Sie sind auf Seite 1von 169

AN ENERGY EFFICIENCY APPROACH FOR UNIFIED TOPOLOGICAL AND DIMENSIONAL SYNTHESIS OF COMPLIANT MECHANISMS

by Joel A. Hetrick

A dissertation submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy (Mechanical Engineering) in the University of Michigan 1999

Doctoral Committee: Associate Professor Sridhar Kota, Chair Professor Norboru Kikuchi Assistant Professor Kazuhiro Saitou Assistant Professor John A. Shaw

Joel A. Hetrick All Rights Reserved

2003

For Mom and Dad, Brent, and of course Elizabeth. Thanks for all your love and support.

ii

ACKNOWLEDGEMENTS I would like to thank my advisor, Professor Sridhar Kota for his encouragement and guidance for completing this work. His interest in the topic of mechanism design, including compliant mechanism design, has significantly deepened my own interest in the subject. I am greatly indebted to Professor Kota, who offered me hourly work as an undergraduate student and progress to supporting me during my time at Michigan as a graduate student. While a graduate student, Professor Kota was instrumental in finding many exceptional and rewarding projects to work on. These projects were not only challenging from an academic perspective, but also allowed me to gain practical hand-on experience which will be invaluable to me in my future career. I can firmly state that I would not have progressed nearly as far with out the enthusiasm, instruction, and guidance provided by Professor Kota. I also greatly appreciate the support and encouragement I have received from him to pursue and academic career. I would like to thank Professor Noboru Kikuchi for his invaluable input on compliant mechanism design and structural optimization techniques. In addition, I would like to thank Professor Kikuchi for offering the Homogenization Design Method course this winter. This course proved to be extremely valuable; thoroughly covering the past, present, and future of structural optimization. I am grateful to my other committee members, Professor Kazu Saitou and Professor John Shaw, who have also offered helpful suggestions and posed insightful questions regarding this work. I would also like to thank both of them for offering sage advice for my future academic career plans. I have learned a great deal about compliant mechanisms and other related topics through my collaboration with Dr. G. K. Ananthasuresh, Dr. Mary Frecker, Dr. Laxman iii

Saggere, Dr. Zhe Li and Jin-Young Joo. The ground work laid by Suresh and Marys work has been essential to the early development of my own understanding of the compliant mechanism design problem. I have also greatly benefited from Sureshs suggestions and critiques through our communication. And finally, I am grateful for the financial support of the National Science Foundation Design (grant #9622261), and the Air Force Office of Scientific Research (contract #F49620-96-1-0205). I would also like to thank Wright Patterson Research Labs and Sandia National Labs for offering their resources to further investigate new and innovative compliant devices.

iv

TABLE OF CONTENTS DEDICATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ii ACKNOWLEDGEMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xii LIST OF APPENDICES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xiii CHAPTER I INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 1.1 Definition of Compliant Mechanisms. . . . . . . . . . . . . . . . . . . . . . . . . 14 1.2 Classification of Compliant Mechanisms . . . . . . . . . . . . . . . . . . . . . . 15 1.3 Advantages of Compliant Mechanisms . . . . . . . . . . . . . . . . . . . . . . . 16 1.4 Applications of Compliant Mechanisms. . . . . . . . . . . . . . . . . . . . . . . 17 1.4.1 Actuator Tailoring . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 1.4.2 Smart Structures. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18 1.4.3 Materials with Prescribed Constitutive Parameters . . . . . . . . . . 19 1.4.4 Micro-Electro-Mechanical Systems . . . . . . . . . . . . . . . . . . . . . 20 1.4.5 Design-for-No-Assembly. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21 1.5 Mechanism Synthesis and Structural Synthesis . . . . . . . . . . . . . . . . . 22 1.6 Scope of the Investigation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23 1.7 Organization of the Dissertation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24 CHAPTER II LITERATURE REVIEW . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26 2.1 Compliance in Design. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26 2.2 Kinematic Approaches . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26 2.2.1 Flexure Devices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 2.2.2 Flexible Link Mechanisms. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28 v

2.2.3 Pseudo Rigid Body Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 2.3 Structural Optimization. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31 2.3.1 Historical Foundations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32 2.3.2 Structural Optimization of Compliant Mechanisms . . . . . . . . . 34 CHAPTER III PROBLEM FORMULATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40 3.1 Design Approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40 3.2 Introduction to the Compliant Lever Model . . . . . . . . . . . . . . . . . . . . 41 3.3 Research Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43 3.4 Energy Efficiency Objective Formulations . . . . . . . . . . . . . . . . . . . . 45 3.4.1 The Force-Displacement Efficiency Formulation . . . . . . . . . . . 46 3.4.2 The Spring-Efficiency Formulation. . . . . . . . . . . . . . . . . . . . . . 52 3.4.3 Optimization of the Compliant Lever using the ForceDisplacement Efficiency Formulation . . . . . . . . . . . . . . . . . . . . 55 3.4.4 Optimization of the Compliant Lever Using the Spring Efficiency Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62 3.5 Characteristics of Optimal Designs . . . . . . . . . . . . . . . . . . . . . . . . . . 67 3.6 Advantages of the Objective Formulation . . . . . . . . . . . . . . . . . . . . . 68 CHAPTER IV NUMERICAL IMPLEMENTATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70 4.1 Optimization Approach. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70 4.2 Modeling Approach and Design Parameterization . . . . . . . . . . . . . . . 70 4.3 Gradient-Based Optimization Algorithms . . . . . . . . . . . . . . . . . . . . . 76 4.3.1 Sequential Linear Programming . . . . . . . . . . . . . . . . . . . . . . . . 80 4.3.2 Sequential Quadratic Programming. . . . . . . . . . . . . . . . . . . . . . 81 4.4 Analytic Calculation of Design Sensitivities . . . . . . . . . . . . . . . . . . . 81 4.4.1 Differentiating with respect to the Element Stiffness Matrix . . 83 4.4.2 Sensitivity Analysis of the Efficiency Formulations . . . . . . . . . 85 The Force-Displacement Efficiency Formulation . . . . . . . . . . . . 85 The Spring Efficiency Formulation . . . . . . . . . . . . . . . . . . . . . . . 88

vi

4.4.3 Sensitivity Analysis of the Volume Constraint . . . . . . . . . . . . . 90 4.4.4 Sensitivity Analysis of the Mechanical Advantage Constraint . 91 4.4.5 Sensitivity Analysis of the Mean Positive Stress Constraint. . . 91 CHAPTER V TOPOLOGICAL SYNTHESIS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96 5.1 Topology Synthesis Procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96 5.2 Convergence of Topology Designs. . . . . . . . . . . . . . . . . . . . . . . . . . . 97 5.2.1 Design Discretization. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98 Fixed-Node Verses Floating-Node Ground Structures . . . . . . . . 99 Discretization Resolution. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100 5.2.2 Comparison of Objective Formulations. . . . . . . . . . . . . . . . . . 102 5.2.3 Effect of Element Lower Bound . . . . . . . . . . . . . . . . . . . . . . . 106 5.2.4 Effect of the Total Volume Constraint . . . . . . . . . . . . . . . . . . 107 5.2.5 Truss Elements Verses Frame Elements . . . . . . . . . . . . . . . . . 108 5.3 Topology Synthesis Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109 5.3.1 Spring Efficiency Optimized Topologies . . . . . . . . . . . . . . . . 109 Compliant Gripper . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109 Compliant Wrench. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112 5.3.2 Force-Displacement Efficiency Optimized Topologies . . . . . 114 Piezo-Stack Amplifier . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114 MEMS Displacement Magnifier . . . . . . . . . . . . . . . . . . . . . . . . 117 5.4 Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119 CHAPTER VI DIMENSIONAL SYNTHESIS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121 6.1 Dimensional Synthesis Procedure . . . . . . . . . . . . . . . . . . . . . . . . . . 121 6.2 Design Convergence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122 6.2.1 Sizing Verses Force/Displacement and Spring Parameter . . . 122 6.2.2 Principle Nodes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124 6.2.3 Effect of Element Discretization . . . . . . . . . . . . . . . . . . . . . . . 125

vii

6.2.4 Effect of Stress Constraints . . . . . . . . . . . . . . . . . . . . . . . . . . . 126 6.2.5 Effect of Element Discretization on Stress Constraints. . . . . . 129 6.3 Optimized Examples. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130 6.3.1 Spring Efficiency Formulation . . . . . . . . . . . . . . . . . . . . . . . . 131 Compliant Gripper . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131 Compliant Wrench. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133 6.3.2 Force-Displacement Efficiency Formulation. . . . . . . . . . . . . . 136 Piezo-stack Amplifier . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136 MEMS Displacement Amplifier . . . . . . . . . . . . . . . . . . . . . . . . 139 6.4 Alternative Topology Synthesis Procedures. . . . . . . . . . . . . . . . . . . 141 6.5 Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149 CHAPTER VII CONCLUSIONS AND FUTURE WORK . . . . . . . . . . . . . . . . . . . . . . . . . . . 151 7.1 Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151 7.2 Contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155 7.3 Future Research Directions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157 APPENDICES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159 BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166

viii

LIST OF FIGURES Figure 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9 2.1 2.2 3.1 3.2 3.3 3.4 3.5 3.6 3.7 3.8 3.9 A Conventional Rigid Link Mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 Illustration of a Compliant Gripper Mechanism . . . . . . . . . . . . . . . . . . . . . . . . 15 Classification of Compliant Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 Force and Stroke Trade-off using a Compliant Mechanism . . . . . . . . . . . . . . . 18 A Shape-Changing Compliant Smart Structure. . . . . . . . . . . . . . . . . . . . . . . . . 19 A Material Microstructure with a Poissons Ration of -0.65 (Fonseca, Kikuchi, 1997) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20 A Compliant Micro-Gripper (Ananthasuresh, Saggere, Kota, 1994) . . . . . . . . 21 Rigid Stapler Verses One-Piece Compliant Stapler (Ananthasuresh, Saggere, Kota, 1994) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21 The Generalized Single-Input, Single-Output Compliant Mechanism Problem 23 Illustration of the Pseudo Rigid Body Model . . . . . . . . . . . . . . . . . . . . . . . . . . 31 A Cantilever-like Michell Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 The Two-Thickness Compliant Lever . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42 Illustration of the Various Solutions to the Compliant Lever Problem . . . . . . . 42 A Mechanism Pushing Against a Constant External Force . . . . . . . . . . . . . . . . 47 The Input and Output Work History (Force-Displacement Formulation) . . . . . 47 A Mechanism Grasping a Soft Workpiece. . . . . . . . . . . . . . . . . . . . . . . . . . . 53 The Input and Output Work History (Spring Formulation). . . . . . . . . . . . . . . . 54 The Compliant Lever Measurement Techniques. . . . . . . . . . . . . . . . . . . . . . . . 57 Plot of h1 and h2 With Respect to the Flexibility / Stiffness Parameter . . . . . . 61 Measurement of Performance Using the Spring Formulation . . . . . . . . . . . . . . 62

3.10 Plot of h1 and h2 With Respect to the Spring Parameter . . . . . . . . . . . . . . . . . 66 4.1 Local Dimensions of the 2-D Truss and Frame Element. . . . . . . . . . . . . . . . . . 72

ix

4.2 4.3 4.4 4.5 4.6 4.7 5.1 5.2 5.3 5.4 5.5 5.6 5.7 5.8 5.9

An Element Array having N Elements and Two Principle Nodes . . . . . . . . . . 75 Flow Chart of a Sequential Gradient-Based Optimization Algorithm . . . . . . . 79 The Two Energy Efficiency Problems for Compliant Mechanism Design . . . . 85 Performance Measurement for the Force-Displacement Formulation . . . . . . . 87 Performance Measurement for the Spring Formulation . . . . . . . . . . . . . . . . . . 89 Dummy Loads Used for Dusi and Duri . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93 Illustration of the Topology Synthesis Procedure . . . . . . . . . . . . . . . . . . . . . . . 97 Compliant Mechanism Design Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98 Two Approaches for Design Domain Discretization. . . . . . . . . . . . . . . . . . . . . 99 Fixed-Node Approach Verses Floating-Node Approach. . . . . . . . . . . . . . . . . 100 Optimized Topologies Resulting From Increased Node Discretization . . . . . 101 Optimized Topologies Verses Spring Parameter. . . . . . . . . . . . . . . . . . . . . . . 103 Optimized Topologies Verses Force/Displacement Parameter . . . . . . . . . . . . 105 Optimization Convergence Verses Element Lower Bound . . . . . . . . . . . . . . . 107 Optimization Convergence Verses Total Volume Constraint . . . . . . . . . . . . . 108

5.10 Design Problem for the Compliant Gripper. . . . . . . . . . . . . . . . . . . . . . . . . . . 110 5.11 Optimized Compliant Gripper Topologies (Fixed-Node Approach). . . . . . . . 111 5.12 Optimized Compliant Gripper Topologies (Floating-Node Approach) . . . . . 111 5.13 Design Problem for the Compliant Wrench . . . . . . . . . . . . . . . . . . . . . . . . . . 112 5.14 Optimized Compliant Wrench Topologies (Fixed-Node Approach) . . . . . . . 113 5.15 Optimized Compliant Wrench Topologies (Floating-Node Approach) . . . . . 113 5.16 Design Problem for the Piezo-Stack Amplifier . . . . . . . . . . . . . . . . . . . . . . . . 114 5.17 Optimized Piezo-Stack Amplifier Topologies (Fixed-Node Approach) . . . . . 116 5.18 Optimized Piezo-Stack Amplifier Topologies (Floating-Node Approach) . . . 116 5.19 Design Problem for the MEMS Displacement Magnifier . . . . . . . . . . . . . . . . 117 5.20 Optimized MEMS Magnifier Topologies (Fixed-Node Approach) . . . . . . . . 118 5.21 Optimized MEMS Magnifier Topologies (Floating-Node Approach) . . . . . . 119 6.1 6.2 Size and Geometry Optimization for Half of a Compliant Gripper . . . . . . . . 122 Optimized Element Sizing for the Compliant Lever Problem. . . . . . . . . . . . . 123

6.3 6.4 6.5 6.6 6.7 6.8 6.9

Illustration of Node Re-Location During Optimization . . . . . . . . . . . . . . . . . 125 Effect of Increased Element Discretization on Mechanism Design . . . . . . . . 126 Feasible Design Space for Spring Stiffness and Input Force. . . . . . . . . . . . . . 127 Feasible Design Space for External Force and Input Displacement . . . . . . . . 129 Effect of Stress Constraint with Increased Element Discretization . . . . . . . . . 130 Optimized Compliant Gripper Size and Geometry (Fixed-Node Topology Design). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132 Optimized Compliant Gripper Size and Geometry (Floating-Node Topology Design). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132

6.10 The Optimized Compliant Gripper Mechanism . . . . . . . . . . . . . . . . . . . . . . . 133 6.11 Optimized Compliant Wrench Size and Geometry (Fixed-Node Topology Design). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134 6.12 Optimized Compliant Wrench Size and Geometry (Floating-Node Topology Design). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135 6.13 An Optimized Compliant Wrench Mechanism . . . . . . . . . . . . . . . . . . . . . . . . 135 6.14 Optimized Piezo-Stack Amplifier Size and Geometry (Fixed-Node Topology Design). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137 6.15 Optimized Piezo-Stack Amplifier Size and Geometry (Floating-Node Topology Design). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138 6.16 An Optimized Piezo-Stack Amplifier . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138 6.17 Optimized MEMS Amplifier Size and Geometry (Fixed-Node Topology Design) 140 6.18 Optimized MEMS Amplifier Size and Geometry (Floating-Node Topology Design). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141 6.19 An Optimized MEMS Displacement Amplifier . . . . . . . . . . . . . . . . . . . . . . . 141 6.20 A Sample of Compliant Mechanism Topology Building Blocks . . . . . . . . . . 143 6.21 Building Block Topology Design for the 20:1 Amplification Mechanism . . . 145 6.22 Illustration of the 20:1 Amplification Mechanism Prototype . . . . . . . . . . . . . 146 6.23 Building Block Topology for MEMS Displacement Amplifier . . . . . . . . . . . 147 6.24 Optimized MEMS Amplifier Size and Geometry (Building Block Topology Design). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148 6.25 Illustration of the Original Direct-Drive Design Versus the MEMS Displacement Amplifier Design (Courtesy of Sandia National Labs) . . . . . . . . . . . . . . . . . . 149

xi

LIST OF TABLES Table 5.1 5.2 6.1 6.2 Efficiencies of Various Topologies Verses Spring Parameter. . . . . . . . . . . . . 103 Efficiencies of Various Topologies Verses Force/Displacement Parameter . . 105 Feasible Compliant Mechanism Design Ranges for the Spring Efficiency Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127 Feasible Compliant Mechanisms Design Ranges for the Force-Displacement Efficiency Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128

xii

LIST OF APPENDICES Appendix A. DIFFERENTIATION OF THE FRAME ELEMENT STIFFNESS MATRIX WITH RESPECT TO NODE COORDINATES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160

xiii

CHAPTER I INTRODUCTION 1.1 Definition of Compliant Mechanisms A mechanism is defined as a mechanical device which transfers forces and motions from an input source to output sink (Erdman, Sandor, 1991). As shown in Figure 1.1, a conventional rigid-link mechanism is designed to obtain mobility exclusively from rigid body translations and rotations using revolute joints, prismatic joints, and higherorder pairs which couple the relative motions of the members.
input revolute joint

output prismatic joint Figure 1.1: A Conventional Rigid Link Mechanism

Traditional mechanisms are engineered to be strong and rigid. The use of flexibility has primarily been avoided due to the increased difficulty in accounting for flexibility and the large potential for performance degradation, such as decreased efficiency, positional inaccuracies, and structural instabilities. In contrast, biological designs such as the thorax of a honeybee or the leg of a grasshopper are strong and compliant. By opposing traditional engineering techniques, nature has found clever

14

15 methods to produce bio-mechanical devices which achieve functionality by taking advantage of structural compliance (Vogel, 1998). To parallel biological design, a compliant mechanism is defined as a mechanism which exploits flexibility from one or more of its members to achieve controlled transmission of forces and motions. Figure 1.2 illustrates a monolithic compliant gripper which closes and exerts force on an object in response to an applied input force. The main challenge in designing a compliant mechanism is to effectively synthesize the most efficient structural form (topology, size, and shape) given the functional requirements.

Fin1

object

Fin

(a) Initial Configuration

(b) Deformed Configuration

Figure 1.2: Illustration of a Compliant Gripper Mechanism

1.2 Classification of Compliant Mechanisms Shown in Figure 1.3, compliant mechanisms can be classified by the various rigid and flexural elements which make up the device (Ananthasuresh, 1994). Partially compliant mechanisms contain rigid mechanical members and conventional joints combined with compliant members. Fully compliant mechanisms contain no mechanical joints and gain mobility strictly from elastic deformation. A compliant member can facilitate elastic deformation by concentrating strains in small regions or more evenly distributing strains throughout the structure. Lumped compliant mechanisms which concentrate strains in small regions are characterized by thin flexural segments which can mimic the behavior of revolute joints. Distributed compliant mechanisms do not contain

16 the characteristic thin flexible segments, but rather distribute the strain more-or-less uniformly along the flexible members. Distributed compliant mechanisms are advantageous over lumped compliant mechanisms in that stress concentrations are generally avoided.

(a) Partially Compliant Mechanism Lumped Compliance

(b) Partially Compliant Mechanism Distributed Compliance

(c) Fully Compliant Mechanism Lumped Compliance

(d) Fully Compliant Mechanism Distributed Compliance

Figure 1.3: Classification of Compliant Mechanisms

1.3 Advantages of Compliant Mechanisms For applications requiring moderately small displacements and rotations, compliant mechanisms have many advantages over conventional rigid mechanisms. As noted by Tuttle (1967), the advantages of flexures (compliant mechanisms) include the elimination of friction, wear and backlash common with conventional mechanical joints. Sevak and McLarnan (1971) added that compliant mechanisms are easier to assemble, take advantage of stored elastic energy to eliminate components such as stroke return springs, and potentially offer weight savings over conventional mechanism designs. Ananthasuresh (1994) also concluded that fully compliant devices are much easier to

17 fabricate than conventional mechanisms and offer better resilience to mechanical overloading and shock absorption. 1.4 Applications of Compliant Mechanisms Due to the many advantages of compliant mechanisms, these devices have found numerous applications in areas such as actuator tailoring, smart structures, material microstructures, micro-electro-mechanical systems (MEMS), and general product design. 1.4.1 Actuator Tailoring Piezoelectric materials, shape memory alloys, and magnetostrictive materials are a few examples of a new breed of solid-state mechanical actuators. These smart materials can effect physical change by altering their chemical, electrical, or thermal state. One drawback of these devices is that the stroke produced by smart materials actuators is often severely limited in range. As noted by Chopra (1996), for smart materials actuators to be effective in aerospace applications, a need exists to efficiently amplify their displacements by a factor of 300 to 500 percent. Due to the elimination of backlash, compliant mechanisms can effectively work with the very small displacements (1-100 m) generated by these devices and can offer a means to efficiently tailor the force and stroke of the actuator to suit specific needs. Shown in Figure 1.4 is an illustration of a compliant mechanism trading force for displacement from a piezoelectric actuator. Here, the forcedisplacement capability of the piezoelectric actuator is chosen at a point, P, and transformed to an alternative force-displacement range, point Q. Furukawa (1991) and Hall (1996) both have developed compliant mechanism-based piezo-actuator amplification devices using trial and error design methods. However, these devices have obtained limited success due to the complexity involved in designing high performance displacement amplification mechanisms. Using newly developed optimization

18 techniques, it is possible to design compliant mechanisms which efficiently amplify the displacements of these smart materials by a factor of 50 or more (Hetrick, Kota, 1999).
force-displacement characteristics of actuator

input operational point force f1 f2 d1 displacement Figure 1.4: Force and Stroke Trade-off using a Compliant Mechanism compliant mechanism P Q d2 desired operational point

1.4.2 Smart Structures Smart structures are another area where compliant mechanisms are likely to provide significant impact. Smart structures utilize controllers, sensors, and actuators to sense and effect structural changes, such as changes in vibration modes, stiffness, or shape. Actuators, sensor, and control systems for these devices must work harmoniously with surrounding structure to sense and provide the appropriate mechanical change. Present smart structure design has primarily focused on designing sensors, actuators, and control systems for a given structure. Compliant mechanisms offer a method to design a structure with enhanced functionality that will harmoniously work with these devices to greatly improve performance. Static shape change of smart structures is one very promising application for compliant mechanisms. As noted by Saggere and Kota (1997), by utilizing distributed compliance rather than distributed actuation, compliant mechanisms can significantly reduce the number of actuators and the complexity of the control system required to achieve static shape change of smart structures. Shown in

19 Figure 1.5 is a smart structure which incorporates a distributed compliant mechanism. Using a single actuator, the compliant mechanism deforms to distribute a localized deformation to effect a net shape change.
shape-changing compliant mechanism

flexible skin

actuator

position feedback sensor

Figure 1.5: A Shape-Changing Compliant Smart Structure

1.4.3 Materials with Prescribed Constitutive Parameters Tailoring the constitutive parameters of periodic material micro-structures represent a new approach for creating new materials with unusual mechanical properties. Instead of selecting an existing homogeneous material with set mechanical properties, materials with customized moduli and poissons ratio can be designed using material micro-structure design techniques. Past researchers (Almgren, 1985, Milton, 1992, Sigmund, 1994, Kikuchi, 1998) have shown that materials with Poissons ratios ranging from 0.499 to -0.999 are possible. These extreme materials can be described as periodic compliant mechanism micro-structures. As the materials deform, the micro-structure mimics a mechanism by expanding or contracting in response to strain. Applications of these materials range from better shock absorbing materials to advanced piezoelectric composite materials. Shown in Figure 1.6 is an example of a periodic microstructure designed with a Poissons ratio of -0.65.

20

Figure 1.6: A Material Microstructure with a Poissons Ration of -0.65 (Fonseca, Kikuchi, 1997)

1.4.4 Micro-Electro-Mechanical Systems Micro-electro-mechanical systems (MEMS) are another field where compliant mechanisms have tremendous advantages over conventional rigid mechanisms. As noted by Ananthasuresh et al. (1994), because of the small scale of MEMS devices, assembly operations are difficult if not impossible. In addition at the MEMS level, frictional forces at mechanical joints can significantly impede the functionality of conventional mechanism designs. Thus, the elimination of joint friction inherent in fully compliant mechanisms is a major advantage over conventional rigid mechanisms. In addition, the extreme aspect ratio of MEMS devices (length scale in the 100s of m, height scale 1-6 m) often causes them to be unavoidably flexible. As noted by Kota (1994), at the MEMS level, compliance in design is not only advantageous to effectively achieve functionality, it is often a necessity. Shown in Figure 1.7 is an example of a micro-gripper mechanism. Thus, compliant mechanisms offer much simpler and more manufacturable methods for creating sophisticated and high-performance micro-mechanisms which can be easily batch produced using standard integrated circuit manufacturing techniques.

