:- i... . . ‘ . i. I: . h. an. .. h .m... F. flan .5. . :u t... .: ....§n.,Wn.Naw.. “25...: a ,. a 9: {an . . R 4.1mm»... z 5‘3?! t»: fifidfbfiwi . up 1. yr”. u. but... .\V o... 18 h... 03.1.! x, a . xi. .utnwauhiflv. 1... ,e... i .u... .5. ‘taIC‘! 4.. hiihflHWrflifii «53.5.2.9 x :1 .3. .A A. . K I ‘ £13,“. uwmfi c ufi....f3........ l n... .3“...- n9.- , itil‘llé.‘ ’ill ’61.. - V. v! I: Livmrxvlia... . .uh.;1h.....i\.é{ A ‘1.f..\0.¢p 9‘ f i .2}. . s .36.. )3I1. .131. 71). 11...!329. 5,. 1...... it: . :1. I... . .11.! .27. .4. P33nuild.y..i . gfixpfimfi . it» ~ .918... {1.631.113 6.-.} ‘tll 1m LIBRARY Michigr“ l State University This is to certify that the dissertation entitled Conjugated Polymer Actuators And Sensors: Modeling, Control, And Applications presented by Yang Fang has been accepted towards fulfillment of the requirements for the PhD. degree in Electrical Engineering Major Professor’s Signature 82/; //2.00 9 Date MSU is an Affirmative Action/Equal Opportunity Employer PLACE IN RETURN Box to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 5/08 K:IProj/Aoc&Pres/ClRC/DateDue.indd CONJUGATED POLYMER ACTUATORS AND SENSORS: MODELING, CONTROL, AND APPLICATIONS By Yang Fang A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Electrical Engineering 2009 ABSTRACT CONJUGATED POLYMER ACTUATORS AND SENSORS: MODELING, CONTROL, AND APPLICATIONS By Yang Fang Conjugated polymers are soft actuation and sensing materials with promising applica- tions in biomedical devices and micromanipulation systems. However, their sophisticated electro-chemo-mechanical dynamics and nonlinear behaviors present significant challenges in such applications. This dissertation is focused on using systems and control tools to ad- dress these challenges. In particular, a systems perspective is taken to model the dynamic and nonlinear behavior of conjugated polymer actuators and sensors, an adaptive control scheme is developed to handle model uncertainties, and a conjugated polymer actuated- micropump is explored both as an interesting application and a platform for validating and extending proposed models. Experimental results on trilayer polypyrrole (PPy) actuators and sensors are presented throughout the dissertation to support the modeling and control studies. On the modeling aspect, a linear, partial differential equation (PDE) is first used to capture the ion transport dynamics in actuation and sensing. With proper boundary con- ditions, the PDE is solved to derive an infinite-dimensional transfer function model that is geometrically scalable and amenable to model reduction. Nonlinear behaviors, in both electrical and mechanical domains, are also modeled. In particular, on the electrochem- ical side, a reduction-oxidation (redox) level-dependent impedance model is obtained by applying perturbation analysis to a nonlinear PDE. On the mechanical side, a nonlinear elasticity theory-based framework is proposed to capture the mechanics involved in large deformation. The framework has not only been effective in modeling a bending actuator, but also motivated the study of a novel, torsional actuator that is based on a fiber-directed PPy tube. On the control aspect, a robust adaptive control scheme is proposed to tackle the time- varying behavior of conjugated polymer actuators. Based on a reduced, linear model, a self-tuning regulator with parameter projection is designed and implemented. The robust adaptive control scheme has shown in experiments its superiority to traditional PID control schemes and a fixed model-following scheme. Finally, the application of conjugated polymer actuators to micropumps is explored. A conjugated polymer-actuated diaphragm micropump is designed and fabricated. In contrast to a typical design of clamping an actuation membrane at all edges, a novel, petal-shaped diaphragm design is proposed to alleviate the effect of edge constraints. Transfer function models from the actuation voltage to the diaphragm curvature and to the flow rate are obtained and verified experimentally. A flow rate of 1260 uL/min is achieved for the new diaphragm design, which represents a significant improvement over the traditional design. DEDICATION TO MY PARENTS AND MS. FANG XIE. iv ACKNOWLEDGMENT I am very grateful to my advisor Dr. Xiaobo Tan for his careful and kind guidance. His vision and expertise have influenced me a lot during my PhD study, and will also benefit me in the future. I also would like to appreciate his generous help on my job search and career development. I would like to thank Professor Hassan Khalil, Professor Nin g Xi, and Professor Thomas Pence for serving on my advisory committee. Their insightful comments on my research are greatly appreciated. I would like to thank Professor Thomas Pence in particular. His class of Nonlinear Elasticity and his personal guidance enabled me to finish the work on modeling mechanical nonlinearity in this dissertation. I would like to thank Prof. Giirsel Alici at University of Wollongong for providing PPy samples and valuable advice on my research, which contributed to successful completion of this dissertation. I am grateful to my colleagues in Smart Microsystems Lab: Zheng Chen, Dr. Mart An- ton, Stephan Shatara, Freddie Alequin, Alex Esbrook, Dawn Hedgepeth, Chris Gliniecki, John Thon, Andrew Temme, Qingsong Hu, Ernest Mbemmo, and Alex Will. I also would like to thank friends at Michigan State who have offered me generous help: Hua Deng, Guokai Zeng, Li Sun, Yue Huang, Zongliang Cao, Yixin Wang and many others. I would like to acknowledge the financial support of my research by NSF CAREER grant (ECCS 0547131) and MSU IRGP (05-[RGP-418). Finally I would sincerely thank my parents Xiaolong Fang and Linhua Liu, and Ms. Fang Xie. It is their support and love that encouraged me to finish this dissertation success- fully. 3 List of Tables .................................... viii List of Figures ................................... ix Introduction 1 1.1 Background on Conjugated Polymers .................... 1 1.2 Fabrication of Trilayer Conjugated Polymer ................. 5 1.3 Research Objectives .............................. 7 1.4 Contributions ................................. 8 1.4.1 Linear Models for Actuation and Sensing .............. 8 1.4.2 Modeling Nonlinearities in Conjugated Polymer Actuators ..... 10 1.4.3 Control of Conjugated Polymer Actuators .............. 12 1.4.4 Conjugated Polymer Micropump ................... 13 1.5 Chapter Descriptions ............................. 14 Linear Models for Conjugated Polymer Actuators and Sensors 16 2.1 Electrical Admittance Module ........................ 16 2.1.1 Review of Diffusive-elastic-metal Model and Adaptation to the Tri- layer PPy Actuator .......................... l. 7 ' 2.1.2 Scaling Laws for Double-layer Capacitance and Circuit Resistance . 21 2.2 Electromechanical Coupling ......................... 23 2.3 Mechanical Output .............................. 25 2.4 Complete Actuation Model and Experimental Verification .......... 29 2.4.1 Complete Actuation Model ...................... 29 2.4.2 Experimental Verification ...................... 30 2.5 Conjugated Polymer Electromechanical Sensor ............... 33 2.5.1 Full Sensing Model .......................... 36 2.5.2 Simplified Sensing Model ...................... 39 2.5.3 Experiments and Discussions .................... 42 2.6 Chapter Summary ............................... 43 Modeling Nonlinearities in Conjugated Polymer Actuators 46 3.1 Redox Level-dependent Impedance Model .................. 46 3.1.1 The Governing Partial Differential Equation ............. 46 3.1.2 Perturbation Analysis ......................... 48 3.1.3 Impedance Model ........................... 50 3.1.4 Experiments and Discussions .................... 53 3.2 Nonlinear Mechanical Model Based on a Swelling Framework ....... 58 TABLE OF CONTENTS vi 3.2.1 Finite Strain Tensors ......................... 61 3.2.2 Stresses ................................ 67 3.2.3 Equilibrium .............................. 70 3.2.4 Experimental Verification ...................... 72 3.3 Fiber-directed Conjugated Polymer Torsional Actuator ........... 76 3.3.1 Nonlinear Mechanical Modeling Framework ............ 80 3.3.2 Boundary Conditions ......................... 85 3.3.3 Nonlinear Mechanical Model .................... 87 3.3.4 Sample Preparation and Experimental Setup ............. 88 3.3.5 Experimental Results and Discussions ................ 90 3.4 Chapter Summary ............................... 90 Control of Conjugated Polymer Actuators 102 4.1 Model Reduction ............................... 102 4.2 Design of Robust Adaptive Controller .................... 109 4.2.1 Self-Tuning Regulator ........................ 109 4.2.2 Parameter Projection ......................... l 11 4.3 Experimental Results ............................. 112 4.3.1 Measurement Setup .......................... 1 12 4.3.2 Results and Discussions ....................... 114 4.4 Chapter Summary ............................... 123 Application: Conjugated Polymer Micropump 125 5.1 Design and Fabrication of the Micropump .................. 125 5.1.1 Diaphragm Design .......................... 125 5.1.2 Fabrication and Assembly of the Micropump ............ 126 5.2 Physics-based, Control-oriented Model for the Pump ............ 128 5.2.1 Electrical Admittance Module of PPy ................ 129 5.2.2 Electromechanical Coupling of PPy ................. 132 5.2.3 Mechanical Module of the Micropump ............... 133 5.2.4 Complete Model ........................... 141 5.3 Experimental Results ............................. 142 5.3.1 Admittance .............................. 142 5.3.2 Displacement ............................. 144 5.3.3 Flow Rate ............................... 149 5.4 Chapter Summary ............................... 149 Conclusion 152 Bibliography .............................................................. 155 vii 2.1 3.1 3.2 3.3 4.1 LIST OF TABLES Parameter values used for the model (2.53). ................. 44 Estimated values for C and R1. ....................... 55 Estimated values for C0 and R2. ....................... 56 Geometric parameters of the samples. .................... 89 Typical values of parameters in the actuation model .............. 106 viii 1.1 1.2 1.3 1.4 1.5 1.6 2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 LIST OF FIGURES The chemical structure of polypyrrole ..................... 2 The chemical structure of polyaniline. .................... 2 Illustration of the actuation mechanism of a trilayer polypyrrole actuator. Left: the sectional view of the trilayer structure; right: bending upon appli- cation of a voltage. .............................. 4 The setup for fabricating trilayer conjugated polymer ............. 6 Fabricated sample of trilayer conjugated polymer. .............. 6 Thickness of layers of a trilayer PPy actuator measured with a microscope. . 7 The complete model structure for conjugated polymer actuators. ...... 16 (a) Illustration of double-layer charging and diffusion for a conjugated polymer film with one side in contact with electrolyte; (b) equivalent circuit model for the polymer impedance. ...................... 18 Definitions of dimensional parameters used in actuation model ........ 20 Double-layer capacitance versus actuator size ................. 24 Resistance versus actuator width (length = 20 mm). ............. 24 Resistance versus actuator length (width = 3 mm) ............... 25 Standard linear solid model for capturing polymer viscoelasticity. ..... 28 Geometric relationship between the beam curvature and the tip displacement. 29 Geometry of the trilayer actuator ........................ 31 ix 2.10 2.11 2.12 2.13 2.14 2.15 2.16 2.17 2.18 2.19 2.20 3.1 3.2 3.3 3.4 3.5 Schematic of the experimental setup for joint force-displacement measure- ment. ..................................... 32 The setup for force/displacement measurement. ............... 32 Force versus displacement under an actuation voltage of 0.4 V (width = 3.5 mm) ....................................... 33 Force versus displacement under an actuation voltage of 0.4 V (width = 6 mm) ....................................... 34 Electrical admittance spectrum (size: 30 x 5 mm) ............... 35 Electrical admittance spectrum (size: 40 x 5 mm) ............... 36 Dynamic displacement response (size: 30 x 5 mm). ............. 37 Dynamic displacement response (size: 40 x 5 mm). ............. 38 Comparison between the numerical solution of the full model and the ana- lytical solution of the simplified model (2.50) when C0 is 0.05 M. ..... 43 Dynamic response of conjugated polymer sensors: Experimental measure- ment (marks) versus model prediction (line). Three samples with different widths (fixed length: 30 mm). ........................ 44 Dynamic response of conjugated polymer sensors: Experimental measure- ment (marks) versus model prediction (line). Three samples with different lengths (fixed width: 5 mm). ......................... 45 Equivalent circuit for the impedance model. ................. 51 The relationship between transferred charges and applied voltage for dif- ferent samples. ................................ 56 Model predictions versus experimental impedance for Sample 1 (0.35 M electrolyte, 0 V DC voltage). ......................... 57 Model predictions versus experimental impedance for Sample 2 (0.35 M electrolyte, 0.5 V DC voltage). ........................ 58 Model predictions versus experimental impedance for Sample 3 (0.35 M electrolyte, 1 V DC voltage). ......................... 59 3.6 Model predictions versus experimental impedance for Sample 4 (0.25 M electrolyte, 0 V DC voltage). ......................... 3.7 Model predictions versus experimental impedance for Sample 5 (0.25 M electrolyte, 0.5 V DC voltage). ........................ 3.8 Model predictions versus experimental impedance for Sample 6 (0.25 M electrolyte, l V DC voltage). ......................... 3.9 Illustration of the reference configuration and the deformed configuration. . 3.10 Definition of the bending radii at different locations .............. 3.11 The relationships between the input voltages and the transferred charges for the two different samples .......................... 3.12 Quasi-static bending under different actuation voltages for Sample 1 (13 x 5 mm) ....................................... 3.13 Quasi-static bending under different actuation voltages for Sample 2 (33 x 6 mm) ....................................... 3.14 Computational results on the change of a versus the applied voltage for Sample 1 (13x5 mm). ............................ 3.15 Computational results on the change of (1 versus the applied voltage for Sample 2 (33 x6 mm). ............................ 3.16 Computational results on the changes of thickness of the PPy layers with the applied voltage for Sample 1 ........................ 3.17 Computational results on the changes of thickness of the PVDF layer with the applied voltage for Sample 1 ........................ 3.18 (a) Illustration of the conjugated polymer-based torsional actuator; (b) pic- ture of the fabricated torsional actuator. ................... 3.19 The experimental setup to fabricate the fiber-directed conjugated polymer actuator ..................................... 3.20 Illustration of the actuator configuration. Left: original configuration; right: deformed configuration ............................. xi 60 61 62 63 67 73 74 75 77 77 80 81 3.21 3.22 3.23 3.24 3.25 3.26 3.27 3.28 3.29 3.30 3.31 3.32 4.1 4.2 4.3 4.4 Experimental setup to measure It“. The c0pper film is attached perpendicu- lar to the outer surface at the tube bottom. .................. 89 The torsional displacement of Sample 1 with 0.005 Hz sinusoidal voltage input ..................................... 91 The change of inner radius of Sample 1 with 0.005 Hz sinusoidal voltage input ..................................... 92 The change of tube length of Sample 1 with 0.005 H2 sinusoidal voltage input ..................................... 93 The torsional displacement of Sample 2 with 0.005 H2 sinusoidal voltage input. ..................................... 94 The change of inner radius of Sample 2 with 0.005 H2 sinusoidal voltage input. ..................................... 95 The change of tube length of Sample 2 with 0.005 Hz sinusoidal voltage input. ..................................... 96 The torsional displacement of Sample 3 with 0.005 Hz sinusoidal voltage input. ..................................... 97 The change of tube length of Sample 3 with 0.005 H2 sinusoidal voltage input. ..................................... 98 The torsional displacement of Sample 4 with 0.005 Hz sinusoidal voltage input. ..................................... 99 The change of inner radius of Sample 4 with 0.005 H2 sinusoidal voltage input. ..................................... 