337
Development of Active Artificial Hair Cell Sensors Bryan Steven Joyce Dissertation submitted to the faculty of the Virginia Polytechnic Institute and State University in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Mechanical Engineering Pablo A. Tarazaga, Chair J. Wally Grant Mary E. Kasarda Donald J. Leo Michael K. Philen May 6, 2015 Blacksburg, Virginia Keywords: Bioinspired, Cochlear Amplifier, Artificial Hair Cell, Biomimetic Sensor, Nonlinear Sensor, Nonlinear Dynamics, Feedback Control Copyright 2015, Bryan S. Joyce

Development of Active Artificial Hair Cell Sensors€¦ · Figure 1.2. Simplified schematic of the cochlea. Here the spiral has been “unrolled” for visual clarity. From Dallos

  • Upload
    others

  • View
    16

  • Download
    0

Embed Size (px)

Citation preview

  • Development of Active Artificial Hair Cell Sensors

    Bryan Steven Joyce

    Dissertation submitted to the faculty of the

    Virginia Polytechnic Institute and State University

    in partial fulfillment of the requirements for the degree of

    Doctor of Philosophy

    in

    Mechanical Engineering

    Pablo A. Tarazaga, Chair

    J. Wally Grant

    Mary E. Kasarda

    Donald J. Leo

    Michael K. Philen

    May 6, 2015

    Blacksburg, Virginia

    Keywords: Bioinspired, Cochlear Amplifier, Artificial Hair Cell, Biomimetic Sensor,

    Nonlinear Sensor, Nonlinear Dynamics, Feedback Control

    Copyright 2015, Bryan S. Joyce

  • Development of Active Artificial Hair Cell Sensors

    Bryan Steven Joyce

    ABSTRACT

    The cochlea is known to exhibit a nonlinear, mechanical amplification which allows the

    ear to detect faint sounds, improves frequency discrimination, and broadens the range of sound

    pressure levels that can be detected. In this work, active artificial hair cells (AHC) are proposed

    and developed which mimic the nonlinear cochlear amplifier. Active AHCs can be used to

    transduce sound pressures, fluid flow, accelerations, or another form of dynamic input. These

    nonlinear sensors consist of piezoelectric cantilever beams which utilize various feedback

    control laws inspired by the living cochlea. A phenomenological control law is first examined

    which exhibits similar behavior as the living cochlea. Two sets of physiological models are also

    examined: one set based on outer hair cell somatic motility and the other set inspired by active

    hair bundle motility. Compared to passive AHCs, simulation and experimental results for active

    AHCs show an amplified response due to small stimuli, a sharpened resonance peak, and a

    compressive nonlinearity between response amplitude and input level. These bio-inspired

    devices could lead to new sensors with lower thresholds of sound or vibration detection,

    improved frequency sensitivities, and the ability to detect a wider range of input levels. These

    bio-inspired, active sensors lay the foundation for a new generation of sensors for acoustic, fluid

    flow, or vibration sensing.

  • iii

    Acknowledgments

    I would like to thank my advisor, Dr. Pablo Tarazaga. This work would not be possible

    without his insight, guidance, and good humor. I am also grateful for my committee members

    (Dr. Wally Grant, Dr. Mary Kasarda, Dr. Donald Leo, and Dr. Michael Philen) for their

    feedback, technical insight, and career advice.

    I must thank the ME support staff, particularly Beth Howell, Cathy Hill, Linda Vick, and

    all of the guys in the ME machine shop. Our department would grind to a halt without them. I

    would also like to thank my past and current lab mates in the Center for Intelligent Materials

    Systems and Structures (CIMSS); the Vibrations, Adaptive Structures, and Testing (VAST) lab;

    and the Virginia Tech Smart Infrastructure Lab (VT-SIL). I have learned a great deal from the

    brilliant minds around me. I would particularly like to thank Mathieu Vandaele whose help was

    instrumental in the acoustic tests of artificial hair cells.

    Finally, I would like to acknowledge the generous support from the U.S. Department of

    Education GAANN Fellowship (Award No. P200A1000136), the Davenport Fellowship, and my

    departmental research and teaching assistant positions. Their support is gratefully

    acknowledged.

  • iv

    Table of Contents

    Chapter 1. Introduction and Literature Review 1

    1.1. Introduction and Research Motivation............................................................................... 1

    1.2. The Auditory Periphery ..................................................................................................... 2

    1.2.1. Overall Structure ....................................................................................................... 3

    1.2.2. Inner and Outer Hair Cells ........................................................................................ 8

    1.3. The Cochlear Amplifier ................................................................................................... 12

    1.3.1. History of the Cochlear Amplifier .......................................................................... 12

    1.3.2. Characteristics of the Cochlear Amplifier .............................................................. 14

    1.3.3. Mechanisms of Amplification................................................................................. 18

    1.4. Mimicking the Cochlea through Passive Devices ........................................................... 19

    1.4.1. Passive Artificial Hair Cells.................................................................................... 20

    1.4.2. Passive Artificial Basilar Membranes and Cochleae .............................................. 23

    1.5. Mimicking the Cochlea through Active Devices ............................................................. 24

    1.6. Active Artificial Hair Cells .............................................................................................. 28

    1.6.1. Basic Design of Active Artificial Hair Cells .......................................................... 28

    1.6.2. Resonance Based Sensors ....................................................................................... 30

    1.6.3. Feedback Control and Nonlinearity ........................................................................ 31

    1.7. Dissertation Overview ..................................................................................................... 32

    1.7.1. Contributions........................................................................................................... 33

    1.7.2. Chapter Summary ................................................................................................... 34

    Chapter 2. Passive Piezoelectric Hair Cell Models 35

    2.1. Model Derivation: Distributed Parameter Model ............................................................ 35

  • v

    2.1.1. Governing Equations .............................................................................................. 36

    2.1.2. Galerkin-Finite Element Approximation Method ................................................... 40

    2.1.3. Modal Decomposition and Added Damping .......................................................... 43

    2.1.4. One Mode Model .................................................................................................... 47

    2.1.5. One Mode Model for a Bimorph Beam .................................................................. 48

    2.2. Model Derivation: System Identification Approach ........................................................ 51

    2.3. Frequency Response and Tuning Curves ......................................................................... 53

    2.4. Conclusions ...................................................................................................................... 55

    Chapter 3. Models of Artificial Hair Cells using Cubic Damping 56

    3.1. Nonlinear Oscillator at a Hopf Bifurcation ...................................................................... 56

    3.2. Control Law and Closed Loop Response......................................................................... 60

    3.3. Simulations of Cubic Damping Systems ......................................................................... 64

    3.3.1. Harmonic Balance Method ..................................................................................... 65

    3.3.3. Numerical Simulations............................................................................................ 69

    3.4. Effect of Higher Modes ................................................................................................... 75

    3.5. Filtering and Time Delays ............................................................................................... 79

    3.5.1. Simulations with a Butterworth Filter..................................................................... 79

    3.5.2. Time Delays in the Feedback Path.......................................................................... 82

    3.6. Conclusions ...................................................................................................................... 85

    Chapter 4. Experimental Results of Active Artificial Hair Cells using Cubic Damping 86

    4.1. Proof-of-Concept Artificial Hair Cell .............................................................................. 87

    4.1.1. Design ..................................................................................................................... 87

    4.1.2. Model Development................................................................................................ 89

  • vi

    4.1.3. Controller Design .................................................................................................... 92

    4.1.4. Experimental Results .............................................................................................. 93

    4.2. Experimental Tests for an AHC in Fluid ......................................................................... 97

    4.2.1. Preliminary Tests in Water ..................................................................................... 98

    4.2.2. Experimental Setup and Model Development ...................................................... 101

    4.2.4. Experimental Results for the Active AHC in Water ............................................. 110

    4.3. Artificial Hair Cell Accelerometer................................................................................. 112

    4.3.1. Design ................................................................................................................... 113

    4.3.2. Model Development.............................................................................................. 114

    4.3.3. Experimental Results ............................................................................................ 116

    4.3.4. Tuning Curves ....................................................................................................... 122

    4.4. Split Bimorph Artificial Hair Cell Design ..................................................................... 123

    4.4.1. Design ................................................................................................................... 124

    4.4.2. Direct Feedthrough Coupling ............................................................................... 125

    4.4.3. Frequency Response Functions............................................................................. 127

    4.4.4. Numerical Simulations of Active AHC ................................................................ 131

    4.5. Conclusions .................................................................................................................... 133