21

Figure 1.7: A Compliant Micro-Gripper (Ananthasuresh, Saggere, Kota, 1994)

1.4.5 Design-for-No-Assembly General product design is another area where compliant mechanisms are likely to make a significant impact. By integrating structural and mechanical functionality, compliant mechanisms are useful for simplifying product design, reducing the total partcount, and greatly reducing fabrication requirements. Monolithic compliant mechanisms naturally lend themselves to injection molding processes which can eliminate assembly procedures altogether. This paradigm of design-for-no-assembly is likely to significantly affect how future products are designed and fabricated. Shown in Figure 1.8 is a rigid multi-component stapler compared to a one-piece compliant stapler. The one-piece compliant stapler (two metal insets) can be manufactured in one injection molding process and replaces the twenty separate components used by the conventional stapler design.

Figure 1.8: Rigid Stapler Verses One-Piece Compliant Stapler (Ananthasuresh, Saggere, Kota, 1994)

22 1.5 Mechanism Synthesis and Structural Synthesis Conventional mechanism design techniques are typically based on the assumption of infinite rigidity and can be performed using purely geometric principles. The task of mechanism design can be broken into topology synthesis (or type synthesis) and dimensional synthesis. Topology synthesis consists of determining the number and order of links (binary, tenerary, etc.) and the interconnection of these links through kinematic pairs. Topology synthesis can be accomplished using graph theory or kinematic chains which map topologies based on degrees-of-freedom determined from Grublers equation. Dimensional synthesis involves adapting a topology for a specific task by determining joint types, sizing the link lengths and cross-sections to satisfy functional and structural requirements. Traditional structural design problems have primarily been based on finding the lightest, stiffest, and strongest structure. A structural form can be described using topology, shape and size where topology refers to the number of members and their connectivity, shape (or geometry) describes the spatial location and lengths of these members with respect to one another, and size consists of the relative thicknesses of the members. Similar to mechanism design, structural synthesis can be divided into topology or layout design and size and geometry design. Topology design is typically posed as a material distribution problem where a finite volume of material is redistributed throughout a pre-determined design domain to maximize the performance of the structure. Topology design is usually one level abstracted from the detail design as the best topology is typically invariant of the magnitude loading conditions and modulus of the design material. The detailed structural design is typically posed as a boundary design problem where the perimeter of a fixed topology is designed to maximize performance given specific material properties and load magnitudes.

23 Because compliant mechanisms are both structures and mechanisms, the goal is to develop a systematic approach to accomplish topological synthesis and dimensional synthesis. In order to achieve this, structural design techniques can be combined with kinematic principles in order to systematically design compliant mechanisms. Figure 1.9 shows a single-input, single-output compliant mechanism design problem. The design domain is represented by d which represents the unknown topology, size, and shape of the mechanism. The input force, Fin, represents a force applied to the body to produce the desired mechanism motion. The local displacement, uin, represents the magnitude of deformation of the input point along the direction of the input force. The external resistive force, Fex, represents forces generated by the environment of the mechanism which resist the desired motion. The displacement, uout, represents the magnitude of deformation of the output point along the direction of the required motion, . The compliant mechanism design problem can be summarized: determine the best structure, d, in order to most efficiently achieve the required force or motion transformation subject to the applied boundary conditions.
Fe mechanism output

uout uin mechanism input Fi d

structure

ground Figure 1.9: The Generalized Single-Input, Single-Output Compliant Mechanism Problem

1.6 Scope of the Investigation The primary objective of this research is to develop a generalized methodology and an automated approach for designing sophisticated and efficient compliant

24 mechanisms. The investigation will primarily focus on developing a two-stage structural design procedure consisting of (i) topology synthesis and (ii) dimensional synthesis. These two design phases are numerically implemented using classical structural optimization techniques using topology optimization to perform topological synthesis and size and geometry optimization to accomplish dimensional synthesis. Central to the development of these procedures is the formulation of a robust and unified optimization approach which can be effectively applied to both topology optimization and size and geometry optimization problems. Structural optimization techniques are developed using the assumptions of linear, static structural response which will be shown to be reasonable for many compliant mechanism design problems. Finally, to illustrate the effectiveness of the developed design tools, several novel compliant mechanism designs are created using the automated two-stage approach. 1.7 Organization of the Dissertation The dissertation is divided into seven chapters. Chapter one, introduces the compliant mechanism and defines the advantages and uses of these devices. Chapter one also sets the stage for the design problems associated with generating a high-performance compliant mechanism and the justification for automating such an approach. Chapter two introduces the history of the compliant mechanism including an overview of flexure and flexible link mechanisms. Chapter two surveys past and present methods to design compliant mechanisms including kinematic-based approaches and more recent structural optimization based approaches. Chapter three defines two new optimization formulations based on maximizing the energy efficiency of a compliant device. Central to this chapter is the analytic solution of a simple compliant lever, which illustrates that the solution to the energy efficiency problems are well-posed and consequently provide robust optimization of topology and size and geometry. Chapter four extends the optimization problem to a numerical form and introduces a finite element based approach for modeling

25 and optimizing compliant mechanisms. Here, the analytic calculation of the gradients of the objective function and the constraints (also known as the sensitivity analysis) are detailed. Chapters five and six detail numerical results of the topological and dimensional synthesis of many sophisticated and high performance compliant devices. Chapter seven summarizes the achievements of the work and presents future directions and extensions of the dissertation.

CHAPTER II LITERATURE REVIEW 2.1 Compliance in Design Traditional engineered devices have been predominantly designed to be strong and rigid. Often, strength and rigidity are critical to achieve designs which are practical, safe, and efficient. However, the inclusion of flexibility in conventional mechanism design can offer many potential benefits including ease of assembly, greater resistance to overloads, and elimination of backlash and friction associated with conventional mechanical joints. A question then arises: how can compliance can be included as a preferred effect to achieve controlled transmission of forces and motions? Engineers have yet to fully utilize compliance in design, mainly due to the increased complexity that deformation imposes on conventional kinematic analyses. The main challenge has been to integrate the fundamentals of kinematics and structural mechanics to develop a rigorous and systematic methodology to design compliant mechanisms. Over the years, researchers have developed two primary approaches: kinematics-based approaches, which build on traditional rigid-link design techniques, and structural optimization-based approaches which seek to synthesize the optimum compliant structure. 2.2 Kinematic Approaches Because the deformation of elastic members greatly complicates design procedures, the development of rigid-link models to approximate structural response was practical to simplify the design of compliant devices. These methods have been successfully used to design lumped compliant mechanisms as well as mechanisms with

26

27 cantilever-like flexible links. Kinematic approximation techniques build on the wealth of traditional rigid-link mechanism design. Classical design techniques using kinematic chains, Burmester theory and graphical synthesis methods have been modified for design of compliant mechanisms. The weakness of these methods is that these approximation may not adequately model the structure under all conditions. Never the less these kinematic approximation techniques have been successfully used to design certain types compliant mechanisms. 2.2.1 Flexure Devices Aside from energy absorbing springs, devices known as flexures were the first use of compliance for mechanism design. Lumped compliant mechanisms obtain motion through thin flexible segments which can be approximated by revolute hinges. These simple devices have been primarily used in force transducers and precise positioning devices where the absence of backlash makes their performance superior to rigid-link mechanisms. Generally, flexures devices are designed with the assumption that external forces (which may oppose motion) are negligible. The disadvantage of concentrating the flexion within a small region is that lumped compliance concentrates stresses at the flexural segments. This limits the range of flexibility and the load carrying capacity of the mechanism. Synthesis for flexures has been primarily performed by using classical methods developed for rigid-link mechanisms. Revolute joints are then replaced by the appropriate lumped compliant segment. The dimensions of these compliant hinges have mainly been designed using trial and error methods. Paros and Weisbord (1965) provided a detailed analysis and approximation of the behavior of flexural pivots. They developed equations to calculate the critical spring rates for angular and linear deflections produced by loads on all three hinge axes of various types of flexure hinges. Tuttle (1967) and Weinstein (1965) addressed the advantages of flexure devices and provided many example micro

28 positioning mechanisms. The drawback of most of these works is that they concentrate on a specific application while failing to address mechanism design in a generalized sense. Due to the interest in amplifying smart materials actuators, flexures have received more recent attention. Fukuarwa et al. (1991), Hall et al. (1995), and Huang et al. (1996) presented flexure devices which were designed to amplify the displacements of piezoelectric actuators. Using many of the techniques described above, these mechanisms achieved limited results meeting performance goals set by the designers. 2.2.2 Flexible Link Mechanisms In 1966, Burns and Crossley reported the first approach for type synthesis of flexible link mechanisms. According to their definition, a flexible link mechanism primarily consisted of no more than four revolute joints and at least one flexible link. Type synthesis methods categorized structural permutations of flexible-link four-bar chains with revolute pairs. Burns and Crossley considered the problem designing these types of mechanisms in order to achieve stable configurations under the influence of external forces at various position settings. Continuing their research, Burns and Crossely (1968) presented an overlay technique for the design of a four-bar linkage with an axiallyflexible couple link which was designed to achieved constant output torque. Included in their work is the approximation of the tip motion a cantilever beam using a circular arc with the center a located at one-sixth of the beam length from the fixed end. This enabled synthesis of a flexible-link mechanism using conventional techniques. The flexible-link mechanism was then realized by replacing of one of the rigid members with a flexible member 1.2 times longer. Shoup and McLarnan (1971a) presented type synthesis enumeration schemes for both planar and spatial flexible link mechanism and developed classifications for such mechanisms. Shoup and McLarnan (1971b) also used the equations of undulating elastica for approximating the large deflection analysis of flexible link mechanisms. Winter and

29 Shoup (1972) presented a displacement analysis of a four-bar mechanism with a flexible crank and developed equations for tracing coupler curves based on boundary conditions experience by the coupler. Sevak and McLarnan (1974) used non-linear finite element analysis combined with gradient based programming approach to minimize the error between a candidate function and the deformed shape of a flexible coupler link. 2.2.3 Pseudo Rigid Body Model Her and Midha (1987) provided one of the first works which defined the term compliant mechanism. Their research defined link compliance and the degree of compliance concept which modified Grublers equation to address the total degrees-offreedom of a compliant mechanism. The concepts of link compliance was additionally extended to perform type synthesis of compliant mechanisms. Midha et al. (1994) refined the classification of compliant mechanisms and separated the functionality of compliant mechanisms from flexible structures. Midha addressed the classification and behavior of segments and links and discussed potential methods to further abstract compliant mechanism design. Murphy et al. (1996) investigated topology synthesis of compliant mechanisms and developed an approach using graph theory combined with vertex-vertex adjacency matrix method to identify all non-isomorphic (unique) compliant mechanism topologies. By adding flexure pivots and compliant segments to traditional graph theory, Murphy illustrated that 1689 non-isomorphic topology permutations could be derived from a four-bar mechanism configuration. Research efforts have also concentrated on compliant mechanisms mobility. By addressing segment compliance and creating virtual rigid segments, Ananthasuresh and Howell (1996), provided a more systematic means to estimate the degrees of freedom of compliant mechanisms. In 1990, Hill and Midha implemented a simple and efficient large deformation analysis method, called the chain algorithm, for fast analysis of compliant mechanisms. In this method a compliant mechanism is treated as a cantilever beam fixed at one end and

30 free at the other end. If a mechanism has a closed configuration (both ends fixed), boundary conditions are imposed using shooting methods. Her et al. (1992) presented trial and error methods which were used in conjunction with the chain algorithm to design compliant mechanisms for function synthesis. Salamon and Midha (1992) studied the mechanical advantage aspects of compliant mechanisms. They concluded that as the energy storage of a compliant mechanism is minimized, the mechanical advantage will remain nearly constant over a broad range of operating conditions. Their work also proposed techniques to measure mechanical advantage operating under three boundary conditions cases. One of the cases is based on a compliant crimper grasping a flexible workpiece where the mechanical advantage of the mechanism changed depending on the spring-rate approximation of the workpiece. Howell and Midha (1994) introduced an approach to design lumped compliant mechanisms. The method is based on conventional rigid-link techniques to design compliant mechanisms for a desired mechanical advantage. A simple optimization approach was utilized to determine the sizing of the small length flexural pivots, subject to stress, and geometry constraints. Howell et al. (1996a), introduced the pseudo rigid body model to approximate the behavior of cantilever-like flexible segments. As shown in Figure 2.1, the tip displacement of an end loaded cantilever beam subject to non-linear geometric displacements can be approximated a simple rigid link mechanism. Using this approach, a compliant mechanism can be modeled as a rigid-link mechanism using torsional springs to model the energy storage. More complex mechanisms can be builtup by combining pseudo rigid body models together. Howell and Midha (1996b) combined the Burmester theory with the pseudo rigid body model to perform motion generation for compliant mechanisms with prescribed energy storage. Here the energy equilibrium and the kinematic equations are weakly or strongly coupled depending upon the number of precision points. The pseudo rigid body model concept has also been

31 extended to initially curved beams and beams with axial loads (Howell and Midha, 1996c).
F F

torsional spring

cantilever beam

pseudo rigid body model flexural pivot flexible link

rigid link torsional springs revolute joint

rigid-link mechanism (synthesis)

pseudo-rigid-body-model (analysis & design)

compliant mechanism (dimensionalized design)

Figure 2.1: Illustration of the Pseudo Rigid Body Model

2.3 Structural Optimization Past structural optimization techniques have primarily been developed to find the lightest, stiffest, and strongest structure. Researchers have developed both analytical and numerical techniques to solve a multitude of structural design problems. Only recently has structural optimization been employed for compliant mechanism design. The generality of structural optimization methods allow for a more fundamental approach for compliant mechanism design by integrating principles of kinematics and structural mechanics. Past compliant mechanism research has mainly concentrated on developing topology optimization techniques to automate topology synthesis of compliant mechanisms.

32 2.3.1 Historical Foundations The origin structural optimization can be traced back to Galileo who obtained the optimum shapes of variable depth beams to design solids of equal resistance. Barnett (1966) and Hemp(1973) illustrate highlights in the history of the subject from Galileo to the 1960s. For the most part, structural design techniques have mainly concentrated on the idea of obtaining the stiffest or strongest structure given a fixed volume or weight limitation. In general, structural design methodologies have been approached using material distribution problems and parametric geometry based problems. Material distribution problems seek to find the optimal location of material within a designated design domain. These approaches are primarily used for topology or layout synthesis problems. Parametric geometry based design techniques involve the design of the shape and size or boundary of a structure. Both techniques offer advantages and disadvantages. For example, topology optimization techniques offer the largest potential for increases in structural performance; however, shape refinement techniques are still required when local constraints such as stress constraints and buckling constraints are to be considered during optimization. Topping (1983) and Levy et al. (1987) present comprehensive reviews of techniques developed to perform topology optimization of discrete structural systems. One of the first researchers to examine the optimal layout of structures, Michell (1904) developed analytical techniques to perform layout design of frame structures. Michell structures are essentially truss-like continua with an infinite number of members and joints. This development is based on the premise that the members of a minimum weight structure should follow the directions of principle strains. Figure 2.2 illustrates a cantilever-like Michell structure. Several researchers including Hemp (1973), and Prager (1974) (1977) (1978) have presented techniques to modify and make practical uses of Michells theory. Pederson (1972) presented an iterative technique for the design of plane

33 trusses subject to a single loading case using a sequence of linear programming with move limits. Both member areas and nodal coordinates are used as design variables. Prager and Rozvany (1977) investigated the optimal layout of grillages (beam grids), which were designed for minimum weight (or volume) subject to a constraint on the compliance. Bendsoe et al. (1994) developed an optimality criteria for the large-scale optimization of truss structures. Using single, multiple and self-weight loading, Bendsoe derived a design variable update techniques to produce the minimum compliance (stiffest) structure. Also presented in the paper is an investigation of node coordinate optimization (geometry change). An example reveals that the alteration of node coordinates can further improve the performance of topology designs.

Figure 2.2: A Cantilever-like Michell Structure

Bendsoe and Kikuchi (1988) developed a perforated microstructure model to solve continuum-based structural optimization problems. Bendsoe and Kikuchis technique, call the homogenization method, is capable of producing the optimal topology and shape of a continuum structure given a fixed-grid design domain, and prescribed loading conditions. The foundations of this work is supported by the concept of the relaxation of

34 the design domain and is based on designing the orientation and sizing of micro-voids located within the material continuum. An alternative approach for structural design is to manipulate or vary the geometric boundary or shape of the structure. These approaches are practical for reducing the number of design variables while considering local constraints, such as stress and local buckling constraints (Haftka and Gurdal, 1992). Haftka and Grandhi (1986) present a detailed overview of shape optimization techniques that were developed in the 1970s and early 1980s. One of the earliest works on shape optimization was reported by Zienkiewicz and Campbell (1973) who presented an optimization technique using the nodal coordinates of planar continuum finite elements as design variables. Researchers following this work have implemented polynomial, spline and orthogonal functions which provide smoother solutions while simplifying the number of design variables needed to describe the boundary. Kristensen and Madsen (1976) defined the boundary using a linear combination of orthogonal functions which modify the starting design. Here, the coefficients of the orthogonal functions are the primary design variables. Imam (1982) presented a technique to optimize the shape of a structure using key (or master nodes). Intermediate nodes were positioned using polynomial blending functions based on the positions of the key nodes. Braibant and Fleury (1984) developed Bezier and B-spline based techniques for shape design. A difficulty which can arise in shape optimization is the need to re-mesh the design domain during the optimization in order to maintain an accurate finite element model. Diaz et al. (1983), demonstrated that the use of a good adaptive mesh refinement strategy can avoid improper (jagged and non-sensible) optimization results. 2.3.2 Structural Optimization of Compliant Mechanisms Research in structural optimization of compliant mechanisms has primarily concentrated on the mathematical formulation of optimization problems suitable for

35 topology synthesis of compliant mechanisms. Problems have ranged from designing material microstructures that exhibit a prescribed elastic response, to determining structures having maximum ratio of flexibility to stiffness. Parametric geometry based techniques have not yet been examined although a technique called image based design which extracts a smoothed image of the homogenized design has been developed Nishiwaki et al. (1998b). The first investigation of topology optimization for compliant mechanism design was investigated by Ananthasuresh (1993) (1994). Using the homogenization design method, Ananthasuresh proposed six different optimization formulations for compliant mechanism design. Although these formulations showed limited usefulness in terms of successfully optimizing compliant mechanism topologies, the work highlighted many issues fundamental for compliant mechanism design and offered a foundation for further development of optimization problems. Based on the design for deflection problems investigated by Chern and Prager (1970), Shield (1970) and Huang (1971), Ananthasureshs first formulation posed the compliant mechanism synthesis problem as one of design for deflection, where the goal is to achieve a specified displacement at a particular point. The problem is posed by minimizing the weight of a structure, subject to a displacement constraint at the output. Sigmund (1995a) implemented this formulation using a truss element ground structure and a sequential linear programming approach where the desired displacement at the output port was prescribed using a penalty method. Sigmund noted that optimal topologies for the truss based mechanisms tended to be singular for simple motion requirements and consequently could collapse under imperfect loading. For this reason, Sigmund investigated multiple loading cases where one of the loadings was used to avoid these types of instabilities. Sigmund (1994) (1995b) also proposed an optimization method to design material microstructures having prescribed constitutive parameters. Using truss and frame elements as well as a density based homogenization techniques, Sigmund proposed an

36 optimization problem to minimize the weight of a material microstructure subject to a constraint on the constitutive parameters. Using the stationary conditions of the Lagrangian, an iterative update scheme was developed to solve the constrained optimization problem. Both two and three dimensional material microstructures were synthesized. Sigmund noted that the frame element ground structures produced the same topologies as the truss element ground structures. Kikuchi et al. (1997) also investigated material microstructure formulations using the minimum weight subject to a prescribed elasticity tensor constraint. Their work used the homogenization approach to synthesize material microstructures with prescribed modulus and poissons ratios. A Gaussian filtering method was introduced to avoid hinging and checkerboard patterns which can lead to impractical microstructure designs. By varying filtering parameters, either detailed or course microstructures could be generated. Larsen et al. (1997), incorporated a density based homogenization approach to design the topology and shape of compliant mechanisms. Larson proposed a least squared formulation to minimize the discrepancy between a measure of the mechanical advantage with a measure of the geometric advantage. This formulation allows the designer to creates mechanisms which minimize the error between the measured and required mechanical advantage (geometric advantage). Resulting topologies achieve the required kinematic relationships; however, many of the optimal designs exhibit thin flexure-like components. Ananthasuresh (1994) also investigated a weighted-sum, multi-criteria formulation where the mechanism should be flexible to deform appropriately and should be stiff to exert maximum force while grasping a workpiece. Here, the measure of a mechanisms flexibility was described as maximizing the output deflection along a desired direction vector given an input load. The measure of a mechanisms flexibility along a desired direction is defined as the mutual potential energy (Shield, 1970). Conversely, the mechanisms stiffness can be measured by constraining the input and applying an external

37 load which opposes the motion of the output. The strain energy or compliance under this state equates to a measure of the mechanisms stiffness. The two objectives were combined using a weighted sum multi-criteria formulation where the goal was to maximize the mutual potential energy (flexibility) while minimizing the strain energy (maximizing stiffness). Designs using the linear weighting method could generally be characterized as having inadequate flexibility. It was postulated that the linear combination was inappropriate for optimization due to scaling differences between the two objectives. Frecker (1997) and Frecker et al. (1997a), posed the multi-criteria formulation as the ratio between the mutual potential energy and the strain energy. Using a ground truss approach, Frecker used a sequential linear programming algorithm to design topologies having an optimal flexibility to stiffness ratio. Freckers results showed improved optimization convergence over those by Ananthasuresh; however, some problems met with convergence difficulties. For this reason, a non-linear material penalty was implemented to aid optimization convergence (Frecker et al. 1997b). This improved the topology synthesis results, but provided minimal insight as to original nature of the convergence difficulties. In order to investigate fully distributed compliance, truss optimized truss topologies were further refined to equalize the distribution of strain energy throughout the mechanism. Results indicated a general decrease in mechanism performance. Nishiwaki et al. (1998a) (1998b) and Nishiwaki (1998), performed a thorough investigation of the Pareto optimality conditions of the multi-criteria compliant mechanism formulations. The Pareto conditionality states that (Stadler, 1988): The set of all Pareto optimal designs represents a boundary of the design variable vector x* which satisfies the multi-criteria optimum condition. This condition essentially states that x* will be Pareto optimal if there does not exist another feasible x that would cause a decrease in one criteria without causing a simultaneous increase in at least one other criteria.

38 Nishiwaki utilized the homogenization approach and investigated several different formulations for topology optimization of compliant mechanisms. One of Nishiwakis formulations posed the problem as the sum of a weighted logarithmic multi-criteria formulation which is a more general form of Freckers original formulation. The advantage of this method was that it could determine all Pareto optimal designs. However, it was noted that not all points along the Pareto boundary will result in an acceptable compliant mechanisms topology. Nishiwaki also proposed optimization formulations for dynamic response (design for a given natural frequency) and multiflexibility (multiple input and output) type problems. Ananthasuresh (1994) also proposed a spring model to capture the resistance of a workpiece being grasped by the mechanism. The workpiece was modeled by incorporating a spring at the output which resists the desired output motion. In this case, the objective was to maximize the work done on the spring by the compliant mechanism. Sigmund (1997) presented an approach which maximized the mechanical advantage of a compliant mechanism subject to a total input constraint. Using a density based homogenization method and a sequential linear programming approach, Sigmund optimized compliant mechanisms for three different cases of a mechanism grasping a workpiece. The first case was where the workpiece was infinitely rigid. The second case was where the workpiece possessed a finite stiffness. The last case was where there existed a gap between the workpiece and the mechanism. Results of the optimization illustrate that varying the input displacement constraint as well as the spring stiffness had a pronounced change on compliant mechanism topologies. Saxena and Ananthasuresh (1998) proposed several formulations which modified the standard mutual potential energy to strain energy ratio proposed by Frecker. In addition, they also proposed a formulation which maximized the output work performed on a linear spring while minimizing the total work performed at the input; however, they presented this formulation as a variation of the original multi-criteria formulation. Their research

39 focused on developing an optimality criteria and an update scheme for the large scale optimization of compliant mechanisms. Saggere and Kota (1997) and Saggere (1998) proposed a new approach for the design of compliant adaptive structures by minimizing the least squared error between the desired and measured shape change. Synthesis techniques were developed for the case of a simple four-bar compliant mechanism where the moments and dimensions of the structure are determined to account for the desired structural change. Techniques were also extended to complex structural systems built-up using networks of beam elements. To perform topology synthesis, the location of the input moments was first determined to minimize the number of actuators to effect a prescribed shape change. Dimensional synthesis was then accomplished by determining the size of each individual beam crosssection from the built-up structure.