100 The change of tube length of Sample 4 with 0.005 Hz sinusoidal voltage input. ..................................... 101 Ratio as a function of D ................. 107 maX(lp1|,|p3|, Izzl) Comparisons between the infinite-dimensional and reduced model ...... 108 Illustration of the robust self-tuning regulator. ................ 109 Schematic of the experimental setup ...................... 113 xii 4.5 4.6 4.7 4.8 4.9 4.10 4.11 4.12 5.1 5.2 5.3 5.4 5.5 Experimental results on trajectory tracking (Batch One), I = 0 h. (a) Achieved trajectories versus desired one under the three controllers; (b) instantaneous tracking errors under the three schemes. ................... l 15 Experimental results on trajectory tracking (Batch One), I = 4 h. (a) Achieved trajectories versus desired one under the three controllers; (b) instantaneous tracking errors under the three schemes (note the different vertical-axis scales). .................................... l 16 Normalized average error ea and maximum error em under the three con- trol schemes (Batch One experiments). (a) Evolution of ea; (b) evolution Of em. ..................................... I 17 Experimental results on trajectory tracking (Batch Two), I = 3 h. (a) Achieved trajectories versus desired one under the three controllers; (b) instantaneous tracking errors under the three schemes. ................... 119 Normalized average error ea and maximum error em under the three con— trol schemes (Batch Two experiments). (a) Evolution of ea; (b) evolution Of em. ..................................... 120 Evolution of voltage input magnitude under the three schemes. (a) Batch One experiments; (b) Batch Two experiments ................. 121 Identified poles and zero under the robust adaptive control scheme in Batch One experiment. (a) Evolution over four hours; (b) Evolution over 100 seconds ..................................... 122 Experimental results of tracking a non-persistently exciting signal under the robust adaptive controller. (a) Trajectory tracking at t = 0 h and t = 4 h; (b) Tracking errors at t = 0 h and t = 4 h. ................. 124 Schematic of a petal-shape pumping diaphragm (top view). Left: before actuation; right: upon actuation. ....................... 126 The assembly schematic of micropump. ................... 128 The mechanism of flap check valves ...................... 129 The microfabrication process to make a flap valve. ............. 130 (a) The assembled micropump (top view); (b) The assembled micropump (bottom view) .................................. 131 xiii 5.6 5.7 5.8 5.9 5.10 5.11 5.12 5.13 5.14 5.15 5.16 5.17 5.18 5.19 A magnified view of the micrOpump to show the structure. ......... 131 The complete model structure for conjugated polymer actuators. ...... 132 Definition of the principal strains. ...................... 133 The coordinates and geometry definition. .................. 135 Illustration of the geometrical relationship in (5.12) and (5. 13), R = i. . . . 136 Geometric relationship in the calculation of the volume V and the work done by the diaphragm ............................. 139 Illustration of the volume change under the actuation voltage for the petal- shaped diaphragm. .............................. 140 Comparison of model prediction from (2.12) with the experimental results for the whole diaphragm and the petal-shaped diaphragm ........... 143 Comparison of model prediction from (5.33) with the experimental results for the petal-shaped diaphragm when there is no water (actuation voltage amplitude 4 V). ................................ 145 Comparison of model prediction from (5.33) with the experimental results for the petal-shaped diaphragm when pumping water (actuation voltage amplitude 4 V). ................................ 146 Comparison of model prediction with the experimental results for the whole diaphragm when there is no water by numerically solving (5.20) (actuation voltage amplitude 4 V) ............................. 147 Comparison of model prediction with the experimental results for the whole diaphragm when pumping water by numerically solving (5.20) (actuation voltage amplitude 4 V) ............................. 148 Flow rate of the micropump at different frequencies when input voltage magnitude is 3 V. ............................... 150 Flow rate of the micropump at different frequencies when input voltage magnitude is 4 V. ............................... 151 xiv Chapter 1 Introduction 1.1 Background on Conjugated Polymers Electroactive polymers (EAPs) are promising actuation and sensing materials with poten- tial applications in robotics and biomedical systems [1—6]. One class of EAP materials are conjugated polymers or conducting polymers [7—9], which are unusual organic materials that exhibit electronic conductivity. In 1977, Hideki Shirakawa, Alan G. MacDiarmid, Alan J. Heeger and coworkers reported high conductivity conjugated polymer, which eventually rewarded them with the 2000 Nobel prize in Chemistry. Polypyrrole (PPy) and polyaniline (PAN I) are two of the most investigated conjugated polymers, because of their good chem- ical stability and substantial strains. The chemical structures of PPy and PANI are shown in Fig. 1.1 and Fig. 1.2. The backbones of conjugated polymers have alternating single and double carbon-carbon bonds (conjugation), which results in positive charge carriers and thus electrical conductivity. Those charge carriers must overcome a band gap in order to be delocalized and thus enable conduction. The band gap is reduced by doping, which involves the addition of charges (dopants) to or removal of charges from the polymer chain. The doping is generally performed chemically or electrochemically. In chemical doping, dopants are donated to or repelled from the polymer backbone through chemical reactions 1 with dopant molecules. /\ /\ \/ \/ \/ Figure 1.1: The chemical structure of polypyrrole. 030,020,113 Figure 1.2: The chemical structure of polyaniline. 33—2 A conjugated polymer can be electrochemically doped by placing it in contact with an electrode in the electrolyte. When a sufficiently positive potential is applied, electrons are removed from the polymer, and negatively charged anions are incorporated into or positively charged cations are repelled from the polymer backbone to maintain the charge neutrality. Application of a sufficiently negative potential can reverse the process and re- pel the trapped anions out of or return the cations back to the polymer (reduction). The oxidation/reduction process is called redox, and these cations or anions are called dopants. The mass transport induced by ion movement during redox is considered to be the pri- mary mechanism responsible for volumetric change and thus the actuation of conjugated polymers [7, 8]. The redox process of PPy involving anions can be described as Oxidation: PPy + A" —) PPy+A— + 6“ Reduction: PPy+A"+e_ —» PPy+A—, where PPy represents the neutral state of polypyrrole and PPy+ is the oxidized state, A‘ 2 represents the dopant anions, PPy+A’ indicates that A" is incorporated into the polymer, and e‘ denotes an electron. Note that an anion-transporting conjugate polymer will expand during oxidation and contract during reduction, while a cation-transporting polymer will demonstrate the opposite behavior. Conjugated polymer actuators have the following attractive features: 0 can be actuated under very low voltages (typically 2 V or less). 0 have large strains (3% in-plane to 30% out—of-plane) and considerable stress. 0 can operate in liquid electrolytes, including body fluids. 0 lightweight and biocompatible. Linear extenders have been made by immersing a single piece of conjugated polymer in electrolyte [7,10]. As the potential difference is changed between the polymer and elec- trolyte, the volume of the conjugated polymer will change, so it can be used to generate linear motion. The bending motion can be generated with a bilayer configuration [8]. When the conjugated polymer layer is expanded or shrunk by accepting or repelling ions, the sub- strate layer does not change. Due to the strain mismatch within the two layers, the bilayer structure will bend. However, these two configurations require electrolyte environments. A trilayer bender can work without this limitation [11], as illustrated in Fig. 1.3. In the middle is an amorphous, porous polyvinylidene fluoride (PVDF) layer that serves both as a backing material and a storage tank for the electrolyte. On both sides of the actuator are the PPy layers. When a voltage is applied across the actuator, the PPy on the anode side is oxidized while that on the cathode side is reduced. The oxidized layer absorbs anions and expands, while the reduced layer gives up anions and contracts. The differential expansion thus leads to bending of the actuator, as shown in Fig. 1.3 (right). Conjugated polymers can be used as electrochemical sensors, because of the significant electrical conductivity change over several orders of magnitude in response to change in 3 Polypyrrole PVDF Predoped anion 0+ 0 9 G o‘ ‘3 oxidized reduced anion cation electron solvent state state Figure 1.3: Illustration of the actuation mechanism of a trilayer polypyrrole actuator. Left: the sectional view of the trilayer structure; right: bending upon application of a voltage. pH, humidity, or binding to biomolecules. These electrochemical sensors are based on the property changes of the single and double carbon-carbon structure while interacting with chemical compounds [12]. The conjugate polymers can also be used as electromechanical sensors where an open~ circuit voltage or a short—circuit current is generated under deformation. Comparing with the more extensive work on conjugated polymer actuators, the understanding of the elec- tromechanical sensing property of conjugated polymer is still in its infancy, where theory and models are needed to explain and quantify the phenomenon. We will explore a theory 4 to explain the electromechanical sensing capability of conjugated polymer in this thesis. 1.2 Fabrication of Trilayer Conjugated Polymer Fabrication of trilayer conducting polymer is achieved by electrochemically oxidizing pyr- role monomer from a solution to grow PPy layers on either side of a gold-coated porous PVDF film, which acts as the working electrode. The porous PVDF films are obtained from Sigma-Aldrich, which has the thickness of 110 pm with pore size of 0.45 pm. It is coated with a thin layer of gold (approximately 100 nm) by using the sputtering coater in W. M. Keck microfabrication cleanroom at Michigan State University. This ensures a good conductivity therefore good electrochemical growth of PPy. The electrolyte is a solution of 0.1 M pyrrole, 0.1 M Li+TFSI— in Propylene Carbonate (PC) with 0.5 w/w% water. Note that the solution with Li+TFSI_ will dope the PPy with TFSI’. The dopant can be changed to other ions, such as PFg, by changing Li+TFSI_ to the salt containing these 6— . A potentiostat is used in the electrochemical deposition, which is a control and measur- ions, such as TBA+PF ing device that keeps the potential of the working electrode at a constant level with respect to the reference electrode in the electrolyte. The one used in the Smart Microsystems Lab is Omni 101B from BSA Biosciences Inc. The cell is a cubic glass container by joining glass together with glass sealant. Two stainless steel meshes are used as the counter electrode and the PVDF membrane is sandwiched between them to ensure even PPy deposition. A part of the film is left protruding out of the electrolyte for electrical connection. The exper- iment setup is shown in Fig. 1.4, and one piece of fabricated trilayer conjugated polymer is shown in Fig. 1.5. After the samples are fabricated, the thickness of the PPy layer can be measured by using the digital imaging system in W. M. Keck microfabrication cleanroom at Michigan State University, which consists of a SPOT digital color camera and an Olympus BX60 5 Figure 1.4: The setup for fabricating trilayer conjugated polymer. Figure 1.5: Fabricated sample of trilayer conjugated polymer. optical microscope. The sectional view of the sample is captured by the imaging system, which also includes the measurement tool to obtain the dimensions. One image captured is shown in Fig. 1.6. Figure 1.6: Thickness of layers of a trilayer PPy actuator measured with a microscope. 1.3 Research Objectives In this dissertation, we will focus on modeling, control, and application of conjugated poly- mer actuators and sensors. To fully utilize conjugated polymer in bio/micro manipulation, bio-sensing, and micro robotics, it is crucial to understand the actuation and sensing prop- erties of conjugated polymer. Therefore, the first objective is to obtain a physics-based but control-oriented mathematical model for conjugated polymer, which captures its ma- jor dynamics and is useful for applications as actuator and sensor. In many applications, large deformation of actuators are required to achieve the tasks. Therefore, the second ob- jective is to understand the nonlinearities in conjugated polymer actuators, which become more significant as a higher input voltage is applied to generate larger deformation. The third objective is to utilize feedback control in conjugated polymer-based system to tackle the uncertainties and allow precise control of the actuator, because actuator behaviors are 7 heavily influenced by the materials and processes used during fabrication as well as the environmental conditions during operation. The fourth objective is to explore the use of conjugated polymer in promising applications. In this thesis, the application of conjugated polymer in micropump is explored, because rnicrofluidic devices are in significant devel- opment in recent years, and conjugated polymer actuators provide an important potential mechanism for realizing compact micropumps with low actuation voltages. PPy is used throughout this study. However, we expect that most of the results are directly applicable or can be extended to other conjugated polymer materials. 1.4 Contributions 1.4.1 Linear Models for Actuation and Sensing Conjugated polymer actuators require low actuation voltage (about 1 V), generate consider- able stress and large strain output, and are light and biocompatible. These advantages make them attractive for a wide range of robotic and biomedical applications, such as micro and biomanipulation, biomimetic systems, and biomedical devices [2—4, 8]. For all these appli- cations, it is highly desirable to have quantitative models available that can predict quasi- static and dynamic actuation performance in terms of intrinsic material parameters and actuator dimensions. Such models will be useful in feasibility analysis, design optimiza- tion, and even actuator control. Alici and coworkers investigated modeling and geometry optimization of bending curvature and force output for trilayer PPy actuators [11, 13, 14]. Christophersen et a1. characterized and modeled the bending curvature for bilayer PPy mi- croactuators of different dimensions [15]. All these studies were focused on quasi-static operating conditions, where the bending curvature and/or force output were examined in terms of material stiffness and actuator geometry. In this thesis, we have developed a scalable, control-oriented model for trilayer conju- 8 gated polymer that captures the major electrochemomechanical dynamics. This model con- sists of three modules: 1) electrochemical dynamics module, adapted from the diffusive- elastic-metal model of Madden [16]; 2) stress-generation module relating transferred charges to internal actuation stress; and 3) mechanical dynamics module. Scaling laws for key pa- rameters of the electrical admittance model are developed, rendering the complete model expressible in intrinsic material parameters and actuator dimensions. Experiments are con- ducted on trilayer conjugated polymer actuators of different dimensions to validate the scalable model for quasi-static force and displacement output, electrical admittance, and dynamic displacement response. Comparing with the extensive work on modeling of conjugate polymer actuators {6,16— 19], research on modeling of conjugate polymers as electromechanical sensors has been rel- atively limited. Wu et a1. [20] investigated the sensing behavior of a trilayer PPy beam by considering the perturbation of the Donnan equilibrium of the ion distribution by mechan- ical stimuli as the primary sensing mechanism. A similar viewpoint was also presented by Takashima and coworkers [21]. For better, understanding of the sensing mechanisms, however, quantitative modeling of the sensing dynamics is desired. We have investigated the dynamic sensing behavior of conjugated polymers. The model derivation starts with a partial differential equation (PDE) that governs the ion redistribu— tion dynamics subject to diffusion and migration (due to electrostatic interactions). How- ever, for real-time sensing applications, it is desirable to have an analytical model. For this purpose, we linearized and solved this PDE analytically in the Laplace domain, which leads to a transfer function that relates the open-circuit sensing voltage to the applied me- chanical deformation. Experimental measurement of the open-circuit voltage matches the model prediction reasonably well. This work provides the first step towards fundamental understanding of mechanical sensing mechanisms of conjugated polymers. 