    Chapter 5. Active Artificial Hair Cells Inspired by Outer Hair Cell Somatic Motility 134

    5.1. Criteria for Implementable Cochlear Models ................................................................ 135

    5.2. Sigmoidal Damping ....................................................................................................... 136

    5.2.1. Model Derivation .................................................................................................. 137

    5.2.2. Numerical Simulations.......................................................................................... 142

    5.3. Feedback through a Tectorial Membrane System ......................................................... 144

  • vii

    5.3.1. Overview ............................................................................................................... 145

    5.3.2. First-order Tectorial Membrane in Feedback Path ............................................... 148

    5.3.3. Second-order Tectorial Membrane in Feedback Path ........................................... 152

    5.3.4. Numerical Simulations.......................................................................................... 157

    5.5. Conclusions .................................................................................................................... 159

    Chapter 6. Active Hair Bundle-based Artificial Hair Cells 161

    6.1. Overview of Active Hair Bundle Motility ..................................................................... 162

    6.2. Active Hair Bundle Model without Inertia .................................................................... 165

    6.3. Active Hair Bundle Model with Inertia ......................................................................... 170

    6.3.1. Model Derivation .................................................................................................. 171

    6.3.2. Behavior of Linearized System and Tuning to a Hopf Bifurcation ...................... 172

    6.3.3. Nonlinear Response from the Harmonic Balance Method ................................... 178

    6.4. Numerical Simulations for Active Hair Bundle-based AHCs ....................................... 180

    6.5. Monostable Active Hair Bundle Model ......................................................................... 185

    6.6. Active Hair Bundle Model Tuned to DC ....................................................................... 188

    6.7. Implementation of Active Hair Bundle Model .............................................................. 190

    6.7.1. Nonlinear stiffness ................................................................................................ 190

    6.7.2. Controller Development........................................................................................ 191

    6.7.3. System Identification by Linear System Approximation...................................... 193

    6.7.4. Nonlinear System Identification through Least Squares Regression .................... 196

    6.8. Conclusions .................................................................................................................... 197

    Chapter 7. Comparisons of Active Artificial Hair Cell Designs 199

    7.1. Cases of Active Artificial Hair Cells ............................................................................. 199

  • viii

    7.1.1. Cubic Damping ..................................................................................................... 201

    7.1.2. Sigmoidal Damping .............................................................................................. 202

    7.1.3. Organ of Corti (OoC) with ζz = 0.1 ....................................................................... 203

    7.1.4. Organ of Corti (OoC) with ζz = 0.01 ..................................................................... 204

    7.1.5. Active Hair Bundle ............................................................................................... 205

    7.2. Performance Metrics ...................................................................................................... 206

    7.2.1. Total Harmonic Distortion .................................................................................... 206

    7.2.2. Settling Time ......................................................................................................... 209

    7.3. Numerical Comparisons between Active AHC Cases ................................................... 210

    7.4. Sensor Recommendations and Conclusions .................................................................. 218

    Chapter 8. Conclusions and Future Work 221

    8.1. Brief Summary of Dissertation ...................................................................................... 221

    8.2. Summary of Contributions ............................................................................................. 223

    8.3. Areas for Future Work ................................................................................................... 224

    8.3.1. Response to Complex Inputs and Stochastic Resonance ...................................... 224

    8.3.2. Self-sensing Hair Cells ......................................................................................... 225

    8.3.3. System Identification and Miniaturization............................................................ 226

    8.3.4. Active Artificial Hair Cell Arrays and Active Artificial Basilar Membranes ...... 227

    Bibliography 228

    Appendix A. Distributed Parameter Model of an Artificial Hair Cell 251

    A.1. Piezoelectric Constitutive Laws .................................................................................... 251

    A.2. Mechanical Domain Equations ..................................................................................... 254

    A.2.1. Free-body Diagram and Newton’s Laws ............................................................. 254

  • ix

    A.2.2. Considerations for the Composite Beam ............................................................. 258

    A.2.3. Piezoelectric Coupling Factor .............................................................................. 263

    A.3. Electrical Domain Equations ........................................................................................ 265

    A.3.1. Charge and Current through a Piezoelectric Actuator ......................................... 265

    A.3.2. Voltage Through a Piezoelectric Sensor .............................................................. 268

    A.3.3. Direct Feedthrough Term from In-plane Coupling .............................................. 269

    A.4. Boundary Conditions .................................................................................................... 273

    A.5. Simplifications for a Bimorph Configuration ............................................................... 273

    A.6. Base Excitation ............................................................................................................. 274

    Appendix B. Galerkin Method for the AHC Distributed Parameter Model 275

    B.1. Overview of the Galerkin Method ................................................................................ 275

    B.2. Galerkin Method for the AHC Model ........................................................................... 276

    B.3. Finite Element Method .................................................................................................. 282

    B.4. Base Excitation .............................................................................................................. 288

    B.5. Forcing Vector from an Applied Pressure ..................................................................... 289

    Appendix C. Modal Decomposition and Frequency Response Functions of the AHC Model

    292

    C.1. Modal Decomposition ................................................................................................... 292

    C.2. Adding Damping ........................................................................................................... 296

    C.3. Frequency Response Functions from General Forcing ................................................. 297

    C.4. Open Circuit Voltage .................................................................................................... 299

    C.5. Frequency Response Functions from Base Accelerations ............................................ 300

    C.6. Frequency Response Functions from Piezoelectric Actuator ....................................... 302

  • x

    C.7. Frequency Response Functions from Plane Wave Pressure ......................................... 303

    Appendix D. Acoustic Tests of Passive Artificial Hair Cells 305

    D.1. PZT Artificial Hair Cell Design and FRF ..................................................................... 305

    D.2. PZT Artificial Hair Cell Tuning Curves ....................................................................... 307

    D.3. PVDF Artificial Hair Cell ............................................................................................. 311

    D.4. Comparison to Biology ................................................................................................. 312

  • xi

    List of Figures

    Figure 1.1. Diagram of the ear. From Dallos (1992),with permission of The Journal of

    Neuroscience [21]. .......................................................................................................................... 4

    Figure 1.2. Simplified schematic of the cochlea. Here the spiral has been “unrolled” for visual

    clarity. From Dallos (1992),with permission of The Journal of Neuroscience [21]. ..................... 5

    Figure 1.3. Cross-section of the organ of Corti on the basilar membrane. From Raphael and

    Atlschuler (2003),with permission of Elsevier Limited [30]. ......................................................... 7

    Figure 1.4. Rows of hair cell stereocilia. From Raphael and Atlschuler (2003), reproduced with

    permission of Elsevier Limited [30]. .............................................................................................. 9

    Figure 1.5. Diagrams of the inner and outer hair cells. From Dallos (1992),with permission of

    The Journal of Neuroscience [21]. ................................................................................................ 10

    Figure 1.6. Measurements showing the cochlear amplifier in a guinea pig cochlea. (a) Basilar

    membrane (BM) displacement of versus frequency and sound pressure level. (b) Basilar

    membrane displacement normalized by the input sound pressure level. These curves would

    overlap for a linear system. Figure from Johnstone et al. (1986), reproduced with permission of

    Elsevier Limited [40]. ................................................................................................................... 15

    Figure 1.7. Displacement of the basilar membrane versus sound pressure level. The

    displacement shows a linear trend at low and high sound pressures levels and a nonlinear

    compression at intermediate sound pressure. Figure from Johnstone et al. (1986), reproduced

    with permission of Elsevier Limited [40]. ................................................................................... 16

    Figure 1.8. Schematic of a simple, active artificial hair cell. (a) Physical layout of the artificial

    hair cell (AHC). (b) Block diagram of the closed-loop system. .................................................. 29

    Figure 1.9. Frequency response function for an example sensor. ................................................ 30

  • xii

    Figure 2.1. Cantilever beam with piezoelectric elements (actuators or sensors). ........................ 36

    Figure 2.2. Piezoelectric bimorph beam. Arrows on the piezoceramic indicate their polarization

    direction. ....................................................................................................................................... 48

    Figure 2.3. First natural frequency of bimorph beam as a function of beam length. The plot uses

    parameters from the small scale artificial hair cell discussed later in Chapter 4. A 1 mm change

    in length changes the natural frequency by about 36 Hz (about 8% change). .............................. 52

    Figure 2.4. Example response of a single degree of freedom system. (a) System response versus

    input level and frequency. (b) Response amplitude versus frequency for different input levels.