CHAPTER III PROBLEM FORMULATION 3.1 Design Approach Many of the applications for compliant mechanisms require the device to provide efficient and controlled transmission of forces and motions. Successfully designing a functional compliant mechanism can require the consideration of many, if not all, of the following criteria (Hetrick, Kota, 1999): Required kinematic motion Desired mechanical or geometric advantage Appropriate stiffness and flexibility trade-off Material properties (modulus, strength, etc.) Geometric non-linearities and buckling instabilities Dynamic considerations Weight limitations Manufacturability It is easy to generalize that producing a high-performance compliant mechanism for a specific application is a complex problem with many trade-offs. Design criteria often conflict with each other such that satisfying one criteria requires sacrifices from another. This leads to the notion that problems must be optimized while considering many potentially conflicting design criteria and constraints. The question then is how to appropriately measure the performance of a compliant mechanism and how to formalize

40

41 the desired characteristics of an optimal compliant mechanism in a mathematically appropriate sense. The challenge in designing compliant mechanisms lies in synthesizing an ideal structural form (i.e. the topology, size and shape) that best performs a desired mechanical function. For consistency here and throughout the rest of the dissertation, the term compliance shall refer to the work performed as a force acting on an elastic body moves through its corresponding displacement, which is equivalent to twice the strain energy (units of force*length). The term flexibility shall refer to ease at which the structure deforms when subject to an input force or displacement (units of length/force). The term stiffness shall refer to the relative stiffness of the structure between the input and output ports (units of force/length). Typically the stiffness is measured by constraining the input and applying an external force acting against the desired output motion. 3.2 Introduction to the Compliant Lever Model When practical, it is advantageous to construct a simple model to check if the solution is well-posed before beginning larger, more complex problems. A two-thickness compliant lever is presented in order to investigate the convergence behavior of both topology optimization and size and geometry optimization problems. Note that the topology of the structure has been chosen a priori such that the basic kinematic requirements are satisfied; however, both topology optimization and size and geometry optimization are based, at least in part, on a resizing procedure. While the model is a simplification of the optimization process, it is an appropriate simplification which can be shown to mimic the behavior of more complex topology optimization and size and geometry optimization problems. Figure 3.1 shows the two-thickness compliant lever. The output port is located at the first third of the total beam length while input port is located at the end of the beam.

42

input force or displacement l h1 2l E h2 desired output motion b

external resistive force Figure 3.1: The Two-Thickness Compliant Lever

In order to develop a robust formulation for compliant mechanisms design, the optimization formulation in question must be feasible under all possible variations of objective function and constraint parameters, and capable of providing a unique and meaningful solution for the two element thicknesses. Figure 3.2 illustrates the potential for non-convergence of the compliant mechanism solution where 3.2 (b), (c), and (d) represent unacceptable solutions. This behavior leads to the following postulate: If an optimization problem displays a null-solution for the simple compliant lever, then the formulation is likely to fail to provide appropriate convergence for larger, more complex compliant mechanism design problems.

h1 > 0

h2 > 0

h1 = 0

h2 > 0

(a) Acceptable Compliant Mechanism Solution

(b) Infinitely Flexible Solution

h1 > 0

h2 = 0

h1 = 0

h2 = 0

(c) Maximum Stiffness Solution

(d) Infeasible Solution

Figure 3.2: Illustration of the Various Solutions to the Compliant Lever Problem

43

While many of the past compliant mechanism optimization formulations do appropriately solve the compliant lever problem, all of these formulations can fail due to constraint violations, or the objective function becoming ill-posed under certain design parameters. For example, the least-square mechanical/geometric advantage optimization problem proposed by Larson (1997) does not account for energy storage due to an external force and consequently produces infinitely flexible mechanisms designs. The maximum mechanical advantage formulation proposed by Sigmund (1997) can produce acceptable compliant mechanism designs; however, the constraint limiting the maximum input displacement can produce infeasible optimization problems if the structure cannot appropriately satisfy this constraint. The multi-criteria approaches proposed by Ananthasuresh (1994), Frecker (1997) and Nishiwaki (1998) are also prone to convergence difficulty due to the fact that not all Pareto optimum points represent feasible compliant mechanism structures (either infinitely flexible or maximum stiffness structures). Research has shown that the choice of an acceptable weighting factor may depend upon the type of finite element used to model the design domain as well as the geometry of the problem. 3.3 Research Objectives Because compliant mechanisms are structures, the principles which govern the ability of a structure to mimic a mechanism (ideally an efficient mechanism) must be fully understood before a robust optimization problem can be developed. In order to determine a basis for understanding compliant mechanism design, the qualities of an ideal compliant device must be understood, including how compliant mechanism should trade-off stiffness and flexibility, and how this trade-off relates to the efficient transmission of forces and motions. Of primary importance is the development of a simple and robust optimization formulation which can be applied for linear static compliant mechanism design. The

44 optimization formulation should be equally applicable to topology synthesis problems as well as size and shape refinement problems, and should work effectively for a variety of modeling and design parameterization techniques. The developed approach should also allow the designer to optionally constrain the kinematic behavior of the mechanism. While such a constraint may impose limits on the feasible design space, the ability to control mechanisms mechanical or geometric advantage would allow greater flexibility and control by the designer. The topology synthesis and the size and shape synthesis design tools can then be used in conjunction to develop an automated two-stage approach for the synthesis of high-performance compliant devices. The goals of this research are summarized: 1. Determine a fundamental understanding of compliant mechanism design including the characteristics of an ideal compliant device. 2. Formulate a robust optimization problem which appropriately solves the simple compliant lever which will be equally applicable for topology synthesis problems as well as size and shape optimization problems. 3. Implement these formulations for topology synthesis and demonstrate the improved convergence robustness. In addition, investigate the ability to optionally control a mechanisms mechanical or geometric advantage. 4. Implement these formulations for size and shape optimization to design practical mechanisms subject to optional stress and kinematic constraints. 5. Using the developed two-stage approach, demonstrate the synthesis of practical, high-performance compliant mechanism designs

45 3.4 Energy Efficiency Objective Formulations Considering a linear, static structure, two of the most important criteria to be satisfied during compliant mechanism design are: kinematic requirements and structural requirements. The first and foremost principle of a mechanism is to generate appropriate output motions when subjected to a given input force or displacement. Thus the output should move in the appropriate direction and magnitude given an input force or displacement. Compliant mechanisms must also satisfy structural requirements. A compliant mechanism must be designed with adequate stiffness and sufficient flexibility. Adequate stiffness is required to provide resistance to external forces. Sufficient flexibility is required for minimal energy expenditure as the mechanism deforms under the applied loading. A novel approach for designing compliant mechanisms can be based on maximizing the transfer of work from an input source to an output sink

out CM = ------------------- .

Work Work in

(3.1)

In order to formulate this type of problem, work must be appropriately measured at the input and output ports. The force-displacement history for both the input and output ports can then be combined to measure the energy efficiency of the mechanism. If the efficiency of a compliant mechanism is maximized, the mechanism will satisfy kinematic requirements while simultaneously minimizing the stored strain energy under the imposed boundary conditions. The advantage of this approach is that multi-criteria methods are avoided, and the optimal design will depend directly on the boundary conditions applied to measure input and output work.

46 3.4.1 The Force-Displacement Efficiency Formulation Considering a static, linear elastic body, the mechanism characteristics can be measured at both the input and output ports by simulating actuation of the structure and measuring its response. This is performed by applying the following boundary conditions: The input is actuated by controlling the displacement of the input port on the body and measuring the reaction forces under equilibrium. The input displacement in effect simulates the maximum range of travel of the mechanism. Work performed at the output is measured by applying an external resistive load which opposes the desired direction of the output port on the body and measuring the output displacement under equilibrium. The external load can be considered to be a load arising from the environment of the mechanism. As shown in Figure 3.3, these boundary conditions are applied in two separate stages. First, the external load is applied while the input is held fixed (the body is considered to be unactuated). The output displaces in the opposite direction of the

desired motion while the actuator generates a reaction force to sustain the position of the input. With the external load applied, an additional force is applied to the input which serves to actuate the mechanism; moving the output in the desired direction and magnitude.

47

load

Fex = mg uout1

load uout2 actuator

Fex = mg

actuator

uin

Fin1 (a) Boundary Conditions Stage1

Fin2 (b) Boundary Conditions Stage2

Figure 3.3: A Mechanism Pushing Against a Constant External Force

The force/displacement history for both input and output ports then displays the behavior shown in Figure 3.4. At the input port, uin represents the applied input displacement. The variables Fin1 and Fin2 represent the initial unactuated input force and the final actuated input force respectively. At the output port, the force, Fex, represents the applied external load. The displacement uout1 represents the loaded, unactuated position of the output port; uout2 represents the loaded, actuated position of the output port. Additionally, uout0 represents the unloaded, actuated displacement of the output port.
uout0 Fin2 Force Force Fin1 Fex unloaded

uin displacement (a) Input History

uout1 uout2 displacement (b) Output History

Figure 3.4: The Input and Output Work History (Force-Displacement Formulation)

48 Unlike rigid-link mechanisms, the kinematic properties of a compliant mechanism are a function of the applied boundary conditions. The ratio of Fex / Fin1 is defined as the initial mechanical advantage. The ratio of Fex / Fin2 is the secondary mechanical advantage. The ratio of (uout2-uout1) / uin is the loaded geometric advantage. The ratio of uout0 / uin is defined as the unloaded geometric advantage. Given an input force or displacement, the compliant mechanism must deform properly such that the output moves in a specified direction. For a linear elastic structure (for which superposition holds), work performed by moving a set of loads through their corresponding displacements due to the application of an alternative set of loading is governed by the Reciprocal Work Theorem (Gere and Timoshenko, 1984) which states: The work done by the forces in the first state of loading when they move through their corresponding displacements in the second state of loading is equal to the work done by the forces in the second state of loading when they move through their corresponding displacements in the first state of loading. The reciprocal equation then states

Pi Pi
i=1

Qj Qj
j=1

(3.2)

Where the force Pi are the forces applied in the first state of loading and the displacement Qj are the corresponding displacements. Conversely, the forces Qj are applied in the second state of loading and the displacements, Pi are the secondary corresponding displacements. The Reciprocal Work Theorem holds for any number of forces acting on the body, even forces that are superposed on one another, as long as the measured displacements are colinear with respect to the applied forces. Thus, according to the theorem, the shaded rectangles shown in the force-displacement history of the compliant mechanism have equal areas. Additionally, the theorem can also be used to show that the

49 initial mechanical advantage is always equal to the inverse of the unloaded geometric advantage. The reciprocal work equation can be rewritten in terms of the input and output force and displacement variables used for the force and displacement measurement of the compliant mechanism,

F in 1 u in = F ex u out 2 .

(3.3)

The triangular regions shown in Figure 3.4 in both the input and output forcedisplacement history represent the change in strain energy or compliance due to a set of loads moving through their corresponding displacements. The total compliance stored in the mechanism can be divided into work due to loading and work due to the applied strain or actuated motion of the mechanism. Using the force-displacement history, the total compliance of the mechanism can be written

1 1 - F ex u out 1 + -- ( F F in 1 ) u in . C total = -2 2 in 2

(3.4)

As noted by Kikuchi (1999), there is a duality between the compliance and the stiffness of a structure. Given an applied force, minimizing the compliance produces a structure which possesses greatest resistance (stiffness) to the applied force. Conversely, given an applied displacement, maximizing compliance produces a structure which possesses greatest resistance to the applied displacement. Much of the work on structural optimization has focused on designing the stiffest structure; however, the principles of compliance can also be extended to the case where flexibility is desired. If an input force is applied, maximizing the compliance will produce the structure possessing maximum flexibility. Conversely, if an input displacement is applied, minimizing the compliance will produce the structure possessing maximum flexibility.

50 For the linear static system, the strain energy or compliance stored in the elastic body limits the available energy which can be transmitted to the output. The goal is to minimize the total strain energy or compliance stored in the mechanism which will allow the greatest amount of energy to be transmitted from input to output. As pointed out by Salamon and Midha (1988), the mechanical advantage of a compliant mechanism will remain nearly constant over a wider range of boundary conditions as the mechanisms internal strain energy is lowered. Thus compliant mechanism that store a minimal amount of strain energy, will not only be efficient, but will also have near constant mechanical and geometric advantages. The optimal flexibility to stiffness trade-off is posed as a minimum compliance problem. The flexibility of the mechanism is measured using an applied displacement. The stiffness of the mechanism is measured by fixing the input and applying an external force to measure the stiffness of the structure. Thus, the problem of minimizing the total strain energy stored in a compliant mechanism can be posed using an input displacement and an external force.

1 1 min C total = -- F ex u out 1 + -- ( F F in 1 ) u in 2 2 in 2

(3.5)

A measure of the energy efficiency of a compliant mechanism can be formulated by placing the reciprocal work performed at the output over the reciprocal work performed at the input plus the total compliance. If the efficiency of a compliant mechanism is maximized, the mechanism will satisfy kinematic requirements while simultaneously minimizing the stored strain energy under the imposed force-displacement boundary conditions

51 F ex u RW RW in + C total

out 2 out FD = ----------------------------------------------------------------------------------------------- = -------------------------------- .

1 1 F in 1 u in + -- ( F F in 1 ) u in + --F u 2 in 2 2 ex out 1

(3.6)

Since the reciprocal work at the input and output are equal, the objective may be simplified by dividing by the reciprocal work. Maximizing the energy efficiency can be shown to be equivalent to maximizing the reciprocal work at the output over the total compliance

RW out RW out -------------------------------- -------------- . RW in + C total C total

(3.7)

Because the values of Fex and 1/2 only serve to scale the magnitude of the objective function, the formulation can be further simplified to maximizing the output displacement while minimizing the total compliance

F ex u out 2 u out 2 - . ------------------------------------------------------------------------- ---------------------------------------------------------------1 ( F in 2 F in 1 ) u in + F ex u out 1 -- ( ( F in 2 F in 1 ) u in + F ex u out 1 ) 2

(3.8)

Maximizing uout2 for a given uin is equivalent to maximizing the geometric advantage of the compliant mechanism. This equivalency leads to the following conclusion:

For the force-displacement energy formulation, the mechanism possessing maximum energy efficiency will maintain the largest ratio of the geometric advantage to the total compliance.

Using concepts of linear superposition the following relationships can be used to simplify the measurement of mechanism response:

52 u out 2 = u out 0 , (3.9)

( F in 2 F in 1 ) u in = F in u in .

(3.10)

Where uout0 is the unloaded output displacement given the fixed input displacement uin. The force Fin is used to represent the difference between the initial reaction force Fin1, and the final reaction force, Fin2, and is measured as the input force required to achieve the fixed displacement, uin, in the absence of the external force, Fex. Thus, the objective formulation can also be expressed:

u out 0 F FD = ----------------------------------------- . F in u in + F ex u out 1

(3.11)

3.4.2 The Spring-Efficiency Formulation To measure the energy throughput of a compliant mechanism, an external spring is applied at the output point to resist the motion of the mechanism. This is similar to formulations proposed by Ananthasuresh (1994) and Sigmund (1997); however, the inclusion of the input and output energy in order to calculate the mechanical efficiency of the mechanism is a new approach to the formulation. The spring efficiency formulation presented in this thesis was first published by Saxena and Ananthasuresh (1998). However, the formulation was independently developed and the resulting conclusions and contributions from this thesis are different than in their work. Specifically, their work failed to address fundamental issues concerning the convergence of optimal designs. In addition, they did not differentiate the spring-energy formulation from earlier developed multi-criteria formulations which possesses distinctly different convergence behavior. Consequently, although the formulation of the optimization problem is the same, the

53 research contribution by implementing and studying behavior of the formulation is different and unique to this dissertation. The efficiency of a compliant mechanism can be measured by appropriately measuring the force-displacement history for the input and output points on the body. Figure 3.5 shows the elastic body pressing against a soft object which could represent the stiffness of a workpiece that the mechanism is grasping.

workpiece

Ks

uout

actuator

uin

Fin Figure 3.5: A Mechanism Grasping a Soft Workpiece

Here, the deflection, uout, due to the input force, Fin, is taken in the direction of the desired output motion which is directly opposed by the spring, Ks. The input deflection, uin, due to the input force, Fin, is taken in the direction of Fin. The force/displacement history for both input and output ports then displays the behavior shown in Figure 3.6. At the input port, uin represents the applied input displacement and Fin represent the input force required to compress both the compliant mechanism and the external spring. At the output port, Ks, represents the stiffness of the spring opposing the output motion. The displacement uout represents the final output displacement as the compliant mechanism deforms and presses against the workpiece. The ratio of Fex / Fin is defined as the springloaded mechanical advantage (where Fex is equal to uout*Ks). Conversely the ratio of uout / uin is defined as the spring-loaded geometric advantage. Due to energy storage in the mechanism, the spring-loaded mechanical advantage is not equal to the inverse of the

54 spring-loaded geometric advantage. Finally, if the resistive spring is removed from the output, the unloaded output displacement over the input displacement is defined as the unloaded geometric advantage.

Fin force force Ks

uin displacement (a) Input history

uout displacement (b) Output history

Figure 3.6: The Input and Output Work History (Spring Formulation)

The mechanical efficiency of a compliant mechanism pushing against an external spring can be expressed in equation 3.12. Due to energy storage, the objective function can never be greater than one. Note that the absolute value of uout is introduced to prevent ambiguity in systems which may attempt to produce output motion opposite to the desired direction. The absolute value can be numerically imposed by using a square/square root combination

1 --K u u 2 s out out S = ------------------------------- . 1 --F u 2 in in

(3.12)

The formulation may be further simplified due to the fact that the variables Ks and Fin merely scale the magnitude of the objective function.

55 u out u out F S = --------------------u in

(3.13)

This simplification leads to the following conclusion: For the spring efficiency formulation, the mechanism possessing maximum transmission efficiency will maintain the largest ratio of the square of the output displacement to the input displacement.

3.4.3 Optimization of the Compliant Lever using the ForceDisplacement Efficiency Formulation Returning to the two-thickness compliant problem, the optimization problem may be formulated by maximizing energy efficiency of the mechanisms subject to a maximum volume constraint and local lower bound constraints on the two thicknesses.

max F ( h 1, h 2 ) subject to: blh 1 + 2 blh 2 V max 0 h 1, h 2 0 (3.14)

For the simple cantilever problem, the energy efficiency must be appropriately measured. The response of the system is calculated using the classical approximation for linear elastic beams subject to small displacements and rotations

(M(x)) (M(x)) U = ------------------- dx + ------------------- dx . 2 EI 1 EI 2


0 l

3l

(3.15)

56 Where U is the strain energy stored in the structure, M(x) is the moment loading along the length of the two-thickness cantilever beam, and I1 and I2 are the moments of inertia for each thickness. For some calculations, methods of superposition are employed in order to measure structural response under simultaneous loading conditions. To calculate the displacement at a point, Castiglianos second theorem is utilized where a dummy load, , is applied at the point of interested along the desired direction of motion. Taking the partial derivative of the strain energy with respect to the dummy load, and setting the dummy load to zero produces the deformation of the structure along the direction of the dummy load.

= lim U 0

(3.16)

where i represents the displacement of the structure at a point. An illustration of the two measurement schemes is shown in Figure 3.7 (a) and (b). The flexibility of the structure is calculated by simultaneously applying the vertical input displacement, in and measuring the strain energy or compliance of the structure under a fixed displacement. For simplicity, E, b, and l are set to unity.

57

h1

h2

out0

in

(a) Geometric Advantage and Flexibility Measurement

h1

h2

Pex (b) Stiffness Measurement Figure 3.7: The Compliant Lever Measurement Techniques

Given the fixed input displacement, in, the strain energy of the lever due to the applied input strain can be expressed:

U in

in h 1 h 2 = ---------------------------------- . 3 3 4 ( 8 h 1 + 19 h 2 )

3 3

(3.17)

The output deflection, out0, is measured along the specified direction of motion along the same line as the dummy load . Taking the partial derivative of the strain energy under these two loads and setting the dummy load to zero yields the magnitude of the output displacement along the dummy load direction due to the applied input strain,

out 0

16 in h 2 = -------------------------- . 3 3 8 h 1 + 19 h 2

(3.18)

58 The stiffness of the structure is calculated by constraining the input, and applying an external force resisting the output motion, Pex. Solving the model for the strain energy with respect to the external force, Pex, yields

U Pex

4 P ex ( 8 h 1 + 3 h 2 ) = ---------------------------------------. 3 3 3 h 1 ( 8 h 1 + 19 h 2 )

(3.19)

Substituting the equations 3.17, 3.18, and 3.19 into the objective formula yields

16 in h 1 h 2 max ---------------------------------------------------------------------------- . 2 6 3 2 3 2 3 in h 1 h 2 + 128 P ex h 1 + 48 P ex h 2

3 3

(3.20)

Writing the Lagrangian of the cantilever problem in negative null form results in the following optimization problem,

16 in h 1 h 2 L = ---------------------------------------------------------------------------- + 1 ( h 1 + 2 h 2 V max ) + 2 ( h 1 ) + 3 ( h 2 ) 2 6 3 2 3 2 3 in h 1 h 2 + 128 P ex h 1 + 48 P ex h 2 . (3.21)

3 3

The necessary Karush-Kuhn-Tucker (KKT) optimality condition states that when the gradient of the Lagrangian vanishes, while satisfying constraints on the Lagrange multipliers, the solution represents either a candidate interior optimum point, or a candidate constrained optimum point (Pierre, 1986). Proving sufficiency of the point in question may be accomplished by computing the second order partial derivative of the Lagrangian, termed the Hessian. If the Hessian remains positive definite at the potential optimum point, then the point is at least a local minimizer of the Lagrangian (second order

59 sufficiency). However, if the Hessian is not positive-definite the point may still represent a potential optimum point. Consequently, an optimal point for the optimization problem may be found by solving for all KKT points which satisfy the necessary conditions, and then directly comparing objective function values for all candidate points. For the simple, compliant lever, the necessary KKT conditions are then expressed in equations 3.23 through 3.26

L =

48 in h 1 h 2 ( in h 1 48 P ex ) --------------------------------------------------------------------------------- + 1 2 2 6 3 2 3 2 3 2 ( in h 1 h 2 + 128 P ex h 1 + 48 P ex h 2 ) --------------------------------------------------------------------------------- + 2 1 3 2 6 3 2 3 2 3 2 ( in h 1 h 2 + 128 P ex h 1 + 48 P ex h 2 )


6 2 2 6144 in h 1 h 2 P ex

3 6

= 0 , 0

(3.22)

1 ( h 1 + 2 h 2 V max ) + 2 ( h 1 ) + 3 ( h 2 ) = 0 ,

(3.23)

i 0 ,

(3.24)

h 1 + 2 h 2 V max h1 h2

0 0 . 0

(3.25)

Since all of the terms, h1, h2, Pex, and in, are positive, the following condition is true for L : h2

3072 in h 1 h 2 P ex 1 = --------------------------------------------------------------------------------- . 2 6 3 2 3 2 3 2 ( in h 1 h 2 + 128 P ex h 1 + 48 P ex h 2 )

6 2 2

(3.26)

60 Which means that the total volume constraint is always active at the optimum point. This L relationship can then be substituted into to determine h1

48 in h 1 h 2 ( in h 1 48 P ex ) + 3072 in h 1 h 2 P ex ------------------------------------------------------------------------------------------------------------ 2 = 0 . 2 6 3 2 3 2 3 2 ( in h 1 h 2 + 128 P ex h 1 + 48 P ex h 2 )

3 6

6 2 2

(3.27)

This indicates that a potential KKT solution exists where 2 is active and h1 = 0. However, by inspection, the objective function will be zero for all h1 equal to zero. Clearly, the objective function will be less than zero for h1 greater than zero. Because 1 is greater than zero, a second KKT solution also exists when 1 is active and therefore

48 in h 1 h 2 ( in h 1 48 P ex ) + 3072 in h 1 h 2 P ex = 0 .