9 1.4.2 Modeling Nonlinearities in Conjugated Polymer Actuators Nonlinearities in conjugated polymer actuator become significant as a relatively high input voltage is applied. N onlinearity exist in both electromechanical and mechanical domains. In the electromechanical domain, the redox level, i.e., the amount of ions incorpo- rated in the conjugated polymer during reduction-oxidation, has significant impact on the electromechanical properties and dynamics of the material and consequently the actuation performance. While the effect of redox level on conductivity and Young’s modulus has been documented [15.22—26], its influence on dynamics of ion transport is more subtle and has received inadequate attention in terms of fundamental understanding and modeling. We have developed a redox level-dependent impedance model for conjugated polymers, which is physically based yet has a compact, explicit form. We start with a governing par- tial differential equation (PDE) that incorporates the dynamics of ionic diffusion, ionic migration, and redox reactions. The PDE is linearized around a given redox level via per- turbation analysis, and an exact, analytical solution is obtained by converting the PDE into the Laplace domain and enforcing appropriate boundary conditions. This model captures double-layer charging, diffusion, and migration effects. Experiments have shown that the proposed model are comparable to the diffusive-elastic-metal model [16] when the redox level is low, but shows clear advantage in predicting the impedance at higher redox levels. In the mechanical domain, the classical beam theory has been used to model the bending curvature for bilayer PPy actuators [15,27] and trilayer actuators [17]. However, this linear elastcity-based model is only valid when the ion movement-induced swelling is small so that: l) the strain and the stress can be related linearly; 2) the geometry nonlinearities can be neglected. Alici and coworkers modeled the bending curvature and force output for trilayer PPy actuators by using finite element analysis in analogy with a thermally driven beam, but no analytical model was presented [28]. The tensile strength experiments have shown that the strain and stress relationship of PPy film becomes clearly nonlinear as the 10 strain is increased over 4% [29,30]. These results indicate that the linear model based on the assumption that the stress varies linearly with the strain becomes invalid as the deformation gets large. Therefore, it is imperative to develop a nonlinear model to predict the actuator performance for the applications where large actuation is involved. In this thesis, a nonlinear elasticity-based method is proposed to capture the mechan- ical deformation induced by transferred ions. Neo-Hookean type strain energy functions are used for both PPy and PVDF to capture the nonlinear stress-strain relationship, which incorporates the effect of swelling [31]. The actuation-induced stresses are derived from the strain energy functions. The equilibrium configuration under a quasi-static actuation voltage is then obtained by solving the force and moment balance equations simultane- ously. Experiments have shown that when the applied voltages are small, predictions by both models are close to experimental data. But the nonlinear elasticity model fits the ex- perimental data better as the input voltage increases, which shows the superiority of the method in modeling large deformations of conjugated polymers. Different configurations of conjugated polymer actuators have been exploited, includ- ing bilayer and trilayer benders [8,32], linear extenders [7]. However, more complicated actuator motions will be required. Fiber-directed material can generate torsional motion be- cause of the anisotropy associated with the interaction between the fibers and the material matrix [33]. We have fabricated a fiber-directed conjugated polymer tube by integrating platinum wires into conjugated polymer during the deposition process of conjugated polymer. We have further proposed a nonlinear elasticity-based model to capture the relationship be- tween the actuation voltage and the torsional motion of the conjugated polymer tube. Com- pared with a linear elasticity-based model, this model is able to capture the complicated deformations and large deformation. The model has been verified with experimental mea- surements by using samples with different sizes. 11 1.4.3 Control of Conjugated Polymer Actuators It is critical to precisely control the force and/or displacement output of conjugated poly- mer actuators in many of their intended applications such as manipulation of single cells and micro-surgical operations. There has been extensive work on understanding the actua- tion mechanism of conjugated polymers as well as improving their actuation performance (strain output, strain rate, force output, work per cycle, lifetime, etc.) [7—9, 34]. However, control and control-oriented modeling of conjugated polymers remain largely unexplored. A proportional controller was used by Qi er al. to speed up the transient responses of an polyaniline actuator [35]. P. Madden treated the actuation dynamics as a first-order system and designed a PID controller for a polypyrrole actuator, where his main interest was to demonstrate a feedback loop consisting of polypyrrole actuator and sensor [17]. Taking again a first-order empirical model, Bowers did simulation studies on PID and adaptive control of conjugated polymers, but no experimental results were presented [36]. The primitive state of conjugated polymer control study is mainly dictated by the sophisticated electrochemomechanical processes during redox reactions, which makes it challenging to have a physical (non-empirical) model suitable for real-time control. In addition, actuator behaviors are difficult to characterize since they are heavily influenced by the materials and processes used during fabrication as well as the environmental conditions (temperature, humidity, electrolyte, etc.) during operation. In this thesis, we have developed a robust adaptive control scheme for conjugated poly- mer actuators with demonstrated performance in trajectory tracking experiments. A key component of the work is a simple model structure reduced from a full, infinite-dimensional physical model through model reduction, which captures essential actuation dynamics and is amenable to efficient real-time control. It enables compact, embedded controller imple- mentation for various micro, robotic, or biomedical applications. The resulting model is of second order with a zero, after we further discard a pole and a zero considering the typical 12 range of physical parameters. A recursive least-squares algorithm is then used to identify online the parameters of the reduced model. A self-tuning regulator [37] is designed based on the identified parameters to make the closed-loop system follow a reference model. A parameter projection step ensures that the parameter estimates stay within the physically- meaningful region, and thus makes the system robust against measurement noises, and unmodeled dynamics and nonlinearities. Experiments are conducted when the conjugated polymer actuator is operating in air, where its actuation behavior shows significant varia- tion over time because of the solvent evaporation. Experimental results have shown that the proposed scheme is superior to the commonly used PID scheme and to the fixed model- following scheme in terms of both tracking accuracy and required control effort. 1.4.4 Conjugated Polymer Micropump Microfluidic devices have been a t0pic of extensive research in recent years because of the significant development of biomedical industry. For many microfluidic systems, microp- umps are desired to integrate and simplify complex analytical procedures. The methods of actuation that have been investigated mainly include electroosmotic pumps, centrifugal pumps, and diaphragm pumps. An electroosmotic pump induces motion of liquid by ap- plying a voltage across a capillary tube, which drags the clustered ions in the liquid and consequently the fluid. However, its application is limited by the drawback that fluid flow cannot be generated when the ion concentration is below 10_5 M or above 10-2 M [38]. Centrifugal pumps are not very effective for fluid flows with low Reynolds numbers and have only been miniaturized to a limited size [39]. Diaphragm pumps have shown promise because of their compact sizes and the ability to deliver different kinds of fluid precisely. Different actuation mechanisms and materials have been investigated for diaphragm mi- cropumps. A thermopneumatic micropump utilizes an electric heater to increase the cham- ber temperature and therefore the pressure which leads to fluid flow [40], but the bandwidth of the device is limited by the heating and cooling process. Electrostatic actuation [41], 13 piezoelectric films [42], and P(VDF—TrFE) based electroactive polymer [43] have been ex- plored, but the high actuation voltages required may change the properties of the samples. Shape memory alloy thin film was also proposed [44], but the possible damage to the fluid sample due to the high transition temperature hindered its use in biological applications. We present a novel, conjugated polymer petals-actuated diaphragm micropump, which is in contrast to the typical whole diaphragm design [43,45,46]. We show through modeling analysis that the new design, by alleviating the edge constraints, can provide significantly larger diaphragm deformation and consequently the higher flow rate. The modeling of mechanical module for the diaphragm is developed using the energy method [47], which incorporates the elastic energy stored in the diaphragm and the work done on the fluid by the diaphragm. The final model for the petal-shape diaphragm pump, after model reduction, is represented as a finite-dimensional transfer function that captures the fundamental physics of conjugated polymer actuators and their interactions with flexible diaphragm and fluid. The rest of the pump is fabricated through PDMS-based MEMS processes. Experiments are conducted to evaluate the pump performance and the effectiveness of the model.The measured admittance, deformation curvature, and flow rate are found to match the models well. The complete model predicts that there is an optimal operating frequency to generate the largest flow rate, which is verified in experiments. 1.5 Chapter Descriptions The electrochemomechanical dynamics of conjugated polymer actuators and sensors are investigated in Chapter 2, where the nonlinearities are ignored and transfer function mod- els are presented to capture the actuation and sensing dynamics. The nonlinearities in both electromechanical and mechanical domains are modeled in Chapter 3. The redox level is introduced to capture the nonlinearity in the electrical domain, while nonlinear elasticity is used to model the mechanical nonlinearity. A fiber-directed conjugated polymer actuator 14 that can generate torsional motion is also proposed in Chapter 3. Its deformation is mod- eled by using nonlinear elasticity-based method. In Chapter 4 a robust adaptive control methodology is presented based on the proposed actuation model, and experimental results are provided to verify the effectiveness of the control strategy. The application of conju- gated polymer actuator in a micropump is studied in Chapter 5, where both modeling and experiments are conducted to investigate the micropump performance. The conclusion of this dissertation is provided in Chapter 6. 15 Chapter 2 Linear Models for Conjugated Polymer Actuators and Sensors The complete actuator model consists of three modules: 1) the electrical admittance mod— ule relating the current (and thus the charge transferred) to the voltage input; 2) the elec- tromechanical coupling module expressing the generated stress in terms of the transferred charge; and 3) the mechanical module connecting the stress to the displacement or force output. This structure is shown in Fig. 5.7, where V is the input voltage, I is the current, 8 is the corresponding strain, and Y is the mechanical output. V Admittance 1 Electromechanical 5 Mechanical Y Module _ Module f Module Figure 2.1: The complete model structure for conjugated polymer actuators. 2.1 Electrical Admittance Module It is desired to obtain the electrical admittance module relating the current (and thus the charge transferred) to the voltage input in order to obtain the complete actuation model. 16 Consequently, one will be able to predict the mechanical output of the actuator under the applied voltage. There are two possible mechanisms for the accumulated anions to enter the polymer matrix, diffusion and migration. J. Madden proposed a diffusive-elastic-metal model for one piece of PPy immersing in the electrolyte, where it was assumed that the polymer matrix is perfectly conducting and the ion transport within the polymer is solely detemiined by diffusion [16]. This model can be adapted to model the trilayer PPy actuator in our work and shows good agreement with experimental data for a wide frequency range from 10‘4 to 105 Hz. 2.1.1 Review of Diffusive-elastic-metal Model and Adaptation to the Trilayer PPy Actuator Firstly, the method in [16] is briefly described. And then we present a different and more straightforward derivation of the final admittance expression than the one in [16]. The voltage input of the actuator is applied across the two PPy layers. This produces a potential difference between each PPy layer and the electrolyte, as illustrated in Fig. 2.2(a). Note that only one side of the PPy is in contact with the electrolyte. Under the potential dif- ference, the anions in the electrolyte migrate toward the polymer, which results in double- layer charges at the polymer/electrolyte interface - like a double-layer capacitance with an equivalent thickness of 5. Fig. 2.2 (b) shows an equivalent circuit model of the polymer impedance, where the faradaic current has been ignored. C denotes the double-layer capac- itance at the polymer/electrolyte interface, R is the electrolyte and contact resistance, and ZD represents the “diffusion impedance”, which will be clarified in the derivation below. In the Laplace domain the total current [(3) in the circuit is the sum of the double-layer charging current and the diffusion current into polymer: [(5) =IC(S)+ID(S). (2.1) 17 Figure 2.2: (a) Illustration of double-layer charging and diffusion for a conjugated polymer film with one side in contact with electrolyte; (b) equivalent circuit model for the polymer impedance. The Kirchhoff ’s voltage law gives 1 V(s) = [(3) -R + s—-C -IC(s). (2.2) Let x denote the thickness direction of the polymer, with x = 0 representing the poly- mer/electrolyte interface. Assume that the ion concentration varies only in the x-direction, which will be denoted as c(x, s). From Fick’s law of diffusion, one has a Qfifi=—F¢Ldggtfiu:0 (am where A is the surface area of the polymer, F is the Faraday constant, (1 is the diffusion coefficient, and x = 0 represents the gradient of ion concentration at the interface. To compute [c(s), one first calculates the charges Qc(s) stored in the double-layer capacitor. Assume that the double layer has a thickness 5 and that the ion concentration within the (thin) double-layer is uniform, which equals c(0,s). Then Qc(s) = F -A - 6 - c(0,s), which leads to [C(s) : s-Qc(s) = F-A - 5 ~s-c(0,s). (2.4) The last equation needed for the derivation is the diffusion equation, which reads in the time domain: ac 32c =d — — 0 h —h 2.5 at 8x2, ¢s/—d>+¢sl = . (2.12) 3§+R.s3/2+R-l/s—d-s-tanhflhz—th/S—M) It is highly desirable to have quantitative models available that can predict quasi—static and dynamic actuation performance in terms of intrinsic material parameters and actuator dimensions. Such models will be useful in feasibility analysis, design optimization, and even actuator control. A scalable electrochemomechanical model is presented for trilayer conjugated polymer actuators, by developing the scaling laws for key parameters of the electrical admittance model. 