thickness, enabling closer match with the design value. Electrical isolation of three

terminals for the differential capacitance measurement scheme, is achieved during this DRIE

step. Both alignment marks needed for front to back alignment for the sensor release and

the dicing marks for die separation, are etched during this DRIE step.

The challenge during this fabrication step is the ability to etch high aspect ratio, narrow,

parallel-walled trenches defining the comb fingers. Narrow trenches increase capacitance and

also allow more comb fingers in the available space increasing the overall static capacitance.

Large static capacitance is critical to mitigate detrimental effects of parasitic capacitance,

discussed in Chapter 3. Parallel walls enable a uniform electric field between the comb

fingers, which is the theoretical basis of the sensing scheme analyzed in Chapter 3. Figure 5-

3 shows scanning electron microscopy (SEM) images indicating the quality of the etch and

the asymmetric gaps (Chapter 3) of the floating element structure.

123

Page 124: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

TetherComb

Fingers

Floating

Element

Stationary

Substrate

Asymmetric

Gaps

200

20

Figure 5-3. SEM images of comb finger structure etched using DRIE, giving 3D perpective.

5.1.2 Seedless Electroplating of Nickel

Seedless electroplating simplifies the fabrication process by eliminating the need to de-

posit a conductive metal seed layer [79]. The use of a seed layer also adds additional fabri-

cation steps. The interdigitated comb fingers have a separation distance of 3.5 µm resulting

in deep - narrow trenches. For sufficiently thin metal layers, only evaporating or sputtering

may suffice, eliminating the electroplating step. However, evaporating or sputtering metal

uniformly onto the high aspect ratio trench sidewalls poses a significant fabrication chal-

lenge due to poor side wall coverage. If used as a seed layer, such a deposition will result in

non-uniform metal plating. This may result in non-uniform gaps causing difference in the

capacitance between adjacent comb fingers, promoting mismatch between C10 and C20. Also

124

Page 125: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

the possibility of charge trapping on the exposed silicon surface raises concerns regarding

drift. Electroplating being a pre-release step, a seed layer will also result in metal deposition

at the bottom of the trenches. This adds a metal removal step to isolate the capacitive

electrodes of the sensor. Nickel electroplating is performed on the previously micromachined

highly doped device layer during this step. Nickel also provides good corrosion resistance

that allows the use of the sensor in harsh environments [80].

A commercially available high speed nickel sulfamate bath from Technic Inc is used

for the electroplating process [81]. The solution is maintained at a constant temperature of

90 F during the plating process with constant stirring. The electroplating process is based

on the work presented in [79]. A nickel electrode is used as the anode and the highly doped

conducting device layer of the SOI wafer serves as the cathode. Previous studies claim that

from an electroplating standpoint, highly doped silicon has conductivities similar to that of

normally doped silicon with a conductive layer [79].

Silicon surface preparation prior to electroplating is essential to ensure good plating

quality. Organic residue left from previous processing is removed using a standard piranha

(3 : 1, H2SO4 : H2O2) clean procedure. The wafers are immersed in the piranha solution

at 120C for 10 min. A thin layer of native oxide is formed on the silicon surface in the

piranha and if exposed to air. The oxide creates an insulating non-conducting surface in-

hibiting the electroplating process. Buffered oxide etch (6 : 1) (BOE) is used to remove

this oxide. The wafer is immersed in the BOE solution for 5 min. The BOE immersion

makes the silicon surface hydrophobic [79]. This is immediately followed by a 5 min dip in

2-propanol. The propanol treatment wets the silicon surface and makes it hydrophillic. The

propanol treatment provides a more uniform growth of Ni during the plating process [79].

Following the propanol treatment the wafer is immersed into the plating bath with the power

supply switched on. Immersion into the aqueous plating bath, without current switched on,

facilitates oxidation inhibiting the plating process [79].

125

Page 126: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

The plating is done in two phases, the rapid nucleation phase called striking and the

fine plating phase for uniformity. A high current density for a short time interval in the

nucleation phase called striking, helps rapid nucleation of Ni on the silicon surface. This

creates locally metalized regions facilitating further electroplating. A lower current density

and longer plating time in the fine plating phase gives finer grain size and a more uniform

surface. This electroplating step creates a 0.2 µm thick coating of Ni on the micromachined

structure created in the previous step.

Vertical sidewalls resulting from the etching process and the uniformity of the elec-

troplating process is verified by looking at SEM images of a cleaved sample of the wafer.

Figure 5-4 shows both vertical sidewalls and uniformity of the plated metal.

5.1.3 Backside Release

This is the final and the most important phase of fabrication. Most of the yield depends

on this release phase of sensor fabrication. A “front to back alignment” enables back side

patterning of the wafer followed by the backside etch.

The trenches defining the floating element in the device layer are filled with photoresist

to provide additional mechanical strength and prevent breaking of tethers during release.

Photoresist 10 µm in thickness is spun on the device layer to serve as a protective coating.

The back side of the wafer is now spun with photoresist and patterned using the front to

back mask aligner.

The wafer is attached with its face down on to a carrier wafer, pre-spun with photoresist.

The carrier wafer provides mechanical support and heat dissipation path during the back

side etch. DRIE is used to etch away the bulk silicon using the BOX as the etch stop.

5.1.4 Die Separation

The processed wafer is separated from the carrier wafer using photoresist stripper, and

cleaned for the die separation step. Heat sensitive, thermal release tapes (REVALPHA) by

Nitto Denko Corporation are used in this step. The front side of the wafer is covered with

a single sided tape for protection. The wafer is mounted on another carrier wafer using a

126

Page 127: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

(a)

(b)

SiliconNickel

100 20

0.143 Figure 5-4. SEM image of a cleaved wafer sample depiciting a) vertical sidewalls and b)

uniformly plated nickel.

double sided heat sensitive tape. Using a dicing saw, the wafer is diced along the dicing lines

etched on to the wafer in the initial DRIE step. The wafer is then placed on a hot plate

at 80C to remove the tape, resulting in individual die. Each individual die is immersed in

BOE for 20 min to release the sensors. The sensors are immediately immersed in methanol

and then supercritically dried in CO2 to avoid stiction. Figure 5-5 shows an optical image

of a fully processed, released floating element sensor die, etched on the SOI wafer.

127

Page 128: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

5.1.5 The MEMS Shear Stress Sensor

Prior to dicing, the wafer consists of 200 die. Only one wafer out of the four processed

wafers is released. The released die measured 5 mm×5 mm in size. The seemingly large size

of the die for a MEMS sensor is not a limitation, but is for better handling and to withstand

repeated tests, thus ensuring first-run success of the proof-of-concept sensor. Figure 5-6

shows the sensor die placed on a Florida quarter to get a perspective of the size of the

sensor.

500 500

Surrounding

Substrate

Bond

PadsBond

Pads

Figure 5-5. Microscopic image of a released floating element sensor structure from Design 3(Table 4.4).

Figure 5-6. A 2 mm× 2 mm sense element on a 5 mm× 5 mm sensor die.

128

Page 129: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

5.2 Packaging

The fabricated sensor die needs to be packaged to characterize the sensor performance.

A recess is cut into a printed circuit board (PCB) such that the sensor is flush with the PCB

surface which again is mounted flush in a Lucite plug that fits with the calibration setup.

Figure 5-7 shows the schematic showing the overall packaging scheme for the sensor. The

contacts to the board are made using gold wire bonds. The Knowles SiSonic amplifier is

placed on the back side of the PCB in close proximity to the sensor to minimize parasitic

attenuation arising from the connection traces on the PCB (Figure 5-8). SMB connectors on

the backside of the PCB allow electrical contact to the sensor for biasing, input and output

signals, also shown in Figure 5-8. The detailed drawings of the PCB and Lucite plug are

provided in Appendix D along with vendor details.

Sensor

Die

Printed

Circuit Board

(PCB)Lucite

Plug

Figure 5-7. Schematic of sensor package for shear stress characterization.

129

Page 130: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Flush mounted

sensorSiSonic

TMvoltage

follower

SMB

connectors

Figure 5-8. Photographs of sensor packaged on a 30 mm× 30 mm PCB.

130

Page 131: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

CHAPTER 6EXPERIMENTAL CHARACTERIZATION

In this chapter, the differential capacitive shear stress sensor, designed and realized in

Chapter 3 and Chapter 5, is experimentally characterized. Design 1 in Table 4-3 is experi-

mentally studied for impedance, static, dynamic, and noise characteristics. The experimental

setup for these measurements is described. Experimental results and discussions that follow

give physical insight into the sensor performance characteristics.

6.1 Experimental Setup

This section describes the setup for experimental characterization of the shear stress

sensor. The setup for die-level impedance measurements on the sensor for estimating its

nominal capacitance is presented. The laminar flow cell used for mean shear stress mea-

surement is illustrated. The dynamic characterization using a plane wave tube (PWT) is

explained, in which dynamic shear stress is generated using acoustic plane waves. Lastly,

the setup for noise floor measurements on the post-packaged sensor in a Faraday cage, is

explained.

6.1.1 Impedance Measurements

Impedance measurements on the pre-packaged sensor are performed in a probe station

using a four point measurement technique. Since the packaged sensor has the amplifier con-

nected to it, which alters the overall impedance of the system, the impedance measurement

is made at the die level. The schematic for this measurement is shown in Figure 6-1. The

measurement setup consists of four probes, two of which (Hp and Hc) are connected to one

bond pad / electrode and the other two (Lp and Lc) are connected to the other bond pad

/ electrode. Hp and Hc stand for high potential and high current and Lp and Lc stand for

low current and low potential, respectively. The impedance of the sensor is measured using

131

Page 132: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Voltage

Current

Lc Hc

Lp HpProbe

Station

HP 4294 A

Impedance

Analyzer

Hp – High Potential

Hc – High Current

Lp – Low Potential

Lc – Low Current

Figure 6-1. Schematic showing the die-level impedance measurement setup for the sensor.

an HP 4294A impedance analyzer, which uses an I-V method where

Zsensor =V

I=

VHp−Lp

IHc−Lc

(6–1)

and Zsensor = R− jXc = R +1

jωC1,2

. (6–2)

The impedance measurement helps to estimate the nominal sensor capacitances, inclusive of

the parasitics i.e., C10+p1 and C20+p2. The subscripts p1 and p2 represent parasitic capaci-

tances in addition to the nominal values C10 and C20, respectively. In case of a mismatch

i.e., C10 6= C20, the capacitor electrodes experience different electrical forces when biased.

As a result, the floating element settles at a new mean position and therefore a new nom-

inal capacitance value at each bias voltage. The sensor nominal capacitance is therefore

estimated as a function of the applied bias voltage. This test also helps to quantify the

mismatch between C1 and C2.

132

Page 133: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Biasing

Circuit

Heise Pressure

Gauge

PC

(LabView)

Mass Flow

Controller

Source

Meter

P1 P2

xh

P

Bandpass

Filter

DAQ

Sensor

Figure 6-2. Schematic showing the mean shear stress/static calibration setup using Poiseuilleflow in a 2-D channel.

6.1.2 Mean Shear Stress Measurement

Mean shear stress measurements are performed in a laminar flow cell, which consists of

two aluminum plates separated by a known distance to form a channel. Figure 6-2 shows

the schematic of the sensor calibration setup for static/mean shear stress inputs. Two

dimensional Poiseuille flow is assumed in the channel, 330 mm long, 100 mm wide, and

1 mm in height, which may be varied if necessary. The expression of shear stress for steady,

fully developed, 2-D flow in a channel is [82]

τw =h

2

dP

dx, (6–3)

where dP/dx is the pressure gradient driving the flow and h is the channel height. As long as

the pressure drop along the length of the channel is linear, measuring the pressure differential

between two pressure taps that are separated by a known distance, ∆x, results in

τw =h

2

P2 − P1

∆x= −h

2

∆P

∆x. (6–4)

The linear pressure drop was previously verified by [83] and [84]. The Poiseuille flow as-

sumption is confirmed by checking for compressibility effects via the Mach number, M , and

ensuring laminar flow via the Reynolds number, Re. The average velocity in terms of the

133

Page 134: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

pressure gradient and channel height is [82]

V = − 1

12µ

dP

dxh2, (6–5)

The maximum velocity is given as

Vmax =3

2V . (6–6)

Based on the velocities, M and Re are expressed as,

M =Vmax

c(6–7)

and Re =V h

ν, (6–8)

where c is the isentropic speed of sound in air and ν is the kinematic viscosity of air. For

the maximum tested shear stress based on Equation 6–4 is 2.9 Pa, M ≈ 0.06 and Re ≈ 442.

This validates the Poiseuille flow assumption, which is true if Re < 1400 and M < 0.3 [82].

+

-RbCp+Ci

Vout

Vac sin(ωt)

-Vac sin(ωt)

C1

C2

-Vcc

+Vcc

Sensing Element

Bond pad for

sense element

through tethers

Substrate on

either side of bond

pad form

electrodes

Figure 6-3. Schematic with optical image of sensor die (5 mm × 5 mm) indicating float-ing element, contact pads, and interface circuit (voltage buffer) for mean shearcharacterization.

134

Page 135: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

The sensor plug is mounted on the flow cell and biased with sinusoidal bias sources for

modulating the sensor output voltage. Figure 6-3 shows a picture of a fabricated sensor and

a schematic of the interface circuit connections for this measurement. For static shear stress

characterization, the flow rate through the channel is varied and sensor output voltage is

measured. A change in flow rate alters the wall shear stress via a change in the pressure

gradient along the length of the channel. The flow rate is varied electronically using an

AALBORG GFC47 mass flow controller, and the corresponding pressure differential and

sensor output voltage are recorded. A Keithley 2400 sourcemeter serves as the voltage source

to control the flow controller. The pressure differential is measured using the Heise ST-2H

pressure gauge using a differential pressure module (50”H2O). An SR560 low noise amplifier

with unity gain bandpass filters the sensor output (around the biasing frequency) for noise

reduction. This output is then fed to a DAQ card. Pressure drop from the Heise pressure

gauge and the sensor output voltage are acquired using LabView on a PC. Equation 6–4 and

the recorded data are used to estimate the sensor’s mean shear stress sensitivity.

6.1.2.1 Biasing Scheme Implementation

The initial plan to implement the MOD-DMOD scheme discussed in Section 3.1.3.3,

is not implemented due to issues like dc offsets and drift associated with the demodulation

circuitry. Only the modulation scheme is implemented for mean shear stress measurements.

This section explains the reasons and the practical implementation of the current biasing

scheme. The configuration shown in Figure 6-3 results in a modulated voltage at the amplifier

input. A shear induced mean capacitance change produces a corresponding change in this

voltage amplitude. The maximum voltage that may be applied to the input of the SiSonic

voltage buffer is limited to 100 mV . Without an applied shear stress, ideally matched

sense capacitors and out-of-phase (180) bias voltages result in no voltage at the input node

(Section 3.1.3.3) [42]. Mathematically, this may be expressed as (Equation 3–64)

vin = Vacsin(ωt)C1 − C2

C1 + C2 + Cp + Ci

= 0, for C1 = C2. (6–9)

135

Page 136: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

In case of a mismatch between the capacitors, C2 = λC1, a proportional sinusoidal voltage

exists at the input node given as

vinmismatch= Vacsin(ωt)

C1 − C2

C1 + C2 + Cp + Ci

(6–10)

or vinmismatch= Vacsin(ωt)

(1− λ)C1

C1 + C2 + Cp + Ci

, for C2 < C1 and λ < 1. (6–11)

This mismatch voltage limits the dynamic range for the static shear stress measurement since

its upper end is limited at 100 mV . The mismatch voltage also increases with increase in

the bias voltage amplitude, Vac, limiting Vac to smaller values and thus lowering sensitivity.

The mismatch voltage may be nulled by using bias voltages of different amplitudes. For

illustration, let Vac1 = Vacsin(ωt) and Vac2 = −βVacsin(ωt) be the bias voltages applied to

C1 and C2, then Equation 6–11 may be rewritten as

vinmismatch=

C1Vac1 − C2Vac2

C1 + C2 + Cp + Ci

= Vacsin(ωt)(1− λβ)C1

C1 + C2 + Cp + Ci

. (6–12)

Thus, if λβ = 1 in Equation 6–12, then vin = 0 i.e., with no shear stress input, the nominal

output voltage of the biased sensor, with mismatched capacitors, is zero. To achieve λβ = 1,

a biasing circuitry is required that can independently control both amplitude and phase of

the two bias voltage signals. Note, that the mismatch is a practical aspect which depends

on fabrication process and is usually quantified via electrical characterization.

A custom made biasing circuit is designed and implemented, which consists of four

AD4898 operational amplifiers, shown in Figure 6-4, with the required phase adjustment

and biasing capabilities. A sinusoidal signal from a function generator is input to two

op-amps, which serve as phase adjust circuits. They are connected in an all pass filter

configuration [85], having variable capacitors, Cvar, to independently adjust the phase of

their outputs. The outputs of the two filter op-amps are input to the voltage buffer and the

inverting op-amp with an adjustable gain (β = R2

R1). The output from the non-inverting and

inverting amplifiers serve as bias voltages for the sensor capacitances, C1 and C2 (assuming

C2 = λC1). Thus, this biasing scheme implementation helps to null out the initial sinusoidal

136

Page 137: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

+

+

+

Vac

+

R1

R1

R1

Cva

r

R1

R1

R1

Cva

r

R1

R2

AD

4898

AD

4898

AD

4898

AD

4898

Phase Adjust

Stage

Inverter

Voltage

Buffer

Vac sin(ωt)

-Vac sin(ωt)

C1

C2

Figure 6-4. Schematic of the biasing circuit scheme to control phase and amplitude of biasingsignals to null out output from a potential mismatch in sensor capacitances.

voltage at the middle node via phase and amplitude adjustment. This helps to use higher

bias voltages, improving sensitivity and dynamic range.