    (c) Response amplitude versus input amplitude at resonance. (d) Input level versus frequency for

    different response levels, i.e. tuning curves. ................................................................................. 54

    Figure 3.1. Sample time responses of the prototypical Hopf bifurcation system. (a) μ = 1 > 0.

    (b) μ = -1 < 0. For both systems, b = 1, c = 2π 10 rad/s, and the initial condition is z(0) = 0.5. 58

    Figure 3.2. Sample phase portraits of the prototypical Hopf bifurcation system. (a) μ = 1 > 0.

    (b) μ = -1 < 0. For both systems, b = 1, c = 2π 10 rad/s, and the initial condition is z(0) = 0.5. 58

    Figure 3.3. Example frequency response of the prototypical Hopf bifurcation system. Here μ =

    0, b = 1, and c = 1 rad/s............................................................................................................... 60

    Figure 3.4. Time response at resonance ( 1 ) of a linear versus nonlinear example. For the

    linear system, 01.0 . For the nonlinear case, a3= 1x10-4

    . ......................................................... 70

    Figure 3.5. Example frequency response of a cubic damping system for various excitation

    amplitudes. (a) Magnitude of the response. (b) Response amplitude normalized by the input

    amplitude. For the nonlinear system, = 0 and a3 = 1x10-4

    . ....................................................... 71

    Figure 3.6. Example frequency response of a cubic damping system for various values of a3.

    Here = 0 and F = 1. .................................................................................................................... 71

  • xiii

    Figure 3.7. Response amplitude versus input amplitude at = 1 for various values of a3. Here

    = 0 for the nonlinear system. A linear system with ζ = 0.01 is shown for comparison. .............. 72

    Figure 3.8. Response amplitude versus input amplitude for both the linear and nonlinear systems

    at (a) = 0.999 and (b) = 0.99. For the nonlinear cases, = 0. .............................................. 74

    Figure 3.9. Response amplitude versus input amplitude at = 1 for (a) = 0.001 and (b) =

    0.01................................................................................................................................................ 74

    Figure 3.10. Effect of multiple modes on the frequency response. ............................................. 78

    Figure 3.11. Simulation results showing a limit cycle oscillation for the two mode system given

    a small initial displacement. .......................................................................................................... 78

    Figure 3.12. Block diagram of the closed-loop system with a filter to reduce spillover. ............ 79

    Figure 3.13. Frequency response functions (FRFs) for examples Butterworth filters. The

    examples use corner frequencies (c) of 3 and 5 and filter order (n) of 2 and 6. ......................... 80

    Figure 3.14. Effect of the using a Butterworth filter in the feedback loop. The filter parameters

    are (a) c = 3, n = 2; (b) c = 3, n = 6; (c) c = 5, n = 2; and (d) c = 5, n = 6. ...................... 81

    Figure 3.15. Effect of a time delay in the feedback path. (a) Peak response versus input

    excitation for different time delays T. (b) Backbone curves for different time delays. These plots

    are generated from the harmonic balance method in Equations 3.46 and 3.47. For these curves,

    a3 = 41.7, ζ = 0.1, and a1 = 2ζ = 0.2. The linear case (a1 = a3 = 0) is shown for reference. ........ 84

    Figure 4.1. Proof-of-concept artificial hair cell design. (a) Photograph of the experimental

    setup. (b) Schematic of the experiment. ....................................................................................... 88

    Figure 4.2. Velocity from control voltage frequency response functions for the proof-of-concept

    AHC from the data, a single degree of freedom fit (SDOF Fit) around the first mode, and a finite

    element model (FE Model) with one mode and with five modes. ................................................ 92

  • xiv

    Figure 4.3. Control algorithm for the AHC. The cantilever beam transfer function is defined in

    Equation 4.2. ................................................................................................................................. 93

    Figure 4.4. Frequency response of the proof-of-concept artificial hair cell for different

    disturbance levels. The left column shows model predictions, while the right column shows

    experimental results. ..................................................................................................................... 95

    Figure 4.5. Amplitude of the tip velocity versus disturbance level at the resonance frequency

    (10.8 Hz). ...................................................................................................................................... 95

    Figure 4.6. Experimental setup for preliminary water tests. An aluminum cantilever beam is

    partially submerged in water. ........................................................................................................ 98

    Figure 4.7. Magnitude of the frequency response function for different water depths. ............ 100

    Figure 4.8. Variation in natural frequency and damping of the first mode with increasing water

    depth. 95% confidence intervals are also shown at each data point. ......................................... 100

    Figure 4.9. Artificial hair cell in water experimental setup. (a) Photograph and (b) schematic of

    the experiment. ............................................................................................................................ 101

    Figure 4.10. Frequency response function (FRF) of velocity to control voltage for the beam in

    air. ............................................................................................................................................... 104

    Figure 4.11. Schematic of the artificial hair cell with an added tip mass and viscous damper to

    account for the added inertia and damping due to the fluid damping. ........................................ 105

    Figure 4.12. Velocity from control voltage frequency response function for the passive sensor

    partially submerged in water. The figure shows the FRF from the data (Data), single degree of

    freedom fit (SDOF fit), and the finite element (FE) model. ....................................................... 108

    Figure 4.13. FRF of velocity to disturbance voltage for the beam in air and in water. ............. 109

  • xv

    Figure 4.14. Frequency response function of velocity from disturbance voltage for the passive

    artificial hair cell partially submerged in water. The figure shows the data (Data), single degree

    of freedom fit (SDOF Fit), and finite element model (FE Model). ............................................ 110

    Figure 4.15. Model and experimental frequency response functions (FRFs) for the active

    artificial hair cell in water. The FRFs are plots of the velocity at the measurement point from

    disturbance signals at different voltage levels and frequencies. ................................................. 111

    Figure 4.16. (a) Input-to-output relationship for the disturbance voltage to the velocity at the

    natural frequency. (b) Fit of the damping ratio versus the amplitude of the velocity. ............... 112

    Figure 4.17. Artificial hair cell accelerometer. The AHC consists of a piezoelectric bimorph

    beam under a base excitation. ..................................................................................................... 113

    Figure 4.18. Velocity to control voltage FRFs for the passive AHC accelerometer from the data,

    a single degree of freedom fit (SDOF Fit) around the first mode using the system identification

    method, and a finite element model (FE Model). Here a 1 V control signal was used. ............ 115

    Figure 4.19. Velocity from base acceleration FRFs of the passive AHC accelerometer. (a)

    Magnitude of the velocity versus frequency for different voltages to the shaker. (b) FRFs of

    velocity with respect to the base acceleration for different shaker voltages............................... 116

    Figure 4.20. Velocity from base acceleration FRFs of the active AHC accelerometer. (a) and (b)

    show the magnitude of the velocity versus frequency for different base accelerations. (c) and (d)

    show the velocity normalized by the base acceleration for different shaker voltages. (a) and (c)

    show results from the model, while (b) and (d) show results from the experiment. .................. 117

    Figure 4.21. (a) Amplitude compression for the active and passive AHC accelerometer. (b)

    Backbone curves for the active and passive AHC. For reference, an ideal model at the Hopf

    bifurcation (μ = 0) is also shown. ............................................................................................... 119

  • xvi

    Figure 4.22. Butterworth filter with a corner frequency of 1500 Hz and filter order n = 2 and n =

    6................................................................................................................................................... 120

    Figure 4.23. Experimental results for the active AHC accelerometer using a poor filter design.