3 6

6 2 2

(3.28)

Provided that 3.28 is true, then h2 can be solved in terms of h1 from the active volume constraint:

V max h 1 - . h 2 = ---------------------2

(3.29)

Setting Vmax equal to one and simplifying the equations yields:

3 h 1 ( h 1 6 h 1 + 15 h 1 20 h 1 + 15 h 1 6 h 1 + 1 ) + P ex 6 5 4 3 2 48 ------- ( 61 h 1 110 h 1 + 19 h 1 60 h 1 + 45 h 1 18 h 1 + 3 ) = 0 . 2 in
2

(3.30)

61 Unfortunately, the roots of the polynomial equation 3.31 cannot be solved in closed form. The polynomial has the unique characteristic of having roots at zero and one (which correspond to h1 = 0 and h1 = 1). The polynomial also has an additional root which falls in between zero and 0.32 and depends on the parameter of Pex/in. Examination of the KKT solutions leads to the conclusion that the intermediate root is the optimum point. In order to gain further understanding of the objective criterion, the values of h1 and h2 can be plotted against the ratio of Pex/in for a maximum volume of one. Shown in Figure 3.8 is a plot of the thicknesses, h1 and h2 versus varying values of Pex/in.
0

log(element height)

h2

h1
5 6

8 10

log(Pex/in) more flexible compliant mechanism designs stiffer compliant mechanism designs

Figure 3.8: Plot of h1 and h2 With Respect to the Flexibility / Stiffness Parameter

For small values Pex/in, h1 becomes close to zero; however h1 remains greater than zero provided Pex/in > 0. For larger values of Pex/in, the stiffness of h1 increases and then remains constant. This represents the maximum stiffness design that still retains the appropriate compliant mechanism design characteristics. Intuitively, this makes sense

62 as designs which flex farther and overcome less force will require greater flexibility (h1 smaller) than designs which flex a smaller distance and overcome a large external force. For all values of Pex/in, between the limits of zero and infinity, the values of h1 and h2 remain greater than zero. Consequently. the objective formulation is well-bounded given only a simple volume constraint. As a result, the optimization problem is well-posed for topology, size, and shape synthesis of compliant mechanisms. 3.4.4 Optimization of the Compliant Lever Using the Spring Efficiency Formulation Shown in Figure 3.9 is the cantilever lever as the input force presses down on the lever and the lever simultaneously compresses the spring. To apply the spring to the analytic model, a compatibility equation must be posed to relate the force of the spring (acting at the output point) to the displacement at the output.
Pin l h1 Ks out in 2l h2

Figure 3.9: Measurement of Performance Using the Spring Formulation

Using Castiglianos theorem, the output displacement, out, can be expressed in terms of the input force, Pin, and external spring, Ks

16 P in out = -------------------- . 3 h1 + 4 Ks

(3.31)

63 The strain energy can be written as a function of the applied input force, Pin and the spring stiffness Ks,

U Pin

4 P in ( 8 h 1 + 19 h 1 h 2 + 32 K s h 1 + 12 K s h 2 ) = ------------------------------------------------------------------------------------------------------------- . 3 3 3 h1 h2 ( h1 + 4 Ks )

(3.32)

Writing out the efficiency equation produces

16 P in 2 K s ---------------------- h 1 3 + 4 K s F = ------------------------------------------------------------------------------------------------------------------- . 2 6 3 3 3 3 4 P in ( 8 h 1 + 19 h 1 h 2 + 32 K s h 1 + 12 K s h 2 ) - ------------------------------------------------------------------------------------------------------------3 3 3 h1 h2 ( h1 + 4 Ks )

(3.33)

The objective function can be simplified

64 K s h 1 h 2 max ------------------------------------------------------------------------------------------------------------------------------- . 3 6 3 3 3 3 ( h 1 + 4 K s ) ( 8 h 1 + 19 h 1 h 2 + 32 K s h 1 + 12 K s h 2 )

(3.34)

This objective function displays a unique independence of the input force Pin. The derivation can be similarly used to show that the same design independence exists if an input displacement is used in place of the input force. Thus, the optimal design of h1 and h2 will strictly depend on the value of Ks and are independent of the magnitude of the input force or displacement. Writing the Lagrangian of the proposed cantilever optimization problem yields

64
3 3

64 K s h 1 h 2 L = ------------------------------------------------------------------------------------------------------------------------------- + 1 ( h 1 + 2 h 2 V max ) 3 6 3 3 3 3 ( h 1 + 4 K s ) ( 8 h 1 + 19 h 1 h 2 + 32 K s h 1 + 12 K s h 2 ) + 2 ( h1 ) + 3 ( h2 ) . (3.35)

The necessary KKT conditions are then expressed in equations 3.37-3.40:

192 K s h 1 h 2 ( 16 h 1 + 19 h 1 h 2 + 64 h 1 K s 48 K s h 2 ) -------------------------------------------------------------------------------------------------------------------------------------- + 1 2 3 6 3 3 3 3 2 ( ( h 1 + 4 K s ) ( 8 h 1 + 19 h 1 h 2 + 32 K s h 1 + 12 K s h 2 ) ) ---------------------------------------------------------------------------------------------------- + 2 1 3 6 3 3 3 3 2 ( 8 h 1 + 19 h 1 h 2 + 32 K s h 1 + 12 K s h 2 )


6 2 1536 K s h 1 h 2

2 3

6 3

2 3

= 0 =, 0

(3.36)

1 ( h 1 + 2 h 2 V max ) + 2 ( h 1 ) + 3 ( h ub ) = 0 ,

(3.37)

i 0 ,

(3.38)

h 1 + 2 h 2 V max h1 h2

0 0 . 0

(3.39)

Again, clearly because h1, h2, and Ks > 0

768 K s h 1 h 2 1 = ---------------------------------------------------------------------------------------------------- . 6 3 3 3 3 2 ( 8 h 1 + 19 h 1 h 2 + 32 K s h 1 + 12 K s h 2 ) L yields h1

6 2

(3.40)

substituting into

65 768 K s h 1 h 2 ( 4 h 1 + 4.75 h 1 h 2 + 16 K s h 1 12 K s h 2 ) -------------------------------------------------------------------------------------------------------------------------------------- 2 = 0 . 3 6 3 3 3 3 2 ( ( h 1 + 4 K s ) ( 8 h 1 + 19 h 1 h 2 + 32 K s h 1 + 12 K s h 2 ) )


6 2 9 6 3 6 2 3

(3.41)

This indicates that a potential KKT solution exists where 2 is active and h1 is equal to zero. By inspection, the objective function will be zero for all h1 equal to zero. Clearly, the objective function will be less than zero for h1 and h2 greater than zero. Because 1 is non-zero, a second KKT solution also exists when 2 is zero and the numerator of equation 3.42 is zero.

4 h 1 + 4.75 h 1 h 2 + 16 K s h 1 12 K s h 2 = 0

(3.42)

Provided that 3.42 is true, then h2 can be solved in terms of h1 from the active volume constraint:

V max h 1 - . h 2 = ---------------------2

(3.43)

Setting Vmax equal to one, this substitution leads to the following polynomial expression:

h 1 ( 45 h 1 + 142 h 1 35 h 1 252 h 1 + 285 h 1 114 h 1 + 19 ) + 512 K s h 1 ( h 1 2 h 1 + 1 ) 16 K s ( 61 h 1 + 110 h 1 19 h 1 + 60 h 1 + 45 h 1 18 h 1 + 3 ) = 0 .


2 6 5 4 3 2 1

(3.44)

Again, the roots of equation 3.45 cannot be solved in closed form. Nearly identical to the earlier formulation, the polynomial has roots at zero and one (which correspond to h1 = 0 and h1 = 1). The polynomial also has an additional root which falls in between zero and 0.32 and depends on the parameter of Ks. Examination of the KKT solutions leads to

66 the conclusion that the intermediate root is the optimum point. In order to gain further understanding of the objective criterion, the values of h1 and h2 can be plotted against the value of Ks. Shown in Figure 3.10 is a plot of the thicknesses, h1 and h2 versus varying values of Ks.
0

log(element height)

h2

h1

8 10

log(Ks)

more flexible compliant mechanism designs stiffer compliant mechanism designs Figure 3.10: Plot of h1 and h2 With Respect to the Spring Parameter

For small values of Ks, the thickness of h1 becomes close to zero, but remains positive as long as Ks > 0. For larger values of Ks, the stiffness of h1 increases and then remains constant. Similar to the force-displacement objective formulation, this configuration represents the maximum stiffness compliant mechanism design. This agrees with the earlier objective formulation as increasing the parameter Ks is equivalent to increasing Pex/in. Designs which press against a soft object will require greater flexibility (h1 smaller) than designs which press against a very stiff object. For all values of Ks, between the limits of zero and infinity, the values of h1 and h2 remain greater than zero. Again, the objective formulation is well-bounded given only a simple volume

67 constraint. Consequently, the optimization problem is also well-posed for topology, size, and shape synthesis of compliant mechanisms. 3.5 Characteristics of Optimal Designs The above exercise was restricted to optimizing the element size for a fixed geometry and topology. However, the formulation is relevant to perform size, shape and topology optimization. Depending on the boundary conditions, volume resource, and material modulus (and thickness); designs will exhibit similarities. Ignoring local constraints such as stress constraints, designs with the same ratio of the external load to the input displacement Fex/uin or similarly the same external spring Ks will yield the same design and the same efficiency. Designs with differing Fex/uin ratio or Ks stiffness will have a differing optimal element sizing, unless the thickness or modulus is adjusted accordingly. It can be noted from Figures 3.8 and 3.10 that as the ratio of Fex/uin or Ks increases, designs will achieve a compliant mechanism configuration of maximum stiffness. Beyond which, increasing the Fex/uin or Ks has negligible change on the design. Conversely for very small values of Fex/uin or Ks, designs will generally exhibit very slender sections as designs tend towards maximum flexibility. Further decreasing these parameters will exhibit a reduced effect on optimal designs. Relating optimal designs under each objective function is difficult as the optimal design for a given Fex/uin and the optimal design for a given Ks exhibit similar trends; however, the curves will be slightly different. Correlating the design relationship between the Fex/uin and the Ks parameters is a topic for future research.

68 3.6 Advantages of the Objective Formulation There are three primary advantages to formulating the objective function based on maximizing the energy efficiency: 1. For both the force-displacement efficiency formulation and the spring efficiency formulation, the solution to the simple cantilever problem is well bounded with only a simple volume constraint. Both optimization formulations avoid inappropriate compliant mechanism solutions (infinitely flexible, maximum stiffness, and infeasible solutions). This creates a wellposed and robust objective formulation for the optimization of topology, size and shape of compliant mechanisms. 2. The stiffness and flexibility trade-off is tied to one specific optimization parameter, Fex/uin or Ks which indirectly control the flexibility and stiffness of the design. These parameters can be correlated to dimensionless factors which can be used to correlate a wide range of designs for any material, thickness, and boundary condition parameters. Optimal designs under these conditions minimize energy storage within the structure and maximize the energy efficiency of the device. 3. The resulting optimized design can be characterized by an efficiency value. This essentially produces a practical tool to help discriminate between competing designs and help provide insight to improve existing designs. The efficiency value can also be used to determine where and under what conditions will a compliant mechanism be practical for specific applications.

CHAPTER IV NUMERICAL IMPLEMENTATION 4.1 Optimization Approach The goal of this research is to produce a robust and efficient method for automating compliant mechanism design. For structural optimization techniques to work appropriately, optimization must address: Formulation of an appropriate optimization problem Technique used to model the structural response Design parametrization Choice of optimization technique or algorithm The purpose of this chapter is to address numerical implementation of the last three criteria listed above. The numerical implementation should be developed in a generalized manner so that the techniques can be equally applied for topology synthesis and dimensional synthesis problems. 4.2 Modeling Approach and Design Parameterization For topological and dimensional synthesis optimization algorithms, a need exists to efficiently calculate the structural response which is then used to compute the objective function and the constraint values. Optimization of structures using a relatively small volume resource typically produces slender frame structures (Bendsoe et al., 1994). For compliant mechanisms this has an added advantage in that topology designs may more evenly distribute strains among members and are less likely to possess stress

70

71 concentrations. Consequently, the finite element method utilizing either rectangular frame elements or truss elements was chosen to approximate the response of the compliant mechanism structure. The advantage of this approach lies in its computational simplicity and robust modeling performance (i.e. elaborate re-meshing algorithms are not required to maintain reasonable modeling accuracy). These two properties are attractive for optimization because this lowers computational requirements and improves optimization speed. Both truss and frame elements were selected for topology optimization to compare element behavior on topology designs. For size and geometry optimization, only frame elements were utilized. This is due to the fact that the structure is often singular (deformation is governed by rigid body motion alone) when modeled using pin-jointed truss elements. Frame elements avoid this behavior because the structure is coupled at the joints and the inclusion of bending stiffness more accurately models the structural response of a fixed compliant mechanism topology. The global stiffness matrix for a structure is composed of many individual element stiffness matrices. Each individual truss or frame element can be written using a local stiffness matrix which contributes to the overall stiffness of the structure. Shown in Figure 4.1 is the standard dimensional form for two-dimensional truss and frame elements with local displacements q1 through q2 for the truss element; and q1 through q6 for the frame element.

72

(x2, y2) A

q4

q1

(x1, y1) (a) Truss Element


q5

(x2, y2)
q6

q4

b
q2 q1 q3

(x1, y1) (b) Frame Element Figure 4.1: Local Dimensions of the 2-D Truss and Frame Element

For the truss element, the local unrotated stiffness matrix is simply (Bathe, 1997)

EA 1 1 k e = -----. l 1 1

(4.1)

Where the local-global rotation matrix is expressed

L =

cos sin 0 0 . 0 0 cos sin

(4.2)

The rotated element stiffness matrix may be expressed simply

73 ke = L k e L .
T

(4.3)

Computing the global stiffness matrix is accomplished using the direct approach where each element contributes to the global model degrees-of-freedom. For the frame element using Hermitian shape functions, the local (unrotated) stiffness matrix can be expressed

h -l 0 0 k e = Eb h -l 0 0

0 h ---3 l
3

0 h -----2 2l h ---3l 0
3 3 3 3

h -l 0 0 h -l 0 0

0
3

0
3

h h ---- -----3 2 l 2l h h ------ ---2 2l 6l 0


3 3 3

h -----2 2l 0

(4.4)

0
3

h h ---- -----3 2 l 2l h -----2 2l


3

h h ---- -----3 2 l 2l h h ------ ---2 2l 3l


3 3

h ---6l

In order to calculate the global stiffness matrix, the local stiffness matrix coordinates must be transformed to the global (cartesian) stiffness coordinates. This is accomplished in the same manner as equation 4.3; however, the rotation matrix L is slightly different due to the increased number of degrees of freedom,

74

cos sin sin cos 0 0 L = 0 0 0 0 0 0

0 0 1 0 0 0

0 0 0 0 0 0 cos sin sin cos 0 0

0 0 0 0 0 1

(4.5)

Two-dimensional frame and truss structures, may be described using element thickness, element modulus, and nodal coordinates. Because the described structures are assumed to be of constant width (out-of-plane thickness) and composed of a simple homogeneous material, the modulus and width are kept constant. This leaves the element thickness and node locations as design variables. To avoid further complication of the optimization methodology, interpolation schemes such as Bezier curves were avoided due to the increased complexity when calculating the design sensitivities. The design variables describing the structure are directly parameterized with respect to each element thicknesses and selected nodal coordinates. For stability and simplicity reasons, each segment of the compliant mechanism topology may be defined having N number of elements with only two principle end nodes. Where N = 1 during topology synthesis problems and N > 1 for size and geometry problems. During optimization, each individual height of an element can change, while only activated principle nodes of a topology section are allowed to adjust position. The arrays of elements which model a given segment are kept equidistant and collinear which aids numerical stability during the optimization procedure. Figure 4.2 shows an element array composed of N number of elements.

75

hN h3 h2 h1

(X2, Y2) (xN+1, yN+1)

(xN, yN)

(x3, y4) (x2, y2)

(X1, Y1) (x1, y1)

Figure 4.2: An Element Array having N Elements and Two Principle Nodes

The generalized compliant mechanism optimization problem may be formalized max Efficiency ( h, X ) subject to: u out 0 1 h 1 = ----------- ----------- 1 = 0 u in GA ul
N

hi li b
=1 g1 = i -------------------10 V max

(4.6)

i g i + 1 = ----------- 1 0 , i = 1...N max h min h i h max X min X i X max

Where h represents each element thickness of the structure, and X represents geometric coordinates of the principle nodes. The equality constraint expressed by h1 is an optional constraint and can be imposed to control the mechanical or geometric advantages of the mechanism. The variable GAul represent the desired unloaded geometric advantage; however, other kinematic properties such as the final mechanical advantage, the loaded geometric advantage have also been successfully applied to compliant mechanism

76 optimization problems. If this constraint is not imposed, the optimization algorithm will find the kinematic relationship that maximizes the energy throughput. The total volume constraint, expressed by g1 is applied to maintain a limited total volume, Vmax, (and thus weight) of the mechanism. The stress constraints represented by the inequality constraints gi+1 through gN+1 can only be applied to the size and geometry problems to limit the maximum stress, max, in the structure. As with the kinematic constraint, these constraint are also optional and serve to help create more practical compliant mechanism designs. The remaining local variable constraints on the element thickness and geometric wandering range are generally given to reduce the searchable design space of the optimization algorithm (MATLAB, 1997). However, these local limits can be given specific ranges to improve the manufacturability of resulting designs. 4.3 Gradient-Based Optimization Algorithms A generic constrained optimization problem can be written in the form:

min F(x) subject to: h(x) = 0 g(x) < 0

(4.7)

Many simple types of constrained optimization problems incorporating a small number of design variables can be solved in closed-form using optimality conditions. For complex, large scale problems, computer-based numerical techniques are attractive. For large scale problem, optimality criteria methods, which use heuristics to solve first order optimality conditions, have been derived for certain classes of structural optimization problems. However, these approaches are generally inflexible in that the update scheme is problemdependent. Altering the objective or constraints requires reanalysis of the optimality conditions and re-derivation of the update scheme. An alternative approach to the

77 optimality criteria methods are mathematical programming approaches. These approaches seek to directly solve the optimization problem using gradient-based search algorithms and can easily be implemented for nearly any type of optimization problem. Gradient-based mathematical programming algorithms typically determine the direction of ascent (decent) from an initial starting point and are one of the most efficient techniques for solving problems formulated using continuous design variables (i.e. F, h, and g are everywhere continuous and differentiable) (Haftka, Gurdal, 1992). Optimization problems can be classified according to the relationship of the design variables with respect to the objective function and constraints. The solution to large-scale linear problems, known as linear programming (LP), is well understood and can be easily solved using algorithms such as the simplex method. Large-scale problems with quadratic or higher order relationship with respect to the objective and constraints are defined as nonlinear optimization problems. The solution to this class of problems is non-trivial and generally requires a non-linear programming algorithm. There are a wealth of non-linear programming algorithms available, although an all-powerful optimization algorithm that works well for all types of problems does not exist. Many of the algorithms break the problem into a series of sequential moves. Given a starting point, the algorithm may sequentially move towards a solution using a simple update scheme, xk + 1 = xk + k dk . (4.8)

Where xk is the starting point, dk is the step direction, k is the step length, and xk+1 is the updated design variable. While there are many types of sequential optimization algorithms available, the overall flow structure of these algorithms is fairly similar. A flowchart of a sequential optimization algorithm is shown in Figure 4.3. Among the best algorithms, the sequential quadratic programming (SQP) algorithm generally outperforms most other algorithms in terms of speed and accuracy

78 (Schitowski et al., 1994). A disadvantage of the SQP algorithm is the need for a large amount of memory to store an approximation of the Hessian (second order partial derivative of the Lagrangian) such that the algorithm only works well for problems using a medium to small number of design variables (usually less than 1,000 design variables). Typically, SQP is better suited to handle smaller design problems with many constraints. Slightly behind the SQP algorithm in convergence speed and accuracy is the sequential linear programming (SLP) algorithm. This algorithm is simpler than SQP and consequently posses far less memory requirements on the computer. Consequently, SLP is quite capable of handling problems in excess of 10,000 design variables, although SLP is not as attractive for problems which must satisfy many nonlinear constraints. For topology optimization problems, both the SQP and SLP algorithms were incorporated due to the large number of design variables. For the size and geometry problems, only the SQP algorithm was utilized due to the increased performance when handling smaller problems with a large number of nonlinear constraints (e.g. stress constraints).

79

Begin at xk

Calculate F(xk), h(xk), g(xk)

yes Reach termination criteria? end

no Calculate F(xk), h(xk), g(xk)

Determine k and dk Determine xk+1

xk = xk+1

Figure 4.3: Flow Chart of a Sequential Gradient-Based Optimization Algorithm

The converged solution of a non-linear programming algorithm typically represents a local optimum point, although in some cases the algorithm may converge to a saddle point (Papalambros, Wilde, 1988). In order to state that the point found is a global optimum point requires proving that the objective and constraints are globally convex. In general, the function and constraints posed for the compliant mechanism problem are not globally convex except for very simple problems such as the two-thickness compliant lever. Thus, gradient-based optimization algorithms can effectively find local optimum points; however, proving that the design is globally optimum design is much more difficult and may require a search of the entire design space.

80 4.3.1 Sequential Linear Programming Sequential linear programming algorithms linearize the objective and constraints using a Taylor series expansion about the current iteration point, xk. The optimization problem in equation 4.7 is expanded to
n

F ( x ) F ( xk ) +

i=1

( xi xki ) xi
g

,
xk

(4.9)

g ( x ) g ( xk ) +

i=1

( xi xki ) xi
xk

(4.10)

The LP problem can then assume the form min cTxk subject to: . Ax < b xlow < xk < xup

(4.11)

Where the LP algorithm is solved using the Simplex method (MATLAB, 1997). The linearization is usually appropriate only in the neighborhood of xk; consequently, movelimits, xlow and xup are enforced to keep the linearized approximation appropriate for the nonlinear optimization problem. Determining a good move-limit strategy is critical for obtaining an acceptable rate of convergence. Move limits which are too small require many iterations and take excessive computational time. Large move limits may oscillate about the optimal point, x* and never converge (Haftka, Grudal, 1993). The employed move limit strategy is described by Sigmund (1997) where each design variable is initially given a 30% move limit (from the initial value). As the algorithm cycles, the move limit for each design variable is increased by a factor of 1.3 if xi has the same step direction for

two consecutive iterations and is decreased by a factor of 0.6 if xi has alternating step directions. This move limit strategy continues until the change in the objective becomes small. Generally 10-6 was found to be an appropriate termination criteria for most problems. 4.3.2 Sequential Quadratic Programming The SQP subproblem is solved in an iterative manner similar to the SLP problem. The objective function is approximated using a quadratic function and the constraints are linearized. Hk refers to the kth approximation of the Hessian which is calculated using the BFGS formulation (MATLAB, 1997). The QP subproblem, shown in equation 4.12, is used to solve for the step direction dk. Where 1...me refers to the number of active constraints and me+1...m refers to inactive constraints. A line search is used to then solve for the step length, k which is based on minimizing a merit function. Each iteration for x continues until the difference between xk+1 and xk is less than 10-6 and all constraints are not violated within a tolerance of 10-6.