6.1.3 Dynamic Measurements

The sensor is characterized for dynamic shear sensitivity, pressure sensitivity, and fre-

quency response in the PWT. Stokes’ layer excitation from propagating or standing acoustic

plane waves is used to estimate the dynamic sensitivity, linearity and frequency response

of the sensor. The sensor configuration for this measurement is the same as shown in Fig-

ure 6-3, but with dc biasing for the sensing capacitors. A known oscillating shear stress

input is generated using acoustic plane waves in a duct [51]. The oscillating acoustic field

in conjunction with the no-slip boundary condition at the duct wall results in an oscillating

velocity gradient, generating a frequency-dependent shear stress. This enables a theoretical

estimate of the wall shear stress if the acoustic pressure is known at a given axial location

137

Page 138: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

in the duct. For plane progressive waves, the input shear stress, τin, corresponding to the

amplitude of the total acoustic pressure, p′, at a given frequency, ω, is theoretically given

as [86]

τinprog = −p′√

jων

cej(ωt−kx0)tanh

(a

√jω

ν

), (6–13)

where k = ω/c is the acoustic wave number, a is the duct width and x0 is the axial location

along the length of the duct. The term η =√

a2ω/ν is the non-dimensional Stokes number.

Equation 6–13 is a 1-D shear stress solution and may be used for 2-D acoustic fields for

η > 2 [86]. The 1-D solution in Equation 6–13 is used in this dissertation because the lowest

frequency for measurements in the PWT is 1 kHz, which corresponds to η = 520.

There are three different measurement setups, each of which vary in terms of the termi-

nation used in the PWT and the sensor location. The first setup uses a rigid termination at

the end of the PWT for shear stress measurements, the second setup has the sensor mounted

at the end for normal acoustic incidence for pressure measurement, and the last setup uses an

anechoic termination for combined shear and pressure measurements. There may be acoustic

reflections, which vary as a function of the PWT termination. For standing acoustic plane

waves, which has reflected waves, the input shear stress is

τinstanding=

(−1

c

√jων tanh

(a

√jω

ν

)ejkds −Re−jkds

ejk(ds−δ) + Re−jk(ds−δ)

)p′ejωt, (6–14)

where R is the complex acoustic reflection coefficient. The sensor and a reference microphone,

which measures acoustic pressure amplitude, are at distances, ds and dps from the termination

(Figure C-1). Appendix C provides the theoretical derivation for the dynamic wall shear

stress based on acoustic reflections from the termination.

6.1.3.1 Dynamic Shear Stress Measurement Setup

This section describes the calibration of the sensor for dynamic shear stress inputs. The

measurement setup and the resulting standing acoustic wave pattern in the PWT is shown

in Figure 6-5. Plane waves are generated by a BMS 4590P compression driver (speaker),

138

Page 139: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Techron 7540 Power

Supply Amplifier

B & K Pulse

Analyzer System

PC

Acoustic Plane

Wave

Shear Stress

SensorSpeaker

Microphones

Rigid

Termination

R=1

PressureVelocity

Standing Wave Pattern

Figure 6-5. A schematic of the dynamic calibration setup for measuring shear sensitivityusing rigid termination with the sensor located at pressure node and velocitymaximum.

mounted at one end of the PWT, while the other end is fitted with a rigid termination. The

PWT consists of a 1” × 1” cross section square duct, with a cut-on frequency of 6.7 kHz

for higher order modes in air, above which Equation 6–13 is no longer valid. The sensor

being tested and a reference microphone (B&K 4138, 1/8”) are flush mounted at known

positions along the length of the tube and the acoustic frequency is chosen such that the

sensor is at a velocity maxima and the microphone is at a pressure maxima. A B&K PULSE

Multi-Analyzer System (Type 3109) acts as the microphone power supply, data acquisition

unit, and signal generator for the compression driver.

For a perfectly rigid termination (R = 1), the acoustic wave is reflected in phase with

the incident wave, resulting in a pressure maxima (doubling) and a velocity minima at the

termination [87]. The reference microphone is placed adjacent (< 3 mm away) to the rigid

termination (pressure maxima) without appreciable phase errors at the testing frequencies

(6.6 kHz maximum). At a pressure node the sensor response is dominated by shear stress,

which is a function of the velocity gradient. Another microphone, located at the same

139

Page 140: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

axial location as the sensor, monitors the pressure at its node to confirm minimum pressure

contribution to the sensor output.

6.1.3.2 Dynamic Pressure Calibration Setup

When used in a real application, the sensor output may have additional components due

to the out-of-plane motion of the sensor caused by forces such as pressure (Section 1.1.3). The

out-of-plane sensitivity of the sensor is a result of a mismatch (non-ideal case) in the sensor

nominal capacitance. For out-of-plane motion both sensor capacitors change in the same

sense i.e., increase or decrease simultaneously. Thus the differential output voltage is zero

for matched capacitors. However, a capacitance mismatch results in a proportional output

voltage. This may be simply illustrated using the example in Section 6.1.2.1. Consider

C2 = λC1, where λ < 1, and with dc bias voltages, ±Vb. Consider the dynamic out-of-

plane motion is sinusoidal i.e., sin(ω1t). The resulting input voltage to the amplifier due to

sinusoidal variation in the capacitance and based on Equation 6–11 is

vinmismatchp= Vb

(1− λ)C1sin(ω1t)

C1 + C2 + Cp + Ci

. (6–15)

Thus, if λ = 1 (no mismatch) the output is zero while if 0 < λ < 1, the sensor has a finite

sensitivity to out-of-plane motion due to forces such as pressure.

To study the out-of-plane sensitivity, if any, the sensor is subjected to normal forces

(pressure). The pressure response of the sensor is directly measured by mounting the pack-

aged sensor at the end of the PWT to impart normal acoustic incidence (Figure 6-6).

6.1.3.3 Combined Shear and Pressure Measurement Setup

As discussed in the previous section, with mismatched capacitors, the sensor output

voltage may be a composite result of the possible multi-axis (in-plane and out-of-plane)

motion of the floating element. In this section the response of the sensor under simultaneous

shear stress and pressure inputs is investigated. To emphasize the importance of this study,

consider a shear stress of 0.5 Pa at 4.2 kHz. The corresponding acoustic pressure (based

on Equation 6–13) is approximately 142.7 dB or 273 Pa. This pressure is over two orders of

140

Page 141: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Techron 7540 Power

Supply Amplifier

B & K Pulse

Analyzer System

PC

Acoustic

Plane Wave

Shear Stress

SensorSpeaker

Microphone

Figure 6-6. A schematic of the dynamic calibration setup for measuring pressure sensitivityusing normal incidence acoustic waves.

magnitude higher than the shear stress value. Let Sshear and Spressure be the respective shear

and pressure sensitivities and let Vshear = Sshearτw and Vpressure = Spressurep be the respective

voltages due to shear and pressure inputs. Now, if Sshear ≈ 100 Spressure, then for the current

example, Vshear ≈ Vpressure. Thus the pressure sensitivity may be two orders of magnitude

smaller than shear sensitivity but, if the pressure input is two orders of magnitude higher

than the shear input, the pressure contribution to the total output is the same as that from

shear.

To investigate this, the PWT is fitted with an anechoic termination (a 30.7” long fiber-

glass wedge), which minimizes acoustic reflections to ensure plane propagating waves in the

duct. The composite sensitivity (shear and pressure) and the frequency response of the

sensor are studied with this setup, shown in Figure 6-7. The setup in Section 6.1.3.1 is not

used to study the frequency response of the sensor because the fixed sensor location (nodal

position) for pressure minima in the standing wave pattern limits the number of acoustic

frequencies.

6.1.4 Noise Measurement

A noise measurement of the packaged sensor provides an estimate of its shear stress res-

olution. It is most relevant to measure the noise floor in the actual shear stress measurement

setup. However, the noise floor of the data acquisition system for the dynamic measurements

141

Page 142: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Techron 7540 Power

Supply Amplifier

B & K Pulse

Analyzer System

PC

Acoustic Plane

Wave

Anechoic

Termination

Shear Stress

Sensor

Speaker

Microphone

Figure 6-7. A schematic of the dynamic calibration setup for shear stress measurement withplane progressive acoustic waves.

is at 1× 10−6 V/√

Hz, which is much higher than that of the sensor, limiting its utility for

noise measurements. Also, the acquisition system may be different based on a measurement

setup. The measurement in the Faraday cage helps to determine the absolute noise floor

of the sensor itself independent of the acquisition system used in a real application. The

packaged sensor, biased with known voltages, is placed in a double Faraday cage without

a physical input, as shown in Figure 6-8. The output of the SiSonic buffer is measured

using a SR785 spectrum analyzer. The details of the sources of noise and the experimental

setup for this measurement can be found in [88]. This experiment is repeated both with

(total noise) and without the sensor (setup noise) to isolate the sensor noise from that of the

measurement system.

6.2 Results and Discussion

This section provides the results of the experiments described in the previous sections.

The settings for each characterization setup is provided. The impedance, static, dynamic,

and noise characteristics of the sensor are explained via experimental plots.

6.2.1 Impedance Measurements

This section presents the impedance measurement results performed on the pre-packaged

sensor die using the probe station and the impedance analyzer. The impedance is measured

142

Page 143: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Spectrum

Analyzer

(SR 785)

Shielded Box

Shear Stress

Sensor

Shielded Box

+

-Rb Cp+i

Vout

Vb

-Vb -Vcc

+Vcc

Computer

(LabView)

Figure 6-8. A schematic of the noise measurement setup for the packaged shear stress sensorusing a double Faraday cage.

across C1 and C2 to estimate their nominal capacitances. The dc bias voltage across the

capacitors is swept from 0 − 18 V with a 400 mV (peak) source signal at 50 kHz. The

bandwidth for this measurement is set to 5, which precisely controls the sweep parameter

(dc bias voltage). The averaging in the system is turned off and the measurement is repeated

31 times with 100 points for each measurement. Ensemble averages are used to estimate the

mean and standard deviation of the measured capacitance as a function of bias voltage. Fig-

ure 6-9 shows the results from the impedance measurements. The results are significantly

different from predicted capacitance value in Chapter 4 due to large parasitic capacitances.

The predicted capacitance is 1.56 pF while the measured values are approximately 14.8 pF

and 16.3 pF for C10+p1 and C20+p2, respectively. The parasitic capacitance is formed be-

tween the highly doped device layer and the float zone bulk silicon substrate separated by

the dielectric BOX layer. The difference between C10+p1 and C20+p2 is roughly 1.5 pF , which

143

Page 144: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

0 5 10 15

16.3

16.31

16.32

16.33

Bias (V)

C20

(pF

)

0 5 10 15

14.8

14.805

14.81

14.815

Bias (V)

C10

(pF

)

Figure 6-9. Capacitance from impedance measurements on the pre-packaged sensor die.

is approximately same as the predicted sensor capacitance. Based on the fabrication pro-

cess and the sensor geometry such a difference is unlikely in the sensor capacitance itself,

suggesting that the difference must be due to the unexpectedly large parasitic capacitance.

Furthermore, a suspected drift due to the probe station was a limitation in obtaining an

accurate quantitative measurement of C10+p1 and C20+p2. This restricted further character-

ization to estimate C10 and C20 and also to quantify their mismatch. The details of the

measurement drift, unexpected parasitic capacitance, and the consequent non-idealities in

the sensor performance are discussed in Section 7.2.2.1 and Section 7.2.1.1.

6.2.2 Mean Shear Stress Characteristics

This section presents the results from the experiments performed in the flow cell for

static calibration of the sensor. The sensor is biased with sinusoidal voltages having am-

plitudes ±2.5 V and frequencies of 10 kHz, respectively. Even with fine tuning using the

biasing circuitry, an ≈ 40 mV amplitude signal remains at the sensor output, while the

permissible input voltage to the SiSonic buffer is 100 mV . If a good capacitance match is

144

Page 145: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

achieved during fabrication, higher bias voltages as per design may be applied. During the

measurement, a time delay of 20s is set between every change in state of the flow controller

and data acquisition to enable the flow to reach a steady state. The data acquisition from

the Heise is limited due to its slow updating speed while the sensor signal is at 10 kHz,

requiring higher sampling rates. Therefore, the pressure and voltage were sampled inde-

pendently, with the pressure sampled at the highest read rate allowed by the serial port

and the sensor signal sampled at 40 kHz and a uniform window is applied with no overlap.

Although simultaneous signal acquisition is ideal, because the measurement is static/mean,

enough averages in the data yield reasonable results. The acquisition time is limited to

the time taken for the pressure measurement due to the slow data acquisition. It is ideal

to have long acquisition times, since it is a mean measurement. However, due to drift in

the biasing voltage, resulting from potential charge build up on the on-the-board variable

capacitors, Cvar, the data acquisition times are kept relatively small. This measurement

uses autospectral estimates of the sensor output voltage at 10 kHz. The equipments and

settings for this measurement are summarized in Table 6-1. The number of averages and

the sampling times are chosen such that, for each flow rate setting, the pressure and sensor

voltage measurements approximately end at the same time.

Table 6-1. Measurement settings for static calibration in the flow cell.

Pressure differential data Number of averages 200

Sensor voltage data Sampling Frequency (kHz) 40

Number of Samples 2000

Number of linear spectral averages 500

Keithley 2400 sourcemeter compliance (mA) 100

SR560 Coupling AC

Filter frequency band (kHz) 3− 30

Gain 1

The shear stress input varies from 0.18 Pa to 2.9 Pa. Figure 6-10 shows the sensitivity

plot for this mean shear measurement. The initial voltage due to the capacitive mismatch is

145

Page 146: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

subtracted from the measured values. The normalized random uncertainty in the measured

voltage is 4.5% (Section 6.2.5). A linear curve fit on the data set is used to estimate the

sensor sensitivity (R2 = 0.9958). The sensor has a linear sensitivity of 0.94 mV/Pa up to

the testing limit of 2.9 Pa.

0.5 1 1.5 2 2.5 3

1.2

1.4

1.6

1.8

2

2.2

2.4

2.6

2.8

3

3.2

Shear Stress (Pa)

Vol

tage

(m

V)

Figure 6-10. Linear sensor output voltage as a function of mean shear stress at bias voltageamplitude of 2.5 V at 10 kHz.

The plot indicates a large uncertainty in the shear stress value compared to the output

voltages but the fit is based only on the mean values of shear stress and sensor output

voltage. Such large error values warrant further analysis to deduce the range of sensitivities

within the measured uncertainty limits. This is achieved via a Monte Carlo simulation

assuming a Gaussian profile for the uncertainty at each point. Each measurement point is

randomly perturbed from its mean value within its error bounds in both shear stress and

output voltage. A linear fit on the resulting curve is used to estimate the sensitivity. This

procedure is repeated 50000 times to build a distribution of the overall shear stress sensitivity

represented in the form of a histogram. The corresponding distribution of the normalized

residuals values (R2) for each fit are also plotted. Figure 6-11 shows the histograms for the

sensitivity and R2. The sensitivity and R2 values and the 95% confidence intervals from this

146

Page 147: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

study are summarized in Table 6-2. The 2D scatter of the sensitivity and R2 indicates that

the sensitivity from the measurement is towards an extreme of the distribution where the

fit is better with R2 = 0.9958. However, the perturbations about the mean measured values

result in a sensitivity that is much different (≈ 1.02 mV/Pa) but with lower mean quality

of fit (R2 = 0.8288). The ellipsoidal shape of the scatter plot provides some intuition into

this effect. An important assumption for this analysis is that the perturbations for each

data point are uncorrelated, which is reasonable since each measured data point should be

a statistically independent value.

0.98 1 1.02 1.04 1.060

500

1000

1500

Sensitivity (mV/Pa)

(a) Sensitivity

0.75 0.8 0.85 0.9 0.950

500

1000

1500

Normalized Residual (R2)

(b) R2

(c) Joint Histogram

0.98 1 1.02 1.04 1.06

0.75

0.8

0.85

0.9

0.95

Sensitivity(mV/Pa)

Nor

mal

ized

Res

idua

l (R

2 )

(d) 2D Scatter

Figure 6-11. Distribution of the measured shear stress sensitivity due to uncertainty in themeasured shear stress and sensor output voltage.

147

Page 148: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Table 6-2. Sensitivity estimates using Monte Carlo technique onthe mean shear stress measurement data.

Parameter Mean Value (µMC) 95% ConfidenceInterval (2σ)

Sensitivity (mV/Pa) 1.017 0.997− 1.037

NormalizedSensitivity(mV/V/Pa)

0.407 0.399− 0.415

Normalized Residual(R2)

0.8288 0.7751− 0.8825

6.2.3 Dynamic Characteristics

As previously described (Section 6.1.3), the sensor characterization in the PWT involves

three different setups. Dynamic shear sensitivity results are obtained using the rigid termi-

nation in the PWT. Pressure sensitivity results are illustrated with the sensor mounted at

the end of the PWT for normal acoustic incidence. The frequency response and combined

sensitivity (shear + pressure) results, using the anechoic termination, are provided. During

each measurement, the recorded quantities are as follows: the respective auto power spectral

densities of the sensor and the reference microphone outputs and the transfer function and

coherence between the sensor and the reference microphone signal. For all sensitivity esti-

mates, the measured autospectra of the respective sensor and reference microphone signals

are used. For the frequency response measurement, the computed sensitivities and the trans-

fer function (sensor to microphone) are used. The measurement settings were kept the same

for all dynamic measurements and are summarized in Table 6-3. The normalized random

error for the autospectrum estimates is 3.2%. The normalized random errors for the mag-

nitude and phase of the frequency response function are each 0.04% (Refer Section 6.2.5).

6.2.3.1 Dynamic Shear Sensitivity

The sensor dynamic sensitivity for shear is measured using a rigid termination (Fig-

ure 6-5), where the sensor is placed at a quarter wavelength (velocity maxima) away from

148

Page 149: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Table 6-3. Measurement settings of B&K PULSE Multi-Analyzer System (Type 3109) for dynamic calibra-tion in the PWT.