    (a) Magnitude of the velocity versus frequency for different voltages to the shaker. (b) Velocity

    normalized by the base acceleration for different shaker voltages. ............................................ 121

    Figure 4.24. Experimental input-to-output and backbone curves for the AHC using a poor filter

    design. (a) Amplitude compression for the active and passive AHC. (b) Backbone curves for the

    active and passive AHC. ............................................................................................................. 121

    Figure 4.25. Tuning curves for the small scale artificial hair cell. The velocity threshold is set to

    1.5 mm/s. ..................................................................................................................................... 123

    Figure 4.26. “Split” configuration bimorph beam. (a) Schematic of the AHC. Arrows on the

    piezoceramic indicate the polarization direction. (b) Test setup. .............................................. 124

    Figure 4.27. Frequency response functions for the split beam sensor. (a) Tip velocity from a

    base acceleration. (b) Velocity from control voltage. (c) Voltage in the sensing element from a

    base acceleration. (d) Sensing voltage from the control voltage. ............................................... 130

    Figure 4.28. Simulation results of the compensated voltage of the split bimorph AHC versus

    frequency for different input accelerations. For comparison, the responses of the system with

    and without the controller are shown. ......................................................................................... 132

    Figure 4.29. Compensated voltage versus input acceleration at resonance. ............................... 132

    Figure 5.1. Schematic of a simple, active artificial hair cell. (a) Physical layout of the artificial

    hair cell (AHC). (b) Block diagram of the closed-loop system. ................................................ 135

    Figure 5.2. (a) Diagram of a hair bundle. (b) Open channel probability (PO) as a function of the

    stereocilia bundle’s deflection y. ................................................................................................. 137

  • xvii

    Figure 5.3. Frequency response functions of an artificial hair cell with sigmoidal velocity

    feedback. ..................................................................................................................................... 143

    Figure 5.4. Output-input relationship at resonance for the system without a controller (a linear

    system with ζ = 0.1), cubic velocity feedback tuned to the Hopf bifurcation, and the sigmoidal

    velocity feedback tuned to the Hopf bifurcation. ........................................................................ 143

    Figure 5.5. Cross-section of the organ of Corti on the basilar membrane. From Raphael and

    Atlschuler (2003),with permission of Elsevier Limited [30]. ..................................................... 146

    Figure 5.6. Simplified relationship of the components of the organ of Corti ............................. 146

    Figure 5.7. Frequency response functions for various first-order tectorial membrane systems. (a)

    zy . (b) zy . Parameter values and the input force to the tectorial membrane system are

    given in the figure legend. .......................................................................................................... 151

    Figure 5.8. Frequency response functions for various second-order tectorial membrane systems.

    (a) Stereocilia deflection proportional to the tectorial membrane displacement (y = z). (b)

    Stereocilia deflection proportional to tectorial membrane velocity ( zy ). The legend gives

    the form of the input force on the tectorial membrane system. For all of these cases, the tectorial

    membrane damping ζz is 0.01. .................................................................................................... 155

    Figure 5.9. Frequency response functions between displacement x and stimulus f for an active

    artificial hair cell with a tectorial membrane system. Here bf = 1, ζ = 0.1, ζz = 0.1, = 1, δ = 0.1,

    and a = 0.16. The parameters are chosen to achieve a Hopf bifurcation and an equivalent cubic

    damping coefficient a3 is 41.7. ................................................................................................... 158

    Figure 5.10. Displacement x to forcing f frequency response functions for the sensor with a

    smaller bandwidth of amplification (ζz = 0.01). Here bf = 1, ζ = 0.1, ζz = 0.01, = 1, δ = 1, and

  • xviii

    a = 0.016. The FRF for F = 1x10-1

    overlaps the linear response. The parameters are chosen to

    achieve a Hopf bifurcation and an equivalent cubic damping coefficient a3 is 41.7. ................. 159

    Figure 6.1. Diagram of a hair bundle. ......................................................................................... 163

    Figure 6.2. Data showing the nonlinear stiffness of a hair bundle from a bullfrog saccular hair

    cell. Image from Martin, et al (2000). Copyright (2000) National Academy of Sciences, U.S.A

    [62]. ............................................................................................................................................. 163

    Figure 6.3. Diagram of a hair cell’s adaptation motor. ............................................................... 164

    Figure 6.4. Frequency response functions of several example, linear, third-order systems. The

    system parameters are listed in the legend key. Note bf is set to one for simplicity. ................. 177

    Figure 6.5. Magnitude and phase of the FRF at resonance ( ) as a function of the third pole

    ρ. Here ξ = 0.1 and = 1............................................................................................................ 178

    Figure 6.6. Comparison of FRFs computed by harmonic balance method (HBM) and by ode45

    numerical solution. Both methods show nearly identical results. Here β = 1, ζ = 0.1, = 0, and

    F = 1x10-2

    . ................................................................................................................................... 181

    Figure 6.7. Harmonic balance method results of an active hair bundle. Here β = 1, ζ = 0.1, and

    = 0. .............................................................................................................................................. 182

    Figure 6.8. (a) Input-to-output relationship and (b) backbone curve with β = 1 and = 0. The

    model shows a compressive nonlinearity with a slope of 0.33 dB/dB and a small deviation of the

    resonance frequency with response amplitude. .......................................................................... 183

    Figure 6.9. FRF for the active hair bundle model for varying β. The forcing input (bf F) was kept

    at 1x10-3

    . ..................................................................................................................................... 184

    Figure 6.10. Numerical results of an active hair bundle sensor tuned to a new resonance

    frequency of = 2 . Here β = 1, ζ = 0.1, and = 0. ................................................................ 185

  • xix

    Figure 6.11. Frequency response functions of a monostable active hair bundle with = 2 and

    β = 1. As before, ζ = 0.1, and = 0. ............................................................................................ 188

    Figure 7.1. Example signals and their frequency spectra. (a) Example signals as a function of

    time. (b) Magnitude of the Fourier transforms of the signals in (a). (c) Example signals of the

    form sin(2πt)+THD sin(6πt). (d) Magnitude of the Fourier transforms of the examples in (c). 208

    Figure 7.2. Example of computing the settling time. .................................................................. 210

    Figure 7.3. Frequency response functions for different active AHC cases. Here the input

    amplitude (bf F) was 1x10

    -3. The model parameters re set such that the systems are tuned to the

    bifurcation (a1 = 2ζ) and have an equivalent cubic damping coefficient (a3) of 41.7. A linear

    system, Equation 7.2 with no control force, is also shown for comparison. Note the cubic

    damping and sigmoidal damping cases overlap in these plots. .................................................. 211

    Figure 7.4. Input-to-output relationship at resonance ( = 1) for the different active artificial hair

    cell cases. The cubic damping and active hair bundle models overlap. The plots for the

    sigmoidal damping and two organ of Corti (OoC) cases also overlap. ...................................... 212

    Figure 7.5. Input-to-output relationship at resonance ( = 1) for active AHC models that have

    been mistuned (μ ≠ 0). ................................................................................................................ 213

    Figure 7.6. Displacement of the basilar membrane versus sound pressure level. Figure from

    Johnstone et al. (1986), reproduced with permission of Elsevier Limited [40]. ....................... 214

    Figure 7.7. Maximum control force (buU) for the active AHC cases driven at resonance ( = 1).

    As with the input-to-output curve, the cases with saturating nonlinearity (sigmoidal damping and

    the two OoC cases) require the same maximum output force. ................................................... 215

  • xx

    Figure 7.8. Total harmonic distortion (THD) in decibels for the active AHC cases driven at

    resonance ( = 1). The THD for the cases with saturating control forces nearly overlap. For

    comparison, a line for 2% (-34 dB) distortion is also shown...................................................... 216

    Figure 7.9. 1% settling time in cycles (Ns) for the active AHC cases driven at resonance ( = 1).

    ..................................................................................................................................................... 217

    Figure 8.1. A simple circuit for self-sensing, piezoelectric actuators. ........................................ 226

    Figure A.1. Cantilever beam with piezoelectric elements (actuators or sensors). ...................... 255

    Figure A.2. Forces and moments acting on a differential element of a beam. ........................... 255

    Figure A.3. Deformation of a differential element of a beam. ................................................... 255

    Figure A.4. Cross-section of the composite beam. The arrows over the piezoelectric elements

    indicate the direction of polarity. ................................................................................................ 259

    Figure A.5. “Split” configuration bimorph beam. Arrows on the piezoelectric actuator and

    sensor indicate the polarization direction. .................................................................................. 269

    Figure B.6. Element of a beam between nodes xi and xi+1. ......................................................... 283

    Figure B.7. Shape functions for an Euler-Bernoulli beam. ......................................................... 284

    Figure D.1. Experimental setup for acoustic excitation of the piezoelectric bimorph. ............. 306

    Figure D.2. Frequency response function (FRF) for the bimorph under acoustic excitation. (a)

    FRF in dB over 100 Hz to 10,000 Hz frequency sweep. (b) FRF around the first natural

    frequency (shown here in a linear scale). .................................................................................... 306

    Figure D.3. (a) Schematic and (b) photograph of the experimental setup for generating tuning

    curves for acoustic excitation...................................................................................................... 308

    Figure D.4. PID control algorithm for generating tuning curves. ............................................... 309

  • xxi

    Figure D.5. Tuning curves from the PZT artificial hair cell. The left column shows curves of

    constant velocity, and the right column plots curves of constant voltage. The bottom row shows

    the tuning curves in the top row normalized by the response level. ........................................... 310

    Figure D.6. PVDF bending sensor. ............................................................................................ 311

    Figure D.7. Tuning curves from the PVDF artificial hair cell. (a) Velocity tuning curve. (b)

    Velocity tuning curve normalized by the velocity level. ............................................................ 312

    Figure D.8. Comparison of tuning curves from (a) PZT AHC sensor and (b) biological

    measurements from a guinea pig cochlea. Biological tuning curves are from Sellick et al.