1 T min -- d H K d + f ( x k ) d n 2 d (4.12) g i ( x k ) d + g i ( x k ) = 0 g i ( x k ) d + g i ( x k ) 0
T T

i = 1, m e i = m e + 1, m

4.4 Analytic Calculation of Design Sensitivities In order to streamline the optimization process, gradient information for both the objective and constraints must be provided to the optimization algorithm. While finite difference methods can be used to approximate the gradients of the objective function and constraints, these method are computationally costly and often suffer from numerical

81

82 difficulties (Haftka, Gurdal, 1992). Since a linear finite element method is used to compute the structural response, several techniques are available to analytically derive the function and constraint gradients. The most straight forward method is known as the dummy load method or the adjoint variable method (Haug et al., 1986). The goal is to describe the derivative of a displacement or a force at a point on the body in terms of the global stiffness matrix. This allows the global stiffness matrix to be derivative with respect to local design variables producing a portion of the objective function and the constraint gradients. Given the stiffness matrix, K, which is a function of the design variables, x, the load vector, F1, and the displacement field, U1; the goal may be to compute the partial derivative of a local displacement, u, with respect to each design variables, xi. The first step is to solve for the deflection field, U1, using standard finite element equilibrium equations shown in equation 4.13. The adjoint method is performed by applying a unit load at the point and direction of interest (represented by Fd), and then solving for a separate deflection field, Ud shown in equation 4.14. The local displacement at a point, u, is expressed in terms of the global stiffness matrix using equation 4.15. The derivative of the displacement with respect to local design variables can then be posed in equation 4.16. Note that the adjoint method works for single unit loads which results in the displacement of the point of interest along the direction of the dummy load. The method also works to calculate the relative displacement between two points of interest where the dummy load vector may consist of equal and opposite unit loads applied at separate locations on the body. This method was used to compute all function and constraint gradients for the topology synthesis problems and the size and geometry optimization problems. K ( x ) U1 = F1 ( x ) (4.13)

K ( x ) Ud = Fd

(4.14)

83 u = Ud K ( x ) U1
T

(4.15)

u T T = [ U F ( x ) Ud K ( x ) U1 ] xi xi d 1

(4.16)

4.4.1 Differentiating with respect to the Element Stiffness Matrix By differentiating the global stiffness matrix with respect to each design variable (for both element sizing and node position), a portion of the analytic sensitivities can be determined. As can be noted from equations 4.1 and 4.4, the frame element combines the axial behavior of the truss element with the bending behavior of the beam element. For this reason, the derivation of the truss element stiffness matrix with respect to the design variables is a subsection of the frame element. In order to simplify the derivative description, methods will be presented using the frame element only, noting that they may be applied in the same manner for the truss element. The derivative of the frame element with respect the individual height variables can be expressed:

1 -l 0 0 1 -l 0 0

0
2

0
2

1 -l 0 0 1 -l
2

0
2

0
2

3h 3h --------------3 2 l 2l 3h -------2 2l 0
2 2

3h 3h --------------3 2 l 2l 3h h ----------2 2l 2l 0
2 2 2

k e --------- = Eb h

h ---l 0

(4.17)

0
2

3h 3h ------- -------3 2 l 2l 3h -------2 2l


2

0 0

3h 3h ------- -------3 2 l 2l 3h h ----------2 l 2l


2 2

h ---2l

84 Note, the rotation matrix does not contain height variables, so the rotation matrix does not affect the derivative of the original stiffness matrix. Thus, the derivative of the global stiffness matrix can be assembled by computing the derivative of the local stiffness matrix, multiplying by the rotation matrix, and then assembling the individual stiffness terms into the global stiffness matrix. The sensitivity analysis with respect to geometric coordinates is slightly more complex because of the alteration of geometric coordinates which effects both the unrotated elemental stiffness matrix and the global rotation matrix. Both the length of an individual element and the angular orientation of an individual element can be expressed in terms of the end positions of the two principle nodes. Thus the length of an individual element can be written

( X2 X1 ) + ( Y2 Y1 ) l = ------------------------------------------------------------ . N

(4.18)

The angular rotation of every element belonging to an element array can then be transformed in terms of the principle node coordinates X2 X1 cos = ------------------------------------------------------------ , 2 2 ( X2 X1 ) + ( Y2 Y1 )

(4.19)

Y2 Y1 sin = ------------------------------------------------------------ . 2 2 ( X2 X1 ) + ( Y2 Y1 )

(4.20)

These expressions are then substituted into the original unrotated stiffness matrix, ke, and the rotation matrix, L. The derivative for the rotated individual stiffness matrix is taken by calculating ke and then differentiating the expression with respect to X1, X2, Y1 and Y2 shown in equations 4.21 and 4.22. Note that the derivatives with respect to X1 and Y1 are

85 simply the negative of the derivatives with respect to X2 and Y2. These derivatives are detailed in Appendix A. ke T -------- = -------- ( L k e L ) X1 X1 ke T -------- = -------- ( L k e L ) X2 X1

(4.21)

ke T -------- = -------- ( L k e L ) , Y1 Y1

ke T -------- = -------- ( L k e L ) Y2 Y1

(4.22)

4.4.2 Sensitivity Analysis of the Efficiency Formulations In order to calculate the gradients of the objective function and constraints, these equations must be expressed in terms of the global stiffness. Returning to the two energy efficiency problems shown in Figure 4.4, the goal is to design the compliant mechanism which maximizes the energy efficiency; either lifting a static external force, or pressing against a soft object.

load uout2 actuator

Fex = mg workpiece

Ks

uout

uin

actuator

uin

Fin2

Fin

Figure 4.4: The Two Energy Efficiency Problems for Compliant Mechanism Design

The Force-Displacement Efficiency Formulation The structural responses required to calculate the objective function and constraints for the force-displacement efficiency formulation are shown in Figure 4.5.

86 Note that U1 represents the structure under prescribed input displacement, uin. U2 is the displacement field of the structure with a unit load applied in the desired direction of output motion and U3 is the displacement field of the structure with a unit load applied in the desired direction of motion. For purposes of simplicity, all structural responses were calculated by using the penalty approach to prescribe the input displacement (which was either zero or uin). This method involves adding a large stiffness term, Cp, to the appropriate degree of freedom of the stiffness matrix. The force array is then updated by the penalty term multiplied by the prescribed displacement. ( K + C p ) U = F + ( C p ) u in

(4.23)

As C p then u in u in . Where uin* represents the desired displacement boundary condition and u in represents the actual finite element response. The appropriate choice for Cp is made by selecting a penalty value large enough such that uin is sufficiently converged. However, Cp must not be too large such that the calculation of the inverse of the stiffness matrix encounters numerical difficulty. Some general guidelines for the value are 104 to 106 higher than the highest element stiffness (Chandrupatla, Belegundu, 1991). The reaction force required to cause the displacement can also be calculated F R = C p ( u in u in ) .

(4.24)

87

uin

Cp

Cp

Cp

KU1 = F1

KU2 = F2

KU3 = F3

Figure 4.5: Performance Measurement for the Force-Displacement Formulation

Note, for some function calculations such as the stress constraint it may be necessary to compute the combined response under the both the input displacement and the external force. The combined displacement field may be calculated by superposing the first two load cases U 4 = U 1 F ex U 2 . (4.25)

Using the unit load method, the elements of the efficiency formulation may be expressed in terms of the global stiffness matrix
T

u out 0 = U 2 KU 1 ,

(4.26)

T u in F in = u in C P ( U 3 KU 1 u in ) ,

(4.27)

u out 1 F ex = F ex ( U 3 KU 3 ) .

(4.28)

The efficiency formulation can be written

88 ( U 2 KU 1 ) = ------------------------------------------------------------------------------------------- . T T u in C P ( U 3 KU 1 u in ) + F ex ( U 3 KU 3 )
T

F FD

(4.29)

The partial derivative of the formulation can then be expressed in equation 4.30. Where xi is a generic design variable representing either height or nodal position variables.
T T T K ( u in C P ( U 3 KU 1 u in ) + F ex ( U 3 KU 3 ) ) U 2 U 1 xi F FD ------------ = -------------------------------------------------------------------------------------------------------------------------------+ 2 T T xi [ u in C P ( U 3 KU 1 u in ) + F ex ( U 3 KU 3 ) ] T T K T K ( U 2 KU 1 ) u in C P U 3 U 1 + F ex U 3 U 3 xi xi ----------------------------------------------------------------------------------------------------------------------2 T T [ u in C P ( U 3 KU 1 u in ) + F ex ( U 3 KU 3 ) ]

(4.30)

The force-displacement efficiency formulation may also be calculated by avoiding the penalty approach and by using principles of linear superposition. Avoiding the penalty approach aids the numerical stability of the optimization, preventing the stiffness matrix from becoming ill-conditioned. This has been primarily implemented for topology synthesis problems, since the derivatives for the some of the constraints (stress) become increasingly cumbersome to work with. The Spring Efficiency Formulation Figure 4.6 illustrates the loading conditions applied to the spring efficiency problem. Where U5 and U6 are the displacement fields resulting from applying a unit load to the input and output respectively.

89

Ks

Ks 1

KU5 = F5

KU6 = F6

Figure 4.6: Performance Measurement for the Spring Formulation

Note, for some constraint calculations such as the stress constraint it may be necessary to compute the response under the exact input force. This may be achieved by appropriately scaling the displacement field by the magnitude of the input force. U 7 = F in U 5 (4.31)

The quantities uout and uin can be expressed in the unit load form:
T

u out = U 6 KU 5 ,

(4.32)

u in = U 5 KU 5 .

(4.33)

The spring efficiency formulation can be expressed as a function of the global stiffness matrix, where the square/square root function is employed to prevent ambiguity in sign.
1 -2 2 T T ( ( U 6 KU 5 ) ) ( U 6 KU 5 ) -------------------------------------------------------T ( U 5 KU 5 )

FS =

(4.34)

90 The partial derivative of the formulation can then be expressed:


1 - 2 2 T T T K ( U 5 KU 5 ) 2 ( ( U 6 KU 5 ) ) U 6 U 5 xi

F S = ------------------------------------------------------------------------------------------------ 2 T xi ( U 5 KU 5 ) .

1 -2 2 T T T K ( ( U 6 KU 5 ) ) ( U 6 KU 5 ) U 5 U 5 xi -----------------------------------------------------------------------------------2 T ( U 5 KU 5 )

(4.35)

4.4.3 Sensitivity Analysis of the Volume Constraint The sensitivity analysis with respect to the volume constraint must also be separated for both size and shape variables. The volume constraint is written in the normalized form
N

g1 =

i=1

hi li b ----------- 1 . V max

(4.36)

The derivative with respect to the height variable can be expressed simply. li b g1 -------- = ---------- hi V max

(4.37)

The derivative with respect to geometric position variables, Xi, can be expressed, hi b - ( l ) ------ Xi i ----------------------------- . V max

g1 -------- = Xi

i=1

(4.38)

Where the generic node coordinate variable represents X1, X2, Y1, or Y2.

91 4.4.4 Sensitivity Analysis of the Mechanical Advantage Constraint The constraint for the initial mechanical advantage is posed,
T

U 6 KU 5 h 1 = -----------------------1 . u in ( MA )

(4.39)

The derivative is then expressed:


T K U6 U5 h1 xi -------- = ----------------------- . xi u in ( MA )

(4.40)

Note that equation for the final mechanical advantage, the loaded geometric advantage and the spring-loaded mechanical and geometric advantage may be similarly expressed using the described techniques. 4.4.5 Sensitivity Analysis of the Mean Positive Stress Constraint Because the shape function of the frame element is cubic, the bending stress distribution is linear along the length of the element. Superposing the bending stress with the axial stress yields: ( q 4 q 1 ) Eh ( 12 z 6 l ) ( 12 z 6 l ) (6z 4l) ( 6z 2l ) - q 2 ---------------------- ( z ) = E --------------------- ------ q 5 ----------------------- + q 3 -------------------- + q 6 -------------------3 3 2 2 l 2 l l l l . (4.41)

Where z is the length along the frame element (from 0 to l), qi represents the local displacements for the body with both the external load and the input displacement simultaneously applied. Solving for the stress along each frame element, requires the calculation of the local displacements at each element. To derive the sensitivity of the

92 stress constraint, dummy loads for each displacement pair (q1 and q4, q2 and q5, and q3 and q6) must be separately calculated using finite element equilibrium equations and applied to derive the gradient of the stress constraint. In order to lower the total number of constraints for the problem, only the stress at the midpoint was calculated. Although this constrains the average stress in each element, this reduces the number of displacement pairs required for each stress calculation and reduces the number of dummy load cases applied to the finite element model. The equation for the stress at the midpoint along the frame element simplifies to

l E Eh -= -( q q ) ------ ( q q ) . 4 1 2 l 2l 6 3

(4.42)

In addition, because the sign of the stresses can flip positive or negative depending on the local strains in the element, the absolute value of the midpoint stress is taken as the design constraint. This complicates the stress calculation slightly, but reduces the total number of constraints. Thus, the constraint limiting the mean positive stress is expressed
1 -1 --

E 2 2 Eh i 2 2 -- ( ( u si ) ) + ------- ( ( u ri ) ) li 2 li -1 . g i = -------------------------------------------------------------------- max

(4.43)

Noting the following substitution u si = q 4 q 1 , (4.44)

u ri = q 6 q 3 .

(4.45)

The quantities uri and usi may be obtained by applying the following simultaneous dummy loads for each element shown in Figure 4.7.

93

cos sin 0 cos sin 0 0 0 1

0 0 1

node n+1

node n+1 node n


Dri = K-1Sri

node n
KUsi = Fsi

Figure 4.7: Dummy Loads Used for usi and uri

The quantities usi and uri can be expressed using the global stiffness matrix where Usi and Uri represent the individual unit forces applied to each element and U4 (U7) represents the displacement field under the exact boundary conditions. The variable, N, represents the total number of elements that comprise the finite element model. u si = U si KU 4 for i = 1 ... N
T

(4.46)

u ri = U ri KU 4 for i = 1 ... N

(4.47)

The derivative for the stress constraint must be handled differently for size variables and shape variables. The derivatives of usi and uri with respect to the thickness variables may be written T K ( u si ) = U si U 4 , hi hi

(4.48)

T K ( u ri ) = U ri U 4 . hi hi

(4.49)

94 The partial derivative of the mean positive stress constraint with respect to element thickness can be expressed
when i=j 1 - 2 2 E -- (u ) ( usi ) ( u si ) si li hj 1 1 - -Eh 2 2 2 2 E i ------ ( u ) ( ur i ) ( u ri ) -------- ( u r i ) ri 2 li 2 li hj

i = h j

. (4.50) -------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- max Because the dummy load vector (used to compute usi) must rotate along with the

local node coordinate, the derivative with respect to the geometric coordinates must be modified by differentiating the dummy load with respect to the geometric position variables. The derivatives of usi and uri with respect to the node position variables may be written
T

F si T K ( u si ) = U , U 4 U si Xi Xi 4 Xi

(4.51)

T K ( u ri ) = U ri U . Xi Xi 4

(4.52)

The partial derivative of the mean positive stress constraint with respect to geometric variables can then be expressed:

95
1 - 2 2 E -- (u ) ( usi ) ( u si ) s i li Xi 1 -Eh 2 2 i -------- ( u ) ( ur i ) ( u ri ) r i 2 li Xi

i = Xj

---------------------------------------------------------------------------------------------------------------------------------------------------------- max
when i = j 1 - 1 2 2 E ------- -- (u ) X l si 1 --

Eh i 1 2 2 ( uri ) + ------- ------- -- 2 X j l i j i + --------------------------------------------------------------------------------------------------------------------- max

(4.53)

CHAPTER V TOPOLOGY SYNTHESIS The goal of topology synthesis for compliant mechanisms is to identify the optimal number and connectivity of structural elements to achieve specified motion requirements. Topology synthesis is perhaps the most critical stage of the design process, due to the fact that the largest gains in performance can be made. A variety of methods have been employed to perform topology synthesis for minimum compliance problems (the stiffest structural design). More recently, these approaches have been adopted for compliant mechanism topology synthesis. The extension and improvement of this work mainly lies in developing and implementing an improved formulation which will provide robust convergence for both topology optimization and size and geometry optimization design problems. 5.1 Topology Synthesis Procedure To perform topology synthesis the designer must first establish a design problem having a fixed domain with a desired input motion and an external force (or spring) which opposes the desired output motion. The design domain is then discretized using truss or frame elements. The designer chooses a material modulus, thickness, and a total volume constraint. Optimization is performed using sequential linear programming or sequential quadratic programming methods. During optimization, element cross-sections are resized within upper and lower bounds. In addition, node locations can be varied within wandering ranges. To prevent the global stiffness matrix from becoming singular, the thickness lower bound is non-zero and elements which converge to the lower bound are not removed. Upon convergence of the optimization algorithm, lower bound elements are 96

97 assumed to not affect the final solution and are removed from the final topology design. Figure 5.1 illustrates the topology design procedure for a simple compliant mechanism.

uin uout Fex (a) Design Problem lower bound elements (b) Discretized Design Domain

(c) Optimized Design

(d) Interpreted Topology

Figure 5.1: Illustration of the Topology Synthesis Procedure

5.2 Convergence of Topology Designs Stable and robust design convergence is the most important factor when optimizing a system. Stated simply: given the compliant mechanism design problem and the initial starting point, the optimization algorithm should produce the best possible solution given the feasible domain of design variations. In addition, this solution should possess minimal sensitivity to criteria such as the element lower bound or the choice of optimization algorithm. Figure 5.2 illustrates a compliant mechanism design problem which will be used to study solution convergence.

98

Fin uin

uout Fex Figure 5.2: Compliant Mechanism Design Problem

uout Ks

5.2.1 Design Discretization Practical application of topology optimization prevents discretization of the design domain using an infinite number of design variables. Two main approaches are used to generate an initial starting structure: a fixed-node full ground structure and a floating-node modular ground structure. The fixed-node ground structure, sub-divides the design domain using a regular matrix of nodes. Every node is connected to every other node with an element, excluding elements that lie directly on top of one another to prevent ambiguity. The floating-node modular structure also sub-divides the design domain using a discrete matrix of nodes; however, not every node is connected to every other node (although there is nothing preventing this). A subset of four nodes are fully connected which is repeated throughout the design domain. The nodes are then given wandering ranges to allow geometric variation while preventing elements from collapsing to zero lengths. Both methods have advantages and disadvantages. The fixed-node method is easier to implement numerically since the derivatives are only dependent on element sizing; however, this method allows limited geometric freedom. In addition, as the number of nodes is increased, the number of design variables grows exponentially. In

99 contrast, the floating-node approach allows for greater geometric freedom and the number of design variables grows linearly with the number of nodes. Figure 5.3 shows two different design domain discretization using the full ground structure approach and the modular ground structure approach.
corresponding wandering range

activated node

(a) Full Ground Structure 28 elements 28 design variables

(b) Modular Ground Structure 20 elements 38 design variables

Figure 5.3: Two Approaches for Design Domain Discretization

Fixed-Node Verses Floating-Node Ground Structures The solution for the fixed-node ground structure was compared to the solution for the floating-node ground structure. The fixed-node ground structure is shown in Figure 5.4 (a). The design domain was divided connecting 16 nodes using a full ground structure for a total of 89 frame elements/design variables. Figure 5.4 (b) shows the optimized topology. The floating-node ground structure is shown in Figure 5.4 (c). The design domain was also divided into 16 nodes corresponding to 66 design variables (42 frame elements and 24 activated degrees-of-freedom). Figure 5.4 (d) shows the initial structure and the optimized topology. Both optimal topologies have structural similarities and differences. It was generally noted that the floating node optimization procedure provided

100 slightly better energy efficiencies (40.98% for the fixed node design as opposed to 42.47% for the floating node design). Generally the floating node approach allows for slightly higher design flexibility as designed are not constrained to specific geometries.
Initial Guess 10 10

= 40.98%

uin

Fex
0 1 2 3 4 5 6 7 8 9 10

10

(a) Fixed-Node Ground Structure


10 10 9 9

(b) Optimized Topology (Fixed-Node) = 42.47%

uin

Fex
0 1 2 3 4 5 6 7 8 9 10

10

(c) Floating-Node Ground Structure

(d) Optimized Topology (Floating-Node)

Figure 5.4: Fixed-Node Approach Verses Floating-Node Approach

Discretization Resolution The choice of design discretization is one aspect of choosing an initial design to proceed to the topology optimization stage. While a finer discretization offers more design variations and a larger chance of converging to a better solution, more elements slows the calculation of the structural response and slows the optimization algorithm. In addition, more design variables may increase the number of local minima. Consequently, the discretization of the design domain is an important aspect which should not be ignored

101 during topology optimization. The solution dependency of the fixed-node and the floating-node topology designs were examined by increasing the nodal discretization of the design domain. The fixed-node ground structure is shown in Figure 5.5 (a). In this case, The design domain was divided into 25 nodes (16 nodes in the previous example) connecting all nodes for a total of 200 elements/design variables. Figure 5.5 (b) shows the optimized topology which shows similarities to the 89 element design in Figure 5.4 (b).
10 10

= 43.17%

uin

Fex
0 1 2 3 4 5 6 7 8 9 10

10

(a) Fixed-Node Ground Structure


10 10 9 9

(b) Optimized Topology (Fixed-Node) = 44.20%

uin

Fex
0 1 2 3 4 5 6 7 8 9 10

10

(c) Floating-Node Ground Structure

(d) Optimized Topology (Floating Node)

Figure 5.5: Optimized Topologies Resulting From Increased Node Discretization

For the fixed node case, increasing the element density resulted in a small (3.8%) increase in energy efficiency consequently the penalty for increasing the number of design variables was not rewarded with a significantly superior design. The floating node ground structure is shown in Figure 5.5 (c). The design domain was also divided into 25 nodes corresponding to 114 design variables (72 elements and 42 activated degrees-of-freedom).

102 Figure 5.5 (d) shows the optimized topology. The 25 node floating-node ground structure produced a nearly identical topology design compared to the 16 node ground structure confirming that the original discretization was adequate for the topology design problem. 5.2.2 Comparison of Objective Formulations Figure 5.6 illustrates optimal topology designs for various Ks parameters. It was found that the optimized topology can vary somewhat depending on the efficiency parameters. Increasing Ks resulted in a lower energy efficiency as the stiffer output spring reduced the magnitude of the uout. Generally, increasing the stiffness of the external spring, Ks decreased the unloaded geometric advantage of topology designs (designs (a), (b), (c), and (d) had unloaded geometric advantages of 1.4, 1, 0.9 and 0.8 respectively). Intuitively this makes sense as the increased stiffness of the spring requires the mechanism to generate more output force and thus a higher mechanical advantage (inverse of the geometric advantage) to displace the spring. In order to compare the performance of topologies, specific designs where further resized under varying the spring stiffness parameter. Table 5.1 displays the measured efficiencies of specific topologies. Results of the study re-affirm that the best topology design is somewhat affected by the stiffness parameter. However, most of the topology designs (mainly (b), (c) and (d)) obtained fairly close efficiency values over a wide range of spring parameters. Topology (a) was noted to have generally inferior performance to that of the other topologies at moderate to high values of the Ks parameter. However, under the original optimized conditions, this topology has similar performance to the other designs.

103

10

10

Fin

= 98.25%

Fin

= 77.32%

Ks
0 1 2 3 4 5 6 7 8 9 10

Ks
0 1 2 3 4 5 6 7 8 9 10

(a) Ks = 0.01
10 10 9

(b) Ks = 1 = 27.89%
9

Fin

Fin

= 3.79%

Ks
0 1 2 3 4 5 6 7 8 9 10

Ks
0 1 2 3 4 5 6 7 8 9 10

(c) Ks = 10

(d) Ks = 100

Figure 5.6: Optimized Topologies Verses Spring Parameter

Table 5.1: Efficiencies of Various Topologies Verses Spring Parameter

Spring Parameter Ks 0.001 0.01 0.1 1 10 100 1000

Topology a 99.45% 98.25% 92.47% 58.76% 12.54% 1.41% 0.14%

Topology b 99.64% 99.04% 95.82% 77.32% 25.75% 3.35 0.35%

Topology c 99.52% 98.55% 94.97% 77.59% 27.89% 3.76% 0.39%

Topology d 98.10% 98.25% 95.49% 78.18% 28.12% 3.79% 0.39%

104 Figure 5.7 illustrates optimal topology designs for various Fex/uin parameters. It was found that the optimized topology can also vary slightly depending on the forcedisplacement efficiency parameter. Increasing Fex/uin resulted in a lower energy efficiency as increasing the ratio of the external force to the input displacement causes the structure to store more strain energy relative to the available reciprocal work. Table 5.2 indicates specific topologies which were re-sized under varying ranges of forcedisplacement parameters. Results of the study confirm that the best topology design is somewhat affected by the efficiency parameter. However, the unloaded geometric advantages of both topologies were nearly identical at 1. Both topology designs obtained fairly close efficiency through a wide range of Fex/uin parameters. The variation in performance can be attributed to the shift in stored compliance. At low ranges of Fex/uin the primary storage of energy in the structure is due to the applied strain. At higher ratios of Fex/uin the dominant energy storage is due to the external force which opposes the mechanism motion. Consequently, the shift in topology designs can be attributed to the shift in the total compliance stored in the mechanism.

105

10

= 99.73% uin

10

= 87.29% uin

Fex
0 1 2 3 4 5 6 7 8 9 10

Fex
0 1 2 3 4 5 6 7 8 9 10

(a) Fex/uin = 0.01


10 10 9

(b) Fex/uin = 1
9

= 42.24% uin

= 6.82% uin

Fex
0 1 2 3 4 5 6 7 8 9 10

Fex
0 1 2 3 4 5 6 7 8 9 10

(c) Fex/uin = 10

(d) Fex/uin = 100

Figure 5.7: Optimized Topologies Verses Force/Displacement Parameter

Table 5.2: Efficiencies of Various Topologies Verses Force/Displacement Parameter

Force/Disp. Parameter Fex / uin 0.001 0.01 0.1 1 10 100 1000

Topology a 99.82% 99.73% 97.99% 87.29% 40.98% 6.39% 0.62%

Topology c 99.82% 99.73% 97.36% 86.84% 42.24% 6.82% 0.64%

106 It should be pointed out that the convergence of topology designs remains stable, producing acceptable compliant mechanism topologies regardless of the efficiency parameters. The slight variation in topology performance poses an interesting problem as the best topology seems to be a function of the specific boundary conditions applied to the structure. However, other optimized examples have countered this behavior somewhat demonstrating more stable topology designs over a relatively broad range of efficiency parameters. Interestingly enough, the force-displacement formulation typically produced symmetrical structures, while the spring formulation yielded asymmetrical structures. This can be attributed to the spring formulation which contains a squared output displacement term, which tends to shift the geometric advantage of the mechanism as the stiffness of external spring is altered. In contrast, the force-displacement formulation which contains a linear output displacement term generally kept the mechanical advantage closer to one. For completeness, topology optimization behavior should be further studied to quantify the effects of varying the efficiency parameters on optimal topologies. 5.2.3 Effect of Element Lower Bound The effect of the element lower bound on topology convergence was also investigated. Shown in Figure 5.8 are two separate topologies resulting from differing element lower bounds. While increasing the lower bound high enough could be shown to alter topology designs slightly, stable convergence was achieved for cases where the lower bound was less than 100 times the thickness of the converged structure. Typically the upper bound element sizing constraint was set high enough such that the element sizing was limited mainly by the total volume constraint.