FFT mode zoom

Windowing function uniform

Percentage Overlap 0

Frequency span 200 Hz − 6.4 kHz

Number of samples/FFT lines 400

Sampling time 62.5 ms

Time steps 122.1 µs

Total acquisition time 62.5 s

Frequency bin width 16 Hz

Number of linear spectral averages 1000

the termination. The chosen frequency based on the sensor location is 1.128 kHz. The input

shear stress is increased by raising the input sound pressure level (SPL) in steps of 5 dB

from 80 dB to 160 dB. The upper and lower ends of the SPL range are set by the perfor-

mance limits of the compression driver. The shear stress corresponding to the input acoustic

pressure varies from 0.1 mPa to 1.9 Pa. Figure 6-12 shows the dynamic shear sensitivity of

the sensor. The sensitivity of the sensor is deduced using a linear curve fit on the data set

(R2 = 0.99995). The estimated sensor sensitivity is 7.66 mV/Pa at dc bias of Vb = 10 V .

During this measurement, the SPL at the sensor axial location was ≈ 40 dB lower than the

maximum pressure measured by the reference microphone at the termination.

The normalized sensor sensitivities from both the mean and dynamic shear stress mea-

surements are 0.407 mV/V/Pa and 0.766 mV/V/Pa, respectively. The difference between

the two values is roughly 47%, which may be attributed to non-idealities in the sensor and

in the measurement setup several of which are discussed in Chapter 7.

149

Page 150: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

10−3

10−2

10−1

100

10−2

10−1

100

101

Shear Stress (Pa)

Vol

tage

(m

V)

Figure 6-12. Sensor output voltage at 1.128 kHz at a bias voltage of 10 V . Sensor is placedat a quarter wavelength (velocity maxima) from the rigid termination.

6.2.3.2 Pressure Sensitivity

The sensor mounted at the end of the PWT for normal acoustic incidence, is experi-

mentally characterized for pressure sensitivity at 4.2 kHz and dc bias voltages of Vb = 5 V ,

8 V , and 10 V , respectively (Figure 6-13).

The frequency (4.2 kHz) is chosen for comparison with the combined sensitivity mea-

surements that are presented subsequently. The SPL range for the experiment is 80−150 dB

varied in steps of 5 dB, to be consistent with the shear sensitivity tests. The minimal varia-

tion of the output at low SPL is because the sensor output voltage is below the noise floor of

the measurement system. The pressure sensitivity of the sensor at Vb = 10 V is 4.8 µV/Pa

(R2 = 0.9976), which is about 1000 times smaller than the shear sensitivity of the sen-

sor (Figure 6-5 and Figure 6-10). A comparison of the dynamic and shear sensitivities at

Vb = 10 V results in a pressure rejection, Hp, of approximately 64 dB for comparable forcing

in shear and pressure, which is computed as

Hp = 20log(Sshear/Spressure). (6–16)

150

Page 151: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

100

101

102

103

10−3

10−2

10−1

100

101

Pressure (Pa)

Vol

tage

(m

V)

Vb = 10 V

Vb = 8 V

Vb = 5 V

Figure 6-13. Voltage output as a function of pressure at 4.2 kHz at different bias voltages.

6.2.3.3 Combined Sensitivity and Frequency Response

The sensor sensitivity due to a possible multi-axis motion resulting from both shear

and pressure inputs is tested in this study. The anechoic termination is used, which ideally

results in plane progressive acoustic waves. Weak reflections if any are accounted for via

Equation 6–14. Since there are no pressure nodes with progressive acoustic waves (with

weak reflections), the sensor encounters both shear and pressure. The experiments are

conducted at three different dc bias voltages, 5 V , 8 V , and 10 V , at a frequency of 4.2 kHz.

This frequency is chosen to ensure high signal to noise ratio (for shear stress) even at low

SPLs, which is limited by the driving capability of the compression driver (From Equation 6–

13, τ ∼ p′√

ω). At the same time, the frequency is kept well below the estimated sensor

resonance (5 kHz). The input SPL is varied in steps of 5 dB from 80 dB to 150 dB. The

shear stress corresponding to the measured pressure input varies from 0.4 mPa to 1.16 mPa.

The sensitivity plots corresponding to this measurement are shown in Figure 6-14.

Note that although the voltage measurement is plotted as a function of shear stress, cross

axis (out-of-plane) sensor sensitivity due to pressure also contributes to the output voltage.

In all three cases, the sensor exhibits a linear sensitivity (R2min = 0.9995) up to the testing

151

Page 152: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

10−3

10−2

10−1

100

10−2

10−1

100

101

Shear Stress (Pa)

Vol

tage

(m

V)

Vb = 10 V

Vb = 8 V

Vb = 5 V

Figure 6-14. Linear sensor output voltage as a function of shear stress at 4.2kHz at 3 differentbias voltages.

limit of 1.1 Pa. The measured sensitivities are 23 mV/Pa, 19 mV/Pa, and 11 mV/Pa,

respectively at Vb = 10 V , 8 V , and 5 V . The combined sensor sensitivity is directly

proportional to the applied bias voltage as expected (Equation 4–3). A plot of the normalized

sensitivity (output voltage per voltage bias) indicates the same via the collapse in the data

for different bias voltages (Figure 6-15). The normalized sensitivities agree closely to

within 6% and are 2.268 mV/V/Pa, 2.316 mV/V/Pa, and 2.196 mV/V/Pa, at Vb = 10 V ,

8 V , and 5 V , respectively. Comparison of the normalized sensitivity estimates from the

rigid (2.268 mV/V/Pa) and anechoic termination (0.766 mV/V/Pa) experiments indicates

significant difference (66%) in the sensor performance in the presence of large pressure signals.

Note, unlike the anechoic case, for the rigid termination the sensor is primarily forced

by shear stress given that it is at a pressure node and velocity maximum. The monitored

pressure at this node is ≈ 40 dB or two orders of magnitude below the peak acoustic pressure.

Also, with a combined input having large pressure signals in addition to shear, the sensor

shear sensitivity may drop from its nominal value. For example, consider the sensor has

a finite out-of-plane deflection. This reduces the overlapping area of the capacitors, which

152

Page 153: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

10−3

10−2

10−1

100

10−5

10−4

10−3

Shear Stress (Pa)

Nor

mal

ized

Sen

sitiv

ity (

V/V

)

Vb = 10 V

Vb = 8 V

Vb = 5 V

Figure 6-15. Sensor output voltage normalized by bias voltage as a function of shear stressat 4.2 kHz.

lowers the sensitivity to in-plane motion, shown graphically in Figure 6-16. Furthermore, the

effective in-plane compliance may be lower since the deflected (out-of-plane) structure offers

a higher stiffness to in-plane motion. The stiffening of the device may not be an issue for

small out-of-plane deflections, but for large deflections (large pressures) this may significantly

alter device performance. This may need further theoretical and experimental analysis for

better understanding of the sensor response to combined shear and pressure inputs. The

frequency response of the sensor is measured for Vb = 10 V , with the anechoic termination

in place (Figure 6-7). As previously stated a rigid termination may not be used for this

experiment since the pressure nodes in the resulting standing wave pattern are limited. This

restricts the different discreet frequencies available for testing the frequency response. The

sensor and the reference microphone are placed at the same axial location along the length

of the tube. The expression of the frequency response normalized by the input shear stress

is [86]

H(f) =τout

τin

=V (f)

τin

∂τ

∂V, (6–17)

153

Page 154: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Original

Position

Displaced

PositionStationary

Comb Fingers

Figure 6-16. Schematic showing a set of overlapping comb fingers deflection due to both

shear (δs) and due to pressure (δp).

where V (f) is the sensor output voltage corresponding to the known shear stress input, τin,

which is theoretically estimated using Equation 6–13. The term ∂τ/∂V is the inverse of

the flat-band sensitivity magnitude of the sensor. There is an additional phase component

in the transfer function between the sensor and microphone. This component changes as

a function of the termination used, SPL, and the respective locations of the sensor and

reference microphone in the acoustic field (Section 7.2.2.2). Correcting for this phase, θ,

Equation 6–17 may be rewritten as

H(f) =τout

τin

=V (f)

τin

∂τ

∂Ve−jθ. (6–18)

154

Page 155: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

To get a more accurate estimate of the normalized frequency response, the results are com-

puted using the sensitivity magnitude from the rigid termination experiments at Vb = 10 V .

To be consistent with the frequency response measurements, the phase value from the com-

bined sensitivity experiment, with the anechoic termination, is utilized. Figure 6-17 shows

the magnitude and phase of the measured frequency response function (FRF) conducted at

Vb = 10 V for a theoretical shear stress input magnitude of 0.5 Pa. The SPL is approxi-

mately 140 dB± 2 dB over the range of testing frequencies. A single frequency is excited at

a time and the SPL is adjusted to maintain a constant shear stress (theoretical).

0 1000 2000 3000 4000 5000 6000 70000

20

40

60

Frequency (Hz)

|H(f

)| (

τ out/τ

in)

0 1000 2000 3000 4000 5000 6000 7000−150

−100

−50

0

50

Frequency (Hz)

Pha

se(D

eg)

Figure 6-17. Frequency response of sensor at Vb = 10 V using τin = 0.5 Pa as the referencesignal up to the testing limit of 6.7 kHz.

As expected from the design, the FRF shows a clear second-order system response with

a flat band region and a resonance frequency of ≈ 6.2 kHz. The quality factor, Q, is

approximately 79. Assuming a second order system, this translates into a damping ratio,

155

Page 156: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

ζ = 0.0063, which is computed using

Q =1

ζ√

1− ζ2. (6–19)

The phase stays constant at zero and shifts to ≈ 90 at resonance. However, there are certain

aspects of the data, which need closer attention and further investigation. For instance, the

normalized magnitude, |H(f)| is only qualitative since only the shear sensitivity is used for

the estimation while the measurement is a combination of both shear and pressure. Fur-

thermore, beyond resonance the phase starts to flatten out well before reaching 180. Both

these effects may be attributed to the pressure sensitivity of the sensor and its effect on the

combined measurement. Preliminary investigation along those lines and recommendations

to further improve the FRF estimate of the sensor, both during design and characterization

are discussed in Section 7.2.2.2.

6.2.4 Noise Characteristics

The noise of the packaged sensor is measured at three dc bias voltages: Vb = 10 V , 8 V ,

and 5 V . The spectrum analyzer settings for the noise measurement, performed piecewise

over a span of 12.8 kHz, are summarized in Table 6-4.

Table 6-4. Spectrum analyzer settings for noise measurement in a double Fara-day cage.

Parameter Settings

Span (Hz) 100 400 1600 12800

Number of samples/FFTlines

800 800 800 800

Bin width (Hz) 0.125 0.5 2 16

Number of linear spectralaverages

30 240 1200 4800

Channel coupling AC AC AC AC

Windowing function Hanning Hanning Hanning Hanning

Window overlap (%) 75 75 75 75

156

Page 157: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

100

102

104

10−7

10−6

10−5

10−4

10−3

Frequency [Hz]

PS

D [V

rms/r

tHz]

Vb = 10 V

Vb = 8 V

Vb = 5 V

Figure 6-18. Measured output referred noise floor of the packaged sensor in Vrms/√

Hz atdifferent bias voltages.

800 900 1000 1100 1200

0.8

1

1.2

1.4

1.6

1.8

x 10−7

PS

D [V

rms/r

tHz]

Frequency [Hz]

V

b = 10 V

Vb = 8 V

Vb = 5 V

Figure 6-19. Zoomed in plot of the output referred noise floor of the packaged sensor near1 kHz at different bias voltages.

157

Page 158: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

The measured noise power spectral density of the sensor is as shown in Figure 6-18.

The highest measured thermal noise power spectral density is 114 nV/√

Hz at f = 1 kHz

with 1 Hz frequency bin for Vb = 10 V . Figure 6-19 shows a zoomed in plot of the sensor

noise floor around 1 kHz. The measured noise power spectral density increases with the

applied bias voltage. Two potential reasons exist for such a dependence on Vb. First, the

noise contribution from the bias source itself may increase with its output voltage, raising

the sensor noise floor. Second, the nominal capacitances C1 and C2 may change with the

applied bias. At each bias voltage the floating element settles at a new mean position, which

is a result of electromechanical coupling and the initial nominal capacitance mismatch. The

noise at the amplifier input is thus a function of the resulting impedance divider formed

by the sensing capacitors. Table 6-5 shows the integrated noise floor of the sensor over the

entire range of measurement and over specific frequency ranges of interest. This is relevant

for time domain shear stress measurements for real time flow control applications.

Table 6-5. Integrated voltage noise floor of the sensor at differ-ent frequency ranges.

Frequency range Integrated noise floor

15.6 mHz − 12.8 kHz (totalmeasured noise)

1.5 mV

10 Hz − 10 kHz 27.3 µV

100 Hz − 10 kHz 18.2 µV

1 kHz − 10 kHz 10.2 µV

The MDS of the sensor, calculated from Equation 4–5 using the dynamic sensitivity at

Vb = 10 V , is 14.9 µPa (f = 1 kHz @ 1 Hz bin). This corresponds to a mechanical deflection

δ = 0.17 pm and a capacitance change ∆C = 75 zF , which are computed using the design

values for the geometry and assuming parallel plate capacitances. As discussed previously,

the quantitative accuracy of |H(f)| (shear output per unit shear input as a function of

frequency) is questionable due to the combined shear and pressure response. However,

an accurate estimate of |H(f)|, if available, may be used to obtain an MDS spectrum in

158

Page 159: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

µPa/√

Hz. Such a spectrum may then be used with a weighting function, represented by

the shape of the turbulent energy spectrum in a given application, to estimate the integrated

sensor noise floor (MDS) in Pa.

Using the maximum tested linear range of shear stress as the upper limit and the noise

floor at 1 kHz as the lower limit, the dynamic range of the sensor is at least 102 dB. Since

the sensor output is still linear at the maximum available shear stress input, its dynamic

range is expected to be higher than 102 dB. The sensor was designed for τmax = 5 Pa, and

if this design limit were achieved, the dynamic range would be 110.5 dB.

6.2.5 Experimental Uncertainty Estimation

The random uncertainty for the presented measurements is computed for both spec-

tral density and frequency response estimates used in this chapter. For the autospectral

estimates, which are used for computing static, dynamic, and pressure sensitivities, the

normalized random error is computed using [89]

εr =

√V ariance[Gss(f)]

Gss

=1√

ndeff

, (6–20)

where Gss is the estimate of the auto powerspectral density of the sensor output voltage, Gss

is the mean power spectral density from the measurement. The symbol, [∧], stands for the

estimator of a given quantity. The effective number of averages, ndeff, rounded to an integer

value, is given as

ndeff= 1 +

(N − 1)

1− r, (6–21)

where N , is the number of averages during the measurement and r is the percentage over-

lap of the windowing function used. All the measurements in this chapter except noise

measurements are performed with a uniform window, for which r = 0.

159

Page 160: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

The normalized random error for the magnitude,∣∣∣Hms(f)

∣∣∣, and phase, |φms(f)|, of the

frequency response function estimate are [89]

εr||Hms(f)|= |Gms(f)|

Gmm(f)

=

√1− γ2

ms(f)

|γms(f)|√2ndeff

(6–22)

and εr|φms(f)=

√1− γ2

ms(f)

|γms(f)|√2ndeff

. (6–23)

Here Gmm and Gms represent autos and cross power spectral densities, respectively, and γms

represents the coherence function between the sensor and the reference microphone signal.

6.3 Summary

The sensor functionality and performance is demonstrated for both mean and dynamic

shear stress inputs. The results show great promise in the conceptual design of the proof

of concept sensor. At a bias of 10 V , the sensor possesses a linear dynamic sensitivity

of 7.66 mV/Pa up to the testing limit of 1.9 Pa and a noise floor of 114 nV/√

Hz (f =

1 kHz @ 1 Hz bin). This translates into an equivalent MDS of 14.9 µPa (f = 1 kHz @ 1 Hz bin)

and a dynamic range of 102 dB. The sensor has a relatively flat frequency response with

bandwidth defined by its resonance at 6.2 kHz. The sensor functionality is demonstrated

for both static and dynamic shear stress input. The normalized mean and dynamic shear

sensitivities are 0.407 mV/V/Pa (mean shear) and 0.766 mV/V/Pa (dynamic shear), re-

spectively. The sensor has a pressure sensitivity of 4.8µV/Pa which translates to a pressure

rejection of 64 dB. Table 6-6 provides a comparison of the predicted and measured values

for the sensor.

The bandwidth and dynamic range estimates are in close agreement while the sensitivity,

noise floor, and MDS estimates are significantly different. The difference in predicted and

measured sensor performance has several possible explanations. A large parasitic capacitance

exists between the sensor electrodes due to the electrical path through the float zone silicon

substrate, lowering sensitivity. The capacitance mismatch and the noise contribution of the

biasing sources increases the overall noise floor of the system. These two effects are directly

160

Page 161: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Table 6-6. Comparison of predicted and measured sensor performance param-eters.

Parameter Predicted Measured

Normalized SensitivitySoverall (mV/V/Pa)

1.424 0.407 (mean)/0.766(dynamic)

Bandwidth (kHz) 5 6.2

Noise Floor(nVrms/

√Hz

∣∣f=1 kHz @ 1 Hz bin

)20 114 @ Vb = 10 V

MDS(µPa

∣∣f=1 kHz @ 1 Hz bin

)0.562 14.9

Dynamic Range atVb = 10 V andτmax = 1.9 Pa (dB)

107 102

reflected in the higher MDS and lower dynamic range of the sensor. The difference in resonant

frequencies may be due to a different flexural rigidity (EI). The assumed Young’s modulus

of Si may be different from that of the actual wafer. Also, geometric variations during

fabrication may result in a different moment of inertia I. Lastly, the pressure gradient

errors, and forces from complex flow around the sensor may alter the effective force on the

sensor leading to errors (flow different during static and dynamic shear). Recommended

changes to the sensor design and fabrication and an approach for improved characterization

are presented in Chapter 7.

161

Page 162: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

CHAPTER 7CONCLUSIONS AND FUTURE WORK

This chapter summarizes the main research contributions of this work. Recommenda-

tions are provided for future work with suggestions for improvement while highlighting some

of the non-idealities in the current design.