2.1.2 Scaling Laws for Double-layer Capacitance and Circuit Resis- tance Two key parameters in the electrical admittance model (and thus in the force or displace— ment response model) are the double-layer capacitance C and the resistance R. It is of inter— est to know how they scale with the actuator dimensions. From the analogy to parallel-plate capacitors, one expects C to be proportional to the interface area between the PPy and the 21 electrolyte, i.e., C = COWL, (2.13) where C0 is the double—layer capacitance per unit area. The scaling of R is more involved. A transmission line model is proposed considering the nonzero (although low) resistivity of PPy. Here the top and bottom layers represent PPy resistance while the middle layer represents the electrolyte resistance. Let p1 and p2 denote the resistivity of the PPy and that of the electrolyte, respectively. Then the resistances R1 and R2 per unit distance are: 2 h R1: L, R2: 132—1. . (hz—h1)W w From the basic transmission line theory [49], 4P192h1 h —h w2 R: (2 1) h , (2.14) 4 tanh 00, i.e., the polymer becomes resistive for high-frequency inputs. Thus R is obtained from 1 . . . R = . . In ex enments an in ut of fre uenc 250 Hz rs used. 2 y( J a) I P P q 3’ Fig. 2.4 shows the comparison between the measured double-layer capacitance and the predicted value from (2.13) for different actuator sizes (L x W), where C0 is identified to be 7.86 x 10‘7F/mm2. Fig. 2.5 compares the measured resistance R and the predicted value from (2.14) as a function of actuator width, while the length of actuators is fixed to 20 mm. The resistivity values of p1 = 9 x IO‘SQm and p2 = 2.79m are obtained by fitting the experimental data points in Fig. 2.5, where p1 is restricted within the typical range [52] during tuning. As an independent validation step, the identified p1 and p2 are further used to predict the resistance for five 3mm-wide actuators with different lengths. As shown in Fig. 2.6, the comparison between the measured resistances. and their theoretical values based on (2. 14) confirms the prediction capability of the scaling law. The agreement between measurement and simulation in Figs. 2.4-2.6 validates the pro- posed laws for C and R, which will be used in the prediction of quasi-static force/displacement, electrical admittance, and dynamic displacement response in the following subsections. 2.2 Electromechanical Coupling As introduced in Chapter 1, the fundamental actuation mechanism of the conjugated poly- mer actuator is the mass transport due to the ion movement during redox that further in- duces the strain changes in the polymer. Therefore, the early works have focused on the 23 x10-4 1.2 4 fi.‘ ; 9 Experimental 7; 1_ 7 ---Simulatved o - . ,I C ’1 .‘3 x '8 0.8 " "7'9 m j. a. x 8 o 6 ’g’ol 7 ,,,,,,, h- ‘ I a ' _t! 0 4 """""" 91/" """""""""""""""""" 52 ,x’ i g 0.2»-- °’~ -------------------- o / a ,x . : 0O 50 100 150 Size (mm2) Figure 2.4: Double-layer capacitance versus actuator size. 70 1 0 Experimental 605 ............................. ---Simulated , \‘ 2 A \ 9,504; ................................................... i a: \ 0 \\ g 40 ————— -_.,\. ............... é 30~ - '5, ------------------------------------ . \“‘~‘ 20- »-~f~~- ‘3 ~~~~~~~~~~~~~ . ‘~~‘--"‘I-- 102 4 6 8 10 Width (mm) Figure 2.5: Resistance versus actuator width (length = 20 mm). static electro-mechanical relationship between the charges transferred into the conjugated polymer and the corresponding strain generated. It was shown that the strain is propor- 24 N on 0 Experimental ---Simulated 26i‘” ...... : ................ A \ e. j a: 24— \-—-- ..... 2 s '3 \s l .3 22L ...... V,“ ..... a, \s 7 i m ‘sa‘~~ 20" i ------- '9" """""" {p “10 2‘0 31) 40 Length (mm) Figure 2.6: Resistance versus actuator length (width = 3 mm). tional to the transferred charge density in the conjugated polymer [16, 53], which can be described as: e: 19p, (2.17) where 19 is the strain-to-charge ratio, p is the charge density. Equivalently, the induced stress by the transferred charges is where E p py denotes the Young’s modulus of the PPy layer. 2.3 Mechanical Output When a voltage is applied across the actuator, the PPy layer on the anode side is oxidized while the other layer on the cathode side is reduced. The oxidized layer absorbs anions and expands, while the reduced layer gives up anions and contracts. The differential ex- 25 pansion thus leads to bending of the actuator, as shown in Fig. 1.3. The bending of the trilayer beam can be predicted based on the linear elasticity theory after the swelling ra- tios are determined for different layers. One key assumption for the linear model is that the relationship between the stress and strain is linear, given by the small strain Young’s modulus, which is reasonable when the strain is sufficiently small. The elastic strain is obtained through the elementary beam theory by assuming that the strain changes linearly in the thickness direction with the distance from an appropriately located neutral line in the thickness direction [17]: c(x) = K-x, (2.19) where K' is the curvature to be determined. When there is no external load, the beam curvature can be obtained by solving the force and moment balance equations simultaneously _hl hl h2 —h2 Oppy1(x)dx+/_h1 O'pVDF(x)dx+/hl Oppy2(x)dx=0, (2.20) hi [12 _h2 GPPyl (xlxdx + [_hl GPVDFCrlxdx + fhl O'ppy2(X)xdx = 0, (2.21) where I21 and hz are as defined in Fig. 2.3. By superimposing the bending effect term upon the swelling term, the normal stresses in the PPy and PVDF layers along the beam direction are taken to be Gppy1(x) = Eppy8(x) —Eppyl9p, (2.22) GPVDF (x) = EPVDFEOC)’ (2.23) GPPy2 (x) = Eppy£(x) + Eppyfip, (2.24) where E PPy and E PV D F are the small strain Young’s moduli of the PPy and PVDF, and PPyl and PPy2 represent the reduced and oxidized PPy layers, respectively. For a symmet- 26 rical trilayer actuator, with the assumption that the thickness of the two PPy layers remains constant, it follows from (2.20) that the neutral line is the beam center line. Therefore, the actuation-induced curvature K can be obtained as [17]: h2—h12 l -— —l 319 ( + hl ) K:—. . 2’11 (”’72-’11)3+Epvdf_1 hr Eppy p, (2.25) where E pvd f is the Young’s modulus of the PVDF layer, and h pvd f denotes half of its thickness. Note that the mechanical dynamics is not considered in the derivation, which means that the results only hold in the quasi-static condition. One needs to consider the mechanical dynamics of the trilayer beam at high frequencies. The actuation bandwidth of a PPy actuator is typically much lower than its natural frequencies. For example, a sample of dimensions 20 x 5 x 0.17 mm was measured to have an actuation bandwidth of 0.5 Hz while its natural frequency was measured to be 65 Hz. Thus the inertial dynamics will be ignored. Therefore, only the material damping effect is considered. The standard linear solid model, illustrated in Fig. 2.7, can capture the viscoelasticity of solid polymer [54], and is adopted for modeling the dynamic moduli of PVDF and PPy. The equivalent modulus E’ in Fig. 2.7 is l + 5+ + E’(s) :5“ 76. 11x1 76. (2.26) X s + _ X Therefore one can furthermore modify (2.25) to be (2.27) to capture the higher frequency dynanucs h -— h (1+-l——h2—1 319 121 - d (1+—%7—H3+—Eli—i 1 Ema In experiments the beam tip displacement is measured by a laser distance sensor, as 27 Figure 2.7: Standard linear solid model for capturing polymer viscoelasticity. illustrated in Fig. 2.8. One can relate the measured displacement y 2 do — d to the curvature it“ via simple geometric calculations: 1 2y K — — = —, r y2 + 12 where l is the distance between the clamped end and the laser incident point when the beam is at rest. For small bending (y << I), the curvature is approximately linear with respect to the displacement: K x —. (2.28) Actuator '/ Clamp Figure 2.8: Geometric relationship between the beam curvature and the tip displacement. 2.4 Complete Actuation Model and Experimental Verifi- cation 2.4.l Complete Actuation Model One can obtain the complete scalable actuation model as follows by combining Eq. (2.12), (2.18), and (2.27): E _ 319 . v _ 2LWh1(h2—hl) h d (it—2.9.4 sz[§tanh((h2—h1)\/§)+\/§l lE’ . fl , (2.29) (Ir—2)3+ pvdf_1 [5+R.S3/2+R.—-s-tanh((h2-h1)\/E) hr Ebpy C 6 d 29 where C and R follow the scaling laws in (2.13) and (2.14). Considering Eq. (2.28), one can obtain the actuation model with tip displacement as the output: 2___3'EL__. v _Wh1(h2-h1) h (h—2-)2—1 sz[fitanh((h2—hl)\/E)+\/§] 1 5 d E’ -\/§ Jr? s . (2.30) (h_2)3+ PVdf_1 —+R-s3/2+R-—-s«tanh((h2-h1) —) ”I Elvpy C 5 d Consider a trilayer actuator clamped at one end, as shown in Fig. 2.9. Under quasi-static conditions, it can be shown through moment balance (2.21) that the following equation holds to characterize the relationship between force and displacement: F: —71Y+sz. (2.31) where 71 and y; are defined by 3 4WEPPyhpvde1+ h )3 _1+ Epvdf] 1’1 = 3L} hpvd f Eppy 2 Lf hpvdf )2—11. and E pvd f denotes the Young’s modulus of the PVDF layer. 2.4.2 Experimental Verification Quasi—static force/displacement measurement is conducted using an experimental setup il- lustrated in Fig. 2.10. Through the slide, a PVDF micro-force sensor [55] can measure the actuation force at different deflection levels under a given actuation voltage, from the free-bending configuration F = 0 to the fully-blocked configuration (v = 0). A picture of 30 Figure 2.9: Geometry of the trilayer actuator. the actual setup is shown in Fig. 2.11. Two sets of PPy actuators are used with widths of 3.5 and 6 mm respectively. In each set there are three samples with lengths of 20, 30 and 40 mm. Given a constant voltage input, the current can be predicted by (2.12) and thus the charge density p. Therefore the force versus displacement curve can be computed using (2.31), where E p py = 80 MPa, E pvd f = 440 MPa [14], 19 = 1.3 x 10—10 m3 -C"'1 [16]. Fig. 2.12 shows the comparison between the measured force-displacement curve and the model prediction for the set of actuators with width 3.5 mm under an actuation voltage of 0.4 V. Fig. 2.13 shows the results for the set of actuators with width 6 mm. It can be seen that in both figures reasonable agreement between measurement and simulation is achieved. Experiments are also conducted for other actuation voltages with satisfactory agreement with model prediction. Those results will not be presented here due to space limitation. One of the implications of the results in Figs. 2.12 and 2.13 is that the longer the actuator is, the smaller is the force output, and vice versa. This fact is in agreement with the results presented in [14]. In Fig. 2.14, the measured admittance is compared to the predicted one for an actuator of size 30 x 5 mm, while in Fig. 2.15 the same is presented for an actuator of size 40 x 5 mm. In both figures the agreement in magnitude plots is excellent. In the mean time the match for the phase plots is good except for the relatively big discrepancy at the low- frequency end. The low-frequency phase mismatch is likely due to the change of actuation 31 Stepper motor Laser ! sensor " PPy actuator Figure 2.10: Schematic of the experimental setup for joint force—displacement measure- ment Figure 2.11: The setup for force/displacement measurement. behavior in air when the solvent evaporates over time (the high-frequency measurements are done in a short period of time and the behavior change there is minimal). The model (2.30) for the dynamic displacement response is verified for a range of actu- ation frequencies for two different actuator sizes, 30 x 5 mm and 40 x 5 mm. The compari- son between the experimental measurement and the model prediction is shown in Fig. 2.16 and Fig. 2.17. The coefficients related to dynamic modulus in (2.26) for PPy and PVDF are identified based on curve fitting: 17 = 5.02 and x = 0.83 for PPy, and n = 4.68 and x = 0.044 for PVDF. From the figures the agreement between measurement and simu- 32 2 : f . 20mm Exp. 0 30mm Exp.‘ I 40mm Exp. —20mm Sim. --- 30mm Sim. ---40mm Sim. j Force (N) _________________ "o i 2 5 4 5 Tip displacement (mm) Figure 2.12: Force versus displacement under an actuation voltage of 0.4 V (width = 3.5 mm). lation is good for the magnitude plots. However, mismatch exists in the low-frequency portion of the phase plots. The cause of this, other than the possible behavior change due to solvent evaporation, requires further investigation. 2.5 Conjugated Polymer Electromechanical Sensor Research on the possibility to use conjugate polymers as electromechanical sensors has been relatively limited, comparing with the extensive work on modeling of conjugate poly- mer actuators [6, 16—19]. Wu et al. [20] investigated the sensing behavior of a trilayer PPy beam by considering the perturbation of the Donnan equilibrium of the ion distribu- tion by mechanical stimuli as the primary sensing mechanism. A similar viewpoint was 33 x10 .. 20mm Exp. 0 30mm Exp. 1 a 40mm Exp. ,,,,,,, i-.. —20mm Sim. i --- 30mm Sim. ~---40mm Sim.— f Force (N) .3 3 4 Tip displacement (mm) Figure 2.13: Force versus displacement under an actuation voltage of 0.4 V (width = 6 mm). also presented by Takashima and coworkers [21]. For better understanding of the sensing mechanisms, however, more experiments and quantitative modeling of the sensing dynam- ics are desired. A preliminary model is proposed to explain the sensing mechanism, which starts with a partial differential equation (PDE) that governs the ion redistribution dynamics subject to diffusion and migration (due to electrostatic interactions). Two boundary conditions are postulated: l) the ion concentration at the PPy/PVDF interface is proportional to the ap- plied mechanical strain, as the latter directly influences the pore sizes of the PPy layer; and 2) there is no diffusion flux at the interface between the PPy layer and the air. This PDE can be solved numerically. However, for real-time sensing applications, it is desirable to have an analytical model. For this purpose, we ignore the migration term of the PDE, 34 A'20 I 1’ _T ’ -9- Experimental] 1 g 3 q _ L": Simulated J 1 a, '25‘ ' 1 ' ' _- "111] u l. B . 1 ‘ ‘ 'E-30»~-» ‘ ,p l U} j . g _ i 51mm, .;;;. . ... L. . l '35?” o 1 2 1 10 10 10 10 60 . 7V1 ——fi'—' “Via—WT Y 'rr—‘*—v—‘ tr'r—f rr‘fi'} ’3‘ ' l 340' '1 a ' «I 20» 4 " ‘ l “L ‘1 3, G 1 “‘13 10' 10 10 10 Frequency (rad/sec) Figure 2.14: Electrical admittance spectrum (size: 30 x 5 mm). which makes the equation linear. The latter can be solved analytically in the Laplace do- main, which leads to a transfer function that relates the open-circuit sensing voltage to the applied mechanical deformation. The obtained solution shows good approximation to the solution of the original PDE when the ion concentration is relatively low and matches the experimental data for conjugated polymer sensors with different lengths and widths. How- ever, some recent experimental results by using conjugated polymer with different PPy thickness show discrepancy from the model prediction, which implies the model has not captured all the sensing phenomenon. 35 ,— .2o .. 3 . q wig-Experimentalm % , l—Simulated ‘ j TJ-25 -‘*-.' . , + 'u , l .3 ‘ 'j -E I3ol" ‘ .. “ ‘ ‘ ‘ ‘ ‘ ‘ _ w a .«1. * * * » . ,. 2 . . l '35-1L “no; 1 l W2 ‘1‘“ 3 10 10 10 10 10 Phase (deg) 0L... .L.L L 1:13.; L E . .L. . -1 0 1 2 3 10 0 10 1 Frequency (rad/sec) Figure 2.15: Electrical admittance spectrum (size: 40 x 5 mm). 2.5.1 Full Sensing Model The Nemst—Planck equation is used widely to describe the flux of ions under the influence of both an ionic concentration gradient and an electric field [56]: CF 2 — V — . . 2 J d( C+RTV¢) (23) The continuity equation holds: 8C v. : ___ 2.33 J at ( ) Since the thickness of the PPy layer is much smaller than its length or width, one can assume that dominant changes are all restricted to the thickness direction (denoted as x direction). This enables one to drop the boldface notation for these variables. From (2.33), 36 A 40 . 4 ;.____ m ‘ ;-9-Experimenta|] 3 .1 —Simu|ated j 5.! . “E 0» .- Oi . t5 . . E 20 10" 10° 101 102 e . . 1 r ~ 1 E400. .i e a ‘ . ‘ J _150i:1_._k 1-..!) 1 . 1. .11 . . . i . 2 1o 10 10 10 Frequency (rad/sec) Figure 2.16: Dynamic displacement response (size: 30 x 5 mm). wecan derive 2 2 8C 8 C F 8C8¢+ FCa (b) (2.34) a: (m HEB—x 1173,22 One can relate (l) to the ionic concentrations through the Gauss’s law: D — = ——V (2.35) Ke 4’ F- (C+ — C‘), (2.36) where D denotes the electric displacement, E denotes the electric field, and ice is the di— electric permittivity of PPy. Therefore, the following equations can be derived based on 37 -B - Experimental 1 E . 3 ‘ [717; —SirrLulated , § .“ . L ‘ L p. ’E' 0i i a 1 5’ ' ‘5‘ ‘L i g i ,, J -2o_,wA+~~+~- 1- .1 . ~ 4 10 10 1O 10 0‘ .1 7 . ., l a i ' % ~ - l 3 -50. “t L 3 . 3 to 400‘» in E I] 1‘ I ,‘ ‘ , . ‘ l .150 . . - ‘1‘ . 1 11 111~~ . 