107

10

10

uin

uin

Fex
0 1 2 3 4 5 6 7 8 9 10

Fex
0 1 2 3 4 5 6 7 8 9 10

(a) Optimized Topology (hlb = 1E-3)

(b) Optimized Topology (hlb = 1E-5)

Figure 5.8: Optimization Convergence Verses Element Lower Bound

5.2.4 Effect of the Total Volume Constraint The main function of the total volume constraint is to force the optimization procedure to efficiently place material throughout the design domain. This results in simpler topology designs which are easier to manufacture and require less design refinement during the size and geometry optimization stage. Shown in Figure 5.9 are two topologies optimized using different total volume constraints. Leaving all other parameters constant, increasing the total volume had the tendency of increasing the available stiffness of material. Due to the correlation of designs under specific boundary conditions, this had the effect of shifting the topology to a design optimized under a slightly lower efficiency parameter.

108

10

10

uin

uin

Fex
0 1 2 3 4 5 6 7 8 9 10

Fex
0 1 2 3 4 5 6 7 8 9 10

(a) Optimized Topology (Vmax = 2)

(b) Optimized Topology (Vmax =4)

Figure 5.9: Optimization Convergence Verses Total Volume Constraint

5.2.5 Truss Elements Verses Frame Elements It was generally found that optimal topologies resulting from truss and frame elements were identical, although the efficiencies of the two structures were slightly different. The difference in efficiency can be attributed to the increased stiffness due to the bending response carried by the frame element. This behavior initially countered early reasoning that the bending stiffness and additional degrees of freedom possessed by the frame element would affect topology designs. The explanation can be attributed to the fact that topology design problems are composed of highly networked structures. For linear structural analysis, these types of structures can be analyzed by making the assumption that loads are primarily carried through tension and compression of the members. Consequently, for highly networked structures, frame elements qualitatively behave the same as truss elements (transmitting forces primarily through tension and compression); however, the measured energy efficiency of the frame structure will be less due to the slightly increased stiffness. The primary disadvantage of frame elements are their increased degrees of freedom over the computationally simpler truss element. This slows the structural response and design sensitivity calculations. The primary advantages of frame elements are that elements could be removed from the topology design without the structure

109 becoming singular (rigid body motion). In addition, when all elements which surround a given node are removed, degrees of freedom associated with that node must also be eliminated from the global stiffness matrix. Incorporating this into the optimization routine would speed design convergence as the number of design variables and structural degrees of freedom would be reduced throughout topology design procedure. It is also possible that this approach could be used to integrate topology synthesis and dimensional synthesis by eliminating unwanted segments and then adaptively increasing the number of elements (N) along the remaining topology segments. Another advantage of frame elements is that rotational degrees of freedom can be used as input and output motions. Consequently, it is possible to use a rotational displacement (or torque) at the input to drive a rotational external torque (or torsional spring) at the output. These motions can also be combined with linear motions allowing more freedom for the designer. 5.3 Topology Synthesis Examples The following examples were generated to demonstrate the capabilities of the topology optimization procedure. Optimized designs reflect the robustness of the developed formulations as designs typically converged to feasible compliant mechanism topologies which satisfied motion requirements while obtaining optimal energy efficiencies. In all cases, limited experience was required to obtain satisfactory convergence as the results do not require methods which aid convergence, such as techniques which penalize intermediate element thicknesses. 5.3.1 Spring Efficiency Optimized Topologies Compliant Gripper Topology synthesis for the classic compliant gripper mechanism was investigated. Figure 5.10 illustrates the gripper design problem. Upon applying an input force, the

110 gripper jaws should close and exert force against an object represented by the external spring, Ks.

design domain
uout symmetry Fin Ks uout

Design Problem

Figure 5.10: Design Problem for the Compliant Gripper

Using symmetry conditions, half of the design domain was discretized. The topology was designed using the fixed-node ground structure as well as the floating-node ground structure technique. For both designs the modulus was set to 1000, the width was set to 1, the total volume constraint was set to 2, and the lower element bound was set at 1E-4. The fixed-node method was discretized using 18 nodes and 99 elements. Figure 5.11 shows the fixed-node ground structure and the optimized topologies for various spring parameters. The floating-node method was discretized using 18 nodes and 47 elements. A total of 16 nodes were activated and given wandering ranges resulting in a total of 74 design variables. Figure 5.12 shows the floating-node ground structure and the topology design resulting from various Ks values. For all cases, the floating node topology designs possessed slightly higher energy efficiencies.

111
10 10

= 82.40%

Fin
0 1 2 3 4 5 6 7

Ks

0 8 9 10

10

(a) Initial Structure


10 10 9

(b) Optimized Design (Ks = 1)


9

= 41.32%

= 6.60%

0 0 1 2 3 4 5 6 7 8 9 10

10

(c) Optimized Design (Ks = 10)

(d) Optimized Design (Ks = 100)

Figure 5.11: Optimized Compliant Gripper Topologies (Fixed-Node Approach)

112
10 10

= 85.76%

Fin
0 1 2 3 4 5 6 7

Ks

0 8 9 10

10

(a) Initial Structure


10 10 9

(b) Optimized Design (Ks = 1)


9

= 46.67%

= 8.40%

0 0 1 2 3 4 5 6 7 8 9 10

10

(c) Optimized Design (Ks = 10)

(d) Optimized Design (Ks = 100)

Figure 5.12: Optimized Compliant Gripper Topologies (Floating-Node Approach)

Compliant Wrench The compliant wrench design problem is shown in Figure 5.13. As the handles of the wrench are compressed downward, the jaws should close and exert force against an object, similar to the gripper mechanism.

Fin uout symmetry Ks uout Fin Figure 5.13: Design Problem for the Compliant Wrench

113 Using symmetry, the design domain was discretized using 25 nodes. For both designs the modulus was set to 1000, the width was set to 1, the total volume constraint was set to 3, and the lower element bound was set at 1E-4. The fixed-node method was discretized using 186 elements. Figure 5.14 shows the fixed-node ground structure and the topology optimized designs for various Ks parameters. The floating-node method was also discretized using 25 nodes and 69 elements. A total of 23 nodes were activated and given wandering ranges resulting in a total of 111 design variables. Figure 5.15 shows the floating-node ground structure and the topology design resulting from various Ks values. For all cases, the floating node topology designs possessed slightly higher energy efficiencies.

114
14 14

= 56.80%
12 12

Fin
10 10 8 8

Ks

10

12

14

10

12

14

(a) Initial Structure


14 14

(b) Optimized Design (Ks = 1) = 2.08%


12 10

= 17.94%
12 10

overlapping elements

10

12

14

10

12

14

(c) Optimized Design (Ks = 10)

(d) Optimized Design (Ks = 100)

Figure 5.14: Optimized Compliant Wrench Topologies (Fixed-Node Approach)

115
14 14

12

12

= 70.20%

Fin
10 8

10

Ks

0 0 2 4 6 8 10 12 14

10

12

14

(a) Initial Structure


14 14

(b) Optimized Design (Ks = 1) = 5.85%


12 10

= 38.37%
12 10

10

12

14

10

12

14

(c) Optimized Design (Ks = 10)

(d) Optimized Design (Ks = 100)

Figure 5.15: Optimized Compliant Wrench Topologies (Floating-Node Approach)

5.3.2 Force-Displacement Efficiency Optimized Topologies Piezo-Stack Amplifier The goal of the piezo-stack amplifier was to generate a low-profile mechanism to efficiently amplify the displacements of a piezo-stack actuator by a factor of 3 to 5 times. To accomplish this, a constraint on the unloaded geometric advantage was imposed using the penalty method which can be used to express a constraint as part of the objective function. For topology synthesis problems, the penalty method works well with the SLP algorithm to approximately enforce an equality constraint. The penalty factor can imposed in the objective function using the following technique

U 6 KU 5 max F FD ( h, X ) P ----------------------- 1 u in ( GA ul )

(1.54)

116 Where P is a constant large enough to force the constraint to become close to zero. Figure 5.16 illustrates the design domain for the piezo-stack amplifier problem. As the piezoelectric stack actuator expands horizontally, the mechanism produces a vertical displacement pushing outwards against an external load.

Fex uout uin

Fex uout piezo-stack actuator

uin uout Fex uout Fex

Figure 5.16: Design Problem for the Piezo-Stack Amplifier

Using one-quarter symmetry, the design domain is discretized using 24 nodes. Shown in Figure 5.17 (a), the fixed-node method was comprised of 165 elements. For both designs the modulus was set to 1000, the width was set to 1, the total volume constraint was set to 1, and the lower element bound was set at 1E-4. The forcedisplacement parameter was fixed at 4 for all optimized designs. Figure 5.17 (b) shows the fixed-node ground structure and the optimized topology without the geometric advantage constraint. The unloaded geometric advantage of the unconstrained topology design was 2.67. The effect of the constraint was investigated by designing amplifiers having geometric advantages close to 3 and 5. To achieve these geometric advantages, the penalty multiplier was set to 40. The actual unloaded geometric advantages were 2.88 for 5.17(c) and 4.34 for 5.17(d). It was noted that constraining the geometric advantage of designs resulted in lower energy efficiencies. The floating-node method was discretized using 24 nodes and 65 elements. A total of 22 nodes were activated and given wandering ranges resulting in 104 design variables. Figure 5.18 shows the floating-node ground

117 structure and the unconstrained optimized topology. The unloaded geometric advantage of the maximum energy efficiency design was 2.66. In addition, the effect of the constraint was also investigated by designing grippers having geometric advantages close to3 and 5. The actual mechanical advantages were 2.99 for 5.18(c) and 4.97 for 5.18(d). To achieve these geometric advantages, the penalty multiplier was set to 100. For all cases, the floating node topology designs possessed slightly higher energy efficiencies.

118
9 9

Fex

= 62.94%

uin

(a) Initial Structure


9 9 8

(b) Optimized Design


8

= 60.86%

= 49.08%

(c) Optimized Design (GA ~ 3)

(d) Optimized Design (GA ~ 5)

Figure 5.17: Optimized Piezo-Stack Amplifier Topologies (Fixed-Node Approach)

119
9 9

= 68.03%

Fex

uin

(a) Initial Structure


9 9 8

(b) Optimized Design = 57.54%

= 65.44%

(c) Optimized Design (GA ~ 3)

(d) Optimized Design (GA ~ 5)

Figure 5.18: Optimized Piezo-Stack Amplifier Topologies (Floating-Node Approach)

MEMS Displacement Magnifier The purpose of the MEMS displacement magnifier was to generate rectilinear motion, while achieving a large geometric advantage. The design was to be used with an electrostatic comb drive to propel a micro-mechanical locking system designed by Sandia National Laboratories Micro-Mechanical Systems division. Figure 5.19 illustrates the compliant mechanism design problem. Given a positive vertical input displacement, the output is designed to move in the negative vertical direction.

120

Fex uout

uin

Figure 5.19: Design Problem for the MEMS Displacement Magnifier

Using symmetry conditions, half of the topology was designed. The fixed-node method was discretized using 25 nodes and 200 elements. For both designs the modulus was set to 1000, the width was set to 1, the total volume constraint was set to 3, and the lower element bound was set at 1E-4. For all designs the Fex/uin ratio was set to one. Figure 5.20 shows the fixed-node ground structure and the unconstrained topology design. The unloaded geometric advantage of the maximum energy efficiency design was 1.24. In addition, the geometric advantage constraint was also investigated by designing amplifiers having mechanical advantages close to 5 and 10. The unloaded geometric advantages were 4.54 for Figure 5.20(c) and 7.46 for Figure 5.20(d). To achieve these geometric advantages, the penalty multiplier was set to 60. The floating-node method was also discretized using 25 nodes and only 72 elements. A total of 23 nodes were activated and given wandering ranges resulting in 107 design variables. Figure 5.21 shows the floatingnode ground structure and the unconstrained topology design. The unloaded geometric advantage of the maximum energy efficiency design was 1.16. In addition, the geometric advantage constraint was also investigated by designing grippers having geometric advantages close to 5 and 10. The unloaded geometric advantages were 4.98 and 9.99 respectively. To achieve these geometric advantages, the penalty multiplier was set to

121 200. For all cases, the floating node topology designs possessed slightly higher energy efficiencies.
10 10

Fex

= 75.80%

uin
0 1 2 3 4 5 6 7 8 9 10

10

(a) Initial Structure


10 10 9

(b) Optimized Design


9

= 56.07%

= 3.73%

10

10

(c) Optimized Design (GA ~ 5)

(d) Optimized Design (GA ~ 10)

Figure 5.20: Optimized MEMS Magnifier Topologies (Fixed-Node Approach)

122

10

Fex

10

= 74.25%

uin
0 1 2 3 4 5 6 7 8 9 10

10

(a) Initial Structure


10 9

(b) Optimized Design


10 9

= 63.25%

= 46.77%

10

10

(c) Optimized Design (GA ~ 5)

(d) Optimized Design (GA ~ 10)

Figure 5.21: Optimized MEMS Magnifier Topologies (Floating-Node Approach)

5.4 Conclusions Results of the topology optimization indicate the capability and robustness of the topology synthesis procedure. This optimization formulation has the unique ability to produce compliant mechanism topology designs regardless of the choice of the efficiency parameters, Fex/uin or Ks. Clear topology designs which satisfy motion requirements while maximizing the mechanical efficiency of the device may be obtained regardless of the structural element type, the element lower bound, and the total volume constraint. A detailed study of topology designs illustrated that the efficiency parameters, Fex/uin or Ks may quantitatively affect resulting topology designs. It was shown that increasing the Ks parameter often decreased the geometric advantage of designs, while increasing the Fex/ uin ratio had a less pronounced effect on optimized topologies. Topology optimization was demonstrated using a fixed-node ground structure and a floating-node ground

123 structure. These techniques require a minimum number of design variables and minimize computational effort to model and optimize designs. The fixed-node ground structure has the advantage in that it is numerically simpler to implement; however, the number of design variables grow exponentially as the number of nodes increases. The floating-node ground structure has an added advantage in that the number of design variables increases less rapidly and solutions generally obtain higher energy efficiencies. In addition the penalty method was introduced to explore the control of the unloaded geometric advantage. Due to the allowance for geometry change, the floating-node method was shown to be more effective finding topology designs for a given geometric advantage constraint.

CHAPTER VI DIMENSIONAL SYNTHESIS 6.1 Dimensional Synthesis Procedure The goal of dimensional synthesis for compliant mechanisms is to generate a detailed description of the structures size and geometry for a given topology. This can be performed considering the exact operational boundary conditions experienced by the mechanism and accounting for practical performance requirements such as stress and mechanical or geometric advantage constraints. The methods described here coordinate with topology synthesis techniques to complete the task of automating compliant mechanism design. For two-dimensional problems, size optimization shall describe the distribution of material about the medial axis of a frame structure. Geometry optimization refers to the relative locations of element node coordinates used to model structure. To optimize size and geometry, a designer must first begin with a topology which can satisfy basic motion requirements. These topologies can be successful identified using topology synthesis techniques developed in Chapter V; however, the designer may also utilize topologies from rigid link mechanisms which have also been shown to work well for some compliant mechanisms designs. Additionally, the designer may elect to combine simple optimal building block mechanisms topologies to form more sophisticated topology designs. To optimize the size and geometry of a mechanism design, each topology segment is discretized using arrays of frame elements; the number of which is controlled by the designer. The designer applies the appropriate boundary conditions to the mechanism, defines that material modulus, stress and volume limitations. As necessitated by the designer, the geometric coordinates of principle nodes can be

121

122 activated and given finite wandering regions. Element arrays are kept collinear and equidistant for stability reasons. Size and geometry optimization is performed using a sequential quadratic programming algorithm which adjusts the element thicknesses and the skeletal geometry. The optimization seeks to maximize the mechanical efficiency under the applied boundary conditions while satisfying all relevant design constraints. An illustration of the process is shown in Figure 6.1.
medial axis

active principle node element array

wandering ranges Fin Initial Guess Structure

Ks Fin Optimized Size and Geometry

Ks

Figure 6.1: Size and Geometry Optimization for Half of a Compliant Gripper

6.2 Design Convergence Similarly to the topology optimization procedure, the convergence of size and geometry designs was investigated. Given an initial starting size and geometry, the variation of optimized designs was investigated with respect to design boundary conditions, element discretization, principle node selection, and stress constraints. 6.2.1 Sizing Verses Force/Displacement and Spring Parameter As shown in Chapter 3, the simple compliant lever problem illustrated the convergence of designs based on the applied boundary conditions, Fex/uin or Ks. Depending on the values of these parameters, designs will be stiffer or more flexible. Figure 6.2 (a) illustrates a numerical implementation of the simple compliant lever. The

123 design is discretized using a total of 24 elements. Figures 6.2 (b), (c) and (d) demonstrate optimized element sizing for various Fex/uin parameters. For small ratios of Fex/uin, the design produced a more flexible design with relatively lumped compliance located close to the ground. As Fex/uin is increased, designs become progressively stiffer; however, the element sizing remains thinner toward the ground location. This follows intuitive behavior as the optimization is mimicking a revolute joint by thinning the material about a central location. However, increasing the efficiency parameter creates a stiffer joint which requires more energy to bend, but exerts greater stiffness with respect to the external force. This same behavior is also displayed using the spring efficiency objective function by increasing Ks parameter.
10 10

output

input

uin Fex
0 1 2 3 4 5 6 7 8 9 10

10

(a) Initial Design


10 10 9 9

(b) Optimized Design (Fex/uin = 1)

0 0 1 2 3 4 5 6 7 8 9 10

10

(c) Optimized Design (Fex/uin = 10)

(d) Optimized Design (Fex/uin = 100)

Figure 6.2: Optimized Element Sizing for the Compliant Lever Problem

124 6.2.2 Principle Nodes The effect of utilizing principle nodes was investigated by posing the following design problem shown in Figure 6.3 (a). Given a horizontal input displacement, the goal was to maximize the efficiency of the structure pushing against a vertical external force. The design domain was formalized using a circular pattern of 8 elements and 6 active nodes. The geometric locations of the input, output, and ground locations were fixed, while the intermediate nodes were allowed to wander within the shaded wandering range shown in Figure 6.3 (a). Figure 6.3 (b) illustrates the optimized result. With the exception of a minor anomaly, the intermediate nodes collapsed into a straight line pattern between the vertex nodes (input, output, and ground). This behavior is generally seen in linear finite element analysis where frame sections are kept relatively straight placing most of the material in tension and compression (although this observation may not be true for non-linear response). Figure 6.3 (c) and (d) show the same design using element arrays between the primary vertex locations. Here the intermediate nodes are kept collinear and equidistant to the input, output, and ground locations, which are fixed. The efficiencies of the two designs are nearly identical ( = 74.04% for the floating node problem and = 73.74% for the fixed node problem). Using this assumption that the medial axis of a topology sections remain straight, the total number of design variables can be reduced, which is advantageous for stability reasons (helping to keep elements from collapsing to zero length and minimizing the total number of design variables. It should be noted that the methods developed take advantage of using element arrays, although the designer may model a structure using as many or as few principle nodes as desired.

125

10

10

Fex uout uin

10

10

(a) Initial Design


10 9

(b) Optimized Design (free intermediate nodes)


10 9

10

10

(c) Optimized Design (fixed intermediate nodes)

(d) Deformed Geometry

Figure 6.3: Illustration of Node Re-Location During Optimization

6.2.3 Effect of Element Discretization The goal of the element arrays is to describe the distribution of material about the medial axis of the topology structure. Increasing the number of elements in an element array can be shown to affect the final design. Figures 6.4 (a) and (b) illustrate two separate designs under the same Fex/uin parameter, discretized using different numbers of element arrays. Figure 6.4 (a) illustrates an optimized compliant triangle using 10 elements per topology leg. The efficiency of the resulting design was 95.69% and the maximum stress was 10.3 MPa. Figure 6.4 (b) shows the design under the same conditions using 40 elements per topology leg. Increasing the element discretization resulted in a mechanism have an increased efficiency of 95.87% while the maximum stress also increased to 20.5 MPa. Ignoring stress limitations, the design of efficient compliant mechanisms appears to favor lumped compliance. However, it should be noted

126 that the efficiency increased less than 1 percent while the maximum stress nearly doubled. If the design were discretized using an infinite number of elements, the thickness of the structure may approach zero at a point; however, this type of design would not be practical as the axial stress would approach infinity. Consequently, the design should be discretized using a finite number of elements with an N high enough to adequately approximate the stresses along the length of the section.
10 10

0 0 1 2 3 4 5 6 7 8 9 10

10

(a) Optimized Design (N = 10)

(b) Optimized Design (N = 40)

Figure 6.4: Effect of Increased Element Discretization on Mechanism Design

6.2.4 Effect of Stress Constraints The effect of stress constraints was investigated for varying input parameters for both the force-displacement efficiency formulation and the spring-efficiency formulation. The same compliant triangle using in the previous example was investigated by incorporating a maximum stress limitation of 10 MPa. Designs were discretized using 20 elements per topology leg. Table 6.1 illustrates the results for the spring efficiency formulation using various ranges of the spring parameter, Ks, to input force, Fin. Designs that converged successfully are indicated by a corresponding efficiency value. Designs which failed to converge due to the fact that a feasible design could not be found given the stress constraint are indicated using a -.

127
Table 6.1: Feasible Compliant Mechanism Design Ranges for the Spring Efficiency Formulation

Ks Fin 5 10 20 40 60 80 100

1 98.92 -

10 95.41 95.24 94.70 -

100 71.68 71.68 71.68 71.68 71.68 -

1000 20.58 20.58 20.58 20.58 20.58 20.58 -

10,000 2.53 2.53 2.53 2.53 2.53 2.53 -

For low values of Ks, designs tended to be primarily limited by the combination of bending stress and axial stress. For high values of Ks, the compliant mechanism tended to be primarily limited by the axial stress limit. Shown in Figure 6.5, The overall structure of the feasibility range tends to resemble a backwards b rotated 90 degrees counterclockwise. Thus for designs with small input forces and low spring stiffness, designs will generally be feasible; however, increasing the input force quickly pushes the design into an infeasible region. For designs with large spring stiffness, the feasible region tends to level off and is primarily a function of the maximum axial stress.
limited by axial stress limited by bending & axial stress Fin feasible region

Ks Figure 6.5: Feasible Design Space for Spring Stiffness and Input Force

128

A similar matrix was investigated for the force-displacement efficiency formulation. Table 6.2 details the feasible and infeasible design ranges. Designs that converged successfully are indicated by a corresponding efficiency value. Designs which failed to converge due to the fact that a feasible design could not be found given the stress constraint are indicated using a -.
Table 6.2: Feasible Compliant Mechanism Design Ranges for the Force-Displacement Formulation

Fex uin 0.20 0.25 0.30 0.35 0.40 0.45 0.50 0.55

20 90.79 92.34 93.38 94.11 94.63 94.98 95.06 -

25 88.92 90.79 92.06 92.93 93.46 -

30 87.10 89.28 90.73 91.47 -

35 85.34 87.73 -

40 -

Increasing the Fex/uin ratio resulted in stiffer compliant mechanism designs with lower energy efficiencies. Decreasing the Fex/uin ratio resulted in more flexible compliant mechanism designs with higher efficiencies. Thus, for designs with lower external forces and higher input displacements, designs are primarily limited by a combination of bending and axial stress. For designs with small input displacements and large external forces, designs are primarily limited by the maximum axial stress present in the structure. Shown in Figure 6.6, the feasible design range tends to resemble a b shape as designs under a small external load remain feasible for large input displacements; however, the maximum

129 external load can only be increased up to a maximum of 35 N which represents the maximum axial load given the stress limitation.

limited by bending & axial stress

uin feasible region limited by axial stress

Fex Figure 6.6: Feasible Design Space for External Force and Input Displacement

6.2.5 Effect of Element Discretization on Stress Constraints Given a fixed stress limitation, the effect of increasing the element discretization was also investigated. Using the compliant triangle, a design close to the infeasible range having nearly fully distributed compliance was used for the convergence study. Given a 10 MPa stress limit, an input displacement of 0.35 cm, and an external force of 30 N, designs were discretized using 5, 10, 20 and 30 element arrays. Shown in Figure 6.7 (a) is the element height verses the normalized topology length. The curves lie mostly on top of each other with the five element discretization showing the most error. As the element discretization is increased, the element height converges rapidly. Note the stresses along the length of the mechanism are nearly equal along the entire length (with the stresses falling off slightly toward the mechanism output. The slight kink which occurs at the higher element discretizations can be explained due to the reduction of the bending stress

130 at this location. Figure 6.7 (b) illustrates the element sizing for the optimized design. In contrast to the lumped compliance design without stress constraints, the distributed stress design displayed a reverse taper which is similar to a fully stressed cantilever beam. Note that the efficiencies for all of the designs were almost identical at 91.47 percent indicating that in the presence of stress constraints, the efficiency changes little due to increased element discretization.
0.25 10

N=5
0.2

N = 10 N = 20 N = 30

element height

0.15

0.1

0.05

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

10

section length (normalized) (a) Element Height Vs. Section Length (b) Optimized Design (N=30)

Figure 6.7: Effect of Stress Constraint with Increased Element Discretization

6.3 Optimized Examples The following examples demonstrate the effectiveness of the automated two-stage approach for compliant mechanism design. Size and geometry optimization is performed using the optimized topologies from the topology synthesis stage. A detailed compliant mechanism design is then produced under the exact design boundary conditions while accounting for stress limitations and, for some designs, mechanical advantage constraints.