7.1 Conclusions

Direct wall shear stress measurement continues to be of importance for the fundamental

study of turbulence, flow control, and for non-intrusive flow measurement applications. A

measurement standard for wall shear stress and its lack thereof, especially for high Reynold’s

number applications, limits the improvement of existing aerodynamic systems and testing

facilities. This continues to drive the development of micromachined direct wall shear stress

sensors since they are non-intrusive and scale favorably for high speed applications. Micro-

machined floating element sensors have been widely investigated for direct wall shear stress

measurements. The previous research efforts were reviewed in Chapter 2, which highlighted

their capabilities and limitations.

Using some lessons learnt from previous research, three key areas were identified and

focusing on these contributed to the overall improvement in sensor performance. These

key areas, which also give the main contributions of this work, are a simple and novel

sensor structure, an effective and potentially inexpensive fabrication process, and a system

level optimization of the design. The sensor was designed to have an asymmetric comb

finger structure to form capacitors. While achieving a relatively flush surface, this design

allows the use of a single material layer to define the sensor structure and electrical contact

pads. Additional capacitance between the stationary device substrate and both the floating

element and tethers, was also utilized to improve sensitivity. The geometry of the sensor

and a differential sensing scheme help in the attenuation of any output due to out-of-plane

motion resulting from pressure, vibration, etc.

162

Page 163: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

The sensor structure was fabricated on a highly conductive substrate (highly doped Si).

A potentially cost effective technique for mass production of such sensors was developed using

a two mask fabrication process. A seedless electroplating on the highly doped Si helped to

passivate the sensing electrodes with metal to prevent drift arising from charge accumulation

on the sensing electrodes.

In the system level design and optimization, the electrical and mechanical systems and

the interface circuit were included. This resulted in an overall improvement in the predicted

and realized sensor performance to give a high sensitivity and low noise floor. This work

presented the mechanical and electrical modeling, a system level optimization, fabrication,

and characterization of the first differential capacitive sensor to make both mean and dynamic

shear stress measurements with sufficient pressure rejection.

The sensor has the highest dynamic range (> 102 dB) and the lowest noise floor/MDS

(14.9 µPa) reported to date for a micromachined direct shear stress sensor. It is also the first

MEMS sensor to have quantified the fluctuating pressure sensitivity for a floating element

shear stress sensor. The sensor possesses a pressure rejection of 64 dB. The dynamic

characterization revealed a flat frequency response, capturing the resonance of the sensor

structure in both magnitude and phase at ≈ 6.2 kHz, which has never been done before.

Table 7-1 provides a comparison of the sensor performance with some previous work. The

comparison indicates that the present work outperforms the previous best by three orders

of magnitude in MDS and at least an order of magnitude in dynamic range.

Table 7-1. Comparison of measured sensor performance with previous work.

Shear StressSensor

Element Size(mm×mm)

τw Range (Pa) fmax

(kHz)Sensitivity(mV/Pa)

Present Work 2.0× 2.0 1.9−14.9×10−6 6.2 7.66

Zhe [44] 3.2× 3.2 0.16− 0.04 0.531 337

Padmanabhan [50] 0.5× 0.5 10− 1.4× 10−3 16 † 320

Schmidt [37] 0.5× 0.5 13− 0.01† 10 0.47† Theoretically predicted values.

163

Page 164: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

These encouraging results represent a step towards development of instrument grade

direct shear stress sensors. The benefits of the comprehensive design, fabrication, and pack-

aging approach, based on physical insight, are reflected in the demonstrated sensor per-

formance. The ability to make successful measurements with such a sensor also presents

the opportunity to identify some of the non-idealities with the sensor design and in the

characterization. The next section discusses some of these non-idealities.

7.2 Non-Idealities in Sensor Design and Characterization

Despite very encouraging sensor performance results, some issues exist both in the

sensor design and the characterization setup. These need to be addressed for better design

and accurate calibration of the sensor for real world applications. Experimental evidence or

simulations that helped identify some of these issues are presented in this section.

7.2.1 Design Aspects

This section provides a discussion on the non-idealities, which pertain to the sensor

design and may be addressed in subsequent designs. Based on the characterization results,

parasitic capacitance and pressure sensitivity are the two main design aspects that may be

improved upon.

7.2.1.1 Parasitic Capacitance

The first issue is the parasitic capacitance, which results from having a float zone bulk

substrate, instead of an insulator, underneath the device layer. Figure 7-1 shows a schematic

of a single bondpad to explain the sources of parasitic capacitance. During fabrication,

only the parallel plate capacitance, CP , between the bond pad and the stationary substrate

around it was considered. Float zone (FZ) bulk silicon was used in the SOI, assuming that

the electrical path through this silicon would be negligibly small. However, this resulted in

a finite resistive path through the silicon. As a result, the capacitance due to the dielectric

SiO2 between the highly conductive device layer and the weakly conductive bulk substrate

adds to the parasitic capacitance. This capacitance due to the stationary substrate is CoxS

and that from the bond pad is CoxBP .

164

Page 165: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

RFZ RFZ

CoxBP

CoxSCoxS

CP CP

A A

Section View

Top View

Bond Pad

Stationary

Substrate

Bulk

Silicon

Device

Silicon

Oxide

(BOX)

-Vb+Vb

Figure 7-1. Schematic with the top view and section of a sensor bond pad showing theassociated electrical impedances.

The circuit model for the entire sensor, including the additional parasitic impedances,

is shown in Figure 7-2. As an example, a single ended sense capacitance value between

points 1 and 2 is simulated as a function of frequency for substrate resistance, RFZ = 10 KΩ

and RFZ = 1 MΩ. The parallel plate capacitance model is used for computing the various

capacitance values. The substrate and bond pad areas serve as parallel plates separated

by the dielectric BOX thickness. The simulated results are shown in Figure 7-3. The

plot indicates that as the substrate resistance or the frequency increases, the additional

parasitic capacitance from the substrate vanishes and the capacitance reaches a much lower

value (CP ). Thus, by choosing a high resistivity substrate and a high frequency bias signal

the adverse effects of parasitic substrate capacitance may be mitigated, improving overall

sensitivity.

This analysis was based on constant values of CoxS and CoxBP . In reality, these capaci-

tance values are time dependent and for mean shear stress measurements, they may vary

165

Page 166: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

CP CP

CoxS CoxS

CoxBP

RFZRFZ

C2C1

+Vb -Vb

1 2 3

Figure 7-2. Schematic with the top view and section of a sensor bond pad showing theassociated electrical impedances.

102

104

106

108

0

2

4

6

8

10

12

Frequency (Hz)

Cef

f (pF

)

RFZ

=10 KΩ

RFZ

=1 MΩ

Figure 7-3. Effective sense capacitance variation with frequency as a function of substrateresistance.

with the applied sinusoidal bias voltages [90]. These capacitors are similar to metal oxide

semiconductor (MOS) gate capacitors. In the present device, the highly conductive device

layer, the BOX, and the float zone bulk substrate form the MOS capacitors. The capaci-

tance value depends on the depletion layer thickness which varies with the bias voltage value.

A delta-depletion model is used to explain the voltage dependence of the MOS capacitor.

166

Page 167: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Detailed description of this model can be found in [91]. Based on this model, the MOS

capacitance value as a function of the applied bias voltage is given as [91]

C =Cox√

1 + Vbsin(ωt)Vδ

, for π < ωt < 2π. (7–1)

Here, Cox is the substrate capacitance due to the BOX, excluding the depletion width. The

term

Vδ ≡ q

2

Ksx20

K20ε0

NA, (7–2)

where q is the unit charge on an electron, x0 is the BOX thickness, NA is the total number

of acceptors in the p-substrate (FZ), and Ks and Ko are the relative permittivities of silicon

and BOX, respectively. Equation 7–1 indicates that the parasitic MOS capacitance varies

nonlinearly with an applied sinusoidal bias voltage. Thus the attenuation factor Hc (Equa-

tion 3–95), which is a function of the parasitic capacitance, also induces a nonlinear variation

in the sensor sensitivity. Therefore, the overall sensor sensitivity, Sτ,overall, also varies with

an applied fluctuating bias voltage. This may inhibit the use of the sensor for simultaneous

mean and dynamic shear stress measurements with sinusoidal bias voltages. Given all these

different issues, using an insulating substrate instead of float zone silicon will eliminate the

parasitic substrate capacitance (MOS) issues, improving overall sensor performance.

7.2.1.2 Pressure Sensitivity

The output due to out of plane sensitivity of the sensor was mitigated in the design

by adjusting the structural geometry of the design and using the differential capacitance

scheme. The back cavity underneath the sensor, formed by the backside release process, has

an associated compliance/stiffness, Cmecav , due to the compressibility of the fluid in it. In a

lumped sense, the cavity compliance opposes the out of plane motion of the sensor referred to

as the cavity stiffening. If the out of plane compliance of the sensor is Cmeop , the attenuation

167

Page 168: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

of the pressure induced sensor output due to cavity stiffening is

Hcavity =Cmecav

Cmecav + Cmeop

. (7–3)

In addition to this, the back cavity is vented to the atmosphere through the gaps around

the sensor, which are necessary for a released floating element structure. The vent offers a

resistance to fluidic flow across the sensor thickness between the front and back side. The

cavity compliance and vent resistance, Rv, form a high pass filter with a corner frequency of

f =1

2π√

RvCmecav

. (7–4)

Microphone designers commonly design the compliance and the vent resistor to have a low

cut-on frequency. In contrast, in a shear stress sensor, these should be designed to move this

cut-on frequency beyond the operating bandwidth for shear stress. This modeling aspect has

not been explored previously and could greatly benefit floating element shear stress sensor

design.

7.2.2 Characterization Aspects

The characterization of the sensor revealed some interesting results which need fur-

ther investigation for better comparison with results from the predictive tools developed in

this dissertation. This section concentrates on three different aspects of characterization:

impedance measurements, dynamic sensitivity, pressure frequency response, and errors from

pressure gradient effects.

7.2.2.1 Impedance Measurement Drift

As mentioned in Section 6.2.1, an accurate estimate of the difference between C10 and

C20 was not achieved due to drift associated with the probe station. Figure 7-4 shows the

drift in the measurement of one of the sensor capacitors. The plot shows the first and the last

measurement sweep out of the 31 sweeps of the dc bias voltage for ensemble averaging. The

trend in the the drift is also indicated in Figure 7-5, which is a plot of the mean capacitance

168

Page 169: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

for each dc bias sweep vs the sweep/measurement number N . The figure indicates that

initially, the drift in mean capacitance is high but tends to decrease as time progresses.

0 5 10 15 2014.79

14.795

14.8

14.805

14.81

14.815

14.82

Bias (V)

C10

(pF

)

FirstLast

Figure 7-4. Capacitance measurement drift indicated via the first and last measurement priorto ensemble averaging.

However, longer measurements indicate that the drift continues over time. The drift is about

5 fF over the course of 31 bias voltage sweeps, which is roughly 22% of the predicted full scale

capacitance change. Such a high drift is therefore a limitation in effectively quantifying the

nominal capacitance and the change in capacitance as a function of bias voltage. Potential

causes of drift in the probe station include, charge build up on the probes and poor contact

of the probe tips with the sensor bond pads. A lack of a standard substrate was also a

limitation in the complete understanding of the nature and source of the drift.

7.2.2.2 Dynamic Shear Measurements

The dynamic shear stress measurements on the sensor were performed with two differ-

ent kinds of terminations at the end of the PWT, the anechoic termination and the rigid

termination. Note that, for the anechoic termination the reference microphone is at the same

axial location as the sensor, whereas for the rigid termination, it is located at the termination

(Section 6.1.3). The dynamic sensitivity was determined based on the autospectral estimates

169

Page 170: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

0 10 20 30 4014.802

14.804

14.806

14.808

14.81

Measurement Number (N)

C10

mea

n (pF

)

Figure 7-5. Drift in mean capacitance with subsequent dc bias sweeps.

of the microphone and sensor signals while maintaining high coherence values between these

signals. A constant sensitivity implies that the transfer function, Hms, between the sensor

and the microphone signal should be constant. However, contrary to expectation, Hms, varies

with the SPL. Figure 7-6 and Figure 7-7 show the transfer function as a function of SPL for

the calibrations with the rigid and anechoic terminations, respectively, at Vb = 10 V .

The maximum variation in the magnitude, |Hms|, over the range of SPLs is ≈ 221 nV

for the rigid termination while it is ≈ 2.99 µV for the anechoic termination. Similarly, the

maximum variation in the phase, φms, is ≈ 2.5 for a rigid termination and it is ≈ 11

for the anechoic termination. Since the same microphone was used for the reference SPL

measurements, this variation is potentially a function of the termination at the end of the

tube and the SPL at the sensor location. In contrast to the anechoic termination case, for the

rigid termination, the SPL is ≈ 40 dB lower than the peak SPL at the reference microphone.

This intuitively suggests that local SPL at the sensor location significantly affects the sensor

output. While the high SPL variation may be attributed to the acoustic nonlinearity, the

low SPL behavior needs further investigation.

170

Page 171: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

100

102

104

106

6.9

7

7.1

7.2x 10

−6

|Hm

s|(V

/Pa)

Rigid

100

102

104

106

3.9

4

4.1

4.2x 10

−5

Pressure (Pa)

|Hm

s|(V

/Pa)

Anechoic

Figure 7-6. Magnitude of sensor to microphone transfer function as a function of acousticpressure.

100

102

104

106

−24

−23

−22

−21

φm

s(D

eg)

Rigid

100

102

104

106

70

75

80

Pressure (Pa)

φm

s(D

eg)

Anechoic

Figure 7-7. Phase of sensor to microphone transfer function as a function of acoustic pressure.

171

Page 172: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Clearly, the comparison indicates that, using a rigid termination with the sensor placed

at an odd multiple of the quarter wavelength (nλ/4 where n is odd) is a better technique to

measure the dynamic behavior of the sensor. Therefore, for future dynamic measurements

for both sensitivity and frequency response, it is recommended to use a rigid termination

with variable location of the rigid plane for tuning. This will ensure that the sensor may

always be located at a pressure minima and a velocity maxima.

7.2.2.3 Dynamic Pressure Measurements

The sensor demonstrated a high pressure rejection, which has never been quantified in

the MEMS shear stress literature. The study of the pressure frequency response is also im-

portant for quantitative shear measurements, since the physical phenomenon usually consists

of shear and pressure inputs. Similar to the dynamic shear measurement studies, the trans-

fer function between the sensor and the reference microphone was investigated for variation

with the input SPL. Figure 7-8 shows the variation of the transfer function with the input

SPL at Vb = 10 V .

100

102

104

106

3

4

5

x 10−6

|Hm

sp

ressu

re

|(V

/Pa)

100

102

104

106

−150

−140

−130

−120

−110

Pressure (Pa)

|φm

sp

ressu

re

|

Figure 7-8. Transfer function between the sensor and the reference microphone for normalacoustic incidence.

172

Page 173: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

This exercise gives results similar to dynamic shear measurements, where both magni-

tude and phase of the transfer function vary with SPL. This requires further investigation

from both the sensor design and acoustics point of view for a better understanding of these

results.

1000 2000 3000 4000 5000 600010

−6

10−5

10−4

|Hm

s|(V

/Pa)

Shear+pressurepressure

1000 2000 3000 4000 5000 6000

−50

0

50

100

Frequency (Hz)

φm

s(D

eg)

Shear+pressurepressure

Figure 7-9. Comparison of combined (shear and pressure) and pressure transfer functionsmeasured between sensor and reference microphone.

Next, the combined pressure and shear transfer function measured with the anechoic

termination is compared with the pressure transfer function measured using normal acoustic

incidence shown in Figure 7-9. A higher, combined pressure-shear transfer function indicates

that despite similar and higher pressure forces, the sensor remains more sensitive to shear in

the frequency range of measurement than it is to pure pressure. The scatter in the pressure

data is due to the poor performance of the sensor for pressure inputs. Note, that the sensor

was not designed for pressure measurements and relies purely on the capacitance mismatch

(ideally zero) for this measurement. Similar qualitative nature of the two magnitude re-

sponses may be due to the shaping of the overall transfer function by the pressure response

173

Page 174: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

of the sensor. Also, for out-of-plane motion of the sensor, close to its in-plane resonance fre-

quency, a small perturbation in the in-plane direction may be sufficient to excite its in-plane

resonant motion. An FEA simulation is performed to ensure that the pressure and shear

resonances of mechanical structure are not in close proximity. The results reveal that the

first mode, which is the in-plane mode, is at 5.2 kHz and the second mode, which is the

out-of-plane mode, is at 14.4 kHz. These values are to within 4% of the LEM predictions.

Thus, for better understanding of the sensor performance a LEM with combined shear and

pressure inputs is necessary, including the effects of cavity stiffening and the cut-on frequency

for pressure.

7.2.2.4 Mean Shear Stress Measurements

The mean shear stress measurements showed sensitivities that were considerably lower

than the dynamic sensitivity estimates. This may be attributed to the nonlinear variation in

the parasitic substrate capacitance discussed previously. Pressure gradient effects may also

contribute to the errors in this measurement. A theoretical estimate of the effective shear

stress is [37]

τeff = τw

(1 +

2Tt

h+

g

h

), (7–5)

where g is the gap underneath the sensor. The additional terms are due to the force acting

on the side of the sensor and the shearing force underneath the sensor structure. These are

simplified estimates and may be different for each sensor and measurement setup. An attempt

was made to quantify these experimentally. For this the mean shear stress measurements are

repeated at Vb = ±2.5 V , with a different channel height, h (Figure 7-10). The measurement

is done twice for h = 0.5 mm and compared with the results for h = 1 mm. The respective

sensitivities are 0.932 mV/Pa and 0.870 mV/Pa for h = 0.5 mm and 0.879 mV/Pa for

h = 1 mm. The variability of estimated sensitivities for h = 0.5 mm is higher than the

difference between sensitivities for h = 1 mm and h = 0.5 mm. These measurements are

therefore not enough to quantify the errors due to pressure gradient effects. This variability

174

Page 175: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

may also be due to drift in the biasing circuitry due to charging of the variable capacitor on

the PCB (Figure 6-4). This problem also needs to be addressed from the circuits perspective.