1 1 -s-~ 10'1 ° 01 1o2 10 1 Frequency (rad/sec) Figure 2.17: Dynamic displacement response (size: 40 x 5 mm). (2.35) and (2.36): 61> _ F x 9_¢ .97 — e/<)C(5=‘)d5+ax(°*’)’ (2‘37) 32¢ F 3x7 : EC(X,I). (2'38) In mechanical sensing, one can assume that C(x, t) fluctuates about some equilibrium con- centration C0: C(x, r) = C0 + C1 (x, t), (2.39) with C0 > C1 (x, t). Further assume that the electrolyte has high ionic conductivity, which implies :—:(0,t) = 0 in (2.37). Therefore, (2.37) can be approximated by i‘hfl, 3): ice (2.40) 38 Combining (2.34), (2.38), (2.39) and (2.40), and ignoring the C? term, one obtains 32C] F2 8C1], F2 3C1 ( + + (C2+2C C )) (241) — z u x -— . . a: 3x2 RTKe 3.. 0 RTKe 0 0 1 Two boundary conditions are imposed on (2.41): C1 (0,1) = KC0£0(I), (2.42) 8C1 Wlxth—hl —0, (143) where 80(t) represents the strain at the PPy/PVDF interface. Note that x = 0 denotes the PPy/PVDF boundary while x = ’12 — hl denotes the PPy/air interface. The first boundary condition implies that the concentration perturbation at x = 0 has a linear correlation with the applied deformation. The second one means that there is no diffusion flux on the other side of the PPy layer. 2.5.2 Simplified Sensing Model Eq. (2.41) is difficult to solve analytically. However, it is desirable to obtain an analytical model for conjugated polymer sensors. Such a model can facilitate fundamental under~ standing of the sensing mechanisms and be instrumental in sensor design and real-time sensing applications. An analytical model from (2.41) can be obtained by further ignoring the terms involving C0. Equivalently, this is to ignore the effect of electric field-induced ion migration. The approximation is valid when the nominal anion concentration C0 is low. The approximated PDE contains only the diffusion term: 8C1 __ dazcl at — 8x2. (2.44) 39 To solve this PDE, (2.44) is firstly converted into the Laplace domain 32C1(x,s) _ sC1(x,s) 8 2 d (2.45) x One can obtain the following analytical solution to (2.45): 3 KC \/; e\/; 2 +6 d 2 l 5 KC qfiwz—hn .‘f 0503 d . (2.46) + (h h (h h e)/;(2—1)+e—\/d(2— 1) Furthermore one can obtain the potential difference across one PPy layer based on (2.35) and (2.36): AV. ¢(h.s) —¢(0.s) dFKCOEO 2 S [ S S K" (fir/:2 hl) 1012— h.) (hz—h1)\[tanh(( ((—h2 h1:\/:)—1]-l-::e —O(—h2 h])2. (2.47) Due to the symmetry of the trilayer structure, the potential across the other PPy layer is + dFKCOEO 2 3K6 [fithz— h.) —(/§(h2—h1) 012— h1)\/;tanh(( ((-h2 h1) ):/:)—1]+ :——:3(f12—h1)2. (2.48) M + 40 Therefore, an analytical model for the total open-circuit sensing voltage is: v = Av, _sz ZdFKCOSO 2 [ SKe S S c([ng—hnH—(GMZ—h.) (h2—h1)\/§tanh((h2 —hl)\/§) — I]. + (2.49) From (2.49), one can obtain the transfer function relating the sensing voltage V(s) to the applied deformation 80(3) (strain at the PPy/PVDF interface): V(S) _ 2dFKCO - l K 50(5) se e\/;(—)h2 h1)+ :—\/:(h2—) hl)+ (—h2 h1)\/;tanh(( ((112— hi) +:\/;)—1]. (2.50) The following equation holds for the strain and the curvature 8(x) : K'x. (2.51) The strain 80 at the boundary of PPy layer and PVDF layer can be further related to the tip displacement: 2h £0_ _ 721), (2.52) 41 Now taking the applied tip displacement y as the input, from (2.50), one can obtain the transfer function for the sensing dynamics: V(s) 4dFKCOh1 2 — — [ Y(S) SlzKe S S e\/;(h2—h1)+e-\[g(h2-h1) +(h2 —h1)\/§tanh((h2 _h1)\/§)_1]' (2.53) We have compared the numerical solution of the full model with the analytical model derived above. In the computation, CO was taken to be 0.05 M. A sequence of sinusoidal inputs 80(t), up to 100 Hz, was used. For each input, the numerical solution of V(t) was computed, and the gain and phase shift at that frequency were evaluated. Fig. 2.18 com- pares the Bode plots obtained through numerical computation with those of the analytical model (2.50). As one can see, overall the discrepancy is not significant; in particular, as the frequency gets high, the discrepancy vanishes. 2.5.3 Experiments and Discussions Fig. 2.19 and Fig. 2.20 show the experimental characterization of the dynamic sensing behavior of the conjugated polymer samples. Predictions from the model (2.53) are also shown in the figures for comparison purposes. The parameters used for the model are listed in Table 2.1. It can be seen that the predictions from the analytical model in general fit the experimental data. In particular, the magnitude plots show good fitting with the experimen- tal data and predict the decaying trend as the frequency becomes higher. Although there is some discrepancy in the phase plots, the general trend of the experimental data is predicted by the model. The model (2.53) predicts that the sensing voltage is independent of the sample width. This is supported by the experimental data in Fig. 2.19, where one can see that the sensing behaviors of three samples, with different widths, are close to each other. The model also 42 —| O g. 0 , 1 1. l 3 1 , ‘ 1 5 ‘ 1 . L J -20 o f 1.2—W- 3 10 10 10 10 0 i F 1 ; — Full model 33 _20 2 ‘--- Approximated model i “‘ ‘ ‘ ‘ i 2 -40 a . 1 1 . i 1 _60 1 1 1111 1 1 1 1 111 1 1 1 .11. 10° 10‘ 1o2 103 Frequency(rad/sec) Figure 2.18: Comparison between the numerical solution of the full model and the analyt— ical solution of the simplified model (2.50) when C0 is 0.05 M. predicts that the sample length will influence the magnitude but not the phase of the transfer function, which is again verified by the experiments (Fig. 2.20). 2.6 Chapter Summary A full electrochemomechanical model is presented for trilayer PPy actuators. The scaling laws for two important parameters, the double-layer capacitance and the resistance, are pro- posed and experimentally verified. The model can thus be written in terms of fundamental material parameters and actuator dimensions. Through experiments with actuators of dif- ferent dimensions (length and width), the following aspects of the scalability of the model are independently validated: quasi-static force/displacement output, electrical admittance, and dynamic displacement response. The model will be instrumental in feasibility analy- 43 Table 2.1: Parameter values used for the model (2.53). Parameter Value F 9.65 x 104 C/mol K 0.072 ’12 . 85 pm h] 55 pm d 1 x 10—8 m2/s K'e 5.31 x 10—10 C2N_lm_2 CO 0.1 M 01 O _'1 . do 01____ -_fi Magnitude (dB) 0 o C: 73" Lg-ZOL a. . Il‘l n. ‘ ‘ iA. _50 1 i r . 1;: 100 1 2 103 Freguency (rad/$330) Figure 2.19: Dynamic response of conjugated polymer sensors: Experimental measure— ment (marks) versus model prediction (line). Three samples with different widths (fixed length: 30 mm). sis for new applications of conjugated polymer actuators and in optimization of actuator fabrication and design. Besides investigating sensing properties, we also proposed a model to explain the sens- ing mechanism of conjugated polymers sensor. The model accounts for the ion transport dynamics within the polymer. including both ion diffusion and electric field-induced mi- 44 -A <3 G ‘;__"_L_j__l Magnitude (dB) -40' Phase (deg) _5oi We: s .i.. .; 0 1 2 1° Fre1d)uency(rad/s¢1ag) Figure 2.20: Dynamic response of conjugated polymer sensors: Experimental measure- ment (marks) versus model prediction (line). Three samples with different lengths (fixed width: 5 mm). gration. A key assumption of the model is that the applied deformation prescribes the ion concentration at the polymer/electrolyte interface. The model was further simplified by dropping the migration term. In that case, an analytical solution was obtained in the Laplace domain, leading to an (infinite-dimensional) transfer function model for the sens— ing dynamics. Experimental results were also reported to support the modeling effort. 45 Chapter 3 Modeling N onlinearities in Conjugated Polymer Actuators 3.1 Redox Level-dependent Impedance Model It is known that the redox level of conjugated polymer has significant impact on its ma- terial properties [22, 23, 57], which will further influence the actuation performances [34]. However, understanding and modeling the influence of redox level are inadequate in the previous work. Therefore the ions transport phenomenon is modeled as follows by a redox level-dependent impedance model, which is developed by including both ionic diffusion and migration effects and by using perturbation analysis. Experiments are conducted under various redox conditions to verify the proposed model. 3.1.1 The Governing Partial Differential Equation The anion flux J inside the polymer is captured by the Nemst-Planck equation [56], includ- ing both a diffusion term and an electric field-induced migration term: C T F = —d V _ ———V 3.1 J < c + RT m, < > 46 where d is the ionic diffusion coefficient, F is the Faraday’s constant, R is the gas constant, T is the absolute temperature, 4) is the electric potential, and V denotes the gradient. The continuity equation gives 8C’ 7 = ‘V'J’ _ C‘F = d(V-VC +V'('EFV¢))’ (3.2) where V- denotes the divergence. Since the thickness of the PPy layer is much smaller than its length or width, one can assume that, inside the polymer, J, D, E, and other changes are all restricted to the thickness direction (denoted as x direction). This enables one to drop the boldface notation for these variables; in particular, D and E will be used to represent the electric displacement and the electric field (along the x direction). Furthermore, (3.2) can be simplified as aC— azc- _1: ar 34: C‘F 82¢ 71': (79:,2—+RTa—x§;+ira—,2‘)' (33) A pure delay is adopted to capture the oxidation dynamics, i.e., C+(x,t) = C~ (x,t — T0), (3.4) where T0 is the time taken for an anion A’ to react with PPy. Finally, there are two boundary conditions for (3.3): 1) C“ at x = 0 equals the concentration in the PPy/PVDF double layer; 2) The ionic flux J (x,t) at x = hz -— h] (the other boundary of PPy layer) vanishes, assuming no ions leaking outside. Equations (2.35), (2.36), (3.3), (3.4), and the boundary conditions form a complete description of the ionic dynamics inside PPy. 47 3.1.2 Perturbation Analysis a¢ 32¢ From (2.35) and (2.36), one can express 5; and 3x? as 39’ _ F x + — £9 79.; _ ’6 O (c _c we: + ax (0,1) (3.5) 8% F —— = —— CL -C' . 3.6 ax, x.‘ > < > Supposing that the double layer charging process is relatively fast comparing to the ionic dynamics in PPy (which is usually the case), the net charge inside the double layer is zero. Further assume that the electrolyte in PVDF has good ionic conductivity, and thus the electric field within is negligible. Under these assumptions, one can apply Gauss‘s law to a small volume containing the double layer and conclude @(OJ) 2 O in (3.5). Plugging 3x (3.5), (3.6), and (3.4) into (3.3), one obtains aC— 82C“ F2 _ _ _ . 7 = (3x2 —RTKeC (C (x,t—T0)—C (x,t)) F2 3C‘ x __ _ —m—5;/O+me< e )0 ales): < 1 Plugging (3.12) into (3.16), one gets se*5(5)<"2 41)) = (s —dB2(s))(a1(s) — a21s». (3.17) Combining (3.15) and (3.17), one can solve for on and a2: —l3(3)(h2"h1)_( —dB2( )) se s s _ (11(5) = A(S) Cdl(s), (3.18) B(S)(h2-h1)_ _ 2 012(3) = 5" A(s) (3 ““6 (mean). (3.19) with A(s) defined as A(s) = s (eWW'z “h1)+e"[3(~‘)(h2 41)) -—2(s—dfi2(s)). (3.20) 3.1.3 Impedance Model Fig. 3.1 shows the equivalent circuit for the proposed impedance model. The total current 1 consists of three components, the double layer charging current 1C, the diffusion current ID, and the migration current 1M. In the figure C denotes the double layer capacitance, R1 is the ohmic resistance of the oxidized PPy layer, and R2 denotes the resistance of the SO reduced PPy layer, the PVDF (electrolyte) layer, and the contacts. ZD and ZM represent the diffusion dynamics and the migration dynamics, respectively. Each of the current com- ponents is discussed next. We will be concerned with the voltage/current associated with the charge concentration perturbation 8C1— . Figure 3. 1: Equivalent circuit for the impedance model. Double layer charging current 15. Double-layer charges are established at the PPy/PVDF interface in response to an applied voltage, which can be captured equivalently through a double layer capacitance C. The total stored charge in the double layer is EFA5Cd1(s), where A and 6 denote the area of PPy layer and the thickness of the double layer, respec- tively. The time derivative of the stored charge gives IC: [C(S) = EFAast[(S). (3.21) Difi‘usion current ID. The diffusion flux at the PPy/PVDF interface (i.e., x = 0) is given 8C by the Fick’s law and equals —-d£—l—(0,s). But an electron is not released until the 8x 51 oxidation takes place after TO delay. 10 is thus expressed as 10(5) = —£FAde_ST0§a£x1:(O,s) = —eFAde—ST0/3(s)(al(s)—a2(s)) (3.22) 5 -8T 3 [3(5)(h -h )_e—fi(5)(h -h) where the second equality follows from (3. 12). Migration current 1M. [M is induced by the net charges Qp inside the PPy layer. Ap- plying the integral forrn of Gauss’s law to the PPy layer and recalling E (0,s) = 0, one can obtain 8 Qp(s) 2‘ AD(h2 —hl,S) 214195012 —h1,5) = _AK€£(h2 _hl’S) _ h2—111 = ——£FA(l——e 5T0)/0 cl-(§,s)d§ (3.24) = —£FA(1 —e—ST0) ,6 where (3.24) follows from (3.5) and (3.25) follows from (3.12). [M is then expressed as 114(8) = SQP(S) = —8FAS(1 —e_ST0)alefi(h2 —hl) _ (be—[3012 —h1)—(a1 T (”2) [3 £FAds(l — 6—ST0)fi(s)(efi (5)012 - hl) _ e‘5(3)(h2 - ’11))F 2 _ A(s) ”611(5)‘ (3.26) Since the total current is 1(5) =IC(s)+ID(s)+IM(s), (3.27) 52 and the total potential drop (see Fig. 3.1) is [c(S) V(s) 2 SC + R11C(s) + R210), (3.28) the impedance of the trilayer conjugated polymer actuator can be calculated using (3.21), (3.23), and (3.26): 5A(S)(1 +sR1C) SC (6M3) “”3 ‘5)(23—ST0 — W" (W2 ‘ ”1) —e-B(s>

<10-4 3.17 ><10—4 R1 (52) 2 x 10—5 8.3 x 10-6 Experimentally measured impedance spectra have been used to verify the proposed redox level-dependent impedance model. For comparison purposes, the diffusive-elastic- metal model [16] has also been used. Besides the parameters estimated in the previous subsection, the only free parameter for the diffusive-elastic-metal model is the diffusion constant d. For each set of experiments (0.35 M or 0.25 M electrolyte), the value of d is estimated by fitting the diffusive-elastic-metal model to the measured impedance spectrum 55 Charges (Columb) i X Sample1,2and3 i —- Simulated 9 05* o Sample4,5and6 ~ ; X’“ --- Simulated I l 1 ’16 0.4L" ~~—-~fk-’/~r ~J;——-— —1 : ,x’6 ' 0.3~~ -, I! ‘ r, 1 ’2’6 I’I o.2—-- -. 7(6--- - .Iz’ 3 /° 0.1 """ “ 7“ f" ......................... _. .I‘T . o r’ 1‘ l 1 1 , 0 0.2 0.4 0.6 0.8 1 1.2 1.4 DC voltage (V) Figure 3.2: The relationship between transferred charges and applied voltage for different samples. Table 3.2: Estimated values for CO and R2. C0(mol/m3) R2152) Sample 1 1000 34.0 Sample 2 1323 46.3 Sample 3 1646 56.7 Sample 4 1000 26.6 Sample 5 1363 30.2 Sample 6 1726 50.3 under a DC voltage of 0 V. The same value will be used for the proposed model as well. The last parameter, T0, for the proposed model, is obtained by fitting the measured impedance spectrum under a DC bias of 0.5 V. Then the measured spectra under DC voltages of 0 V and 1 V are compared to the predictions by the redox level-dependent model, without fur- ther tuning of any parameters, thus serving as independent checks on the proposed model. Note that the predictions by the diffusive-elastic-metal model do not vary with the DC bias. 56 Fig. 3.3 through Fig. 3.5 show the results for Sample 1 through Sample 3. Using the aforementioned methods, the values of d and T0 are identified as: d = 2.5 x 10' 11m2 /s, T0 = 1.9 ns. It is found that, while the two models are comparable at a low redox level, the proposed model shows clear advantage at higher redox levels, as can be seen in Fig. 3.4 and Fig. 3.5. Consistent results are obtained for Sample 4 - Sample 6. Here d and T0 are estimated to be d = 1.5 x 10; 1 1m2 /s, T0 = 4 ns. It is again seen that the proposed model is able to capture the redox-level dependent impedance, as shown in Fig. 3.6 through Fig. 3.8. This work, for the first time, incorporates the effect of redox level into the dynamics of conjugated polymers in an integrative way, which enhances fundamental understanding of conjugated polymers and and facilitates the use of nonlinear control methods for effectively controlling these materials. The proposed method applies to general conjugated polymers as well. -)(- Experiments --- Diffusive-elastic-metal model — Redox level-dependent model J: tn ‘1 é O Magnitude (dB) 9.: -1 _sLfi___ 10" ° 10 9 e. ’51 0 ‘L 3 . o -2or .. U1 . “I S n. _40 1 11‘1 1 11 1 i 1 1 10'1 10° 10“ 10‘ 10‘ 102 Frequency (rad/sec) Figure 3.3: Model predictions versus experimental impedance for Sample 1 (0.35 M elec- trolyte, 0 V DC voltage). 57 “TH—7' ‘ _F‘fii -)t- Experiments 1 --- Diffusive—elastic-metal model 1 i L:— Redox level-dependent model 0. _— q—"T— ___f ”‘t‘éfif h D Magnitude (dB) _1 Cage 33?: Phase (deg) '8 10° 10‘ 102 Frequency (rad/sec) Figure 3.4: Model predictions versus experimental impedance for Sample 2 (0.35 M elec- trolyte, 0.5 V DC voltage). 