    (1982), with permission of J. Acoust. Soc. Am. [135]. ............................................................... 313

  • xxii

    List of Tables

    Table 3.1. Parameters for the two mode example in Figure 3.10. These parameters are based on

    the small scale AHC in Chapter 4. ................................................................................................ 77

    Table 4.1. Properties of the proof-of-concept AHC setup. ........................................................... 88

    Table 4.2. System parameters for the proof-of-concept AHC based on finite element (FE) and

    system identification (ID) modeling techniques. Damping for the FE model was assumed from

    the curve-fit data. .......................................................................................................................... 91

    Table 4.3. Curve fit parameter values from the model and the data of the proof-of-concept AHC

    results in Figure 4.5. Only data for disturbance amplitudes between 5 and 40 V were used. The

    units of c are mm/s/Vk where k is the compression value in the table. ......................................... 96

    Table 4.4. Properties of the artificial hair cell in water setup. ................................................... 102

    Table 4.5. System identification results for the sensor in air and in water. Most of the

    parameters were estimated from a single degree of freedom fit from the velocity to control FRF.

    The disturbance influence term d1 was estimated from the velocity to disturbance FRF. ........ 108

    Table 4.6. System parameters for the AHC accelerometer based on FE model and system

    identification modeling techniques. ............................................................................................ 115

    Table 4.7. Curve fit parameter values from the AHC accelerometer data in Figure 4.21a. The

    units of c are mm/s/gk where k is the compression value in the table. ........................................ 119

    Table 5.1. Equivalent cubic velocity coefficients for first-order tectorial membrane systems in

    the feedback path. ....................................................................................................................... 152

    Table 5.2. Equivalent cubic velocity coefficients for second-order tectorial membrane systems.

    ..................................................................................................................................................... 156

    Table 7.1. Summary of active artificial hair cell (AHC) models. ............................................... 220

  • xxiii

    Table D.1. PID gains for input frequency for the PZT sensor. ................................................... 309

  • 1

    Chapter 1. Introduction and Literature Review

    This chapter begins by outlining the mechanisms behind hearing in vertebrates and the

    research into mimicking these processes. Next the motivation for active artificial hair cells is

    provided. This chapter concludes by highlighting the major contributions of this work and by

    providing an outline of this dissertation.

    1.1. Introduction and Research Motivation

    Biology has provided inspiration for a number of technologies. Engineers are turning to

    nature for solutions to difficult problems in locomotion, material design, signal processing,

    sensor design, control, and a host of other fields [1-3]. Hearing is one biological mechanism

    which has seen research interest in the past few decades. Several sophisticated components work

    together to give mammals the ability to detect a remarkable range of frequencies and sound

    pressure levels. Healthy human ears can detect a range of frequencies between 20 Hz and 20,000

    Hz [4, 5]. Whales and some species of bats have hearing ranges as high as 200,000 Hz [6].

    Humans can detect sound pressure levels as low as 0 dB (20 μPa RMS sound pressure) and up to

    120 dB (20 Pa RMS) before severe pain occurs. Other mammals, such as cats, can hear as low

    as -15 dB (4 μPa RMS) [4].

    There has been research interest aimed at developing bio-inspired devices which could

    one day replace damaged components of the ear [7-9]. The cochlea is the spiral-shaped portion

    of the mammalian auditory system responsible for transducing sound into electrical signals. The

    cochlea is fully developed upon birth, and its constitutive components are not repaired after they

  • 2

    are damaged [10]. Loss of cochlear hair cells can cause a profound loss in hearing [11-14]. By

    replacing these hair cells with engineered, artificial hair cells could provide some recovery of

    auditory function.

    The cochlea and its hair cells have inspired a number of novel sensor designs. As

    discussed in the next sections, the cochlea is able to detect minute, sound-induced vibrations

    comparable to the Brownian motion of atoms [15]. The cochlea decomposes complex sounds

    into its different frequency components like a Fourier analyzer before sending its electrical

    signals to the brain. Researchers would like to produce acoustic, fluid flow, orientation, and

    vibration sensors which mimic the cochlea’s incredible sensitivity to small input levels and high

    frequency selectivity.

    1.2. The Auditory Periphery

    This section produces an overview of the anatomy and physiology of the auditory

    periphery (the outer, middle, and inner ears) in mammals. These components work together to

    transmit sound information to the cochlea where it is transduced into electrical signals sent to the

    brain. While the central nervous system and the brain form an important component of how

    mammals perceive sound, they are excluded from the discussion presented here (see Schnupp et

    al. for more information [16]). More detailed information about hearing in mammals and other

    animals can be found in [4-6, 16-20].

  • 3

    1.2.1. Overall Structure

    The auditory periphery is often divided into three regions: the outer, the middle, and the

    inner ear [4, 5]. Figure 1.1 shows a diagram of the middle and inner ear. The outer ear consists

    of the pinna and the ear canal (not shown in Figure 1.1). The pinna (also known as the auricle) is

    the portion of the ear on the outside of the head and is responsible for directing sound into the ear

    canal. Sound waves travel down the ear canal to the tympanic membrane (also known as the ear

    drum). The middle ear consists of three connected bones called the ossicles (the malleus, the

    incus, and the stapes) in a space of air called the tympanic cavity. The malleus connects to the

    tympanic membrane while the stapes connects to the oval window, a small membrane on the

    cochlea. The ossicles and their supporting ligaments transmitted sound from the tympanic

    membrane to the oval window. The middle ear structure provides a pressure amplification due to

    the lever action of the ossicles and the decrease in area from the tympanic membrane to the

    smaller oval window. The resulting twenty-fold increase in pressure from the tympanic

    membrane to the oval window creates an acoustic impedance matching scheme [19]. This

    allows for a more efficient transfer of sound waves from the air in the ear canal to the fluid in the

    cochlea. In addition, ligaments connected to these bones will tension under loud sound pressure

    levels [4]. This acoustic reflex attenuates sound transmission to the cochlea and offers some

    protection to the cochlea from dangerous sound pressure levels. The Eustachian tube connects

    the middle ear to the nasal passages and aids equalizing the pressure between the middle ear and

    atmosphere and in draining mucus from the middle ear.

  • 4

    Figure 1.1. Diagram of the ear. From Dallos (1992),with permission of The Journal of

    Neuroscience [21].

    The inner ear consists of the cochlea and the vestibular system. The vestibular system is

    responsible for determining linear and rotational accelerations. These aspects are important for

    balance and sensing spatial orientation. The vestibular system consists of the semicircular canals

    (which determine rotational accelerations) and the otolithic organs (which sense linear

    accelerations). The spiral portion of the inner ear is the cochlea. The cochlea is responsible for

    transducing the sound-induced vibrations into electrical signals.

    Figure 1.2 shows a schematic of the cochlea uncoiled. In humans the cochlea makes a

    little more than 2.5 turns. The human cochlea is about 35 mm long and has a radius around 1

    mm [6, 19, 22]. Sound-induced vibrations of the stapes push on the oval window and generate

    pressure waves in the cochlear fluid. The cochlea is divided by two membranes (the basilar

    membrane and the Reissner’s membrane) to form three fluid-field chambers (the scala tympani,

    the scala media, and the scala vestibuli). At the apex of the spiral the scala tympani and the scala

  • 5

    vestibuli merge at a location called the helicotrema. The scala tympani and scala vestibuli

    contain an ionic fluid called perilymph which is low in potassium ions and high in sodium ions

    [23]. The scala media (also called the cochlear duct) contains endolymph, another ionic fluid

    which is high in potassium ions and low in sodium ions.

    Figure 1.2. Simplified schematic of the cochlea. Here the spiral has been “unrolled” for

    visual clarity. From Dallos (1992),with permission of The Journal of Neuroscience [21].