0.5 1 1.5 2 2.50.5

1

1.5

2

2.5

3

Shear Stress (Pa)

Vol

tage

(m

V)

h=1 mmh=0.5 mmh=0.5 mm

Figure 7-10. Comparison mean shear stress measurements at Vb = 2.5 V at different channelwidths, h, in the flow cell.

7.3 Recommendations for Future Sensor Designs

Future designs of the current sensor may be significantly improved based on experiences

drawn from the design, fabrication and characterization aspects of this work. Several sugges-

tions were included in Section 7.2 from both design and characterization stand points. This

section combines them and recommends an approach for the next generation of the direct

differential capacitive shear stress sensors.

For the sensor design, the pressure response needs to be accounted for in the model.

This will include the effects of the cavity compliance and the damping/vent resistance to vent

the cavity to the ambient pressure. This will considerably improve the pressure sensitivity

issues. A simple LEM for the sensor for pressure forcing is represented in Figure 7-11,

with the various mechanical lumped elements in pressure are indicated with the subscript

175

Page 176: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

‘p’. For the design optimization, the maximum pressure force may be set to be orders

fpressure

ue

Rmep

Mmep

Cmep fp

+

-

Rvent

Ccav

Transduction

Figure 7-11. Mechanical LEM for pressure sensitivity of the sensor.

of magnitude higher than the shear forcing. This may yield some interesting results. The

design with combination of cavity stiffening and cut-on frequency may enable thinner tethers

improving in-plane sensitivity while resulting in sufficient pressure rejection. However, the

trade-off would be lower nominal capacitance. From the circuits standpoint, the optimization

may include the current noise of the amplifier and the thermal noise of the bias resistor.

This allows to understand the trade-off involved in using different amplifiers instead of the

SiSonic. Furthermore, the demodulation and biasing circuits may be included as part of

the optimization to develop a predictive tool for the overall system performance, instead of

piecewise analysis.

The sensor may be fabricated by wafer bonding a silicon substrate to a pyrex wafer.

This will eliminate the parasitic substrate capacitance issues altogether. Using a thin-back

of the floating element prior to bonding to the pyrex may enable higher bandwidths by

lowering the mechanical mass. Backside electrical contacts using electronic through wafer

176

Page 177: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

internconnects (ETWI) in the pyrex will help eliminate wire bonds, which disturb the flow.

A compliant, dielectric, hydrophobic layer on the floating element will eliminate errors due

to forces on the side of the floating element and forces due to flow underneath the sensor

structure.

The packaging used in this dissertation was developed based on packaging and testing

capabilities available in a laboratory setup. However, the package may be significantly

improved using shielded cables, which can carry multiple signals, providing power supply to

the amplifier, biasing signals for the sensor, and the sensor output voltage. This will further

lower the size of the PCB providing a more compact package for using them in shear stress

sensor arrays.

177

Page 178: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

APPENDIX AMECHANICAL ANALYSIS

In this appendix the mechanical analysis for the floating element structure is presented.

First, the expression for linear deflection is derived from the governing differential equations.

Next, the non-linear deflection of the sensor structure is presented using two different solution

techniques, the energy method and the analytical technique. Lastly, the expressions for

lumped mass and compliance are derived based on the principle of conservation of energy.

A.1 Beam Deflection

A.1.1 Governing Equation

The governing differential equation is derived before explaining the solution procedure.

The curvature of a beam is first determined to formulate the governing equations. Consider

an arbitrary curve as shown in Figure A-1.

x

y

φs

Figure A-1. A deflected beam with arbitrary curvature.

The length of the arc for such a curve is

L =n∑

i=1

√(x− xi)

2 + (y − yi)2

=n∑

i=1

√(∆x)2 + (∆y)2

=n∑

i=1

√1 +

(∆y∆x

)2(∆x)

=b∫

a

√1 +

(dydx

)2dx.

(A–1)

178

Page 179: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

The curvature, κ or1

ρfor a curve is defined as,

1

ρ= κ =

∣∣∣∣dϕ

ds

∣∣∣∣ . (A–2)

Considering a positive curvature Equation A–2 is rewritten as

1

ρ= κ =

ds. (A–3)

Using the definition of the slope, the angle, ϕ is

ϕ = tan−1

(dy

dx

). (A–4)

Using Equation A–1 the differential length, ds along the curve is

ds =

√1 +

(dy

dx

)2

dx. (A–5)

The curvature is rewritten in terms of the variation along the x-axis as

1

ρ= κ =

ds=

dx

dx

ds. (A–6)

Substituting Equation A–5 and A–4 into Equation A–6 results in

1

ρ= κ =

ds=

d

dx

(tan−1

(dy

dx

)) 1√

1 +(

dydx

)2

=d2y/dx2

(1 +

(dydx

)2)3/2

.

(A–7)

Now consider a beam under pure bending. The curvature of such a beam is [92]

1

ρ=

M

EI, (A–8)

179

Page 180: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

where M is the resisting moment, E is the Young’s modulus of elasticity, and I moment of

inertia of the beam. Combining Equation A–7 and A–8 gives

Mx

EI=

d2w (x)/dx2

(1 +

(dw(x)

dx

)2)3/2

, (A–9)

where Mx is the moment along the length of the beam, and w (instead of y) is the deflection

in the transverse direction due to the applied load. Equation A–9 is the governing equation

for the deflection of a beam due to pure bending. For both small and large deflections, the

slope dw(x)dx

¿ 1. Therefore, the simplified governing equation is

Mx

EI= d2w (x)

/dx2. (A–10)

Direction

of Flow

Le

Clamped

Boundary

MB

RA

P Q

2Lt

P Q

RA

V

Wt

Q

xx=0

Tt

xx=0

x=Lt

MA

MA Mx

RB

x

z

y

Rigid body motion

P

Figure A-2. Simplified mechanical model of the floating element structure.

180

Page 181: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Consider a beam as shown in Figure A-2. This is same as Figure 3-2, which is repeated

here for convenience. The assumptions for beam deflection were previously stated in Sec-

tion 3.1.1 and therefore not repeated here. In the simplified mechanical model, for the shear

stress sensor, each pair of tethers form a clamped-clamped beam, with a point load P applied

at center by the shear stress, τw, on the floating element. A uniformly distributed load, Q,

along the length of the tethers accounts for the shear force on the tethers. Two such beams

share the load applied on the floating element because of the geometric symmetry as shown

in Figure A-2. Thus, the shear force each tether pair that form beams is

P =τwWeLe

2+

τw (NWfLf )

2, (A–11)

and Q = τwWt. (A–12)

(A–13)

Due to symmetry the solution is obtained by solving for half the beam length. The boundary

conditions for a clamped-clamped beam are mathematically stated as

w(x = 0) = 0, (no deflection)

dw

dx

∣∣∣∣x=0

= 0, (zero slope)

anddw

dx

∣∣∣∣x=Lt

= 0. (symmetry) (A–14)

A.1.2 Small Deflection Analysis

The Euler Bernoulli theory and small beam deflections in response to an applied force are

assumed while solving Equation 3–4. The theory hypothesizes that a straight line transverse

to the neutral axis remains straight, inextensible, and normal before and after deformation

[65]. With these assumptions, the linear solution for small beam deflection is obtained.

181

Page 182: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

The reaction forces, RA and RB are determined first in terms of the applied loads, and

are given as

RA = RB =P

2+ QLt. (A–15)

The resisting moment Mx is

Mx = −MA + RAx− Qx2

2. (A–16)

Substituting Equation A–16 in A–10 gives

d2w (x)

dx2=

1

EI

(−MA + RAx− Qx2

2

). (A–17)

Solving the above equation gives the deflection and its for the beam as

dw (x)

dx=

1

EI

(−MAx + RA

x2

2− Qx3

6

)+ c1, (A–18)

and w (x) =1

EI

(−MA

x2

2+ RA

x3

6− Qx4

24

)+ c1x + c2. (A–19)

Now, the boundary conditions are applied to solve for the constants as follows:

w (0) =c2 = 0, (A–20)

dw (0)

dx=c1 = 0, (A–21)

dw (Lt)

dx=

1

EI

(−MALt + RA

L2t

2− QL3

t

6

)= 0. (A–22)

Simplifying Equation A–22 results in

MA =RALt

2− QL2

t

6. (A–23)

Substituting Equation A–15 into A–23 and simplifying it gives

MA =PLt

4+

QL2t

3. (A–24)

182

Page 183: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

The deflection in terms of the applied forces is obtained by substituting Equation A–15,

A–20, A–21, and A–23 into Equation A–19 and given as

w (x) =1

EI

(−

(PLt

4+

QL2t

3

)x2

2+

(P

2+ QLt

)x3

6− Qx4

24

). (A–25)

The above equation is rewritten in terms of the applied shear stress. Substituting Equa-

tion A–11 and A–12, the simplified expression for the deflection is

w (x) = − τw

EI

((3 (LeWeLt + NWfLfLt) + 8WtL

2t ) x2

48

−(LeWe + NWfLf + 4WtLt) x3

24+

Wtx4

24

).

(A–26)

Using the expression for moment of inertia,

I =TtW

3t

12, (A–27)

Equation A–26 is further simplified using Equation A–27 to give

w (x) = − τw

4ETtW 3t

(3 (LeWeLt + NWfLfLt) + 8WtL2t ) x2

− (2LeWe + 2NWfLf + 8WtLt) x3 + 2Wtx4

. (A–28)

The -ve sign indicates that the deflection is downwards as per the chosen sign convention.

The maximum deflection occurs at the centre i.e., at x = Lt. The maximum deflection is

expressed as,

δ = −w (Lt) =1

ETtW 3t

(τwLeWeL

3t

4+

τwNWfLfL3t

4+

τwWtL4t

2

). (A–29)

A.1.3 Large Deflection Analysis

Large beam deflection causes extensional strain along the neutral axis, previously ig-

nored in the small deflection analysis. Some nonlinear terms in the Green Lagrange strain

displacement relationships, are not negligible for large deformations. Detailed solutions using

two different techniques, the energy method (Section 3.1.2.1) and an analytical solution tech-

nique (Section 3.1.2.2) are illustrated. First, an expression for the total strain in the beam

183

Page 184: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

due to both bending and stretching is derived. The total strain in the beam is expressed as

εt = εbending + εstetching. (A–30)

The bending and stretching strains are [65],

εbending = ε1xx = −z

∂2w

∂x2(A–31)

and εstretching = ε0xx =

∂u

∂x+

1

2

(∂w

∂x

)2

. (A–32)

The total change in length of the beam due to stretching is

δLts =

2Lt∫

0

ε0xxdx

=

2Lt∫

0

(∂u

∂x+

1

2

(∂w

∂x

)2)

dx

= [u (x)]2Lt

0︸ ︷︷ ︸0

(clamped boundary)

+1

2

2Lt∫

0

(∂w

∂x

)2

dx

=1

2

2Lt∫

0

(∂w

∂x

)2

dx.

(A–33)

Thus the strain in terms of the transverse deflection due to stretching of the beam is

ε0xx =

δLts

2Lt

=1

4Lt

2Lt∫

0

(∂w

∂x

)2

dx. (A–34)

Combining Equation A–30, A–31, and A–34, the total strain in the beam is rewritten as

εt = −z∂2w

∂x2+

1

4Lt

2Lt∫

0

(∂w

∂x

)2

dx. (A–35)

184

Page 185: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

A.1.3.1 Energy Method

The energy method solution utilizes the principle of virtual work to arrive at a solution

using the Rayleigh-Ritz method [65]. The principle is represented as

U = Stored energy −Work done, (A–36)

where U is the total energy.

Stored energy. The energy stored in any elastic structure due to displacement is

W (q1) =

q1∫

0

e (q) dq, (A–37)

where e represents effort (Force, Pressure etc.) and f represents a flow variable (displace-

ment, velocity etc.). For a distributed system energy density is a parameter of interest rather

than the energy i.e.,

W = W∆∀. (A–38)

Now in terms of units

W =J

m3=

N −m

m3=

N

m2

m

m, (A–39)

which indicates the energy density may be represented as the product of stress and strain.

Thus, energy density for normal stress may be written as,

W =

ε∫

0

σ (ε) dε, (A–40)

where σ is the stress and ε is the strain. As long as the deflection is within the linear elastic

limit of the material, Equation A–40 may be rewritten as

Wnormal =

ε∫

0

Eεdε =1

2Eε2 =

1

2σε. (A–41)

185

Page 186: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Similarly, the energy density due to shear stress may be written as

Wshear =1

2τγ. (A–42)

The total energy is thus

W =

∫∫∫(Wnormal + Wshear)dxdydz. (A–43)

A consequence of Euler Bernoulli beam theory is, there is no shear in the beam. Thus

considering only the normal energy density and substituting Equation A–41 into A–43 gives,

W =

Tt∫

0

Wt/2∫

−Wt/2

2Lt∫

0

1

2σεtdxdydz. (A–44)

Work Done. The work done by the external forces acting on the beam is

Work = Pδ + QLtδ = (P + QLt) δ. (A–45)

Substituting Equation A–44 and A–45 into Equation A–36 results in

U =

Tt∫

0

Wt/2∫

−Wt/2

2Lt∫

0

1

2Eε2

t dxdydz − (P + QLt) δ. (A–46)

Substituting for P and Q from Equation A–11 and A–12, the above equation is rewritten as

U =

Tt∫

0

Wt/2∫

−Wt/2

2Lt∫

0

1

2Eε2

t dxdydz

︸ ︷︷ ︸I1

−(τw

2(WeLe + NWfLf ) + τwWtLt

︸ ︷︷ ︸I2

. (A–47)

An estimate of the strain εt is needed to evaluate the integral I1 in Equation A–47. In the

solution procedure using the Rayleigh-Ritz method, a trial function for the deflection is used

to evaluate I1, which is

w =δ

2

(1 + cos

(π (Lt − x)

Lt

)). (A–48)

186

Page 187: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Therefore,

dw

dx=

δπ

2Lt

(sin

(π (Lt − x)

Lt

)), (A–49)

andd2w

dx2= − δπ2

2L2t

(cos

(π (Lt − x)

Lt

)). (A–50)

The strain ε is computed by combining Equation A–35, A–49, and A–50 to give

εt = z

(δπ2

2L2t

(cos

(π (Lt − x)

Lt

)))+

1

4Lt

2Lt∫

0

2

π

Lt

(sin

(π (Lt − x)

Lt

)))2

dx

=

(zδπ2

2L2t

(cos

(π (Lt − x)

Lt

)))+

δ2π2

16L3t

2Lt∫

0

(1− cos

(2π(Lt−x)

Lt

))

2dx

=

(zδπ2

2L2t

(cos

(π (Lt − x)

Lt

)))+

δ2π2

32L3t

[x +

Lt

2πsin

(2π (Lt − x)

Lt

)]2Lt

0

= zδπ2

2L2t

cos

(π (Lt − x)

Lt

)+

δ2π2

16L2t

.

(A–51)

The integral I1 in Equation A–47 is evaluated, using the expression for ε from Equation A–51

as follows,

I1 =1

2

Tt∫

0

Wt2∫

−Wt2

2Lt∫

0

E

(zδπ2

2L2t

cos

(π (Lt − x)

Lt

)+

δ2π2

16L2t

)2

dxdydz

=Tt

2

Wt2∫

−Wt2

2Lt∫

0

E

(zδπ2

2L2t

cos

(π (Lt − x)

Lt

)+

δ2π2

16L2t

)2

dxdz

=ETt

2

Wt2∫

−Wt2

(z2δ2π4

8L4t

[2Lt] +δ4pi4

256L4t

[2Lt]

)dz

= ETt

(δ2π4W 3

t

96L3t

+δ4π4Wt

256L3t

).

(A–52)

187

Page 188: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Substituting Equation A–52 into A–47, the total potential energy of the system is expressed

as

U = ETt

(δ2π4W 3

t

96L3t

+δ4π4Wt

256L3t

)

︸ ︷︷ ︸I1

−(τw

2(WeLe + NWfLf ) + τwWtLt

︸ ︷︷ ︸I2

. (A–53)

According to the principle of virtual work,“the total potential energy is stationary with

respect to any virtual displacement” [42]. This implies that the change in the total potential

energy due to the change in any virtual parameter should be minimum (ideally zero) to

maintain the stationary conditions of U i.e.,

dU

dδ= 0. (A–54)

Substituting Equation A–53 in the above equation and knowing that π4

96≈ 1 and π4

128≈ 3

4

results in

δ

(1 +

3

4

Wt

)2)

=τwWeLe

4ETt

(1 +

NWfLf

WeLe

+ 2WtLt

WeLe

)(L3

t

W 3t

). (A–55)

Equation A–55 gives the expression for the deflection using energy method under large

deflection. The accuracy of the solution is however limited by the choice of trial function.

A.1.3.2 Analytical Method

As discussed already under large deflections for a beam in bending, there is extensional

strain along the length of the beam. An axial force, Fa responsible for this strain is computed

in this method. The solution presented here is not a closed form analytical solution and

requires an iterative procedure to solve for the deflection of the beam. The governing equation

is same as Equation A–9 and still retains the assumption d2wdx2 ¿ 1. The schematic to the

right in Figure A-3 represents the free body diagram of the body of one half of the clamped-

clamped beam shown in the figure to the left.

Balancing moments at a given section at a distance x from the free end gives,

M (x) =P

2(x) +

Qx2

2+ Fa (w (x)− w (0))−M0, (A–56)

188

Page 189: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Fa

P/2Q

Lt

M0

Fa

M0

P/2M(x) Q

x

Figure A-3. Schematic of one half of a clamped-clamped beam under large deflection.

where M0 is the resisting moment at x = 0. The governing equation for this method obtained

by substituting Equation A–56 into A–9, resulting in

EId2w (x)

dx2− Faw (x) =

P

2(x) +

Qx2

2− Faw (0)−M0,

d2w (x)

dx2−

(Fa

EI

)

︸ ︷︷ ︸λ2

w (x) =Q

2EI︸︷︷︸X

x2 +P

2EI︸︷︷︸Y

x− (Faw (0) + M0)

EI︸ ︷︷ ︸Z

, (A–57)

ord2w (x)

dx2− λ2w (x) = Xx2 + Y x− Z. (A–58)

To solve Equation A–57, the axial force, Fa, due to the large deflection is determined.