131 6.3.1 Spring Efficiency Formulation Compliant Gripper Size and geometry optimization was performed for two compliant gripper topologies. The mechanisms was designed with a modulus of 2 GPa (Nylon), an out-ofplane thickness of 0.5 cm and a maximum volume of 3 cm3. The external spring was given a stiffness of 100 N/cm and the input force was 50 N. The maximum stress was limited to 20 MPa which corresponds to roughly one-third of the ultimate stress for Nylon. Using the results from the fixed-node topology synthesis from Figure 5.11 (c), the topology was further discretized using 40 elements and 5 principle nodes. Three of the five principle nodes were activated and given the following wandering ranges shown in Figure 6.8 (a). Figure 6.8 (b) illustrates the optimized size and geometry and (c) shows the deformed geometry. Note that the mechanism deforms as intended while satisfying the maximum stress constraint. The optimized design possessed an efficiency value of 91.05 percent. Using the results from the floating-node topology synthesis from Figure 5.12 (a), the topology was discretized using 48 elements and 7 principle nodes. Five of the principle nodes were activated and given the following wandering rages shown in Figure 6.9 (a). Figure 6.9 (b) illustrates the optimized size and geometry and (c) shows the deformed geometry. The mechanism deforms as intended while satisfying the maximum stress constraint. The optimized design possessed an efficiency value of 90.85 percent indicating that, for this particular problem, the fixed-node topology was slightly better than the floating node-topology. Figure 6.10 illustrates the finalized mechanism design (from the fixed-node topology design) with both symmetrical halves combined.

132
15 15

10

10

Fin

Ks
cm

0 0 5 10 15

cm
0 5 10 15

(a) Initial Structure


15

(b) Optimized Design

= 91.05%
10

MAsl = 1.3438 GAsl = 0.6721

max = 20 MPa Volume = 3 cm3


cm

10

15

(c) Deformed Geometry Figure 6.8: Optimized Compliant Gripper Size and Geometry (Fixed-Node Topology Design)

133
15 15

10

10

Fin

Ks
cm

cm
15 0 0 5 10 15

10

(a) Initial Structure


15

(b) Optimized Design

= 90.85%
10

MAsl = 1.661 GAsl = 0.5471

max = 20 MPa Volume = 3 cm3


cm

10

15

(c) Deformed Geometry Figure 6.9: Optimized Compliant Gripper Size and Geometry (Floating-Node Topology Design)

Figure 6.10: The Optimized Compliant Gripper Mechanism

134 Compliant Wrench Size and geometry optimization was performed for two compliant wrench topologies. Both mechanisms were designed with a modulus of 2 GPa (Nylon), an out-ofplane thickness of 0.5 cm and a maximum volume of 3 cm3. The external spring was given a stiffness of 100 N/cm and the input force was 10 N. The maximum stress was limited to 20 MPa. Using the results from the fixed-node topology synthesis from Figure 5.12 (d), the topology was discretized using 87 elements and 10 principle nodes. Eight of the principle nodes were activated and given wandering rages shown in Figure 6.11 (a). Figure 6.11 (b) illustrates the optimized size and geometry and (c) shows the deformed geometry. Note that the mechanism deforms as intended while satisfying the maximum stress constraint. The optimized design possessed an efficiency value of 83.34 percent. The deformations shown in Figure 6.11 (c) are relatively large and may be inadequately modeled using linear finite element theory. Never the less, the linear finite element technique coupled with the optimization procedure represents a good first approximation for compliant mechanism problem which undergo non-linear deformations. Future work should concentrate on extending the size and geometry optimization to handle large deformations. Using the results from the floating-node topology synthesis from Figure 5.15 (d), the topology was further discretized using 71 elements and 15 principle nodes. Thirteen of the principle nodes were activated and given wandering rages shown in Figure 6.12 (a). Figure 6.12 (b) illustrates the optimized size and geometry and (c) shows the deformed geometry. The mechanism deforms as intended while satisfying the maximum stress constraint. The optimized design possessed an efficiency value of 86.05 percent indicating that the floating-node topology design was actually better than the fixed node topology design. Figure 6.13 illustrates the finalized mechanism design (from the floating-node topology design) with both symmetrical halves combined.

135

15

15

Fin
10 10

Ks
cm

cm
15 0 0 5 10 15

10

(a) Initial Structure


15

(b) Optimized Design

= 83.34%
10

MAsl = 3.95 GAsl = 0.211

max = 20 MPa Volume = 3 cm3


cm

10

15

(c) Deformed Geometry Figure 6.11: Optimized Compliant Wrench Size and Geometry (Fixed-Node Topology Design)

136

15

15

Fin
10 10

Ks
cm

cm
15 0 0 5 10 15

10

(a) Initial Structure


15

(b) Optimized Design

= 86.05%
10

MAsl = 3.6 GAsl = 0.238

max = 20 MPa Volume = 3 cm3


cm

10

15

(c) Deformed Geometry Figure 6.12: Optimized Compliant Wrench Size and Geometry (Floating-Node Topology Design)

Figure 6.13: An Optimized Compliant Wrench Mechanism

137 6.3.2 Force-Displacement Efficiency Formulation Piezo-stack Amplifier The goal of the piezo-stack amplifier was to generate a low-profile mechanism to efficiently amplify the displacements of a piezoelectric stack actuator. As the piezoelectric actuator expands horizontally, the mechanism produces a vertical displacement expanding against a downward external load. The performance characteristics of the piezo-stack actuator were taken from Kinetic Ceramics (1999). The selected actuator measured 8 cm by 1 cm (diameter) and could produce a maximum blocking force of 1200 N and an unloaded displacement of 80 m. Assuming a linear force-deflection relationship, the actuator can supply a maximum of 600 N at 40 m displacement (300 N and 20 m per quarter section). The compliant mechanism was designed with a modulus of 200 GPa (Steel), and an out-of-plane thickness of 1.0 cm and a maximum volume of 2 cm3 per quarter. The maximum stress was limited to 200 MPa. In addition, a mechanical advantage constraint was imposed on the design constraining the initial mechanical advantage to 0.333. Using a one to three mechanical advantage, the maximum force the actuator can lift would be 200 N for a 100% efficient compliant mechanism. Allowing for inefficiency in the mechanism and potential reserve power, the appropriate external force was estimated to be 150 N (75 N per quarter). Size and geometry optimization was performed for two piezo-stack amplifier topologies. Using the results from the fixed-node topology synthesis from Figure 5.17 (b), the topology was discretized using 39 elements and 8 principle nodes. Six of the principle nodes were activated and given wandering rages shown in Figure 6.14 (a). Figure 6.14 (b) illustrates the optimized size and geometry and (c) shows the deformed geometry. Note that the mechanism deforms as intended while satisfying the initial mechanical advantage constraint. The optimized design possessed an efficiency value of 81.16 percent.

138 Using the results from the floating-node topology synthesis from Figure 5.18 (c), the topology was discretized using 50 elements and 10 principle nodes. Note this starting topology was simplified during earlier stages as size and geometry optimization produced mechanisms with overlapping members. Eight of the principle nodes were activated and given wandering rages shown in Figure 6.15 (a). Figure 6.15 (b) illustrates the optimized size and geometry and (c) shows the deformed geometry. The mechanism deforms as intended while satisfying the initial mechanical advantage constraint. The optimized design possessed an efficiency value of 81.58 percent indicating that the floating-node topology design was again slightly better than the fixed node topology. Figure 6.16 illustrates the finalized mechanism design (from the floating-node topology design) with all four symmetrical halves combined.
9 9

Fex
7 6

uin

cm
1 1 2 3 4 5 6 7 8 9 1 1 2 3 4 5 6 7 8

cm
9

(a) Initial Structure


9

(b) Optimized Design

= 81.16%
MAi = 0.333 GAul = 3.0 MAf = 0.3260 GAl = 1.6730 max = 56 MPa
cm
1 2 3 4 5 6 7 8 9

Volume = 2 cm3

(c) Deformed Geometry (x 50) Figure 6.14: Optimized Piezo-Stack Amplifier Size and Geometry (Fixed-Node Topology Design)

139

Fex
7 6

uin

cm
1 1 2 3 4 5 6 7 8 9

cm
1 2 3 4 5 6 7 8 9

(a) Initial Structure


9

(b) Optimized Design

= 81.58%
MAi = 0.333 GAul = 3.0 MAf = 0.3245 GAl = 1.7250 max = 85 MPa
cm
1 2 3 4 5 6 7 8 9

Volume = 2 cm3

(c) Deformed Geometry (x 50) Figure 6.15: Optimized Piezo-Stack Amplifier Size and Geometry (Floating-Node Topology Design)

piezo-stack actuator

Figure 6.16: An Optimized Piezo-Stack Amplifier

140 MEMS Displacement Amplifier The purpose of the MEMS displacement magnifier was to generate rectilinear motion, while achieving a large geometric advantage. The design was to be used along with an electrostatic comb drive to propel a micro-mechanical locking system designed by Sandia National Laboratories Micro-Mechanical Systems division. Given a positive vertical input strain, the output is designed to move in the negative vertical direction. The mechanism was to be strained 2 m with the electrostatic comb actuators providing a maximum of 70 N (35 N per half). The specified external force was 3 N (1.5 N per half). The goal of the mechanism was to provide a geometric amplification of 10:1 which was enforced by constraining the initial mechanical advantage to 0.10. The material modulus was 160 GPa (Polysilicon), and the width was 2.5 m. In addition, the minimum in-plane thickness was limited to 1.25 m due to photolithography manufacturing limitations. The maximum stress of polysilicon is 1.2 GPa. The maximum volume of the mechanism was limited to 50 m3. Size and geometry optimization was performed for two MEMS displacement amplifier topologies. Using the results from the fixed-node topology synthesis from Figure 5.20 (d), the topology was discretized using 38 elements and 7 principle nodes. Five of the principle nodes were activated and given wandering rages shown in Figure 6.17 (a). Figure 6.17 (b) illustrates the optimized size and geometry and (c) shows the deformed geometry. Note that the mechanism deforms as intended while satisfying the initial mechanical advantage constraint. The optimized design possessed an efficiency value of 79.46 percent. Due to the lower bound element thickness limitations, the geometry of the structure changed significantly. However, the basic mechanism topology qualitatively satisfies the initial motion requirements and consequently remained appropriate.

141 Using the results from the floating-node topology synthesis from Figure 5.21 (d), the topology was further discretized using 56 elements and 7 principle nodes. Five of the principle nodes were activated and given wandering rages shown in Figure 6.18 (a). Figure 6.18 (b) illustrates the optimized size and geometry and (c) shows the deformed geometry. The mechanism deforms as intended while satisfying the initial mechanical advantage constraint. The optimized design possessed an efficiency value of 79.96 percent indicating that the floating-node topology design was slightly better than the fixed node topology; however, the fixed node topology possessed a slightly higher loaded geometric advantage indicating that it was slightly stiffer with respect to the external force. Both mechanisms required less than the maximum 35 N per actuator requirement achieving efficient transmission of forces and motions. Figure 6.19 illustrates the finalized mechanism design (from the floating-node topology design) with both symmetrical halves combined.
30 30

25

Fex

25

20

20

15

15

10

10

uin
x10 m

x10 m
5 5 0 5 10 15 20 25 30

5 5

10

15

20

25

30

(a) Initial Structure


30

(b) Maximum Efficiency Design

= 79.46%
MAi = 0.10 GAul = 10.0 MAf = 0.0664 GAl = 9.8796
x10 m

25

20

15

10

max = 141 MPa


30

5 5

10

15

20

25

(c) Deformed Geometry

Volume = 50 m3

Figure 6.17: Optimized MEMS Amplifier Size and Geometry (Fixed-Node Topology Design)

142

30

30

25

Fex

25

removable section

20

20

15

15

10

10

uin
x10 m

x10 m
30 5 5 0 5 10 15 20 25 30

5 5

10

15

20

25

(a) Initial Structure


30

(b) Optimized Design

= 79.96%
MAi = 0.10

25

20

15

GAul = 10.0 MAf = 0.0662 GAl = 9.7708


x10 m

10

max = 113 MPa


30

5 5

10

15

20

25

Volume = 50 m3

(c) Deformed Geometry Figure 6.18: Optimized MEMS Amplifier Size and Geometry (Floating-Node Topology Design)

anchoring location
Figure 6.19: An Optimized MEMS Displacement Amplifier

6.4 Alternative Topology Synthesis Procedures An advantage of the described size and geometry optimization is that the designer may experiment with any compliant mechanism topology synthesis technique including

143 topology optimization, rigid-link assimilation, or techniques which combine simpler building block mechanism topologies to create more sophisticated compliant mechanism topologies. The designer can then use the size and geometry optimization procedure to finalize the mechanism design. It should be noted that topology optimization can fail to produce acceptable topologies for extreme amplification mechanisms (generally when the geometric advantage constraint is greater than 10). Consequently, rigid-link assimilation and building block techniques are advantageous to design mechanisms possessing these extreme geometric advantages. One of the goals of topology synthesis for compliant mechanisms is to identify and characterize simple topology building blocks which produce a desired kinematic motion given the design space and the location of input and output. New and more sophisticated compliant mechanisms can be built up from simple mechanism topologies by stacking the building blocks in series or parallel. The resulting built-up structure can then be optimized using the described size and shape optimization. While topology optimization will not always reinforce building block methodologies, the approach represents a practical and intuitive method for designing more sophisticated compliant mechanism topologies. Figure 6. 20 illustrates several simple building block configurations which can be derived from topology optimization. Included in the building block structures are an approximate relationship for the unloaded geometric advantage (or inverse of the initial mechanical advantage). The approximate unloaded geometric advantage is largely dependent upon the geometry of the structure noted by the height and width dimensions a and b. The relationships are approximate due to the energy stored in the mechanisms. Also, the input and output of the topology designs can be reversed which then inverts the relationship for the unloaded geometric advantage.

144

in a out

in a

out b GAul ~ b / a (a) Transverse Displacement Building Block 1 b GAul ~ b / a (b) Transverse Displacement Building Block 2

out in a in b GAul ~ 1 (d) Parallel Guidance Building Block a out

b GAul ~ b / 4a (c) Transverse Displacement Building Block 3

Figure 6.20: A Sample of Compliant Mechanism Topology Building Blocks

To calculate the total unloaded geometric advantage for a mechanism combined using building blocks in series or parallel, the following approximate relationships hold

GA total

GAi
i=1 M

(series),

(6.1)

GAi
=1 GA total i -----------------M

(parallel).

(6.2)

145 To calculate the total energy efficiency for a mechanism combined using building blocks in series or parallel can similarly be estimated

total

i
i=1 M

(series),

(6.3)

i
i=1 total ------------- (parallel). M

(6.4)

These equations are approximate due to the rotational coupling of building blocks (which affects the mechanical and geometric advantages) and due to the energy storage in the structure (i.e the approximation will be more accurate for mechanisms with higher efficiencies). It is important to note that combining building blocks in series rather than parallel produces greater losses in efficiency due to the multiplication of efficiency values. However, stacking mechanisms in series may allow for large geometric amplification while occupying a relatively small area as shown in the following example. Certainly, the arrangement of these types of mechanisms is nearly limitless and can depend on many criteria such as space and manufacturability requirements as well as performance considerations. In order to achieve a large mechanical transformation, the design shown in Figure 6.21 stacked the displacement amplification of eight triangular topologies (four on each side). The triangular mechanism closest to the input has an unloaded geometric advantage of 1.5. The second, third, and fourth mechanisms have unloaded geometric advantages of 1.7, 2.5 and 3 respectively. Multiplying the combined advantages together yields an approximate geometric advantage of 19.1 for the entire mechanism. By designing the topology and geometry of the structure in this manner, the goal was to approximately

146 obtain an unloaded geometric advantage of 20. Figure 6.21 illustrates the building block approach for establishing the mechanism topology.

output relative motion

triangular building block

input Figure 6.21: Building Block Topology Design for the 20:1 Amplification Mechanism

Only half of the mechanism was optimized due to symmetry. The finite element model was composed of a total of forty-five beam elements and optimized using element sizing only (no geometry change). The material modulus was 2 GPa, the mechanism outof-plane thickness was 0.5 cm, the total volume was 8 cm3, the external load was 2 N, and the input displacement was 0.1 cm. The mechanism was intended to be manufactured using layered manufacturing techniques; therefore, due to manufacturing limitations, the lower bound of each beam element height was set to 0.1 cm. The resulting size optimization was able to achieve a mechanical efficiency of 67.6 percent. The unloaded geometric advantage of the finite element model was 18 (compared to the predicted 19.1 which corresponds to 6 percent error) indicating the success of the approximate relationship to calculate the total unloaded geometric advantage. The optimized prototype mechanism is shown in Figure 6.22. Note that the deformed mechanism contains the initial position of the input and output indicated by white lines. The prototype mechanism performs mostly as indicated. The prototype displays an unloaded geometric advantage of approximately 15:1. The error from the

147 predicted 18:1 advantage is most likely due to geometric non-linearities which occur as some legs of the topology undergo large displacements and rotations. The response under an external load could only be measured with forces up to 0.5 Newtons as the mechanism tended to buckle under greater external loads indicating the importance of buckling analysis for compliant mechanism structures.
output

input (a) Undeformed Mechanism

(b) Deformed Mechanism Figure 6.22: Illustration of the 20:1 Amplification Mechanism Prototype

The following building block mechanism topology was developed for the MEMS displacement amplifier for Sandia National Labs. The development of this mechanism preceded much of the topology synthesis work; therefore, building block techniques where used to derive the mechanism topology. To design an appropriate topology, a building block arrangement was developed capable of satisfying the primary motion

148 requirements while obtaining an approximate 10:1 geometric advantage. Figure 6.23 illustrates the symmetrical mechanism topology design.
output motion

input motion Figure 6.23: Building Block Topology for MEMS Displacement Amplifier

Using symmetry, the design was discretized using 35 elements and 8 principle nodes. Six of the principle nodes were activated and given wandering rages shown in Figure 6.24 (a). The design was optimized under the exact conditions of the MEMS displacement amplifiers in section 6.3.2. Figure 6.24 (b) illustrates the optimized size and geometry and (c) shows the deformed geometry. Note that the mechanism deforms as intended while satisfying the initial mechanical advantage constraint. The optimized design possessed an efficiency value of 78.0 percent. Comparing the building block results to the topology optimized designs which were later developed, the building block based design obtained an energy efficiency roughly 2 percent less than that of the topology optimized designs. The resulting performance of these topologies is often slightly less than that of the topology optimized designs; however, it is possible to obtain extreme geometric advantage designs where topology optimization can fail to produce a successful mechanism topology.

149

Optimized Design 30

Fex
25 20

30

25

20

15

15

10

10

uin
x10 m

x10 m
5 5 0 5 10 15 20 25 30

5 5

10

15

20

25

30

(a) Initial Structure


Deformed Geometry 30

(b) Optimized Design

= 78.02%
MAi = 0.10 GAul = 10.0 MAf = 0.0648 GAl = 9.8023
x10 m

25

20

15

10

max = 108 MPa Volume = 50 m3

5 5

10

15

20

25

30

(c) Deformed Geometry Figure 6.24: Optimized MEMS Amplifier Size and Geometry (Building Block Topology Design)

Researchers at Sandia Labs designed an intricate micro engine as a part of their research into micro-mechanical weapon locking and armament systems. The engine is powered by a set of three linear comb-drive actuators as shown in Figure 6.25 (a). The comb-drives are separated 120 degrees apart and convert the linear reciprocating motion of comb-drive actuators into continuous rotational motion of the main drive gear. The comb drive actuators used in the original design occupy too much die area relative to the rest of the locking system (not shown). Figure 6.25 (b) shows the redesigned micro engine with a compact comb-drive integrated to a 10:1 compliant motion amplifier. Compliant displacement multipliers enable them to use shorter stroke comb-drive actuators. Not only do such integrated actuator-transmission devices occupy less real estate, but they provide the opportunity to significantly increase the force generated per unit area (Rodgers, 1999). This is due to the fact that the comb-drive teeth can be more

150 tightly packed due to the shorter stroke. Consequently, the new amplifier-actuator combination is theoretically capable of pushing loads 15 times greater than the original design; all while requiring roughly one-fifth the total die area.

(a) Original Design

(b) Compliant Amplifier Design

Figure 6.25: Illustration of the Original Direct-Drive Design Versus the MEMS Displacement Amplifier Design (Courtesy of Sandia National Labs)

6.5 Conclusions The proceeding examples demonstrate that the developed size and geometry methods are flexible and powerful tools when used in conjunction with topology synthesis techniques. As illustrated in the design examples, the sizing and geometry of the structure often undergo significant changes due to the increased freedom afforded by the expanded node-wandering ranges and practical performance constraints such as stress constraints and geometric advantage constraints. The use of element arrays is shown to be appropriate for linear compliant mechanism design where structures typically produce

151 straight-line topology sections. This allows further simplification of the design problem as principle nodes and element arrays can be used to adequately describe the structure while minimizing the total number of design variables. The application of stress constraints is shown to produce practical mechanism designs which more evenly distribute strains throughout the structure. This is necessary as unconstrained designs tend to favor lumped compliance. Constraining the initial mechanical advantage is shown to allow the successful design of mechanisms with specific kinematic properties. This becomes quite useful for amplifier-type mechanisms which are beneficial for amplifying piezo-electric or other small-displacement actuation devices. The primary limitation of the size and geometry procedure can be attributed to the linear finite element approximation, which may not be accurate for mechanisms which undergo large deformations and rotations. However, the developed techniques are seen as a good first approximation for mechanisms which undergo significant non-linear behavior.