    The width of the basilar membrane increases along the length of the cochlea from 0.1 mm

    at the base to 0.4 mm at the apex [6, 19, 22]. Its thickness decreases from 13 μm at the base to 5

    μm at the apex. Stiffness measurements of the basilar membrane also indicate that the constitute

    fibers of the membrane are stiffer toward the base of the cochlea compared to the apex [22, 24,

    25]. The changing geometry and fiber stiffness give the basilar membrane a spatially varying

    stiffness. The result is that a tone of a particular frequency will induce a larger amplitude of

    vibration in certain location than elsewhere. Thus the frequency of the stimulus can be mapped

    to a particular location along the length of the cochlea. Low frequencies cause larger vibrations

    near the apex of the cochlea while high frequencies induce larger vibrations near the stapes. This

  • 6

    tonotopic mapping allows the cochlea to decompose complex signals into its frequency

    components and allows the ear to distinguish between different frequencies. Researchers have

    shown that this mechanism is highly efficient at performing Fourier analysis on complex signals

    compared to traditional discrete and fast Fourier transform algorithms [26, 27].

    Another byproduct of tonotopic mapping is the appearance of a traveling wave along the

    basilar membrane. When a pure tone is applied, the magnitude and phase of the basilar

    membrane create the appearance of a traveling wave moving from the stapes toward the apex.

    This traveling wave peaks in a region whose location is a function of the stimulus frequency.

    There is still some debate whether the cochlea experiences a true traveling wave (which carries

    energy along the length of the cochlea) or if this is a pseudo-traveling wave with propagation of

    energy [28, 29].

    Figure 1.3 shows a diagram of the interior of the cochlea. Sitting on the basilar

    membrane inside the scala media is a collection of cells called the organ of Corti. The organ of

    Corti and the basilar membrane together are referred to as the cochlear partition as they are the

    main structural division in the cochlea. Reissner’s membrane is believed to have little influence

    on the propagation of waves inside the cochlea and mainly serves to separate the endolymph in

    the scala media from the perilymph of the scala vestibuli [4, 16]. The endolymph is generated by

    a collection of the blood vessels in the stria vascularis.

  • 7

    Figure 1.3. Cross-section of the organ of Corti on the basilar membrane. From Raphael

    and Atlschuler (2003),with permission of Elsevier Limited [30].

    The organ of Corti contains a set of sound-sensing cells called hair cells (the next section

    provides more details on these cells). The tectorial membrane overlays the organ of Corti. The

    tips of the stereocilia of the outer hair cells are connected to the tectorial membrane, while the

    stereocilia of the inner hair cells are not connected. The pillar cells, the Deiter’s cells, and the

    Hensen’s cells serve various roles in supporting the other cells and giving rigidity to the overall

    structure. As discussed in the next section, when fluid motion causes the stereocilia of the inner

    hair cells to deflect, the inner hair cells stimulate neighboring afferent neurons which in turn

    generates a neural spike, or action potential. The bodies of these afferent neurons rest in the

    spiral ganglion which sits inside a conical shaped medulla at the center of the cochlea’s spiral.

    These neural spikes propagate through the neurons of the vestibulocochlear nerve to the brain.

  • 8

    1.2.2. Inner and Outer Hair Cells

    Mammals possess two types of hair cells: inner hair cells and outer hair cells. Humans

    have around 32,000 hair cells in the two cochleae (about 8,000 inner hair cells and 24,000 outer

    hair cells) [6]. The body (or soma) of outer hair cells is around 20 μm tall for cells near the basal

    end of the cochlea and around 50 μm tall for cells at the apex [19]. The size of the inner hair

    cells has more variation, but they are generally longer near the apex than at the basal end of the

    cochlea. Both types of hair cells possess bundles of 20 to 300 stereocilia. Each stereocilia has a

    diameter around 0.2 μm for most of its length, but tappers to less than 0.05 μm at its root in the

    hair cell. Stereocilia lengths vary between 2 μm and 6 μm. The stereocilia are hexagonally

    packed into a V-shaped pattern, and the height of the stereocilia varies linearly across the bundle

    (see Figure 1.4). This geometry and the cross-links between stereocilia cause the bundle to be

    more compliant along the axis of symmetry than in the orthogonal direction. Filament structures

    called tip links connect the tips of adjacent stereocilia in the bundle. The tip links are connected

    at one or both ends to ion channels which open when the tip links are under tension. The

    basolateral side of the hair cells (the portion of the cell wall at the base and sides of the cell) is

    bathed in perilymph with a resting electric potential around 0 mV [21]. The interior of cell body

    has resting potential between -40 mV for the inner hair cells and -70 mV for the outer hair cells.

    The stereocilia is surrounded by endolymph in the scala media with a resting potential around 80

    mV to 100 mV. The result is a voltage difference between 120-160 mV between the endolymph

    around the stereocilia and interior of the cell. Upon deflection of the hair cell bundle toward the

    tallest stereocilia, the tip links are stretched, the ion channels open, and the voltage difference

    drives positively charged potassium ions into the cell. Increasing deflection toward the largest

  • 9

    stereocilia further increases the influx of ions into the cell. Deflection away from the tallest

    stereocilia causes these channels to close and decreases the flow of ions into the cell. This influx

    of ions (i.e. current) increases the voltage inside the cell. When a constant displacement is

    applied to the stereocilia, the current into the cell slowly decreases, or adapts, to a lower level.

    This adaptation process is thought to result from the action of myosin motors which adjust the

    tension in the tip links [20, 31]. Measurements show this adaptation process is dependent on the

    concentration of calcium ions around the hair cells [18, 32].

    Figure 1.4. Rows of hair cell stereocilia. From Raphael and Atlschuler (2003),

    reproduced with permission of Elsevier Limited [30].

    Mammals possess two types of hair cells: inner hair cells and outer hair cells. Figure 1.5

    shows schematics of these hair cell types. The two types of hair cells are characterized by their

    location in the cochlea and their function. The inner hair cells are aligned in a single row along

    toward the inside of the cochlea’s spiral, while the outer hair cells are arranged in three rows

  • 10

    further away from the center of the spiral. For the inner hair cells, the depolarization caused by

    hair bundle deflection opens voltage-gated calcium channels, which in turn activates a release of

    glutamate, a neurotransmitter, at the base of the cell. These neurotransmitters cross the space

    between the hair cell and a nearby afferent nerve cell, trigger a depolarization of the afferent

    neuron, and starts an electrical nerve signal which propagates along the auditory nerve to the

    brain. The resulting electrical pulses encode information about the intensity, duration, and

    frequency of the resulting mechanical stimulus. Therefore, the inner hair cell serves to transduce

    mechanical induced vibration into electrical signals. Increasing stereocilia deflection toward the

    tallest stereocilia increases the rate of glutamate release, which increases the firing rate of the

    spiral ganglion cells. If a particular section of the basilar membrane vibrates more than the rest,

    then there are more neural spikes generated from that region. In this manner, the neurons encode

    sound intensity at a particular frequency into a neural firing rate.

    Figure 1.5. Diagrams of the inner and outer hair cells. From Dallos (1992),with

    permission of The Journal of Neuroscience [21].

  • 11

    Outer hair cells do not stimulate afferent neurons like their inner hair cell counterparts.

    The outer hair cell possesses a motor protein called prestin embedded in their cell walls. When

    there is a voltage change across the cell’s membrane, the prestin protein changes configuration

    [33]. This process causes the cell wall to deform and the body of the outer hair cell to contract.

    These forces push on the basilar membrane and the tectorial membrane. The result is that the

    outer hair cells function like actuators by producing mechanical forces upon application of a

    voltage change. This process is referred to as somatic motility or electromotility and serves to

    boost sound-induced vibration. This increased vibration in turn causes larger deflections of the

    inner hair cell stereocilia and thus amplifies the perception of weak sound pressure levels. While

    outer hair cells are unique to mammals, other animals have also evolved a secondary type of hair

    cell [34-37]. These animals have a high sensitivity and an increased frequency range than those

    with just sensing hair cells.

    The inner and outer hair cells are innervated by afferent and efferent neurons. However,

    the inner hair cells are predominately innervated by afferent neurons which transmit electrical

    signals to the brain. The outer hair cells are innervated primarily by efferent, or motor, neurons.

    Research shows that these efferent neurons cause an inhibitory effect on the outer hair cells

    under the presence of loud sounds [18]. The result is that the brain can “turn down” the gain of

    the cochlear amplifier when it detects high sound pressure levels.