The axial strain under large deflection is same as Equation A–34 with limits changed for one

half of the beam given as

ε0xx =

δLts

Lt

=1

2Lt

Lt∫

0

(∂w

∂x

)2

dx. (A–59)

For the tether the longitudinal stress is expressed as

σxx =Fa

Area= Eε0

xx, (A–60)

where the area of the tether, Area = TtWt. Therefore, the axial force,

Fa = Eε0xxWtTt. (A–61)

Combining Equation A–61 and Equation A–59 gives

Fa = EWtTt

2Lt

Lt∫

0

(∂w

∂x

)2

dx. (A–62)

189

Page 190: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Using this estimate, Equation A–57 is solved, which has a homogenous and a particular

solution written as

w = wg + wp. (A–63)

The homogeneous differential equation is

d2wg (x)

dx2− λ2wg (x) = 0, (A–64)

which has a solution given by

wg (x) = C1 cosh (λx) + C2 sinh (λx) . (A–65)

The inhomogeneous equation to obtain particular solution is

d2wp (x)

dx2− λ2wp (x) = Xx2 + Y x− Z. (A–66)

Assume a particular solution of the form

wp (x) = ax2 + bx + c. (A–67)

Substituting Equation A–67 into A–66 gives

2a− λ2(ax2 + bx + c

)= Xx2 + Y x− Z. (A–68)

Matching coefficients of different powers of x,

a = −X

λ2,

b = − Y

λ2, (A–69)

and c =λ2Z − 2X

λ4(A–70)

Substituting coefficients from Equation A–69 into Equation A–67 gives

wp (x) = −X

λ2x2 − Y

λ2x +

λ2Z − 2X

λ4. (A–71)

190

Page 191: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

The total deflection obtained by combining Equation A–63, A–65, and A–71 is

w (x) = C1 cosh (λx) + C2 sinh (λx)− X

λ2x2 − Y

λ2x +

λ2Z − 2X

λ4, (A–72)

and the deflection gradient is

dw (x)

dx= C1λ sinh (λx) + C2λ cosh (λx)− 2X

λ2x− Y

λ2. (A–73)

Applying the boundary conditions in Equation A–14 gives,

dw (0)

dx= C2λ− Y

λ2= 0 ⇒ C2 =

Y

λ3, (A–74)

anddw (Lt)

dx= C1λ sinh (λLt) +

Y

λ2cosh (λLt)− 2X

λ2Lt − Y

λ2= 0

⇒ C1 =Y + 2XLt − Y cosh (λLt)

λ3 sinh (λLt). (A–75)

Substituting Equation A–74 and A–75 into Equation A–72 the deflection is rewritten as

w (x) =

[Y + 2XLt − Y cosh (λLt)

λ3 sinh (λLt)cosh (λx)

+Y

λ3sinh (λx)− X

λ2x2 − Y

λ2x +

λ2Z − 2X

λ4

].

(A–76)

Now, using values of X, Y , Z, and λ from Equation A–57 the above equation becomes,

w (x) =

[1

Faλ sinh (λLt)

(QLt +

P

2− P

2cosh (λLt)

)cosh (λx)

+P

2Faλsinh (λx)− Q

2Fa

x2 − P

2Fa

x + w (0) +M0

Fa

− Q

λ2Fa

].

(A–77)

Recognizing that w (x = 0) = w (0), the resisting moment M0 is determined and is expressed

as

M0 =Q

λ2− 1

λ sinh (λLt)

(QLt +

P

2− P

2cosh (λLt)

). (A–78)

191

Page 192: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Substituting Equation A–78 into A–79 gives,

w (x) =

[P

2Faλsinh (λx) +

(cosh (λx)− 1)

Faλ sinh (λLt)

(QLt +

P

2− P

2cosh (λLt)

)

− Q

2Fa

x2 − P

2Fa

x + w (0)

].

(A–79)

The last boundary condition i.e., w(Lt) = 0 results in

w (0) =

[− P

2Faλsinh (λLt)− (cosh (λLt)− 1)

Faλ sinh (λLt)

(QLt +

P

2− P

2cosh (λLt)

)

+Q

2Fa

L2t +

P

2Fa

Lt

].

(A–80)

Equation A–80 is the expression for deflection of the beam under large deflections, which

includes the effect of strain in the neutral axis. The gradient of the deflection is written as

dw

dx=

[P

2Fa

cosh (λx) +sinh (λx)

Fa sinh (λLt)

(QLt +

P

2− P

2cosh (λLt)

)

− Q

Fa

x− P

2Fa

].

(A–81)

Using the expressions derived so far the following steps are followed to determine the deflec-

tion in an iterative method.

1. Start with an initial value of axial force Fa

2. Obtain the eigen value given by λ =√

Fa

EI

3. Estimate the deflection w (x) and the gradient dwdx

4. Calculate Fa again using the relation Fa = E WtTt

2Lt

Lt∫0

(∂w∂x

)2dx

5. Repeat step 1 to 4 until |F n+1a − F n

a |/F n+1a ≤ 1e− 8

6. Use Fa to obtain the maximum center deflection w (0)

A.2 Lumped Parameters

The basic concepts of lumped element modeling were explained in Section 3.2 in Chap-

ter 3. In this section, the derivations for lumped mass and lumped compliance of the floating

element sensor structure are presented.

192

Page 193: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

A.2.1 Lumped Mass

The lumped mass is computed by using the total kinetic energy due to the motion of

the beam and equating it to the equivalent kinetic energy of the lumped system. For a linear

system,

W ∗KE =

∫dW ∗

KE =

f0∫

0

pdf =1

2Mlumpedf

20 , (A–82)

where, p and f represent momentum and flow, respectively. The expression for linear de-

flection in Equation A–28 is used. The velocity or flow variable represented in terms of the

deflection is

v (x) = jωw (x) . (A–83)

The velocity at x = Lt is

v (Lt) = jωw (Lt) . (A–84)

Combining Equation A–83 and A–84,

v (x) =v (Lt)

w (Lt)w (x) . (A–85)

Consider a differential element of the tether of mass dm at a location x. The momentum of

the element is then given by,

p = dmv (x) = ρTtWtv (x) dx, (A–86)

where ρ is the mass density of the beam. By definition of co-energy, the differential kinetic

co-energy is expressed as

dW ∗KE = ρTtWtdx︸ ︷︷ ︸

dm

v (x) dv (x) . (A–87)

193

Page 194: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Therefore,

W ∗KE =

∫dW ∗

KE

W ∗KE =

Lt∫

0

v(x)∫

0

ρNiTtWtv (x) dv (x)dx

W ∗KE =

Lt∫

0

ρNiTtWtv2 (x)

2dx.

(A–88)

Combining Equation A–88 and A–85 gives

W ∗KE =

Lt∫

0

ρNiTtWtv2 (Lt)

w2 (Lt)

w2 (x)

2dx

=ρNiTtWt

2

v2 (Lt)

w2 (Lt)

Lt∫

0

w2 (x) dx.

(A–89)

In the above computation, the distributed tether mass is lumped at the center with a dis-

placement equal to the tip deflection. Combining, Equation A–82 and A–89 gives

Mtme = 4ρNiTtWt

w2 (Lt)

Lt∫

0

w2 (x) dx. (A–90)

Note the additional factor 4 in Equation A–90, which accounts for mass of four tethers

suspending the floating element. The total lumped mass of the entire floating element is the

sum of the tether and the floating element mass given as

Mme = Mtme + Melement = Mtme + ρNi (WeLe + NWfLf ) Tt. (A–91)

194

Page 195: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Combining Equation A–28, A–29, A–90, and A–91 and evaluating the integral results in

(solve in MAPLE)

Mme = ρTtLeWe

1 + 19235

(WtLt

LeWe

)+ 3

(NWf Lf

LeWe

)+ 384

35

(WtLt

LeWe

) (NWf Lf

LeWe

)

+1072105

(WtLt

LeWe

)2

+ 3(

NWf Lf

LeWe

)2

+ 1072105

(NWf Lf

LeWe

)(WtLt

LeWe

)2

+3(

NWf Lf

LeWe

)2 (WtLt

LeWe

)+ 2048

315

(WtLt

LeWe

)3

+(

NWf Lf

LeWe

)3

(1 +

NWf Lf

LeWe+ 2 WtLt

LeWe

)2 . (A–92)

A.2.2 Lumped Compliance

The lumped compliance is evaluated by computing the total potential energy of the

system and equating it to the equivalent potential energy of the lumped system. For a linear

system,

WPE =

∫dW ∗

PE =

q0∫

0

edq =1

2

w2 (Lt)

Clumped

, (A–93)

where e and q are effort and displacement, respectively. Consider an element of the beam of

length dx. The total distributed force acting on a half section of the beam is

F =

Lt∫

0

(Q +

P

2δ (x− Lt)

)dx, (A–94)

where P and Q are defined in Equation A–11 and A–12 and δ represents a dirac delta

function. Thus the potential energy in Equation A–93 is rewritten as

WPE =

Lt∫

0

w(x)∫

0

(Q +

P

2δ (x− Lt)

)dw (x) dx. (A–95)

Substituting Equation A–11 and A–12 into A–95 gives the potential energy expressed in

terms of the shear stress, τw, to give

WPE =

Lt∫

0

w(x)∫

0

τw

(Wt +

(WeLe

4+

NWfLf

4

)δ (x− Lt)

)dw (x) dx. (A–96)

195

Page 196: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Equation A–96 requires the forcing shear stress to be expressed in terms of the distributed

system to solve for the total potential energy. Substituting for τw from Equation A–28,

Equation A–96 is expressed as

WPE =

Lt∫

0

w(x)∫

0

4ETtW3t

(Wt +

(WeLe

4+

NWf Lf

4

)δ (x− Lt)

)(

(3(LeWeLt+NWf Lf Lt)+8WtL2t)x2

−(2LeWe+2NWf Lf+8WtLt)x3+2Wtx4

)

w (x)dw (x) dx

=

Lt∫

0

4ETtW3t

(Wt +

(WeLe

4+

NWf Lf

4

)δ (x− Lt)

)(

(3(LeWeLt+NWf Lf Lt)+8WtL2t)x2

−(2LeWe+2NWf Lf+8WtLt)x3+2Wtx4

)

w2 (x)

2dx.

(A–97)

The distributed deflection w (x) is expressed in terms of the central deflection w(Lt) by

eliminating the shear stress τw from Equation A–28 and A–29, to give

w (x) = w (Lt)

((3LeWeLt+3NWf Lf Lt+8WtL2

t)x2

−(2LeWe+2NWf Lf+8WtLt)x3+2Wtx4

)

WeLeL3t

[1 +

NWf Lf

WeLe+ 2 WtLt

WeLe

] . (A–98)

Combining Equation A–97 and Equation A–98 results in

WPE =

Lt∫

0

w2 (Lt)

[2ETtW3

t

(Wt+

(WeLe

4 +NWf Lf

4

)δ(x−Lt)

)

W2e L2

eL6t

[1+

NWf LfWeLe

+2WtLtWeLe

]2

] ((3LeWeLt+3NWf Lf Lt+8WtL2

t)x2

−(2LeWe+2NWf Lf+8WtLt)x3

+2Wtx4

)dx

= w2 (Lt)

(ETtW

4t

L2t WeLe

)

1 +NWf Lf

WeLe+ 32

15

(WtLt

WeLe

)

(1 +

NWf Lf

WeLe+ 2

(WtLt

WeLe

))2

+

ETtW3t

(WeLe

2+

NWf Lf

2

)w2 (Lt)

WeLeL3t

[1 +

NWf Lf

WeLe+ 2 WtLt

WeLe

]

= ETt

(Wt

Lt

)3

1 + 2(

NWf Lf

WeLe

)+ 4

(WtLt

WeLe

)+ 4

(NWf Lf WtLt

(WeLe)2

)

+6415

(WtLt

WeLe

)2

+(

NWf Lf

WeLe

)2

2(1 +

NWf Lf

WeLe+ 2

(WtLt

WeLe

))2 w2 (Lt) .

(A–99)

Since the computation above is only for a single tether, the total potential energy of the

floating element suspended by four tethers is four times the value in Equation A–99, written

196

Page 197: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

as,

WPE = 2ETt

(Wt

Lt

)3

1 + 2(

NWf Lf

WeLe

)+ 4

(WtLt

WeLe

)+ 4

(NWf Lf WtLt

(WeLe)2

)

+6415

(WtLt

WeLe

)2

+(

NWf Lf

WeLe

)2

(1 +

NWf Lf

WeLe+ 2

(WtLt

WeLe

))2 w2 (Lt) . (A–100)

Combining Equation A–93 and A–100 and solving for Cme gives the lumped compliance of

the floating element sensor, expressed as

Cme =1

4ETt

(Lt

Wt

)3

(1 +

NWf Lf

WeLe+ 2

(WtLt

WeLe

))2

1 + 2(

NWf Lf

WeLe

)+ 4

(WtLt

WeLe

)+ 4

(NWf Lf WtLt

(WeLe)2

)

+6415

(WtLt

WeLe

)2

+(

NWf Lf

WeLe

)2

. (A–101)

197

Page 198: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

APPENDIX BTWO PORT ELEMENT MODELING

B.1 Two Port Models

A two-port element is represented in general as shown in Figure B-1, where A & B and

C & D are conjugate power variables (effort and flow). By definition, the power transfer is

always into the two-port element and hence must be conserved [42],

PowerNET = AB − CD = 0 or AB = CD. (B–1)

Two-Port

Element

+ +

− −

A D

B C

Figure B-1. General representation of an ideal two-port element.

Two types of linear elements satisfy the condition in Equation B–1. One is a transformer

and the other is a gyrator, represented as follows:

Transformer

D

C

=

n 0

0 1/n

A

B

(B–2)

Gyrator

D

C

=

0 n

1/n 0

A

B

(B–3)

The transformer and gyrator as circuit elements are shown in Figure B-2. The transformer

model in Equation B–2 represents the transduction of effort/flow in one domain to corre-

sponding effort/flow in a different energy domain via the transduction factor, n. The gyrator

model in Equation B–3 relates effort in one energy domain to flow in a different domain.

For example, electrodynamic speakers are modeled using a gyrator where electrical current

198

Page 199: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

+ +

--

A

B

D

Cn

Gyrator

+ +

--

A

B

D

C1:n

Transformer

Figure B-2. Circuit representation of the transformer and gyrator.

(B) through a coil in a magnetic field produces a proportional actuation force (D) via elec-

trodynamic coupling. A transformer model is applicable for a capacitive microphone where

pressure (A) is transduced into a proportional voltage (D) due to diaphragm deflection.

Two port networks also allow representation using an impedance analogy or an admit-

tance analogy based on definition of A,B,C, andD in Equation B–1. For an impedance

analogy the across variables, A & D are effort variables. For admittance analogy the across

variables A & D are flow variables. This definition usually depends on the specific transduc-

tion problem. As a convention, in the impedance analogy an effort (eg: pressure, voltage,

force) is an across variable and a flow (eg: current, volumetric flow rate, velocity) is a through

variable. An example of an impedance to impedance analogy for a two-port element is shown

in Figure B-3.

Transducer

Element

+

I U

V

_

+

F

Figure B-3. Impedance to impedance analogy representation of a two port element.

B.2 Linear Conservative Transducers

This section presents models for linear conservative transducers (LCT) based on the

discussion for two port models in the previous section. These transducers linearly transduce

energy from one energy domain to the other. The linear transduction maintains spectral

199

Page 200: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

fidelity of the input signal. For an electromechanical transducer, variables in the electrical

domain are expressed in terms of those in the mechanical domain and vice versa. Consider

the transducer element shown in Figure B-3 with the instantaneous values of effort variables

expressed in terms of flow variables as,

v = v (i, u) andf = f (i, u) , (B–4)

where v & i are voltage and current for the electrical domain and f & u are force and

velocity for the mechanical domain, respectively. In the Fourier domain these variables are

represented as,

V = V (I, U) andF = F (I, U) . (B–5)

Applying chain rule to Equation B–4,

dv =∂v

∂i

∣∣∣∣u=0

di +∂v

∂u

∣∣∣∣i=0

du, (B–6)

and df =∂f

∂i

∣∣∣∣u=0

di +∂f

∂u

∣∣∣∣i=0

du. (B–7)

Linearizing about a mean condition and replacing differentials with fluctuations in the Fourier

domain results in,

V =V

I

∣∣∣∣U=0

I +V

U

∣∣∣∣I=0

U = ZEBI + TEMU, (B–8)

and F =F

I

∣∣∣∣U=0

I +F

U

∣∣∣∣I=0

U = TMEI + ZMOU. (B–9)

Thus the characteristic equation of a linear transducer is,

V

F

=

ZEB TEM

TME ZMO

I

U

, (B–10)

where the impedances are defined as follows:

ZEB ≡ V

I

∣∣∣∣U=0

, (B–11)

200

Page 201: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

the blocked electrical impedance,

ZEF ≡ V

I

∣∣∣∣F=0

, (B–12)

the free electrical impedance,

ZMO ≡ F

U

∣∣∣∣I=0

, (B–13)

the open-circuit mechanical impedance,

ZMS ≡ F

U

∣∣∣∣V =0

, (B–14)

the short-circuit mechanical impedance,

TEM ≡ V

U

∣∣∣∣I=0

, (B–15)

the open-circuit electromechanical transduction factor, and

TME ≡ F

I

∣∣∣∣U=0

, (B–16)

the blocked electromechanical transduction factor.

For a reciprocal transducer, TME = TEM = T , which implies that the same velocity, U

can be obtained using both a force and a voltage, and in addition the same current, I can

be obtained using a force and a voltage [66]. If a transducer is reciprocal, the impedance

transformation factor is,

ϕ′ =T

ZMO

. (B–17)

The equivalent circuit of a LCT is shown in Figure 31 where the free electrical impedance is

ZEF = ZEB − T 2

ZMO

=(1− κ2

)ZEB, (B–18)

201

Page 202: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

where κ is the electromechanical coupling factor expressed as [42],

κ2 =T 2

ZEBZMO

. (B–19)

+

F MSZ

U

I

EFZ

+

V

1:φ ′

MOZ

Figure B-4. Equivalent circuit representation of a transducer using impedance analogy.

x′

x

0x x=

0x =

fixed

plates

movable

plates

A

Figure B-5. Schematic of a parallel plate capacitive transducer.