3.2 Nonlinear Mechanical Model Based on a Swelling Frame- work The complete actuation model structure has been discussed in Chapter 2. Extensive works have been done to understand the dynamic relationship between the amount of transferred ions and the applied potential. Another important aspect is to predict the mechanical output based on the swelling introduced by the transferred ions. The classical beam theory has been applied to model the bending curvature for bilayer PPy actuators [15, 27] and trilayer actuators as shown in Eq. (2.27). However, the model is only valid when the ion movement- induced swelling is small so that the following two assumptions hold: 1) the strain and the stress can be related linearly; 2) the geometry nonlinearities can be neglected. The tensile strength experiments have shown that the strain and stress relationship of PPy films becomes clearly nonlinear as the strain is increased over 4% [29,30]. These results indicate 58 A 45 w] -)(- Experiments % 1 1 1 { --- Dittusive-elastic-metal model V ,1 —- Redox level-dependent model 0 40 . 1 1: ‘ “.1 . .. i = . . on ‘ 1 “a 35 W a: . . . g “ 1 1 3o . . 1. 1 1 10'1 10° 101 1o2 103 10‘ 4 _O i Phase (deg) ”e at °I —. _A c i o _A O b 10 10 Frequency (rad/sec) Figure 3.5: Model predictions versus experimental impedance for Sample 3 (0.35 M elec- trolyte, 1 V DC voltage). that the linear model based on the assumption that the stress varies linearly with the strain becomes invalid as the deformation gets large. Therefore, it is imperative to develop a nonlinear model to predict the actuator performance for applications where large actuation is involved. Large swelling and bending present challenges to the mechanical model based on the linear elasticity theory. Firstly, when the deformation is large, the swelling effect and the bending effect are coupled and thus cannot be superimposed as those in (2.22) and (2.24) because of the geometric nonlinearities. Secondly, the strain-stress relationships of PVDF and PPy films become nonlinear as the strain increases [29,30,63,64]. Thirdly, under large swelling, the symmetrical trilayer arrangement generates a neutral line that is no longer at the center line due to the geometric nonlinearity. Therefore, the nonlinear elasticity theory- based method is introduced in this section to model large bending motions. We use v1 and V2 to denote the swellings of the reduced and oxidized PPy layer, respec- 59 8 NEW—aw. (TE—fitments 'o --- Diffusive-elastic-metal model 1 I 35- , — Redox level-dependent model 1 g , 1 . .1 'E 30- . ~3 9 1 '- E 25 L-A_141111.a; gAw17 -1 0 10 10 Phase (deg) .,1_x_1_1_11 , 1 L+A+A4¢J d °- 1 ° 101 Frequency (rad/sec) Figure 3.6: Model predictions versus experimental impedance for Sample 4 (0.25 M elec- trolyte, O V DC voltage). tively. The swelling is defined as the ratio between the volume after electrical actuation and the original volume. If v,- > 1, the volume is increased; if Vi < 1, the volume is decreased. Assuming that the volume change is proportional to the transferred charge density, one can calculate V1 and v; at the steady state V0 — k2PV0 : V1 V0 l—kzp, (3.31) v +k V v2 : 0V—02p0:1+k2p, (3.32) where V0 represents the original volume. The value of k2 can be determined by experimen- tal data, which has the order of 10”10m3 -C_1 [16]. When one PPy layer is reduced and repels dopants, the other PPy layer is oxidized and accepts dopants. Therefore, the PVDF layer is merely an ion conducting layer, and we assume that its volume does not change. 60 A 40 f a . .. -)t- Experiments g --- Diffusive-elastlo-metal model v 35 _, — Redox level-dependent model 8 :‘..~ ‘::':::: :::::;::: :::"::: 3 .3: : .4 " '1: 1: ' : ':.' 'E 30"" “"r"7"“1“*'r“itr: """" . . ."I‘ U? "z : 11': ':.':':: ' ::::: g ‘1:1 1:1 ll llll t|ll 25 .1 4 11110 1 11111 L 1L1112 11 11 3 10 10 10 10 10 0 A a... . V . ca -20- j g 1 D. 31: .. ‘f"f:'.' :.'.L§.'ill i _30 1 .111111 : 1:11;;1'1 ; 11:11:11 1 -1 0 2 3 10 10 10 10‘ 10 Frequency (rad/sec) Figure 3.7: Model predictions versus experimental impedance for Sample 5 (0.25 M elec- trolyte, 0.5 V DC voltage). 3.2.1 Finite Strain Tensors In linear elasticity the infinitesimal strain tensor is used to capture the structural deforma- tion, where “infinitesimal” implies that the theory is only valid as the displacement gradient is vanishingly small. However, the strain can be significant in many applications. Thus the finite strain tensors are introduced in nonlinear elasticity, which will be adopted below. Consider a trilayer beam that bends because of differential swellings in the two outer layers. The reference configuration is in Cartesian coordinates, while the deformed config- uration is in cylindrical coordinates for the convenience of modeling the bending. This is shown in Fig. 3.9, where x represents the thickness coordinate, y represents the length co- ordinate, and 2 stands for the width coordinate. The axes in the cylindrical coordinates are labeled by r, 9 and g, which represent the radius, azimuth, and width coordinates, respec- tively. Because the length of the beam is usually much larger than the width, it is assumed 61 a 45 TTT -)t- Experiments 1 3 :3.‘ --- lefusive-elastlc-metal model 3 40- 1 — Redox level-dependent model ~ :1 ; 1'1" “ ":': I: l ‘ III I I c 35' - 1 . 1.. 0’ . ‘. . : . . .. g l I'l t l l i l t 1 1 30 .1 10 441111 1 L112 1 ll 3 10 10 1O 10 10 A i’ -1 V 3 m -2 .c a :.:' .:1 .1.....: ':':::." ':. 1. _30 : .:.:..11 . :1:.:.:: 1 :.:-.:::: ; 1:11;; -1 0 1 2 3 10 10 10 1 Frequency (rad/sec) 0 Figure 3.8: Model predictions versus experimental impedance for Sample 6 (0.25 M elec- trolyte, l V DC voltage). that the there are no changes along the width direction 2, which means g = z. The bending angle 9 is considered to be proportional to the position along the y axis in the reference configuration, 9 = ay, (3.33) where a is a proportionality constant that will be determined by the analysis. The bending radius r is assumed to be independent of y and z,which means r = r(x). The orthonormal basis vectors in the Cartesian coordinates are denoted by ex, ey, and ez, and they are denoted by er, e9, and eg in the cylindrical coordinates. By definition, these vectors are 62 represented by |/\ l/\ :11 |/\ .5‘ N l/\ .< P‘ |/\ |/\ N|- .3“ N O r=r(x) 6=ay §=Z Figure 3.9: Illustration of the reference configuration and the deformed configuration. For ease of analysis, the dyadic product operation is denoted by the symbol ®. The meaning of ® is illustrated as follows: , , xm X1y2 X1Y3 (x1 X2 X3 ) ®(>’1 Y2 >3 = sz1 ma W3 - (3'34) «W1 X3>’2 x3>’3 This gives the matrix representations CO CO OO O O O (3.36) O O €9®ey = O O O O O O (3.37) O O O eg®ez =- O 0 fl Suppose that the deformation takes X = (x, y,z)T in the reference configuration to ‘1’ = (r, 6, g)T in the deformed configuration. The deformation gradient is defined as 8‘11 : — : F . F 8X (:2) (1‘? dX, (3 38) where (1‘? : dr-er + rd6e9 +dgeg, dX = dxex+dyey+dzez. 64 The deformation gradient is written as dr F = Eer®ex+rae9®ey+eg®ez dr 3; 0 O = 0 rec O - (3.39) O O 1 The left Cauchy-Green deformation tensor is [65] B = FF’= 12er®er+lge9®e9 +eg®eg 3,2 0 0 = 0 Ag 0 , (140) 0 O 1 where A Ag it Are: (341) r_dx’ 9— ' ' For incompressible materials, the deformation satisfies the constraint that the volume is not changed by the bending after swelling, which means det F = v, (3.42) where V is defined to be the swelling ratio in the different layers, v1, for the reduced PPy layer V = 1, for the PVDF layer - (3-43) v2, for the oxidized PPy layer 65 From (3.81), the above equation can be written as dr v — = — 3.44 dx ar’ ( ) implying that it is a function of r, a, and v. One can integrate (3.84) and obtain the following equation for the reduced PPy layer: x r(x)= \/r%+£-/ h Vldx, —h2 —— WPVDFtrz, a) + 1 . v—1(WPP}l(r2,a,v1) —WPPy1(r1,a,V1))- Similar to (3.62), the following equations for hoop stresses in the PVDF layer and the other 69 PPy layer can be derived, USE/BF = C£VDF+r————ar , (3.65) aPPyZ — oPP"z+L~—awppyz (3 66) “96 ‘ n“ v2 ar ' ° 3.2.3 Equilibrium The force and bending moment at any cross section of the beam are zero at the equilib- rium, which is similar to (2.31) and (2.21) but takes different expressions in the cylindrical coordinates: r r [2 O'rggyld +/ 3oggDFdr+/4aggy2dr=o, (3.67) r1 r2 r3 r r fzo oggylrdr+/ 3 oggDFrdr+/;4 O'Ppyzrdr20. (3.68) r1 2 To capture the nonlinear strain-stress relationship of the PVDF and PPy as the strain in- creases [29, 30, 63,64], the energy functions for PVDF and PPy layers are assumed to be of neo-Hookean types [66], which means that they take the general form of w = 521—(11 ~3v2/3), (3.69) 70 where it is the shear modulus. It therefore follows from (3.41), (3.84), (3.53), and (3.54) that 2 3 pryl(r,a,V1) = ’HPPy1(a—lv)2+ +(0‘ W121” 31/1/11 (3.70) vamna) = it‘PVDFl‘a‘r) +1ar>2~21 (3.71) 1 v 2 31 WPP)12(rvavV2) = ‘2‘flPPy[(a_i)2+(ar)2+1”3V2/ 1. (3.72) The shear modulus 1.1,- is related to the small strain Young’s modulus Ei by u,- = Ei/ 3, since the constraint (3.44) implies a Poisson ratio of 0.5. It is assumed that the energy functions for PVDF and PPy layers are of neo-Hookean types, because the neo-Hookean model contains the quadratic terms of the mechanical deformation gradient, and it has been adopted to model the nonlinear strain-stress relationships of different smart materials, such as PVDF [67, 68] and ionic polymer-metal composite (IPMC) [69]. Substituting (3.70), (3.71), and (3.72) into (3.62), (3.65) and (3.66), one obtains V2 PP l UZPP l 1 GM = ——yvlr[a2(32— r2)——(—2+¥)], (3.73) 11 1 l 053” = ———2”VZDF1a2(3r2-r%>—-‘—,(r—2+;§)1+orr1r2, (3.74) 2 PM _ “PPy 232 2 V§1 1 375 099 _ 2V2 [a (r -—r3)—- a2(_ r2 +g)]+0'rrlr3- (- ) From (3.73), (3.74) and (3.75), the left hand side of (3.67) is expressed as /r20_PPy1d + ’3 OPVDFdr+ [’4 (HP/32d ,1‘799 ,2 99 3 “99 r = ;1(WPPy2('4:avV1) " WPPy2(r3, a, V1 )) + Cl'rrlr3 ' (’4 — ’3) +r3(WPVDF(r3ia1vl)“WPVDF(r21a)V1))+Grrlr2 ' (r3 _ r2) “Fez—(pryl (f2, 01, V1) — WPPy1(r1 , (IA/1)). (3.76) 1 71 Substituting (3.70), (3.71), and (3.72) into (3.76), one finds that the force balance (3.67) becomes 2 “PPyV211222HPVDF111222 372—;(3*;§)+a (r4—r3)]+ 2 1&3<:*:)+a(r3—r2)1 4 3 - 3 2 2 1” V l l fl[—1-(— - —) + (120% —r‘17')] = 0. (3.77) 2v 222 lar2 rl Note that (3.77) also ensures the boundary condition 055,204, a, v2) 2 0. From (3.73), (3.74) and (3.75), one can find that the moment balance (3.68) becomes “P133244 V12'1'122V12 2 v1 {20‘ (rz—r1)+(—07) ln6+2(r1—r2)[(arl) +(ar1) 1} 3 v r l 1 +#PVDF{ZO‘2(T§—’g)+(';1)21n%+§(r%—r§)1(a;3)2+(a’2)2l}+ “—Z’Zéazpj — r31) + (3191113 + $63 — r2)[(%)2 + (063)21} = O.(3.78) In summary, the problem is now formulated as solving the nonlinear equations (3.77) and (3.78) simultaneously to obtain the two unknown variables r1 and a. These two variables capture the deformed configuration for given swellings V1 and v2. Note that in the linear elasticity theory-based method, the force balance is automatically satisfied for a symmet- rical trilayer structure. However, (3.77) is required in the nonlinear elasticity theory-based method. Newton’s method is applied to numerically solve (3.77) and (3.78) with the aid of (3.49)—(3.51). Thus a large bending strain model is established for the trilayer conjugated polymer actuator with actuation voltage as the input and bending radius as the output. 3.2.4 Experimental Verification The PPy layer is 30 pm thick, and the PVDF layer is 110 pm thick, which implies that the values of the parameters h] and I12 are 55pm and 85pm, respectively. The electrolyte used 72 is 0.05 M tetrabutylammonium hexafluorophosphate (TBA+PF'6”) in the solvent propylene carbonate (PC). Two actuators of different dimensions are cut from the fabricated sheet. The size of two samples are 12.8 x 5.0 mm2 (Sample 1) and 31.2 x 6.0 mm2 (Sample 2) respectively. Each actuator is stored in the electrolyte before experiments for five hours to ensure that the PVDF layer is sufficiently soaked with electrolyte. The relationships between the input voltages and the transferred charges for the two different samples are shown in Fig. 3.11. The constant k1 in (2.15) can be determined from the experimental data, which is found to be 0.0936 C / V and 0.1801 C / V, respectively, for Sample 1 and 2. The value of the swelling-to-charge ratio K2 is obtained through fine tuning based on the reported value in [16], which results in k2 = 1.12 x 10—10 m3 -C—l. The small strain Young’s moduli of PPy and PVDF are taken to be 60 MPa and 612 MPa [13], respectively. 0.35 1 O ISample1 03-... .- .. . .. _ x Samplez 4L 0.25 ........ ______ I I I I 1 1 1 1 1 1 1 , 1 1 I 1 1 1 I l 1 I l 1 l 1 1 1 l 1 t 1 1 I I I I 1 1 I d "I -------------------------------------------- Charges (C) 0:5 1 1:5 Input voltage (V) Figure 3.11: The relationships between the input voltages and the transferred charges for the two different samples. The predictions of r1 and a on the basis of (3.77) and (3.78) are shown in Fig. 3.12 — Fig. 3.15 for the two samples. The experimental data for the bending radii are also 73 compared with the predictions of r1 in Fig. 3.12 and Fig. 3.13 for the two samples. The predictions by the linear model (2.29) are also shown in the figures for comparison pur- poses. Note that the same parameters are used in the linear model. It can be seen that when the applied voltage is low, (i.e., the volume changes of the PPy layers are small), the predictions by the linear and nonlinear models can both fit the experimental data. But as the applied voltage increases, the prediction of the linear model deviates significantly from the experimental data. On the other hand, the nonlinear model still shows good fitting, thus demonstrating the effectiveness of the nonlinear model over a larger deformation range. 0.35 ; -e- Experimental 0.3- | — NonlinearModel A r --- Linear Model 1 P N or P [0 Bending radius (m 0 b P -t a or O 01 l 0.5 1 1:5 Input voltage (V) C Figure 3.12: Quasi-static bending under different actuation voltages for Sample 1 (13x5 mm). Based on the obtained values of r1 and a, one can calculate r2, r3 and r4 by using (3.49), (3.50) and (3.51). Furthermore the thickness of PPy layers can be obtained by calculating r2 — r1 and r4 — 13 at different actuation voltages. Similarly the thickness of PVDF layer is obtained through r3 -— r2. The results for Sample 1 are shown in Fig. 3.16 and Fig. 3.17. When the voltage increases, the thickness of the reduced PPy layer will decrease and that of the oxidized PPy layer will increase for the oxidized PPy layer. The 74 0.7 l -9- Experimental 0.6 ~~~~~~~ ~ -~ 1-: — Nonlinear Model J --- Linear Model m) v0.51-—- 0.41-- 0.31- 0.2-~44 Bending radius 0'5Voltage (V)1 Figure 3.13: Quasi-static bending under different actuation voltages for Sample 2 (33x6 mm). 4o , , 1 . 35~- AAAAA 30.-- , , ....... W- ,,,,,,,,,,,,,,, a. (rad/ m) N 9 15~~- 11111111111111111111 10» 1 5» . 0o o. 2 Voltaige (V) Figure 3.14: Computational results on the change of a versus the applied voltage for Sam- ple 1 (13x5 mm). 75 25 l V fl T l 1 8.2 0:4 0:6 0:8 1 1:2 114 1.6 Voltage (V) Figure 3.1.5: Computational results on the change of 01 versus the applied voltage for Sam- ple 2 (33x6 mm). change percentage at 2 V is -0.83% for the reduced PPy layer and 0.67% for the oxidized PPy layer. The thickness of the PVDF layer is decreasing but at a slower rate when the input voltage is increasing. The change percentage at 2 V is 0.21%. Note that these results pro- vide interesting insight into the deformed configuration. Such information is not available from linear models, since the latter assume fixed thickness throughout the deformation. 3.3 Fiber-directed Conjugated Polymer Torsional Actua- tor Different configurations of conjugated polymer actuators have been exploited. Bilayer and trilayer benders are investigated in [8, 32], which utilize the strain difference in different layers of a beam to generate bending motion. Linear extenders are investigated in [7], where the conjugated polymer is immersed in electrolyte and the volume is changed as 76 30.2 — Reduced PPy layer ,2” 30.1 ' 5 so- 3 10 .5 _2 29.9 .1: l- 29.8~--- 29'70 0 1.5 2 Volta19e (V) Figure 3.16: Computational results on the changes of thickness of the PPy layers with the applied voltage for Sample 1. 110 1 — PVDF layer 1 109.95“: (11m) d in . l | 1 . . . I . l . . 1 . l l | (I! m 35’ £109.05 ..................... - . .C .