  • 12

    1.3. The Cochlear Amplifier

    This section begins by providing a brief history of the discovery and importance of the

    cochlear amplifier. Next some important characteristics of the cochlear amplifier are discussed

    and how they aid in sound detection.

    1.3.1. History of the Cochlear Amplifier

    In 1928, Georg von Békésy performed some of the first experiments on the cochleae of

    human cadavers [24, 38]. He applied a silver speckle pattern to the basilar membrane and

    observed its motion using a strobe light. While he played tones through a loud speaker, von

    Békésy observed a traveling wave moving longitudinally along the basilar membrane. The

    amplitude of the traveling wave reached a maximum at a position along the length of the basilar

    membrane dependent upon the frequency of the tone. Because of the limited sensitivity of

    optical measurements from that era, von Békésy had to apply sound pressures greater than 100

    dB in order to detect the displacement of the basilar membrane [4]. von Békésy’s work earned

    him the 1961 Nobel Prize in Physiology or Medicine.

    However, the early work of von Békésy and others with cadavers could not explain the

    high pressure sensitivity and sharp frequency tuning seen in responses from auditory nerve fibers

    of living mammals [39]. In addition to the “first filter” of BM’s traveling wave, a “second filter”

    was proposed to aid in frequency selectivity of the cochlea. While the first filter was

    mechanical, the second filter was thought to be electrical in nature. However theories of how the

    proposed second filter works were based on passive filtering and could not account for the

  • 13

    observed responses of the healthy cochlea. Measurements inside the cochlea suggested an active

    mechanical feedback was at work [40].

    The idea of an active mechanism inside the cochlea was first suggested by T. Gold in

    1948 [41]. He proposed the “regeneration hypothesis” in which an active process was present in

    the cochlea to counteract viscous damping from the cochlear fluids. The idea is named after

    “regenerative receivers” used in radio engineering to create positive feedback to counteract

    resistive losses which would normally limit frequency selectivity. However Gold’s work was

    largely forgotten until experimental work in the 1970’s demonstrated the passive “second filter”

    was inadequate for describing the neural responses in mammals [39]. While an electrical

    filtering process is known to exist in the cochleae of turtles, it is widely believed not to be

    present in mammals [4, 18].

    As technology advanced, it became possible to make measurements inside the cochlea.

    In 1971 Rhode was able to use the Mössbauer effect to measure basilar membrane vibrations of a

    living squirrel monkey [42, 43]. The results showed the vibrations underwent compressive

    nonlinearity, i.e. vibrations grew with increasing sound pressure at a rate of less than 1 dB/dB.

    This nonlinearity occurred only for frequencies near the characteristic frequency and disappeared

    upon the death of the monkey. In Rhode’s 1978 work, he indicated that the basilar membrane

    was sharply tuned at low sound pressure levels and was poorly tuned at high sound pressures [4,

    44].

    In the late 1970’s, Kemp was the first to record acoustic emissions from the human ear

    [45]. He detected these acoustic emissions around 10 milliseconds after an impulsive acoustic

    excitation (a “click”). These sounds are now referred to as “transient-evoked otoacoustic

    emissions” (TEOAE) [19]. Kemp noted that the acoustic emissions demonstrated compressive

  • 14

    nonlinearity characteristics (the response amplitude did not scale linearly with input level), and

    these emissions were absent in deafened ears. He theorized these acoustic emissions were

    byproducts of an active mechanism in the cochlea. In 1979, Kemp made measurements of tones

    emitted from the ear without external stimulus [4, 45]. These spontaneous otoacoustic emissions

    (SOAE) are also considered to be byproducts of the feedback processes. In 1980, Mountain

    showed that the distortion product form of the evoked otoacoustic emissions could be shifted by

    changing the endocochlear potential or by stimulating the crossed olivocochlear bundle in the

    brain stem [4]. Thus the OHC’s were linked to frequency tuning in the neural and mechanical

    responses of the cochlea.

    1.3.2. Characteristics of the Cochlear Amplifier

    The cochlear amplifier has four important characteristics: amplification, frequency

    sensitivity, compressive nonlinearity, and spontaneous oscillations [20]. The first three

    characteristics can be summarized in data obtained shown in Figure 1.6 and Figure 1.7. Figure

    1.6a plots the basilar membrane displacement in a guinea pig as a function of driving frequency

    and sound pressure level. Figure 1.6b the basilar membrane displacement from Figure 1.6a

    divided by the driving sound pressure level. Figure 1.7 shows the basilar membrane amplitude at

    the characteristic frequency (14 kHz for this data) as a function of sound pressure level. These

    plots show the basilar membrane responds linearly except near the resonance or characteristic

    frequency. For low sound pressure levels, the cochlear amplifier boosts the response over a

    narrow range of frequencies, creating a sharp resonance peak at the characteristic frequency.

    This amplification of small inputs is evident in Figure 1.6b and Figure 1.7. The characteristic

  • 15

    frequency depends on the measurement location along the cochlea. For a given location along

    the basilar membrane, the frequency which causes the largest deflection is called the

    characteristic frequency. Because of the tonotopic nature of the cochlea, one can also define the

    characteristic place as the location of largest deflection for a given stimulus frequency.

    (a) (b)

    Figure 1.6. Measurements showing the cochlear amplifier in a guinea pig cochlea. (a)

    Basilar membrane (BM) displacement of versus frequency and sound pressure level. (b)

    Basilar membrane displacement normalized by the input sound pressure level. These

    curves would overlap for a linear system. Figure from Johnstone et al. (1986),

    reproduced with permission of Elsevier Limited [40].

  • 16

    Figure 1.7. Displacement of the basilar membrane versus sound pressure level. The

    displacement shows a linear trend at low and high sound pressures levels and a nonlinear

    compression at intermediate sound pressure. Figure from Johnstone et al. (1986),

    reproduced with permission of Elsevier Limited [40].

    This amplification allows the ear to detect low sound pressures. At the threshold of

    auditory deflection (0 dB or 20 μPa RMS sound pressure level), the deflection of the basilar

    membrane is around 0.3 nm [40, 44, 46]. For comparison, the diameter of a hydrogen atom is on

    the order of 0.1 nm; thus mammals are able to detect vibrations comparable to the thermal noise

    [20]. This amplification only occurs in living creatures. A few minutes after death, the threshold

    of auditory response increases by 40 to 60 dB (i.e. sensitivity falls to less than 1% of that of the

    living cochlea) [47]. The narrow bandwidth of the peak allows the cochlea to detect small

    changes in frequency, which gives the cochlea an improved frequency selectivity. This

    frequency discrimination is vital for comprehension for speech and music [48].

  • 17

    It is important to note this amplification is nonlinear in nature. As also shown in Figure

    1.6 and in Figure 1.7, the displacement at the characteristic frequency increases at a less-than-

    proportional rate with increasing sound pressure [18]. Humans have a threshold of sound

    detection around 0 dB or 20 μPa RMS sound pressure level. This corresponds to a basilar

    membrane oscillation around 0.3 nm in amplitude. At the threshold of pain (120 dB or 20 Pa

    RMS), the basilar membrane of many mammals oscillates at an amplitude on the order of 100

    nm [18, 49]. Thus the nonlinearity compresses a large sound pressure range into a smaller range

    of response amplitudes. This compression can be seen in Figure 1.6a and Figure 1.7.

    Compressive growth rates between 0.12 to 0.5 dB/dB have been seen in measurements from

    various mammals [18, 49, 50]. This compressive nonlinearity allows the cochlea to shrink a 120

    dB dynamic range of sound pressure levels into a range around 30 to 40 dB of basilar membrane

    displacement. Similar amplitude compression is observed in neuron firing rates and in

    recordings of inter-cochlear voltage changes [19, 51]. In addition to the nonlinear growth of the

    peak response, the bandwidth of amplification also increases as sound pressure level increases.

    The final characteristic of the cochlear amplifier is spontaneous otoacoustic emissions.

    Studies from several different animals show the ear can actively emit tones of distinct

    frequencies [4, 45, 52, 53]. These tones are considered by many to be byproducts of the active

    feedback process. In addition to spontaneous oscillations, when the cochlea is stimulated with

    two or more tones, additional tones of different frequencies can be measured emanating from the

    ear and can be heard by the listener [4, 54]. These distortion-product otoacoustic emissions are

    also indicators of a nonlinear process. Spontaneous and distortion-product otoacoustic emissions

    are altered or absent in damaged cochleae [45]. Thus these emissions can used be used as a

    diagnostic tool to test hearing in newborns [54].