Example:. Consider a parallel plate capacitor with a fixed plate and movable plate as

shown in Figure B-5.

The nominal gap between the plates is x0 and the change in gap due to a sens-

ing/actuation force is x′ (t) such that,

x = x0 − x′ (t) . (B–20)

202

Page 203: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

The capacitance is computed as,

CE(t) =ε0A

x=

ε0A

x0 − x′(t)=

ε0A

x0

(1− x′(t)

x0

)−1

= CEB

(1− x′(t)

x0

)−1

, (B–21)

where CEB is the blocked capacitance when x′ (t) = 0. The voltage is related to the capaci-

tance and stored charge between the parallel plates by,

V (t) =Q(t)

CE(t)=

Q(t)

CEB

(1− x′(t)

x0

). (B–22)

Equation B–22 is the sum of the voltage at rest and that due to electromechanical coupling.

The electrostatic force for a constant charge between the plates is [42],

FE = −1

2

Q2

CEBx0

= −1

2

Q2

ε0A. (B–23)

The compliance/stiffness of the diaphragm results in a mechanical restoring force,

FM = kx′ (t) =x′ (t)CMO

, (B–24)

where k is the mechanical stiffness and CMO is the open circuit compliance of the diaphragm.

Adding Equation B–23 and B–24, the total force on the diaphragm is,

F = FM + FE =x′(t)CMO

− 1

2

Q2

ε0A=

x′(t)CMO

− 1

2

Q2

CEBx0

. (B–25)

Assume Q(t) = Q0 + Q′(t) and V (t) = V0 + V ′(t) with the constraint that the perturbations

are small compared to the mean, i.e., x′(t)x0

<< 1;Q′(t)Q0

<< 1; and V ′(t)V0

<< 1. Applying these

assumptions, the linearized form of Equation B–22 and B–25 are,

V (t) =Q′(t)CEB

− V0x′(t)

x0

, (B–26)

and F (t) =x′(t)CMO

− V0Q′(t)

x0

. (B–27)

The two-port model representation is generally applicable for transducers that measure per-

turbations about a mean. Thus dropping the ”primes” and representing Equation B–26 and

203

Page 204: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

B–27 in terms of their Fourier components results in,

V =

(1

jωCEB

)I +

(− V0

jωx0

)U (B–28)

and F =

(− V0

jωx0

)I +

(1

jωCMO

)U. (B–29)

The matrix form of the governing equations for a capacitive transducer is,

V

F

=

ZEB TEM

TME ZMO

I

U

=

1jω CEB

−V0

jω x0

−V0

jω x0

1jω CMO

I

U

. (B–30)

Equation B–30 indicates the transducer is reciprocal but indirect in nature. Comparing

Equation B–30 and B–17,

ϕ′ =T

ZMO

= − V

x0

CMO. (B–31)

The equivalent circuit representation for a capacitive transducer is shown in Figure B-6.

This is consistent with the general modeling approach of a linear conservative transducer

discussed in the previous section. Using Equation B–19, the compliance CEF is,

1

jωCEF

=(1− κ2

) 1

jωCEB

, (B–32)

where

κ2 =T 2

ZEBZMO

=

(V

jωx0

)2/

1

(jω)2 CEBCMO

=V 2

x20

CEBCMO (B–33)

.

+

F

U

I

+

VMO

C

EFC

0

0

1:

M

x

C V

0

1:MO

VC

x−

_

Figure B-6. Circuit representation of a capacitive transducer for constant charge biasing.

204

Page 205: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

APPENDIX CSHEAR STRESS IN PWT WITH REFLECTIONS

This chapter shows the solution procedure to correct for effects of reflection for fluctu-

ating wall shear stress measurements in the plane wave tube. The derivation is for a general

impedance termination at the end of the tube having a reflection coefficient, R as shown in

Figure C-1.

Techron 7540 Power

Supply Amplifier

B & K Pulse

Analyzer System

PC

Acoustic Plane

Wave

Shear Stress

SensorSpeaker

Microphone

General

Impedance

Termination

R=Z

PressureVelocity

Standing Wave Pattern

Left Running

Wave (Incident)

Right Running

Wave (Reflected)

dps

ds

Figure C-1. Setup for shear stress in PWT for a general impedance termination.

Assume a fully developed flow in a duct, driven by an oscillating pressure gradient. The

governing differential equation is

∂u

∂t= −1

ρ

∂p

∂xejωt + ν

∂2u

∂y2. (C–1)

205

Page 206: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Based on no-slip boundary condition at the duct wall and the finite velocity at the center of

the duct the expression for velocity by solving Equation C–1 is [93]

u (y, t) =jejωt

ρω

∂p

∂x

1−

cosh

(y√

jων

)

cosh

(a√

jων

)

. (C–2)

Now assume a left running (incident wave) plane acoustic wave along the length of the duct

given by [87],

p1 = p1(x)ejωt = p′1e(jωt−kx)

∴ ∂p1

∂x= −jω

cp′1e

(jωt−kx) (C–3)

where x is the position along the length of the duct. The corresponding velocity due to this

pressure is,

u1 (x, y, t) =jejωt

ρω

∂p1

∂x

1−

cosh

(y√

jων

)

cosh

(a√

jων

)

. (C–4)

Similarly, for a right running wave (reflected wave) the pressure is

p2 = p2(x)ejωt = p′2e(jωt+kx)

∴ ∂p2

∂x= −jω

cp′2e

(jωt+kx) (C–5)

The corresponding reflected velocity is

u2 (x, y, t) =jejωt

ρω

∂p2

∂x

1−

cosh

(y√

jων

)

cosh

(a√

jων

)

. (C–6)

Using linear superposition, the total velocity at a given location, x, along the length of the

duct is

u (x, y, t) = u1 (x, y, t) + u2 (x, y, t) . (C–7)

206

Page 207: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Note: u1&u2 account for the direction of the wave via the sign associated with the pressure

gradient in Equations C–3 and C–5, respectively. Substituting Equations C–4 and C–6 in

Equation C–7 results in

u (x, y, t) =jejωt

ρω

1−

cosh

(y√

jων

)

cosh

(a√

jων

)

(∂p1

∂x+

∂p2

∂x

). (C–8)

The wall shear stress due to this velocity is,

τwall = µdu

dy

∣∣∣∣y=a

. (C–9)

Substituting the expression for velocity from Equation C–8,

τwall =ejωt

j

√jµ

ρωtanh

(a

√jω

ν

)(∂p1

∂x+

∂p2

∂x

). (C–10)

Now, substituting for the pressure gradients from Equations C–3 and C–5 gives

τwall = −1

c

√jωµ

ρtanh

(a

√jω

ν

)(p′1e

j(ωt−kx) − p′2ej(ωt+kx)

). (C–11)

Defining distances from the termination x = l− d, where l is the length of the tube and d is

the distance from the termination, the τwall in the above expression may be rewritten as

τwall = −1

c

√jωµ

ρtanh

(a

√jω

ν

)(p′1e

j(ωt−kl+kd) − p′2ej(ωt+kl−kd)

), (C–12)

or τwall = −1

c

√jωµ

ρtanh

(a

√jω

ν

)p′1e

−jkl

︸ ︷︷ ︸P+

ej(ωt+kd) − p′2ejkl

︸ ︷︷ ︸P−

ej(ωt−kd)

, (C–13)

or τwall = −1

c

√jωµ

ρtanh

(a

√jω

ν

)P+

(ejkd − P−

P+e−jkd

)ejωt. (C–14)

The reflection coefficient at a given termination based on incident and reflected pressures is

defined as,

R =p′2e

jkl

p′1e−jkl=

P−

P+. (C–15)

207

Page 208: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Substituting R from Equation C–15 in Equation C–14 results in,

τwall = −1

c

√jων tanh

(a

√jω

ν

)P+

(ejkd −Re−jkd

)ejωt. (C–16)

Assuming that the shear stress sensor is located at a position ds from the termination,

τwall = −1

c

√jωv tanh

(a

√jω

ν

)P+

(ejkds −Re−jkds

)ejωt. (C–17)

The pressure P+ejωt has to be expressed in terms of total pressure measured by the micro-

phone at a distance dps from the termination. The total pressure at a distance dps is,

p (dps) ejωt = p1 + p2 =(p′1e

j(ωt−kl+kdps) + p′2ej(ωt+kl−kdps)

)

p (dps) ejωt = p′1e−jklej(ωt+kdps) + p′2e

jklej(ωt−kdps)

p (dps) ejωt =(P+ejkdps + P−e−jkdps

)ejωt

pmeasuredejωt = p (dps) ejωt = P+

(ejkdps + Re−jkdps

)ejωt.

(C–18)

Substituting for P+ from Equation C–18 in terms of pmeasured, into Equation C–17,

τwall =

(−1

c

√jων tanh

(a

√jω

ν

)ejkds −Re−jkds

ejkdps + Re−jkdps

)pmeasurede

jωt. (C–19)

Let the separation between the sensor and microphone be δ = ds − dps, then the wall shear

stress may be expressed as

τwall =

(−1

c

√jων tanh

(a

√jω

ν

)ejkds −Re−jkds

ejk(ds−δ) + Re−jk(ds−δ)

)pmeasurede

jωt. (C–20)

208

Page 209: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

APPENDIX DPROCESS TRAVELER AND PACKAGING DETAILS

Wafer:. 100±0.1 mm” p-type (100) SOI wafer of thickness 545±10 µm, 2 µm±5% thick

buried oxide (BOX) layer, 45±1 µm thick device layer with resistivity 0.001−0.005 ω− cm,

and 500± 10 µm thick handle wafer with resistivity > 1000 ω − cm.

Masks:.

• Floating element DRIE mask (FEM)

• Backside/Cavity etch mask (CEM)

Process Steps:.

1. Wafer cleaning

(a) SC-1 (Ion Strip) - 15 : 3 : 2 H2O : H2O2 : NH4OH at 75C for 10 min

(b) SC-2 (Organic Strip) - 16 : 3 : 1 H2O : H2SO4 : H2O2 at 75C for 10 min

(c) Oxi-clean - 50 : 1 H2O : HF at room temperature for 30s

(d) Acetone/Methanol/DI wash

2. Etch on device layer to define sensor structure

(a) Coat with HMDS for 5 min

(b) Spin resist (1 µm, AZ 1512) - spin at 500 rpm 100 rpm/s 5s−4000 rpm 100 rpm/s 50s

(c) Soft bake 60s (hot plate) at 95 C

(d) Expose using mask FEM - EVG620 Ch-A 29.5 mW/cm2 for 1 s (hard contact)

(e) Develop using AZ 300MIF for 60 s

(f) Post exposure bake (oven) - 95C for 60 min

(g) Etch using STS-DRIE - recipe VJ pol50, 42 cycles and recipe BAO2 VJ (avoids

footing), 20 cycles

(h) Acetone/Methanol/DI wash to clean photoresist

3. Nickel electroplating

(a) Piranha clean to remove organic CHF3 passivation - 3 : 1 H2SO4 : H2O2 at

120 C for 10 min

209

Page 210: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

(b) Oxide removal using BOE (6 : 1) for 5 min

(c) Immediate 2-propanol immersion for surface wetting and to prevent oxidation

(d) Immediate immersion in Technic Nickel sulfamate solution at 90 F (Terminals -

Ni electrode positive, SOI wafer negative)

(e) Begin plating simultaneously with immersion - 30s strike with 125 mA (nucle-

ation), 5 min plating with 30 mA (finer grains), 5 min plating with 7 mA (very

fine grains)

(f) Rinse with DI water

4. Backside/Cavity etch

(a) Coat backside with HMDS for 5 min

(b) Spin resist on backside (10 µm, AZ 9260) - spin at 500 rpm 100 rpm/s 5s −2000 rpm 100 rpm/s 50s

(c) Spin AZ 9260 on handle silicon wafer - 2000 rpm 1000 rpm/s ramp rate

(d) Soft bake 30min (oven) at 95 C

(e) Front to back align using EVG620

(f) Expose using mask CEM - EVG Ch-B 62 mW/cm2 for 23 s (hard contact)

(g) Develop using AZ 300MIF for 3 min

(h) Spin resist (5 µm, AZ 9260) - spin at 500 rpm 100 rpm/s 5s−4000 rpm 100 rpm/s 50s

(i) Place SOI wafer face down on handle wafer and apply gentle pressure for uniform

contact

(j) Post exposure bake (Oven) - 95C for 60 min

(k) Etch back side cavity using STS-DRIE - recipe VJ pol50, 470 cycles and recipe

BAO2 VJ (avoids footing), 130 cycles

(l) Remove PR mask - Acetone/Methanol/DI wash

(m) Separate from handle wafer using S-46 PR stripper - heat 80C for ≈ 15 min

(n) Rinse Acetone/Methanol/DI

5. Die separation

210

Page 211: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

(a) Lay down double sided heat sensitive tape REVALPHA (120C) on back side of

wafer (without air bubbles)

(b) Stick wafer, face up on a handle silicon wafer for support during dicing

(c) Lay down single sided heat sensitive tape REVALPHA (80C) on front side of

wafer (without air bubbles)

(d) Dice wafer using dicing saw - Tresser 620 - Kulicke and Soffa blade S1235 - thick-

ness (1− 1.2 mm)

(e) Heat on hot-plate at 80C to peel off tape from front side of wafer

(f) Heat on hot-plate at 120C to separate die from tape and handle wafer

6. Die release

(a) Etch BOX in BOE (6:1) for 20 min

(b) Immediate immersion in methanol

(c) Super critical dry in CO2 (Prof. Ho-Bun Chan’s lab)

Packaging:.

1. Sensor packaging in PCB

(a) Cut recess in PCB for sensor mounting (CNC recipe name) - PCB layout shown

in Figure D-1 - made at Sierra Proto Express

(b) Glue SiSonic to back side of PCB using epoxy (Dualbond 707) - cure at 35C for

10 min

(c) Wire bond to bond pads using ball size 4. (Prof. Ho-Bun Chan’s lab)

(d) Cover bonds with epoxy (Dualbond 707) - cure at 35C for 10 min

(e) Solder SMB connectors for electrical contact and fill through vias with solder

(f) Glue sensor in place with epoxy (Dualbond 707) - cure at 35C for 10 min

(g) Wirebond using ball size 6 and force at 100. (Prof. Ho-Bun Chan’s lab)

2. Sensor PCB packaging in plug

(a) Sensor PCB is press fit in Lucite plug - drawing and dimensions shown in Figure D-

2 - made at TMR engineering (Ken Reed)

211

Page 212: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

Figure D-1. PCB Layout and dimensions.

212

Page 213: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

6

10

60

40

27

,1

R46,12

12,7

10,16

10

6

5

10,16

17,86

1,62

5

SC

AL

E1

,00

0

SC

AL

E1

,00

0

Ce

nte

r of th

rou

gh

ho

le is

R=

75

SC

AL

E1

,00

0

SC

AL

E1

,00

0

Figure D-2. Drawing of Lucite plug.

213

Page 214: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

REFERENCES

[1] E. I. Administration, International energy outlook 2008, DOE/EIA-0484. Washington,DC: U.S. Department of Energy, http://www.eia.doe.gov/oiaf/ieo/index.html, 2008.

[2] M. Gad-el Hak, “Flow control : Passive, active, and reactive flow management.” Cam-bridge University Press, 2000, ch. 10, pp. 209–210.

[3] E. I. Administration, “Petroleum marketing monthly, doe/eia-0380.” Washington, DC:U.S. Department of Energy, 2008, p. 5.

[4] ——, “World crude oil prices.” Washington, DC: U.S. Department of Energy,http://tonto.eia.doe.gov/dnav/pet/pet-pri-wco-k-w.htm, 2009.

[5] R. J. Stiles, “Introduction to aeronautics: a design perspective.” Washington: AIAA,2004, ch. 3, p. 80.

[6] J. W. Naughton and M. Sheplak, “Modern development in shear stress measurement,”Progress in Aerospace Sciences, vol. 38, pp. 515–570, 2002.

[7] H. Schlichting, “Boundary-layer theory.” New York: McGraw-Hill, 1979, ch. 1, pp.6–9.

[8] F. M. White, “Viscous fluid flow.” New York: McGraw-Hill, 1991, ch. 4-6.

[9] K. G. Winter, “An outline of the techniques available for the measurements of skinfriction in turbulent boundary layers,” Progress in Aerospace Sciences, vol. 18, pp. 1–57, 1977.

[10] J. H. Haritonidis, “The measurement of wall shear stress,” Advances in Fluid MechanicsMeasurements, pp. 229–261, 1989.

[11] H. Tennekes and J. L. Lumley, “A first course in turbulence.” Cambridge, MA: MITPress, 1972, ch. 1-5.

[12] H. Ludwieg, “Instrument for measuring the wall shearing stress of turbulent boundarylayers,” NACA, Tech. Rep. NACA TM 1284, 1950.

[13] J. S. J. Mathieu, “An introduction to turbulent flow.” Massachusetts: CambridgeUniv. Press, 2000, ch. 5.

[14] D. B. Spalding and S. W. Chi, “The drag of a compressible turbulent boundary layeron a smooth flat plate with and without heat transfer,” Journal of Fluid Mechanics,vol. 18, no. 1, pp. 117–143, 1964.

[15] R. L. Panton, “Incompressible flow.” New York Chichester: Wiley, 1996, ch. 18, p.526.

[16] S. K. Robinson, “Coherent motions in the turbulent boundary-layer,” Annual Reviewof Fluid Mechanics, vol. 23, pp. 601–639, 1991.

214

Page 215: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

[17] R. Rathnasingham and K. Breuer, “Active control of turbulent boundary layers,” J.Fluid Mech., vol. 495, p. 209233, 2003.