- 109.8 ‘ ,— ______ f .............. L ______________ 109.7 1 i 1 50 0.5 1 1.5 2 Voltage (V) Figure 3.17: Computational results on the changes of thickness of the PVDF layer with the applied voltage for Sample 1. ' ion transfer between the polymer and electrolyte takes place under an applied potential. However, more complicated actuator motions will be required. For example, torsional 77 motion is highly desirable in microsurgery and other robotic applications. Fiber-reinforced elastic material can generate complicated deformation [70—72]. The anisotropy associated with the interaction between the fibers and the material matrix makes the material resist elongation in certain directions, which generates torsional motion when the volume of the material changes [33]. This principle provides a potentially useful approach to fabricating compact torsional actuators using conjugated polymer. Also conjugated polymer is an ideal material to realize the design of fiber reinforced elastic material actuator in practice, since the volume of the polymer can be changed easily with electrical actuation. A conjugated polymer tube was fabricated by Ding et al. [73], where platinum wires were integrated into conjugated polymer during fabrication process. However, the purpose for the platinum wires in that paper was to increase the electrical conductivity along the tube, so the torsional motion introduced by the fiber-directed structure was not investigated. To the best of the author’s knowledge, there has been no detailed investigation on torsional motion in a fiber-reinforced conjugated polymer actuators. In this dissertation, torsional motion of a fiber-directed conjugated polymer actuator is investigated from both theoretical and experimental perspectives. A nonlinear elasticity- based model is proposed to capture the relationship between the actuation voltage and the torsional motion of the conjugated polymer tube. Compared with a linear elasticity-based model, this model is able to capture the large deformations. The model is verified with experimental measurements made on a platinum fiber-directed conjugated polymer tube. Fiber-reinforced elastic materials can generate complicated deformation when subject to swelling. When the fibers are aligned in special ways, the material matrix containing fibers will expand in the direction perpendicular to the fiber [33]. A specific case is studied where the fiber is helically wound about a elastic tube with a fixed pitch angle, which is illustrated in Fig. 3.18. The elastic material we consider here is PPy, the volume of which can be changed through electrical activation. Theoretical analysis shows that the tube will generate torsional motion when subject to swelling. Platinum wires are chosen to 78 be integrated into PPy tube during the fabrication with some pitch angle [3. ‘\\\\\\\\\\\\\\\\\‘\\\\‘\\V‘ \\\\\\\\\\\\\\\\\\\\\\\\\\\\ \\\\\\\l\\\\\\\\\\\\\\\\\\\\ 1111 \l \ \\\\ PPy Platinum ‘~_ _— lll III/ I III III/III Ill/llllI/II/l/I/ll/lll/l / 1/1/1[III/[11111111111111 (a) (b) Figure 3.18: (a) Illustration of the conjugated polymer-based torsional actuator; (b) picture of the fabricated torsional actuator. To fabricate the actuator, it is crucial to integrate the fibers into the material matrix of PPy. This would ensure that, when the PPy tube volume changes, the torsional motion will be generated when the motion in other directions are constrained. For this purpose, firstly a glass cylinder with a diameter of 2.5 mm is used as the substrate and coated uniformly with gold through sputtering. Then a platinum wire with diameter of 25 pm is wound uniformly along the glass cylinder. The pitch angle [3 is 80°. This tube is soaked in the electrolyte (0.05M Li+TFSI" in Propylene Carbonate) and connected to the working electrode of a potentiostat (OMNIlOl from Cypress Systems). An electrochemical deposition process is conducted by following the method in [13]. To maintain a constant potential on PPy when PPy grows on the surface of glass tube, an Ag/AgCl reference electrode (Aldrich Chemical) is used. The auxiliary electrode is a stainless steel mesh. The electrolyte used in experiment is a mixed solution of 0.1M pyrrole, 0.1M Li+TFSI— in Propylene Carbonate 79 with 0.5 w/w% water. The current density is maintained at 2 mA/cm2 for good uniformity. Different PPy tube wall thickness can be achieved by controlling the deposition time. The PPy wall integrates the platinum wire during the deposition. Then because the sputtered gold layer has poor adhesion force with the PPy inner surface, the PPy tube can be easily taken off from the glass cylinder. The inner diameter of the tube can be controlled by choosing different glass cylinder diameters. 1.Reference electrode 1 2 3 ~ 2. Working [UUHMI electrode ’ 3.Auxiliary electrode Figure 3.19: The experimental setup to fabricate the fiber-directed conjugated polymer actuator. 3.3.1 Nonlinear Mechanical Modeling Framework The original and deformed configurations are shown in Fig. 3.20, where F represents the deformation gradient, which can map vectors (expressed in Z, R, 6) domain) in the original configuration to vectors (expressed in z, r, 9 domain) in the deformed configuration. The 80 pitch angle is defined as {3 (0 < [i < 90") [33]. aI’d/IIAIIIlax/Adzaa11144Jdaa aaaaz’a’aazzazz’zzazaazizlzaa "a’aa””p’I—I’I”a,””a"a Z.“ . Kr": ’ ’ ’ ’ ’ ' —-L__.—O’ I r._— r-—’ ‘r~_‘___.—’ ‘J-__.—-’ ‘L ~L I r~~ ‘8- ”~L 1 r ‘ x\\\s\\\\\\\s\\\~\ss\\s~\x\x~ xxxxxsxxxxx\\\\\\\\\\\\\\\\xx \\\\\\\\\\\\\\\\\\\\\\\\\\\\\ KNKNNNNKNNKN\\|\1\NL£RN\1N\K\\\ \_ F aaaaaadaad AdadaaaaaaaAAaaaa azazzzzzIII/Izzzzzzzzza/z’za aaaa’z”’ aaaaaaa’I’aaiaiaal ‘ ‘ .0 ‘- -’ -~‘~ ‘ ‘ ~ ‘ Figure 3.20: Illustration of the actuator configuration. Left: original configuration; right: deformed configuration. The deformation can be characterized by the following equation: (3.79) =®+KZ, 2:212, r: r(R), 9 where K is the twist per unit length, and AZ is the elongation per unit length in the longitu- dinal direction. Suppose that the deformation takes a particle at location X in the reference 81 configuration to the location x in the deformed configuration. The deformation gradient is defined as F = — (3.80) where 3X 2 (1R CR +Rd®e® +dZ eZ, 8x = drer+rd9e9+dzez, which are the orthonormal unit vectors in the original configuration, and the deformed configuration. The deformation gradient tensor is r F— dre ®e + (1]? e9 ®e® + 10139 ®ez + Azez 1862, (3.81) where 18> is the dyadic product. We assume that PPy is mechanically incompressible both before and after swelling. This assumption is based on relevant literature [74]. So the deformation satisfies the constraint that the volume is not changed by the deformation after swelling, which means det F = v. (3.82) This, together with (3.81), implies l d E32717: = v. (3.83) Rewriting (3.83) as rdr = {ERdR and integrating, we get 2 R r2 = r? + 1.2. [R deR, (3.84) z where ’1' and ,Rl- are the inner radii of the tube in the original configuration and the deformed 82 configuration, respectively. The left Cauchy-Green deformation tensor is defined by B = FFT. (3.85) Following (3.81) and (3.83), B can be further expressed as v2R2 1 2 B — Wer®er+(K2+—R—2)r €9®89 +x'lzr(e9 ®ez+ez®e9) +1122ez®ez. (3.86) The principal invariants of B are defined as 11 = tr(B), 12 = I3tr(B—1), 13 = det(B). (3.87) From (3.81), the invariant 11 is obtained as 22 2 R v2 +-r—+K2r2+lzz. (3.88) 2 I'lzr I: l R2 The elastic energy function of PPy can be denoted by (Pm and modeled in the neo-Hookean form as l Chm = Ella] — 3v2/3), (3.89) where u is a material elastic constant that can be taken as Eppy / 3, where Eppy is the Young’s modulus [48]. Let M be the unit vector in the reference configuration that defines the fiber direction in the material matrix: M = sinliee +cosBeZ, (3.90) where [i is the pitch angle defined in Fig. 3.20. Given the deformation gradient F, the unit 83 vector M is mapped into m=FM= (Kcosfi+%sinfi)ye6+zizcosliez. (3.91) The square of the stretch of the fiber under actuation is 14 = tr(m®m) = (KCOSfi+%Sinfi)2Y2+XZ-2C082fl. (3.92) The strain energy function of the fiber can be modeled as crnf = $1414 —1)2, (3.93) where 7 is another material constant that may be taken as proportional to E pl atinum' The proportionality constant will depend on the volume fraction of fibers in the overall compos- ite. Using the connection between the nonlinear theory and the linear theory of anisotropic elasticity [75], we may take 7: SE platinum in this investigation. Therefore, the total energy of the fiber-reinforced PPy actuator is W =m+f. (3.94) The Cauchy stress tensor is [33] _1 8W —FT — I. 3.95 Considering the expression of total energy in (3.94), the Cauchy stress can be written as [331 =-——B+——m®m—pl. (3.96) 84 Considering (3.86) and (3.91), we can express (3.96) as 0' = arrer®er +099e9 ®e9 + 062(96 ®ez +ez ®e9) +O'zzez 81 e3, (3.97) where 2 (NR 0 = _p+ , (3.98) rr Azzrz o - — +“r2(x‘2+-—1—) 99 — P v R2 2 +3§L(Kcosa+%smm2(z4—1), (3.99) 2 022 = _p+%Azz+—VZAZZ(I4—l)coszfi, (3.100) 092 = gxlzr+gglzrcosmkcosfi 1 +§sinfi)(I4—l). (3-101) This framework is built on strain energy functions and general deformation gradients, and it is thus valid for both small and large deformations. 3.3.2 Boundary Conditions The boundary conditions on the conjugated polymer actuator are specified as follows: 0rr1R=Ri =01 0rr1R_—_RO =01 (3.102) which represent that there is no direct loading applied to the inner and outer lateral surfaces. Notice that R is used in (3.102) since r is also a function of R. The equilibrium equation 85 diva = 0 in the directions of r, z, and 6 can be written as follows: 80'" ladre aGrz 1 _ _ _ _ = 3.103 3r +r 36 37. +r(0'rr 096) 0’ ( ) 892.19% 804 91.2- 8r r 80 (92 r _ 80,9 1 c9099 3026 20,9 Tit—:99— a. + 1 —°’ (“05’ 0, (3.104) where are = 0, an = 0 in this case. Considering (3.99), (3.100), and (3.101), one can show that (3.104) and (3.105) reduce to ’89; = 0 and 3%) = 0, which means that p is a function of r only. Given are = 0 and orz = 0, one can write (3.103) as follows by using (3.83): dO'rr _ VR which can be furthermore expanded by considering (3.98) uvd R2 dp _ VR __ __ 222511892). (3.107) d—R; — EZ'TZ'(O'rp-— (799)-1‘ Integrating (3.107) after substituting the expression of (3.98) minus (3.99), one can obtain the expression of p by using the boundary condition Orrl R = R0 = 0: vRZp R0 v2R3 R(R2K2+l) —Zfl(xcosfi + lsirz[3)2(14 — 1))dR. (3-108) AZ R Applying the other boundary condition Orrl R : Ri = 0 in (3.108) gives the following con- straint equation: A20, (3mm 86 where A: — —— — fRi (2( lzzr4+x +R2) +(xcosp + %sin[3)2'y(l4 —1))RdR. (3.110) 3.3.3 Nonlinear Mechanical Model One can see from Fig. 3.20 that three variables quantify the deformed configuration: r, 112, and K. Because the profile of r can be calculated based on (3.84), variable r can be replaced by ’1' to characterize the deformation together with the other two variables. Eq. (3.110) has provided one constraint equation. The other two require knowledge of the moment T and axial load P, which are given by the following expressions: r0 2 T = 27r/ O'QZr dr, (3.111) ri , r0 P = 27r/ri ourdr. (3.112) When the end of the tube is not constrained, it follows that T = 0 and P = 0. These two equations, along with (3.109), provide three nonlinear equations for ri, K, and 42- One finds that: the expression (3.110) is strongly dependent on ri, the expression (3.111) for T is strongly dependent on K, and the expression (3.112) for P is strongly dependent on AZ. In general, the numerical analysis to obtain these variables by solving three nonlinear equations is difficult. Therefore in this thesis, AZ, representing the length change ratio in the axial direction, will be measured and used as a given parameter. Since A and T give the strongest dependence on ’i and K, the equations A = 0 and T = 0 are retained in the analysis and numerically solved to obtain ’1' and K. By substituting from (3.101), one can 87 integrate (3.111) and obtain the following expression: TzuaK+y(b0+b1K+b2K2+b3K3)=0, (3.113) where a, b0, b1, b2 and b3 are terms involving material constants. Similarly one can integrate (3.110) and obtain the expression for A as follows: A=uQ+yF, (3.114) where Q and F are also terms involving material properties and Az, K, and ri, which will not be listed in this thesis due to the page limitation. The two unknown variables ri and K can be obtained by numerically solving the coupled equations A = 0 and T = 0 using (3.113) and (3.114). 3.3.4 Sample Preparation and Experimental Setup Three factors are changed during fabrication to study their influence on the performance of the actuators, which are the PPy thickness, the pitch angle, and the inner radius. Two PPy thicknesses, 0.381 mm and 0.686 mm, are obtained using different deposition duration. Two different pitch angles are obtained during the winding process, which are 60 ° and 80 0. Two glass tubes with different radii are used to change the inner radii of the tube, which are 2.2 and 1.3 mm. Four samples are fabricated that are named Sample 1, 2, 3, and 4. Their parameters are listed in Table 3.3. The length is kept as 33.5 mm for all four samples. A computer equipped with a DSl 104 R&D Controller Board (dSPACE Inc) is used for data acquisition and processing. A voltage is applied across the PPy tube and the counter electrode, which is a stainless steel mesh in the experiments. The PPy tube is soaked in 0.1M Li+TFSI‘ in Propylene Carbonate (PC). Sticky copper film is attached to different 88 Table 3.3: Geometric parameters of the samples. thickness (mm) pitch angles (degree) radius (mm) Sample 1 0.381 80 1.3 Sample 2 0.381 80 2.2 Sample 3 0.686 80 2.2 Sample 4 0.686 60 2.2 positions of the PPy tube for measurement purposes. Three motion signals are captured to characterize the tube deformation, which are the inner radius ri, torsional displacement )1 at the end of the tube as shown in Fig. 3.21, and the length change at the end of the tube that equals to A; - L based on (3.79), where L is the length of the PPy tube. The model-predicted variables ri and 42 can be correlated directly to the measured variables, while K is obtained based on the geometric relationship in Fig. 3.21: y = Rmtan(K-L), where y is the measured displacement, and Rm is 4 mm in this thesis. Figure 3.21: Experimental setup to measure K. The copper film is attached perpendicular to the outer surface at the tube bottom. 89 Laser sensor 3.3.5 Experimental Results and Discussions The quasi-static response is studied by applying a low frequency sinusoidal voltage on the actuator. The frequency is chosen to be 0.005 Hz to allow the actuator to reach the quasi- static condition. The sinusoidal voltage amplitude is 3.6 V. Because only one laser sensor is available, measurements of inner radius, torsional displacement, and length change are per- formed separately and then synchronized through the voltage signal. The Young’s moduli of platinum and PPy are chosen to be 168 GPa and 80 MPa in model prediction [48,76]. The experimental results for Sample 1, 2, 3, and 4 are listed as follows. It is observed in experiments that the inner radius and torsional displacement signals have reversed phases. In particular, when the inner radius decreases, the torsional displacement increases. As discussed in Sec. 5.2, the length change will be used to obtain 42 and taken as a given parameter to reduce the difficulties in numerical analysis. Therefore the model-predicted inner radius and torsional displacement are compared with the experimental data for dif- ferent samples. In general, the model can predict the reversed phase of these two signals. The model can also predict the response change from different samples. The comparison between Fig. 3.22 and Fig. 3.25 shows the trend that a larger tube radius leads to a larger torsional displacement, when other parameters are kept the same. The comparison between Fig. 3.25 and Fig. 3.28 shows the trend that a thinner tube generates a larger torsional dis- placement. The comparison between Fig. 3.28 and Fig. 3.30 shows the trend that a smaller winding angle results in a larger torsional displacement. Therefore the experiments and model prediction suggest that a configuration with larger. radius, thinner tube thickness, and smaller winding angle tend to generate large torsional motion. 3.4 Chapter Summary In the electrochemical domain, a redox level-dependent impedance model was proposed for conjugated polymer actuators. The model was derived based on a nonlinear partial 9O Voltage (V) '30 50 100 150 — Experimental data °-°‘ ~ -u- - Model prediction 0 50 100 150 Time (s) Torsional displacement (mm) Figure 3.22: The torsional displacement of Sample 1 with 0.005 Hz sinusoidal voltage input. differential equation that incorporates the dynamics of ion diffusion, ion migration, and polymer oxidation, which was linearized using perturbation techniques. The linearized PDE was solved analytically in the Laplace domain, with proper boundary conditions en- forced. Comparison with experimental results supported that the proposed model is able to capture the influence of redox level on the impedance spectrum. The model provides an effective way to integrate the diffusion and migration effects of ion flux in conjugated polymers, which used to be treated separately with different models [57, 77]. The pro- posed model not only contributes to fundamental understanding of the complicated, redox level-dependent electrochemical behavior, but also holds potential for nonlinear control of conjugated polymer actuators. In the mechanical domain, a nonlinear elasticity theory-based framework is proposed 91 Voltage (V) o n is via 1 '19 A N f F—J__—-_fi E ‘ — Experlmental data g 1 £24- Model prediction g 1:775 """""""" '__,,,, ————— »_ if???