CHAPTER VII CONCLUSIONS AND FUTURE WORK 7.1 Conclusions The use of compliance in traditional mechanical devices has primarily been avoided due to the increased difficulty accounting for flexibility in kinematic design. A compliant mechanism is defined as a mechanical device that contains one or more flexible members which exploits elastic deformation to achieve controlled transmission of forces and motions. For applications requiring moderately small displacements and rotations, compliant mechanisms have many advantages over conventional rigid mechanisms including the elimination of friction, wear and backlash common with conventional mechanical joints. Compliant mechanisms are also easier to assemble and can take advantage of stored elastic energy to eliminate components such as stroke return springs. The main challenge in designing a compliant mechanism is to exploit compliance and generate an ideal structural form (topology, size, and shape) given the functionality requirements and performance specifications. High-performance compliant mechanisms which efficiently transmit forces and motions are difficult to design using trial and error approaches. Consequently, the development of a systematic approach to design these flexible mechanisms is critical to improve performance, and speed the development of new designs. Past systematic approaches can be divided into kinematic-based approaches and structural optimizationbased approaches. Kinematic based approaches have mainly adapted conventional rigidlink mechanism design techniques for compliant mechanism design. The drawback of this method is that kinematic degrees-of-freedom have limited applicability to structures

151

152 which deform according to the forces applied it. More recently, structural optimization techniques have been developed for compliant mechanism design. The generality of structural optimization methods allow for a more fundamental approach for combining kinematics and structural mechanics. Past structural optimization approaches for compliant mechanisms have focused mainly on topology design problems. A two-thickness compliant lever is presented in order to investigate the convergence behavior of both topology optimization and size and geometry optimization problems. While the model is a simplification of these types of optimization problems, it is an appropriate simplification which can be shown to mimic the behavior of more complex topology optimization and size and shape optimization problems. In order to develop a robust formulation for compliant mechanisms design, the optimization formulation in question must be capable of yielding a feasible compliant mechanism solution under all variations of objective function and constraint parameters. Many of the past compliant mechanism optimization formulations can fail to appropriately solve the compliant lever problem varying ranges of design valueswhich can lead to convergence difficulties for larger more complex problems. In this dissertation, two new, robust optimization formulations have been developed based on a linear, static compliant mechanism (a) lifting an external load and (b) pressing against a soft object. These boundary condition cases and are mathematically formalized as energy efficiency problems. Both energy efficiency formulations remain well-posed for the compliant lever regardless of the efficiency parameter, Fex/uin or Ks which control the flexibility and stiffness of optimized designs. This creates a well-posed and robust objective formulation for the optimization of topology, size and shape (geometry) of compliant mechanisms. Another advantage of the proposed formulation is that optimized design can be characterized by an efficiency value. This provides a practical tool to help discriminate between competing designs and provide insight to improve existing designs. The energy

153 efficiency value can also be used to determine where and under what conditions will a compliant mechanism be practical for specific applications. Using the two energy efficiency formulations, a two-stage approach for automated compliant mechanism design is developed which consists of a topology synthesis stage and a dimensional synthesis stage. These stages are numerically implemented using linear finite element theory along with truss and frame elements. A generalized approach for modeling the structure and calculating the design sensitivities is presented which can be applied to both topology design problems as well as size and geometry design problems. Analytic calculation of design sensitivities is performed using the adjoint variable method and by directly differentiating the element stiffness matrices with respect to element size and node coordinate design variables. Results of the topology optimization demonstrate the capability and robustness of the topology synthesis procedure. These optimization formulations posses the unique ability to produce compliant mechanism topology designs regardless of the choice of the efficiency parameters, Fex/uin or Ks. Topology designs which satisfy motion requirements while maximizing the mechanical efficiency of the device, can be obtained regardless of the structural element type, the element lower bound, and the total volume constraint. A detailed study of topology designs illustrated that the efficiency parameters, Fex/uin or Ks may quantitatively affect resulting topology designs. It was shown that increasing the Ks parameter often increased the mechanical advantage of designs, while increasing the Fex/ uin ratio had a less pronounced effect on optimized topologies. Topology optimization was demonstrated using (i) a fixed-node ground structure and (ii) a floating-node ground structure. These techniques require a minimum number of design variables and minimize computational effort to model and optimize designs. The fixed-node ground structure has the advantage in that it is numerically simpler to implement; however, the number of design variables grow exponentially as the number of nodes increases. The floating-node ground structure has an added advantage in that the number of design variables increases

154 less rapidly and solutions generally obtain higher energy efficiencies. In addition the penalty method was introduced to explore the control of the initial mechanical advantage. Due to the allowance for geometry change, the floating-node method was shown to be more effective when finding topology designs for a given mechanical advantage. Results of the proposed size and geometry methods demonstrate that the approach is flexible and powerful when used in conjunction with topology synthesis. For many cases, the initial structure resulting from topology synthesis is fairly close to the size and geometry optimized design. However, it was demonstrate that the geometry of designs, such as the MEMS displacement amplifier, may be greatly modified due to design constraints, such as the minimum thickness of element members. The use of element arrays was shown to be appropriate for linear compliant mechanism design where structures typically produce straight-line topology sections. This allows further simplification of the design problem as principle nodes and element arrays can be used to adequately describe the structure while minimizing the total number of design variables. The application of stress constraints is shown to produce practical mechanism designs which more evenly distribute strains throughout the structure. This is necessary as unconstrained designs tend to favor lumped compliance. Constraining the initial mechanical advantage is shown to allow the successful design of mechanisms with specific kinematic properties. This is highly beneficial for amplifier-type mechanisms which can amplify piezo-electric or other small-displacement actuation devices. The primary drawback of the size and geometry procedure can be attributed to the linear finite element approximation, which may not be accurate for mechanisms which undergo large deformations and rotations. However, the developed techniques are seen as a good first approximation even for mechanisms which undergo non-linear deformations.

155 7.2 Contributions The main contribution of this work are the development a generalized methodology and an automated approach for unified topological and dimensional synthesis of compliant mechanisms. The primary research contributions can be summarized: 1. The formulation of two new objective functions developed using the concept of maximum energy efficiency. A systematic technique was introduced to quantify the energy efficiency of a compliant mechanism lifting a static load and grasping a soft workpiece. A generic optimization formulation was then developed which could be equally applied to topology optimization and size / geometry optimization problems. 2. A two-thickness compliant lever was introduced to explore the behavior of optimal compliant mechanism designs. Results of the model can quickly predict if the optimization problem is well posed for compliant mechanism design. Analysis of the two energy efficiency optimization problems prove that they remain well-posed for compliant mechanism designs for all positive, finite values of the energy efficiency design parameters, Fex/uin and Ks. 3. Using the developed optimization formulations, numerical implementation was accomplished using truss and frame finite elements to approximate structural response. Analytic calculation of design sensitivities of the objective functions and constraints was accomplished using the adjoint variable method and by directly differentiating the element stiffness matrix. Sensitivity calculations were performed for element sizing and node coordinate variables. Using the developed modeling and sensitivity approach, a unified approach for topological and dimensional synthesis of compliant mechanisms was

156 implemented to automate the design of sophisticated, high-performance compliant devices. 4. Topology optimization was demonstrated using fixed-node and floating-node ground structures. In addition, a constraint limiting the initial mechanical advantage was applied using the penalty method to design compliant mechanisms with required kinematic properties. Size and geometry optimization was demonstrated using the concept of element arrays and principle nodes which simplify the number of design variables required for dimensional synthesis. Both mechanical advantage and stress limitations were applied to size and geometry problems to produce practical compliant mechanism devices. In addition to the contributions of this research, many engineering insights into the fundamentals of designing compliant mechanisms have been gained from this study. These insights are summarized below: If an optimization problem displays a null-solution for the simple compliant lever, then the formulation is likely to fail to provide appropriate convergence for larger, more complex compliant mechanism design problems. Minimizing the strain energy stored in a compliant mechanism not only maximizes the energy efficiency of the device, but also allows the mechanism to provide near constant mechanical and geometric advantages over a broad range of operating (boundary) conditions. For a compliant mechanism pushing against a constant external force, the mechanism possessing maximum energy efficiency will maintain the largest ratio of the geometric advantage to the total compliance. For a compliant mechanism pressing against a soft workpiece, the mechanism possessing maximum transmission efficiency will maintain the

157 largest ratio of the square of the output displacement to the input displacement. Ignoring local constraints such as stress constraints, designs with the same ratio of the external load to the input displacement Fex/uin or similarly the same external spring Ks will yield the same design and the same efficiency. 7.3 Future Research Directions Compliant mechanisms represent a synergy between kinematics and structural mechanics. Research in this area presents many intriguing challenges and offers the potential to greatly benefit the areas of MEMS, smart structures, bio-mechanical devices, and general product design. In order to ensure future success, compliant mechanism design techniques must be expanded using more sophisticated modeling techniques, such as non-linear finite element analysis. In addition new optimization formulations must be developed to handle new types of compliant mechanisms problems, such as multi-material design problems, multiple-input and output mechanisms, and synthesis of non-linear design problems. The techniques developed in this thesis are appropriate for two-dimensional single-input and single-output compliant mechanism design. Extending this case to threedimensional and multiple-input multiple-output compliant mechanism design is critical to design more sophisticated compliant devices. An example of such a sophisticated compliant mechanism can be found in smart structure applications, where the goal may be to reshape an entire surface subject to one or more localized input actuator locations. This represents an extremely challenging problem as the mechanism must deform from one shape to another while maintaining structural integrity during loading. Consequently, developing automated design tools to handle structural and kinematic synthesis for shape deforming structures would represent an important step forwards for smart structures technologies.

158 Traditional structural optimization techniques assume that the material properties are fixed and that the material is passive. The piezo-stack actuator mechanism represented a design case where the energy input from a piezoelectric actuator was used to design an appropriate compliant mechanism. In this particular case, the compliant mechanism design is separated from the piezoelectric actuator. However, a more advanced type of problem could continuously distribute the active (piezoelectric) and passive (compliant) materials throughout the design domain. Multiple material design techniques could model active materials, such as piezoelectric or shape memory alloy materials, or other materials having differing thermal expansion coefficients or poissons ratios. For cases where mechanisms must operate at high speed, dynamic response becomes important. Compliant mechanisms can be designed for cases where the lowest natural frequency must be above a prescribed threshold value to avoid resonance problems. In addition, devices could be designed where both the shape and frequency of vibration modes can be precisely controlled to achieve mechanism functionality via dynamic response. All of the past compliant mechanism optimization techniques have relied on linear structural analysis techniques. However, many compliant mechanism design problems involve nonlinear deformations and nonlinear material response which can produce large errors when modeled using linear techniques. Extending synthesis methods to handle nonlinear material response and nonlinear geometric deformations is critical for expanding the generality and capability of compliant mechanism design.

BIBLIOGRAPHY

166

167 BIBLIOGRAPHY Almgren, R., 1985, An Isotropic Three-dimensional Structure with Poissons ration = 1, Journal of Elasticity, Vol. 15, No. 4, pp. 427-430. Ananthasuresh, G.K., Kota, S. and Gianchandani, Y., 1993, Systematic Synthesis of Microcompliant Mechanisms - Preliminary Results, Proceedings of the Third National Applied Mechanisms and Robotics Conference, Cincinnati, Ohio, Paper No. AMR-93-082. Ananthasuresh, G.K., Kota, S., Kikuchi, N., 1994, Strategies for Systematic Synthesis of Compliant MEMS, DSC-Vol. 55-2, Proceedings of the 1994 ASME Conference on Dynamic Systems and Control, Chicago, IL, pp. 677-686. Ananthasuresh, G.K., 1994, A New Design Paradigm for Micro-Electro- Mechanical Systems & Investigations on the Compliant Mechanism Synthesis, Dissertation, University of Michigan, Ann Arbor, Michigan, 1994. Ananthasuresh, G.K., and Howell, L.L., 1996, Case Studies and a Note on the Degree-ofFreedom in Compliant Mechanisms, Proceedings of the 1996 ASME Design Engineering Technical Conferences, 96-DETC/MECH-1217. Bathe, K, 1996, Finite Element Procedures, Prentice Hall, Inc., Upper Saddle River, NJ. Barnett, R., 1966, Survey of Optimum Structural Design, Experimantal Mechanics, pp. 19A-26A. Bendsoe, M.P., Kikuchi, N., 1988, Generating Optimal Topologies in Structural Design Using a Homogenization Method, Computer Methods in Applied Mechanics and Engineering, Vol. 71, pp. 197-224. Bendsoe, M.P., Ben-Tal, A., Zowe, J., 1994, Optimization Methods for Truss Geometry and Topology Design, Structural Optimization, No. 7, pp. 141-159. Braibant, V., Fleury, C., 1984, Shape Optimal Design Using B-Splines, Computer Methods in Applied Mechanics and Engineering, Vol 44, No. 3, pp. 247-267, Burns, R.H., and Crossley, F.R.E., 1966, Structural Permutations of Flexible Link Mechanisms, ASME Paper No. 66-Mech-5. Burns, R.H., and Crossley, F.R.E., 1968, Kinetostatic Synthesis of Flexible Link Mechanisms, ASME Paper No. 68-Mech-36. Chandrupatla, T., Belegundu, A., 1991, Introduction to Finite Elements in Engineering, Prentice Hall Inc., Englewood, NJ. Chern, J., Prager, W., 1970, Optimal Design of Beams for Prescribed Compliance Under Alternative Loads, Journal of Optimization Theory and Applications, Vol. 5, No.

168 6, 1970, pp. 424-431. Chopra, I., 1996, State-of-the-Art of Smart Structures and Integrated Systems, Proceedings of the 1996 SPIE Smart Structures and Materials Conference, San Diego, CA. Diaz, A., Kikuchi, N., Papalambros, P., Taylor, J., 1983, Design of Optimal Grid for Finite Element Methods, Journal Structural Mechanics, Vol 11, No. 2, pp. 215230. Erdman, A., Sandor, G., 1991, Mechanisms Design: Analysis and Synthesis. Prentice Hall, Englewood Cliffs, NJ. Furukawa, E., Mizuno, M., Terada, K., 1991, Magnifying mechanism for use on piezodriven mechanisms, International Journal of the Sociecty for Precision Engineering, Vol. 25, No. 4, pp. 315-322 Frecker, M., Ananthasuresh, G., Nishiwaki, S., Kikuchi, N., Kota, S., 1997a, Topological Synthesis of Compliant Mechanisms Using Multi-Criteria Optimization, Journal of Mechanical Design, Vol. 119, No. 2, pp. 238-245. Frecker, M.I., Kota, S., and Kikuchi, N., 1997b, Use of Penalty Function in Topological Synthesis and Optimization of Strain Energy Density of Compliant Mechanisms, Proceedings of the 1997 ASME Design Engineering Technical Conference, DETC97/DAC-3760 Frecker, M.I., 1997, Optimal Design of Compliant Mechanisms, Doctoral Thesis, The University of Michigan, Ann Arbor, MI. Furukawa, E, Mizuno, M., Terada, K., 1991, Magnifying mechanism for use on piezodriven mechanisms, International Journal of the Sociecty for Precision Engineering, Vol. 25, No. 4, pp. 315-322. Haftka, R.T., Grandhi, R.V., 1986, Structural Shape OptimizationA Survey, Computer Methods in Applid Mechanics and Engineering, Vol. 57, pp. 91-106. Haftka, R., Gurdal, Z., 1992, Elements of Structural Optimization (3rd Edition), Kluwer Academic Publishers, Boston, MA. Hall, S.R., Prechtl, E.F., 1996, Development of a piezoelectric servoflap for helicopter rotor control, Smart Materials and Structures, Vol. 5, No. 1, pp. 26-34. Hemp, W., 1973, Optimum Structures, Clarendon Press, Oxford, England, 1973. Hetrick, J., Kota, S., 1999, An Energy Formulation for Parametric Size and Shape Optimization of Compliant Mechanisms, Journal of Mechanical Design, Vol. 121, No. 2, pp. 229-234.

169 Her, I., and Midha, A., 1987, A Compliance Number Concept for Compliant Mechanisms, and Type Synthesis, Journal of Mechanisms, Transmissions, and Automation in Design, Transaction of the ASME, Vol. 109, No. 3, pp. 348-355. Her, I., Midha, A., and Salamon, B.A., 1992, A Methodology for Compliant Mechanisms Design: Part II - Shooting Method and Application, Proceedings of the 18th ASME Design Automation Conference: Advances in Design Automation, DE-Vol. 44-2, pp. 39-45. Hill, T.C., and Midha, A., 1990, A Graphical User-Driven Newton-Raphson Technique for use in the Analysis and Design of Compliant Mechanisms, Journal of Mechanical Design, Trans. ASME, Vol. 112, No. 1, pp. 123-130. Howell, L.L. and Midha, A., 1994, A Method for the Design of Compliant Mechanisms with Small-Length Flexural Pivots, Journal of Mechanical Design, Vol. 116, No. 1, pp. 280-290. Howell, L.L., and Midha, A., 1996a, A Loop-Closure Theory for the Analysis and Synthesis of Compliant Mechanisms, Journal of Mechanical Design, Vol. 118, No. 1, pp. 121-125. Howell, L.L., and Midha, A., 1996b, Evaluation of Equivalent Spring Stiffness for Use in a Pseudo-Rigid-Body Model of Large-Deflection Compliant Mechanisms, Journal of Mechanical Design, Vol. 118, No. 1, pp. 126-131. Howell, L.L., and Midha, A., 1996c, Parametric Deflection Approximations for Initially Curved, Large-Deflection Beams in Compliant Mechanisms, Proceedings of the 1996 ASME Design Engineering Technical Conferences, 96-DETC/MECH-1215. Huang, N., 1971, Optimal Design of Elastic Beams for Minimum-Maximum Deflection, Journal of Applied Mechanics, Transactions of the ASME, Vol. 38, Seriese E, No. 4, pp. 1078-1081. Huang, X., Saif, M, MacDonald, N., 1996, A Micromotion Amplifier, Proceedings of the 9th Annual International Workshop on Micro Electro Mechanical Systems, San Diego, CA, IEEE 96CH35856. Imam, M., 1982, Three-Dimentional Shape Optimization, International Journal for Numerical Methods in Engineering, Vol. 18, pp. 661-673. Kikuchi, N., 1999, Generalized Layout Optimization of Elastic Structures, Course notes on the Homegenization Design Method, The University of Michigan, Ann Arbor, MI. Kikuchi, N., Nishiwaki, S., Fonseca, J.S.O., Silva, E.C.N., 1998, Design Optimization Method for Compliant Mechanisms and Material Microstructure, Computational Methods in Applied Mechanics and Engineering, Vol. 151, pp. 401-417.

170 Kinetic Ceramics, 1999, Piezo-Stack Actuator Specifications, on-line product catalog. Kota, S., Ananthasuresh, G.K., Crary, S.B., Wise, K.D., 1994, Design and Fabrication of Microelectromechanical Systems, ASME Journal of Mechanical Design, Vol. 116, pp. 1081-1088. Kristensen, E., Madsen, N. 1976, On the Optimal Shape of Fillets in Plates Subjected to Multiple In-plane Loading Cases, International Journal for Numerical Methods in Engineering, Vol. 10, pp. 1007-1009. Larsen, U., Sigmund, O., and Bouwstra, S., 1997, Design and Fabrication of Compliant Micromechanisms and Structures with Negative Poissons Ratio, Journal of Microelectromechanical Systems, Vol. 6, No. 2, pp. 99-106. Levy, R., Lev, O.E., 1987, Recent Developments in Structural Optimization, ASCE Journal of Structural Engineering, Vol. 113, No. 9, pp. 1939-1962. Matlab Users Guide, 1997, Matlab Optimization Toolbox, The MathWorks, Inc., online users guide. Michell, A., 1904, The Limits of Economy of Material in Frame Structures, Philosophy Magazine, Vol 8., No. 47, pp. 589-597. Midha, A., Norton, T.W., and Howell, L.L., 1994, On the Nomenclature, Classification, and Abstractions of Compliant Mechanisms, ASME Journal of Mechanical Design, Vol. 116, No. 1, pp. 270-279. Milton, G.W., 1992, Composite Materials with Poissons Ratios Close to -1, Journal of the Mechanics and Physics of Solids, Vol. 40, No. 5, pp. 1105-1137. Murphy, M.D., Midha, A., and Howell, L.L., 1996, The Topological Synthesis of Compliant Mechanisms, Mechanism and Machine Theory, Vol. 31, No. 2, pp. 185-199. Nishiwaki, S., Frecker, M., Seungjae, M., Kikuchi, N., 1998a, Topology Optimization of Compliant Mechanisms Using the Homogenization Method,, International Journal for Numerical Methods in Engineering, Vol. 42, No. 3, pp. 535-559. Nishiwaki, S., Min, S., Ejima, S., Kikuchi, N., 1998b, Structural Optimization Considering Flexibility, JSME International Journal, Vol. 41, No. 3, pp. 476-484. Nishiwaki, S., 1998, Optimum Structural Topology Design Considering Flexibility, Doctoral Thesis, The University of Michigan, Ann Arbor, MI. Papalambros, P., Wilde, D., 1988, Principles of Optimal Design, Cambridge University Press, New York, NY. Paros, J.M., and Weisbord, L., 1965, How to Design Flexural Hinges, Machine Design,

171 November 25, 1965, pp. 151-156. Pederson, P., 1970, On the Minimum Mass Layout of Trusses, Advisory Group for Aerospace Research and Development, Conf. Proc. No. 36, Symposium on Strucutal Optimization, AGARD-CP-36-70. Pierre, D., 1986, Optimization Theory with Applications, Dover Publications Inc., New York, NY. Prager, W., 1974, A Note on Discretized Mitchell Structures, Computer Methods in Applied Mechanics and Engieering, Vol. 3, pp. 349-355. Prager, W., 1977, Optimal Layout of Cantilever Trusses, Journal of Optimization Theory and Applications, Vol. 23, No. 1, pp. 111-117. Prager, W., 1978, Nearly Optimal Design of Trusses, Computers and Structures, Vol. 8, pp. 451-454. Prager, W., Rozvany, G.I.N., 1977, Optimal Layout of Grillages, Journal of Structural Mechanics, Vol. 5, No. 1, pp. 1-18. Saggere, L., Kota, S., 1997, Synthesis of Diestributed Compliant Mechanims for Adaptive Structures Application: An Elasto-Kinematic Approach, DETC97/ DAC-3861, Proceedings of the 1997 ASME Design Engineering Technical Conferences, Sacramento, CA Saggere, L., 1998, Static Shape Control of Smart Structures: A New Approach Utilizing Compliant Mechanisms, Doctoral Thesis, The University of Michigan, Ann Arbor, MI. Salamon, B.A., and Midha, A., 1992, An Introduction to Mechanical Advantage in Compliant Mechanisms, Proceedings of the 18th ASME Design Automation Conference: Advances in Design Automation, DE-Vol 44-2, pp. 47-51. Saxena, A., Ananthasuresh, G.K., 1998, An Optimality Criteria Approach for the Topology Synthesis of Compliant Mechanisms, Proceedings of the 1998 ASME Design Engineering Technical Conferences, DETC98/MECH-5937. Schittkowski, K., Zillober, C., Zotemantal, R., 1994, Numerical Comparison of NonLinear Programming Algorithms for Structural Optimization, Structural Optimization, Vol. 7, No 1-2, pp. 1-19. Sevak, N.M., McLarnan, C.W., 1974, Optimal Synthesis of Flexible Link Mechanisms with Large Static Deflections, ASME Paper No. 74-DET-83. Sheild, R.T., 1970, Optimal Structural Design for a Given Deflection, Journal of Applied Mathematics and Physics, Vol. 21. pp. 513-520.

172 Shoup, T.E., McLarnan, C.W., 1971a, A Survey of Flexible Link Mechanisms Having Lower Pairs, Journal of Mechanisms, Vol. 6, No. 3, pp. 97-105. Shoup, T.E., McLarnan, C.W., 1971b, On the Use of the Undulating Elastica for the Analysis of Flexible Link Devices, Journal of Engineering for Industry, Transactions of the ASME, pp. 263-267. Sigmund, O., 1994, Materials with Prescribed Constitutive Parameters: An Inverse Homogenization Problem, International Journal of Solids and Structures, Vol. 31, No. 17, pp. 2313-2329. Sigmund, O., 1995a, Some Inverse Problems in Topology Design of Materials and Mechanisms, Proceedings of the IUTAM Symposium on optimization of mechanical systems, Stuttgart, Germany, pp. 26-31. Sigmund, O., 1995b, Tailoring Materials with Prescribed Elastic Properties, Mechanics of Materials, Vol. 20, pp. 351-368. Sigmund, O., 1996, Design and Manufacturing of Material Microstructures and Micromechanisms, Proceedings of the third international conference on intelligent materials, ICIM96, Lyon, France, pp. 856-866. Sigmund, O, 1997, On the Design of Compliant Mechanisms Using Topology Optimiztion, Mechanics of Structures and Machines, Vol. 25, No. 4, pp. 493-524. Stadler, W., 1988, Multicriteria Optimization in the Engineering and in the Sciences, Plenum Press, New York, NY. Topping, B.H.V., 1983, Shape Optimization of Skeletal Structures: A Review, ASCE Journal of Structural Engineering, Vol. 109, No. 8, pp. 1933-1951. Tuttle, S.B., 1967, Mechanisms for Engineering Design, Chapter 8: Semifixed Flexural Mechanisms, John Wiley and Sons, Inc., New York, NY. Vogel, S., 1998, Cats Paws and Catapults, W. W. Norton and Company, New York, NY. Weinstein, W.D., 1965, Flexural Pivot Bearings, Machine Design, June 10, 1965, pp. 150-157. Winter, S.J., and Shoup, T.E., 1972, The Displacement Analysis of Path Generating Flexible-Link Mechanisms, Mechanism and Machine Theory, Vol. 7, No. 4, pp. 443-451. Zienkiewicz, O., Campbell, J., 1973, Shape Optimization and Sequential Linear Programming, in: Gallagher, R., Zienkiewicz, O., eds., Optimum Structural Design (Wiley, New York, 1973) pp. 109-126.

Das könnte Ihnen auch gefallen