[18] A. Padmanabhan, H. Goldberg, K. D. Breuer, and M. A.Schmidt, “A wafer-bondedfloating-element shear stress microsensor with optical position sensing by photodiodes,”Journal of Microelectromechanical Systems, vol. 5, no. 4, pp. 307–315, 1996.

[19] L. Lofdahl and M. Gad-el Hak, “MEMS-based pressure and shear stress sensors forturbulent flows,” Measurement Science & Technology, vol. 10, no. 8, pp. 665–686, 1999.

[20] H. Alfredsson, A. V. Johansson, J. H. Haritonidis, and H. Eckelman, “The fluctuatingwall-shear stress and the velocity field in the viscous sublayer,” Phys. Fluids, vol. Vol.31, pp. 1026–1033, 1988.

[21] A. Padmanabhan, “Silicon micromachined sensors and sensor arrays for shear-stressmeasurements in aerodynamic flows,” PhD thesis, Department of Mechanical Engineer-ing, Massachussets Institute of Technology, Cambridge, MA, 1997.

[22] Z. W. Hu, C. L. Morfey, and N. D. Sandham, “Wall pressure and shear stress spectrafrom direct simulations of channel flow,” AIAA Journal, vol. 44, no. 7, pp. 1541–1549,2006.

[23] M. Sheplak, L. Cattafesta, T. Nishida, and C. McGinley, “MEMS shear stress sen-sors: Promise and progress,” in 24th AIAA Aerodynamic Measurement Technology andGround Testing Conference, AIAA-2004-2606, Portland, Oregon, 2004.

[24] F. H. Clauser, “Turbulent boundary layers in adverse pressure gradients,” Journal ofthe Aeronautical Sciences, vol. 21, no. 2, pp. 91–108, 1954.

[25] B. Oudheusden and J. Huijsing, “Integrated flow friction sensor,” Sensors and ActuatorsA, vol. 15, pp. 135–144, 1988.

[26] E. Kalvesten, G. Stemme, C. Vieider, and L. Lofdahl, “An integrated pressure-flowsensor for correlation measurement in turbulent gas flow,” Sensors and Actuators A,vol. 52, pp. 51–58, 1996.

[27] C. Liu, J. B. Huang, Z. J. Zhu, F. K. Jiang, S. Tung, Y. C. Tai, and C. M. Ho,“Micromachined flow shear-stress sensor based on thermal transfer principles,” Journalof Microelectromechanical Systems, vol. 8, no. 1, pp. 90–99, 1999.

[28] M. Sheplak, V. Chandrasekaran, A. Cain, T. Nishida, and L. Cattafesta, “Characteri-zation of a silicon-micromachined thermal shear-stress sensor,” AIAA Journal, vol. 40,no. 6, pp. 1099–1104, 2002.

[29] T. von Papen, H. Steffes, H. D. Ngo, and E. Obermeier, “A micro surface fence probe forthe application in flow reversal areas,” Sensors and Actuators A: Physical, vol. 97-98,pp. 264–270, 2002.

215

Page 216: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

[30] S. Groe, W. Schroder, and C. Brucker, “Nano-newton drag sensor based on flexiblemicro-pillars,” Measurement Science and Technology, vol. 17, no. 10, pp. 2689–2697,2006.

[31] C. Brucker, D. Bauer, and H. Chaves, “Dynamic response of micro-pillar sensors mea-suring fluctuating wall-shear-stress,” Experiments in Fluids, vol. 42, no. 5, pp. 737–749,2007.

[32] S. Groe and W. Schroder, “Mean wall-shear stress measurements using the micro-pillarshear-stress sensor mps3,” Measurement Science and Technology, vol. 19, no. 1, 2008.

[33] S. Groe and W. Schrder, “Dynamic wall-shear stress measurements in turbulent pipeflow using the micro-pillar sensor mps3,” International Journal of Heat and Fluid Flow,vol. 29, no. 3, pp. 830–840, 2008.

[34] D. Fourguette, D. Modarress, F. Taugwalder, D. Wilson, M. Koochesfahani, andM. Gharib, “Miniature and moems flow sensor,” AIAA paper, 2001-2982, 2001.

[35] M. Gharib, D. Modarress, D. Fourguette, and D. Wilson, “Optical microsensors for fluidflow diagnostics,” in 40th AIAA Aerospace Sciences Meeting and Exhibit, AIAA-2002-252, Reno, NV, 2002.

[36] M. A. Schmidt, “Microsensors for the measurement of shear forces in turbulent boundarylayers,” PhD thesis, Department of Mechanical Engineering, Massachussets Institute ofTechnology, 1988.

[37] M. A. Schmidt, R. Howe, S.D.Senturia, and J. Haritonidis, “Design and calibration of amicromachined floating-element shear-stress sensor,” Transactions of Electron Devices,vol. 35, pp. 750–757, 1988.

[38] T. Pan, D. Hyman, M. Mehregany, E. Reshotko, and B. Williams, “Characterizationof microfabricated shear stress sensors,” Proceedings of 16th International Congress onInstrumentation in Aerospace Simulation Facilities, WPAFB, 1995.

[39] D. Hyman, T. Pan, E. Reshotko, and M. Mehregany, “Microfabricated shear stresssensors, part 2: Testing and calibration,” AIAA Journal, vol. 37, no. 1, pp. 73–78,1999.

[40] T. Pan, D. Hyman, and M. Mehregany, “Microfabricated shear stress sensors, part1:design and fabrication,” AIAA, vol. 37, no. No.1, pp. 66–72, 1999.

[41] M. Rossi, “Acoustics and electroacoustics.” Norwood, MA: Artech House, 1988, ch. 5.

[42] S. D. Senturia, “Microsystem design.” Boston: Kluwer Academic Publishers, 2001, ch.6, 16.

[43] J. Zhe, K. R. Farmer, and V. Modi, “A MEMS device for measurement of skin fric-tion with capacitive sensing,” in Proceedings of 2001 Microelectromechanical SystemsConference, IEEE Circuits and Systems Society, Berkeley, CA, USA, 2001, pp. 4–7.

216

Page 217: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

[44] J. Zhe, V. Modi, and K. R. Farmer, “A microfabricated wall shear-stress sensor withcapacitative sensing,” Journal of Microelectromechanical Systems, vol. 14, no. 1, pp.167–175, 2005.

[45] M. McCarthy, L. Frechette, V. Modi, and N. Tiliakos, “Initial development of a MEMSwall shear stress sensor for propulsion applications,” in Propulsion Measurement Sen-sor Development Workshop NASA Marshall Space Flight Center, Huntsville, Alabama,2003, pp. 1–10.

[46] N. Tiliakos, G. Papadopoulos, V. Modi, A. O’Grady, M. McCarthy, L. Frechette, J.-P.Desbiens, and R. Larger, “MEMS shear stress sensor for hypersonic aeropropulsion testand evaluation,” in 2006 Annual ITEA Technology Review, 2006.

[47] N. Tiliakos, G. Papadopoulos, A. O. Grady, V. Modi, R. Larger, and L. Frechette,“A MEMS-based shear stress sensor for high temperature applications,” in 46th AIAAAerospace Sciences Meeting and Exhibit, AIAA-2008-274. Reno, NV: AIAA, 2008.

[48] A. V. Desai and M. A. Haque, “Design and fabrication of a direction sensitive MEMSshear stress sensor with high spatial and temporal resolution,” Journal of Micromechan-ics and Microengineering, vol. 14, no. 12, p. 1718, 2004.

[49] A. Padmanabhan, H. Goldberg, M. A.Schmidt, and K. D. Breuer, “A silicon microma-chined sensor for shear stress measurements in aerodynamic flows,” in 34th AerospaceSciences Meeeting and Exhibit, AIAA 96-0422, Reno, NV.

[50] A. Padmanabhan, M. Sheplak, K. D. Breuer, and M. A.Schmidt, “Micromachined sen-sors for static and dynamic shear stress measurements in aerodynamic flows,” TechnicalDigest, Transducers ’97, Chicago, IL, pp. 137–140, 1997.

[51] M. Sheplak, A. Padmanabhan, M. A. Schmidt, and K. S. Breuer, “Dynamic calibrationof a shear-stress sensor using stokes-layer excitation,” AIAA, vol. 39, no. 5, pp. 819–823,2001.

[52] F.-G. Tseng and C.-J. Lin, “Polymer MEMS-based fabryperot shear stress sensor,”IEEE Sensors Journal, vol. 3, no. 6, pp. 812–817, 2003.

[53] S. Horowitz, T. Chen, V. Chandrasekaran, T. Nishida, L. Cattafesta, and M. Sheplak,“A micromachined geometric moire interferometric floating element shear stress sensor,”in 42nd AIAA Aerospace Sciences Meeting and Exhibit, 2004-1042. Reno, NV: AIAA,2004.

[54] S. Horowitz, T. Chen, V. Chandrasekaran, K. Tedjojuwono, T. Nishida, andL. Cattafesta, “A wafer-bonded, floating element shear-stress sensor,” in Technical Di-gest, Solid-State Sensor, Actuator and Microsystems Workshop, Hilton Head, SC, 2004,pp. 13–18.

217

Page 218: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

[55] K.-Y. Ng, “A liquid-shear-stress sensors using wafer-bonding technology,” Masters the-sis, Department of Electrical Engineering and Computer Science, Massachussets Insti-tute of Technology, 1990.

[56] K.-Y. Ng, J. Shajii, and M. A. Schmidt, “A liquid shear-stress sensor fabricated us-ing wafer bonding technology,” in International Conference on Solid-State Sensors andActuators, Transducers ’91, 1991, pp. 931–934.

[57] H. Goldberg, K. Breuer, and M. Schmidt, “A silicon wafer-bonding technology formicrofabricated shear-stress sensors with backside contacts,” Technical Digest, Solid-State Sensor and Actuator Workshop, pp. 111–115, 1994.

[58] A. A. Barlian, R. Narain, J. T. Li, C. E. Quance, A. C. Ho, V. Mukundan, and B. L.Pruitt, “Piezoresistive MEMS underwater shear stress sensors,” in International Con-ference on Micro Electro Mechanical Systems, MEMS 2006. Istanbul, Turkey: IEEE,2006, pp. 626–629.

[59] A. A. Barlian, S. J. Park, V. Mukundan, and B. L. Pruitt, “Design and characterizationof microfabricated piezoresistive floating element-based shear stress sensors,” Sensorsand Actuators a-Physical, vol. 134, no. 1, pp. 77–87, 2007.

[60] Y. Li, V. Chandrasekharan, B. Bertolucci, T. Nishida, L. Cattafesta, D. P. Arnold,and M. Sheplak, “A laterally implanted piezoresistive skin-friction sensor,” in TechnicalDigest, Solid-State Sensors, Actuators, and Microsystems Workshop, Hilton Head, SC,2008, pp. 304–307.

[61] Y. Li, V. Chandrasekharan, B. Bertolucci, T. Nishida, L. Cattafesta, and M. Sheplak,“A MEMS shear stress sensor for turbulence measurements,” in 46th AIAA AerospaceSciences Meeting and Exhibit, AIAA-2008-269. Reno, NV: AIAA, 2008.

[62] R. R. Spencer, B. M. Fleischer, P. W. Barth, and J. B. Angell, “A theoretical-study oftransducer noise in piezoresistive and capacitive silicon pressure sensors,” IEEE Trans-actions On Electron Devices, vol. 35, no. 8, pp. 1289–1298, 1988.

[63] P. R. Scheeper, A. G. H. van der Donk, W. Olthuis, and P. Bergveld, “A review ofsilicon microphones,” Sensors and Actuators A: Physical, vol. 44, no. 1, pp. 1–11, 1994.

[64] Y. Li, M. Papila, T. Nishida, L. Cattafesta, and M. Sheplak, “Modeling and optimizationof a side-implanted piezoresistive shear stress sensor,” in Proceeding of SPIE 13th AnnualInternational Symposium on Smart Structures and Materials, paper 6174-7, San Diego,CA, 2006.

[65] J. N. Reddy, “Theory and analysis of elastic plates.” Philadelphia, PA: Taylor &Francis, 1999, ch. 1, pp. 21–25.

[66] F. V. Hunt, “Electroacoustics; the analysis of transduction, and its historical back-ground.” Cambridge, MA: Harvard University Press, 1954, ch. 6.

[67] R. K. Wangsness, “Electromagnetic fields.” New York: Wiley, 1986, ch. 6.

218

Page 219: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

[68] L. K. Baxter, “Capacitive sensors design and applications.” New York : IEEE Press,1997, ch. 3, pp. 42–43.

[69] B. E. Boser and K. W. Markus, “Design of integrated mems,” in Designing Low PowerDigital Systems, Emerging Technologies (1996), 1996, pp. 207–232.

[70] B. E. Boser, “Electronics for micromachined inertial sensors,” in International Confer-ence on Solid State Sensors and Actuators, Transducers ’97, Chicago, vol. 2, 1997, pp.1169–1172 vol.2.

[71] J. C. Lotters, W. Olthuis, P. H. Veltink, and P. Bergveld, “A sensitive differentialcapacitance to voltage converter for sensor applications,” IEEE Transactions on Instru-mentation and Measurement, vol. 48, no. 1, pp. 89–96, 1999.

[72] D. T. Martin, “Design, fabrication and characterization of a MEMS dual-backplate ca-pactive microphone,” PhD thesis, Department of Electrical and Computer Engineering,University of Florida, 2007.

[73] K. Kadirvel, “Closed loop interface circuits for capacitive transducers with application toa MEMS capacitive microphone,” PhD thesis, Department of Electrical and ComputerEngineering, University of Florida, 2007.

[74] D. Martin, K. Kadirvel, T. Nishida, and M. Sheplak, “An instrument grade MEMS con-densor microphone for aeroacoustic measurements,” in 46th AIAA Aerospace SciencesMeeting and Exhibit, AIAA-2008-257. Reno, NV: AIAA, 2008.

[75] B. E. Boser and R. T. Howe, “Surface micromachined accelerometers,” IEEE Journalof Solid-State Circuits, vol. 31, no. 3, pp. 366–375, 1996, 0018-9200.

[76] P. V. Loeppert and S. B. Lee, “Sisonictm - the first commercialized mems microphone,”in Solid-State Sensors, Actuators, and Microsystems Workshop, Hilton Head, SC, 2006,pp. 27–30.

[77] R. T. Haftka and Z. Grdal, “Elements of structural optimization.” Dordrecht ; Boston:Kluwer Academic Publishers, 1992, ch. 5.

[78] D. T. Martin, J. Liu, K. Kadirvel, R. M. Fox, M. Sheplak, and T. Nishida, “A microma-chined dual-backplate capacitive microphone for aeroacoustic measurements,” Journalof Microelectromechanical Systems, vol. 16, pp. 1289–1302, 2007.

[79] L. D. V. Llona, H. V. Jansen, and M. C. Elwenspoek, “Seedless electroplating on pat-terned silicon,” Journal of Micromechanics and Microengineering, vol. 16, no. 6, pp.S1–S6, 2006.

[80] M. Schlesinger and M. Paunovic, Modern electroplating, 4th ed., ser. ElectrochemicalSociety series. New York: Wiley, 2000.

[81] “http://www.technic.com/chm/nickelel.htm.”

219

Page 220: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

[82] R. W. Fox and A. T. McDonald, Introduction to fluid mechanics, 5th ed. New York:J. Wiley, 1998.

[83] A. M. Cain, “Static characterization of a micromachined thermal shear stress sen-sor,” Thesis (M.S.), Department of Electrical and Computer Engineering, Universityof Florida, 1999., 1999.

[84] Y. Li, “Side-implanted piezoresistive shear stress sensor for turbulent boundary layermeasurement,” Aerospace Engineering thesis, Ph. D., Department of Mechanical andAerospace Engineering, University of Florida, 2008.

[85] A. S. Sedra and K. C. Smith, “Microelectronic circuits.” New York: Oxford UniversityPress, 1997, ch. 11.

[86] V. Chandrasekaran, A. Cain, T. Nishida, L. N. Cattafesta, and M. Sheplak, “Dynamiccalibration technique for thermal shear-stress sensors with mean flow,” Experiments inFluids, vol. 39, no. 1, pp. 56–65, 2005.

[87] D. T. Blackstock, “Fundamentals of physical acoustics.” New York, NY: Wiley, 2000,ch. 4.

[88] R. Dieme, G. Bosman, T. Nishida, and M. Sheplak, “Sources of excess noise in siliconpiezoresistive microphones,” Journal of the Acoustical Society of America, vol. 119,no. 5, pp. 2710–2720, 2006.

[89] J. S. Bendat, “Nonlinear system analysis and identification from random data.” NewYork: Wiley, 1990, ch. 8,9.

[90] G. Bosman, “University of florida, private communication,” 2009.

[91] R. F. Pierret, “Field effect devices,” in Modular series on solid state devices ; v. 4,2nd ed. Reading, Mass.: Addison-Wesley Pub. Co., 1990, ch. 2.

[92] R. C. Hibbeler, “Mechanics of materials.” Upper Saddle River, N.J.: Prentice Hall,1997, ch. 12, pp. 575–579.

[93] V. Chandrasekaran, A. Cain, T. Nishida, and M. Sheplak, “Dynamic calibration tech-nique for thermal shear stress sensors with variable mean flow,” in Aerospace SciencesMeeting and Exhibit, 38th, AIAA-2000-508, Reno, NV, 2000.

220

Page 221: A MICROSCALE DIFFERENTIAL CAPACITIVE DIRECT WALL …

BIOGRAPHICAL SKETCH

Vijay Chandrasekharan grew up in Nagpur, a city located right in the center of India.

After high school, he went to the National Institute of Technology Karnataka (NITK), India

where he received his bachelor’s in mechanical engineering in 2002. After working for a year,

he started graduate school in the Department of Mechanical and Aerospace Engineering at

the University of Florida in 2003. He received his MS in 2006 and is currently completing his

doctoral degree at the Interdisciplinary Microsystems Group working under the guidance of

Prof. Mark Sheplak. Stemming from his inclination towards product development, Vijay’s

research interests include micromachined (MEMS) sensor and actuator design, modeling,

fabrication and characterization.

221