  • 18

    1.3.3. Mechanisms of Amplification

    The exact processes behind the cochlear amplifier are still an issue of debate in the

    auditory community [4, 55-59]. Several theories have been proposed over the years, but current

    research focuses on two experimentally observed mechanisms: hair bundle motility and somatic

    motility.

    Hair bundle motility theories advocate that the stereocilia bundles protruding from the

    hair cells contribute to the amplification of incoming sounds. Measurements of hair cells from

    insects, turtles, and bullfrogs show compressive nonlinearity to an applied stimulus and

    spontaneous oscillations [20, 60-63]. Hair bundle motility is believed to be created by a

    combination of a nonlinear stiffness of the hair bundles (which has a negative stiffness region

    around the vertical position of the hair bundle) and an adaptation process which readjusts the

    tension in the tip links to force the hair bundles toward an unstable vertical position [62, 64-66].

    These effects have also been seen in mammalian hair cells [67]. While this process likely

    underlies active hearing in lower vertebrates, critics argue that active hair bundle motility lacks

    the power required to drive the amplifications recorded in mammals [4, 56].

    In mammals there is strong evidence that somatic motility of the outer hair cells

    (previously mentioned in Section 1.2.2.) plays an important role in amplification. A change in

    voltage across the cell membrane causes the cell body (the soma) to contract. In this manner, the

    outer hair cell acts like a piezoelectric actuator producing a force under an applied voltage

    change [68]. Also like a piezoelectric material, the outer hair cells show a measureable charge

    displacement under an applied force [69]. These measurements indicate that isolated outer hair

    cells have an effective piezoelectric coefficient around 20 μC/N, which is four orders of

  • 19

    magnitude larger than any man-made piezoelectric materials [69-71]. While this effect is

    powerful enough to drive amplification in the mammalian cochlea, critics argue that somatic

    motility behaves linearly for physiologically relevant voltage changes, and therefore somatic

    motility alone cannot explain the nonlinear oscillations observed in the cochlea [56]. In addition,

    the capacitance of the cell membrane and the conductance of the ionic fluid inside the cochlea

    create a low-pass filter with a corner frequency around 1,000 Hz (significantly lower than the

    20,000 Hz upper limit of human hearing) [21, 34, 72, 73]. This low pass filter on its own should

    significantly attenuate high frequency oscillations.

    Some researchers are now advocating that both hair bundle motility and somatic motility

    are needed to explain the observed characteristics of the mammalian cochlea [20, 39, 56, 74].

    The hair bundle motility creates a nonlinear input to the otherwise linearly behaved

    electromotility of the outer hair cell. This in turn creates the nonlinear behavior and frequency

    sensitivity of the cochlear amplifier. The somatic motility serves to boost amplification and aids

    in detecting higher frequencies. Some propose that the membrane filtering problem which

    plagues somatic motility alone could be overcome by its coupling with an active hair bundle [56,

    75, 76].

    1.4. Mimicking the Cochlea through Passive Devices

    The cochlea has severed as inspiration for many researchers looking to develop new

    types of sensors or auditory prosthetics. This section will examine several passive devices which

    mimic the cochlea and its sound transduction. The next section will examine systems which

    attempt to reproduce the cochlea’s nonlinear amplification.

  • 20

    1.4.1. Passive Artificial Hair Cells

    A common, bio-inspired approach to designing a flow or acoustic sensor is to mimic the

    cochlear hair cells. One common approach to creating an artificial hair cell is to treat the hair

    bundle like a single cantilevered beam. Flow- or acoustic-induced vibration of a cantilever beam

    can be transduced into an electrical signal through a variety of methods. The natural frequency

    of a beam is dependent on the material properties, length, width, and tip mass. An array of

    beams of varying geometries and tip masses can mechanically filter the inducing force into a set

    of frequency sub-bands much like the basilar membrane. Simplicity of design, low power

    requirements, and their small size make these artificial hair cells an appealing approach for many

    researchers.

    Researchers have used a variety of transduction methods to transform the mechanical

    vibration of these hair cells into an electrical signal. One method is to use piezoelectric

    materials. Piezoelectric materials produce an electrical charge when subjected to an applied

    mechanical stress. These materials are attractive to many researchers because of the possibility

    of developing a self-powered sensor which does not need an external power source.

    In late 1990s and early 2000s, Mukherjee and colleagues examined the feasibly of using

    polyvinylidene fluoride (PVDF) cantilever beams to build cochlear implants [7, 77-79]. For

    acoustic sensing, piezoelectric polymers like PVDF offer several advantages over ceramic

    piezoelectric materials, such as a greater range of motion, higher voltage sensitivity, higher

    dielectric breakdown voltages, better acoustic impedance matching with air and water, and easier

    manufacturing [78, 80]. Experimental work of their beams showed a sensitivity up to 447 μV/Pa

  • 21

    for underwater testing, which is comparable to some commercially available hydrophones [81].

    However the authors were skeptical whether an unamplified signal from an acoustically excited

    piezoelectric material could generate a sufficient electrical impulse to stimulate the neurons in

    the spiral ganglion.

    Hur et al. fabricated and characterized an array of cantilever beams constructed from

    single crystal lead magnesium niobate-lead titanate (PMN-PT) [82, 83]. The single crystal

    PMN-PT is a newer type of piezoelectric material which offers higher sensitivities and lower

    dielectric losses than PVDF or the traditional lead zirconium titanate (PZT) piezoceramic

    material. The authors applied a 94 dB (1 Pa) constant sound pressure to the array of these PMN-

    PT cantilever beams with lengths varying from 3000 μm to 500 μm. The maximum voltage

    output from a single cantilever was 80.4 mV for the longest beam, while the smaller beams

    produced less voltage. At higher frequencies, the higher natural frequencies of the longer beams

    could be observed. These frequencies were lower than the first natural frequencies of the shorter

    beams. Without filtering the signals, excitation of higher modes would complicate the

    frequency-spatial decomposition. This problem occurs for other arrays of beam-like sensors

    throughout the literature.

    Kim et al. examined several arrays of piezoelectric beams with narrow supports [8].

    Each array consisted of 16 beams that were 200 μm wide and spaced 400 μm apart. The beam

    lengths were varied between 305 and 3200 μm, and the array showed frequency selectivity

    between 2-20 kHz. The system was excited electrically and acoustically and the response was

    measured using a scanning laser-Doppler vibrometer. Sound pressures between 105-110 dB

    were used to acoustically excite the system, however the output voltage levels from the beams

    during this testing was not stated.

  • 22

    Other researchers have developed MEMS-scale hair sensors which utilize piezoresistance

    [3, 84-90]. Piezoresistance is the property of a material to change resistance when strained. This

    method is commonly used in developing strain gauges. For these artificial hair cells, the

    deflection of a hair-like structure deformed a piezoresistor and created a change in its resistance.

    Like their larger strain gauge counterparts, these devices required an external current source in

    order to measure the change in resistance of the piezoesistor under flow-induced strain.

    Several sensors have been constructed using a variable capacitor design [91-93]. Here a

    hair-like structure was connected to one side of a parallel plate capacitor. Deflection of the hair

    changed the separation distance of the electrode plates. The result was a time-varying

    capacitance which under a fixed voltage resulted in a change in current through the capacitor.

    This change in current was translated into an applied deflection on the hair structure.

    Leo, Sarles, and colleagues have examined artificial cell membranes formed by lipid

    bilayers [94-96]. A hair-like structure was inserted into a water-swollen, lipid-encased hydrogel

    droplet. A liquid bilayer was formed with a neighboring lipid-encased aqueous volume. Air

    flow forced the hair and the bilayer to vibrate. This vibration varied the curvature of the lipid

    bilayer and resulted in a change in the membrane’s capacitance. This time-varying capacitance

    generated several picoamperes of current across the membrane. These sensors used soft

    materials commonly found in most cells as opposed to the artificial materials found in other

    designs. This allows for a more biocompatible artificial hair cell. While the transduction

    mechanism is different than that of the natural hair cells, the composition of their artificial

    membranes allows for the possibility of adding ion channels or other molecules to modify or

    enhance transduction in ways similar to mammalian hair cells.

  • 23

    1.4.2. Passive Artificial Basilar Membr