“ ,_ 7 E ._ 1.77 o 1: 5 1.7651 1 50 100 150 Time (s) Figure 3.23: The change of inner radius of Sample 1 with 0.005 Hz sinusoidal voltage input. to analyze the bending configuration of a trilayer conjugated polymer beam under different actuation voltages. Neo-Hookean type strain energy functions are used for the PPy and PVDF layers of the trilayer beam to capture the nonlinearity under large deformations. For a constant actuation voltage, the swelling of each PPy layer is determined by the amount of transferred ions. The bending configuration at the equilibrium is obtained by solving the force and moment balance equations simultaneously. Experimental results have validated the effectiveness of the proposed nonlinear model. The method further provides insight into the deformation details, such as the change of layer thickness. Similar analysis can be used for some other EAPs that involve the local volume change (swelling). This is important for many applications involving large deformations, when the elastic modulus of the material already becomes nonlinear and linear elasticity analysis does not hold. A fiber-directed conjugated polymer actuator is fabricated by integrating platinum wires 92 Voltage (V) A 0.04 F j E E. O 8’ N .c 0 ‘9 Ci . .1 '0'020 50 100 150 Time (s) Figure 3.24: The change of tube length of Sample 1 with 0.005 Hz sinusoidal voltage input. into conjugated polymer during fabrication. Experiments have demonstrated that the tube- shaped conjugated polymer actuator can generate torsional deformation, when its volume is changed by applying an actuation voltage across itself and the electrolyte. Three impor- tant factors are changed during the fabrication process to make conjugated polymer tubes with different configurations, including the tube thickness, the tube radius, and the pitch angle of the platinum wires on the tube. A nonlinear elasticity-based model is utilized to capture the actuator performance, which holds for both small and large deformations. The model prediction is compared with the experimental data, which has verified the effective- ness of the nonlinear elasticity-based model. The performance of actuators with different configurations is compared, which suggests that a configuration with larger radius, thinner tube thickness, and smaller wounding angles will generate larger torsional motion. 93 Voltage (V) P c: a: — Experimental data - -t(- Model prediction 1 _O O .5 b b O A I I l 47:- _.?. 1‘? J; 'o 100 150 Torsional displacement (mm) ON 0 0| C Time (s) Figure 3.25: The torsional displacement of Sample 2 with 0.005 Hz sinusoidal voltage input 94 Voltage (V) 0 50 100 150 — Experimental data ’é‘ 2-2 -x- Model prediction g : m2.19~-----— —~ : 3 ,5 2 ._ 2.18 2 E. . . 2.17 1 0 100 150 Time (s) Figure 3.26: The change of inner radius of Sample 2 with 0.005 H2 sinusoidal voltage input. 95 Voltage (V) Length change (mm) :5 :5 -s o O” 50 ’ 100 150 Time (s) Figure 3.27: The change of tube length of Sample 2 with 0.005 H2 sinusoidal voltage input. 96 Voltage (V) .50 50 100 150 200 x 10'3 — Experimental data 5 r i -x- Model prediction 0 50 100 150 200 Time (s) Torsional displacement (mm) Figure 3.28: The torsional displacement of Sample 3 with 0.005 H2 sinusoidal voltage input. 97 Voltage (V) O N P O _L J o o o 01 l i .5 O O 0| ___________________________ Length change (mm) 9 I D d 0 50 100 150 200 Time (s) Figure 3.29: The change of tube length of Sample 3 with 0.005 Hz sinusoidal voltage input. 98 A 2 — 2, 3’ o :3 3 > .2 - -4 i . i 0 so 100 150 200 E -——Experimentaldata E 0.01 I. . -it- Model prediction — C ‘ ' l I 20001 _ , , ‘m 8-”. § L “i x w. i - l1 — o .......... ' , -- __ ....... Q ”'9 ik‘ ‘11,.) I I l : _.o_oos .............. . ..... «I L- _ g g I : "' -o.01 i 1 1 g 0 so 100 150 200 '- Tlme(s) Figure 3.30: The torsional displacement of Sample 4 with 0.005 Hz sinusoidal voltage input. 99 Voltage (V) O -4 i L e 0 50 100 150 200 2.2 i . —— Experimental data F E . ' «- Model prediction E l l 7; 2.1 ‘ 3 '6 E h 2.18 0 C E i l 2.17 . 1 . 0 50 100 150 200 Time (s) Figure 3.31: The change of inner radius of Sample 4 with 0.005 Hz sinusoidal voltage input. 100 A Voltage (V) O M K: 50 1 00 1 50 200 _L O 0| 7... ___.--_..._—..-__._..4 Length change (mm) O 50 100 150 200 Time (s) Figure 3.32: The change of tube length of Sample 4 with 0.005 Hz sinusoidal voltage input. 101 Chapter 4 Control of Conjugated Polymer Actuators 4.1 Model Reduction The full actuation model (2.30) can be rewritten as with the constant Yb") _ Cm C(1+ Eggtanhflhz ”’11) 3)) 319L[(%:13)2—1] E . 8’11(’12 *h1)W[(7}:-%)3 + 1%; -1] |li> Cm Note that the material dynamics (2.26) is ignored because it is only significant when conju- gated polymer is operated at high frequencies. High frequency response becomes too weak to be useful in applications [78]. This can be observed from Fig. 2.16 and 2.17, where the displacement magnitude drops over 20 dB as frequency increases from 0.3 to 30 rad/sec. Eq. (4.1) is an infinite-dimensional system due to the term tanh((h2 — h1)\/§), and 102 thus is not suitable for real—time control purposes. Utilizing the equality '[16] 1 s V tah_ _ m “(M/z) 1 4x/s_z —n=Os+7r2(2n+ 1)2z and letting z = , one converts (4.1) into 4M2*h02 y(S) : Cm . (4.2) V(S) 1 sR + 2d 2 1 ) C 1 + °°_ ( 30124115 n"05+7r2(2n+l)2d(2(h2—h1))_2 To understand the rationale behind model reduction, take typical parameters d = 2 x 10" 10 l m2/s andhz—hl =30um. Thepole of is at—13.7 for s+ n2(2n+1)2d(2(h2 — ’21))“2 n = 2, at —26.9 for n = 3, and at —44.4 for n = 4, etc. This indicates that one can obtain a low-order approximation to (4.2) by ignoring terms associated with large n. In particular, for a low-frequency input, the approximation will preserve well the behavior of (4.2). We thus discard terms with n 2 2. This results in the following third-order system for the actuator: 2 I I I V(S) s3+a’lsz+a’25+a’3’ 103 where the parameters all have explicit physical meanings: a, _ 40 + 52:20 + 1 _ ' (hz—h1)6 2022—1192 RC’ , 57:202 97:402 57:20 (12 : 35+ 4+ 2 ’ (’12—’11) 16(h2—h1) 2(’12-’11) RC , 971:402 a = l 3 16(h2 — h1)4RC I Cm bl = 7? b, _ 400,, + 57300,, 2 (hz-h1)5R 2(h2—h1)2R’ b3 __ SnzDsz 97:402Cm + . (I12 -—hl)35R 16(h2 —h1)4R For typical parameters, (4.3) has one pole and one zero which are located far to the left of the imaginary axis comparing to other poles and zeros, and therefore the model can be further reduced to second order with one zero. This is shown next. The system (4.3) has three poles with explicit expressions [79]: 9 1, 2 —Qcos(§)—§a1 (4-4) 9+2 1, p2 2\/—Qcos( 3 n)—§a1, (4.5) 9+47r 1 p3 = 2\/-QCOS( 3 )—§a’1. (4.6) Pl where P _Q3 = 30’2—0’12 9 3 I I I I3 P : 9ala2 — 27a3 —— 2al 54 6=c05_1( ), i 104 With typical physical parameters [16], a’l , agz, and a’3 are all relatively large numbers (> > 1). This implies 4’12 >> 3a’2 in Q, and 20,13 > > 9a'l (1:2 and 2a? >> 27a'3 in P, which leadsto Q ~ ill—2- P~ 0,13 9‘ “"F’ M e if"— 3, 6 z cos 1— £1,13/27)=cos l(—-l)--7r a’ (1’ p1 z 2--3—l-cos(§)—-3—l=0, a' 3 a' P2 * 2'?1'°°S(—375)“3"=‘“1’ al 571' a! p3 z 2".T1COS(-3—‘)‘——3-l-=O Clearly I p2| is very large while [ml and |p3| are relatively small. The analysis on the zeros is simpler. The zeros of (4.3) are I I2 I I _b2‘\/b2 ‘4'b1'b3 z] 2b’1 , (4.7) 22 : —b’2 + \/b2:2:’1—4cb’1 -b’3. (4.8) With typical parameters, the following holds: I2 4-b’1- ’3<>b1, 105 which implies I 2 21 z ——I_’ b1 zz z 0 Therefore, |zl| is very large while |22| is relatively small. A numerical example is provided to illustrate the analysis. Table 4.1 lists the typical values for the relevant physical parameters. The corresponding system parameters for (4.3) are: a’l 2.32x103,a'2=9.79x103,a’3=3.38x103, b’l z 0.0667Cm,b’2:71.48Cm, 3:195.1Cm, the three poles are —0.38, —23 19, —3.84, and the two zeros are —- 1069, -2.74. Table 4.1: Typical values of parameters in the actuation model. Parameter Value D 2 x 10’10 mZ/s ’22 ‘hl 30 um R 15 Q 6 25 mm C 5.33 x 10—5 F When operating in air, the trilayer actuator will dry up due to solvent evaporation. This implies that the diffusion coefficient D will decay over time. It is thus of interest to see whether the above analysis on pole/zero locations still holds when D is very small. Fig. 4.1 shows the ratio min(|P2!.|21 I) max(|P1l,|P3iil32l) as a function of D, while other parameters are chosen as in Table 4.]. It is clear that even 106 when D is close to 0, one can safely ignore one pole and one zero of (4.3) [50]. The final reduced model for the trilayer actuator thus has the following structure: (5) 3(5) bls+b2 :_=———. 4.9 (5) A“ s2+a1s+a2 ( ) ‘< < It is expected that despite its simple looking, (4.9) captures the dominant physics of the actuator within the actuation bandwidth. In particular, all parameters of (4.9) can be related to fundamental physical parameters for the full model (2.29). 33m . 320i 310 L 300 _, 290-7 » . 280~ 270-— 16-1‘3'A‘ ‘ A l-12 “ -11 1 . Diffusion coefficient (mZ/s) minilpziiiql) —.— as a function of D. max(ip1iiip3i, izzl) Figure 4.1: Ratio Fig. 4.2 compares the Bode plot of (4.1) and (4.9) using parameters in Table 4.1. The unmodeled dynamics is bounded in general, and the maximum discrepancy is reached around 103 rad/sec. However, considering that the actuator is usually operated in the low frequency range up to a few Hz, the reduced model is a good approximation for control implementation while capturing the main dynamics. 107 '20 r i r E 2""; v’ . 3'30“ ~~~~~~~ ‘-—g-———-— _~ _« m 1 . . .340 """" ’ ' ‘ ” ' -- ~ , a _IMinite-dimensional modeI ___ '3 --'Reduced model 63-50~ ............. .. l6 2 60 10.2 10° 102 10‘ 106 100 j I I -——— .H_.___‘ Phase(deg) 8 10. 10 102 10 10 Frequency (rad/sec) Figure 4.2: Comparisons between the infinite-dimensional and reduced model. 108 4.2 Design of Robust Adaptive Controller 4.2.1 Self-Tuning Regulator Controller adaptation is desirable for conjugated polymers since their actuation behaviors can vary significantly over time. Given the model structure (4.9), a self-tuning regulator is adopted due to its simplicity. The idea is to estimate systems parameters online, and then construct a controller based on these estimates so that the closed-loop system would behave like a model system Gm(s) (model-following). Fig. 4.3 illustrates the major components of a self-tuning regulator. Note that a low-pass filter is used to filter the noises in the output signal before the output is sent to the self-tuning regulator. The bandwidth is chosen to be 30 Hz to cover the actuation frequency range in experiments. Self-tuning regulator r“‘--—f'. ''''''' 1 Specification l ' w v Plant parameters | l Controller 4' Estimation <5 | design | | Controller ______ I | Uparameters 1- Reference I i Conductin . ‘ Controller : : g L_.__,. l V Input polymer Output Figure 4.3: Illustration of the robust self-tuning regulator. In the estimation step, the recursive least-squares algorithm is chosen to identify pa- rameters in (4.9) based on the input into and output from the conducting polymer shown as follows §-=memm (Mm an=im-flman am “2:” = am) —P (4.12) 1.09 where 9 = [01 a‘2 bi 521T d ) %L“‘1va(s>1 L‘11va>1- ‘ A stable transfer function H f is introduced to avoid direct differentiation of the output signal in estimation 1 H = —————————-. (4.13) f 52 + ZS + 1 Note that L"1 [*] means inverse Laplace transform. The desired transfer function is chosen as B b + b C(s) = 1’— : 1ms 2’" . (4.14) Am 52 +a1ms+azm A general linear controller is chosen as RV(s) 2: Tr(s) — Sy(s), (4.15) where r(s) is the reference input. R and S are determined by the Diophantine equation AR+BS=Am (4.16) to track the desired denominator Am, and Bm is tracked by choosing T. Finally the con- troller is obtained as : blms+b2mr(s _ (aim —c’1‘1)s+(a2m _62)y(s). V(S) A ,. ,. ,. (4.17) b1S+b2 bls+b2 In experiments, the desired transfer function parameters are chosen as alm =4,a2m =4,b1m 21,172," 21.5. 110 4.2.2 Parameter Projection If the model (4.9) were exact, the controller (4.17) in combination with the online param- eter identifier would lead to asymptotic model-following [37]. However, higher-frequency dynamics and nonlinearities are not included in (4.9). Measurement noises are not reflected in the model either. These undesirable factors might lead to instability of the closed-loop system if no proper steps are taken. In this thesis parameter projection is adopted as a robustification mechanism for the self-tuning regulator. From the expressions of the parameters in (4.3), one knows that the two poles and the zero of (4.9) are all negative, thus the parameters a) , 02, b1, and b2 should all be positive, which are bounded by a small constant m > 0. Furthermore, based on the given polymer parameters and the knowledge of physical parameters, an upper bound M > 0 of these parameters can be determined. The update rule (4.10) is thus modified to incorporate parameter projection: 0 if 19,-(t) 2M and [P(t)tp(t)e(t)],- >0 t) = 0 iféi(‘)=m and [P(t) 0. In the experiments the values of M and m are chosen to be M = 1x 104, m = 0.001. One can represent the true output y(s) of the actuator under the input V(s) as bs+b y(s)-———-—‘ 2 v — 2 (s)+A(s)V(s)+./V[V(s)]+w(s). (4.19) S +GIS+02 Here A(s) represents the unmodeled higher-frequency dynamics, i.e., the difference be- tween (4.1) and (4.9). Since both (4.1) and (4.9) are stable and have bounded frequency responses, A(s) will be stable and bounded. JV [V(s)] denotes the influence of unmodeled 111 nonlinearity (e.g., hysteresis). The nonlinearities will be bounded due to the dissipative na- ture of the materials. In (4. 19), w(s) denotes the measurement noise, which is also bounded. It can be shown [80, 81] that the proposed adaptive control scheme in Section 4.2.1 with parameter projection (4.18) is robust in the presence of bounded unmodeled dynamics and nonlinearities, measurement noise, and slow and bounded parameter variations; in particular, all signals in the closed-loop system will be bounded. 4.3 Experimental Results 4.3.1 Measurement Setup Tracking experiments are conducted to examine the effectiveness of the proposed robust adaptive control scheme. A trilayer polypyrrole actuator (20 x 5 x 0.17 mm) is clamped at one end, where the actuation voltage is applied. The tip displacement is measured by an OADM 2016441/S 14F laser sensor from Baumer Electric Inc. with resolution of 5 um. The controller is implemented in a PC equipped with dSPACE DS1104. The experimental setup is shown in Fig. 4.4. Before each experiment, an actuator cut with the specified size is soaked in the electrolyte (TBA+PF8 in propylene carbonate). The inner porous PVDF layer thus stores electrolyte, which enables the actuator to operate in air for some time. The time of continuous in-air operation depends on how fast the stored solvent evaporates, and without further packaging, it is about 4-5 hours. The experimental temperature and humid- ity are maintained as 25°C and 27% respectively. For practical applications, the packaging issue will have to be solved so that the actuator can work in air for much longer time. On the other hand, the current actuator demonstrates significant time-varying behavior, which provides a good testbed for verifying the proposed adaptive scheme. For comparison purposes, a PID controller and a fixed model-following controller are also implemented. Actuators with same dimensions and same conditions are used for all 112 4— Laser sensor ___] i“ Amplifier Computer and DSpace Figure 4.4: Schematic of the experimental setup. three controllers. The design of the PID controller and that of the model-following con- troller are also based on the model structure (4.9), while the model parameters are identified in separate experiments shortly before the tracking experiment starts. The model-following controller is constructed as in (4.17), except that the controller parameters will not be up- dated. The PID gains are carefully tuned and verified in Matlab simulation before the experiments to ensure that the initial tracking errors are within the similar range as those under other controllers, and the PID controller is U 6 _ = 6 + — +0.02s. (4.20) E s For each continuous tracking experiment (3-4 hours long), the parameters of the PID con- troller and the model-following controller remain constant (i.e., non-adapting). Throughout the experiments, the reference model Gm(s) is chosen to be s+1.5 G s:——. m() 32+4s+4 113 4.3.2 Results and Discussions In the first batch of experiments (Batch One), the reference input r(t) = 0.Ssin(7rt) + 0.53in(0.27tt)V, and the actuator output y(t) is required to track the desired trajectory ym(t) = Gm(s)[r(.)](t). The latter contains two frequency components (0.1 Hz and 0.5 Hz) with peak-to-peak variation of 0.62 mm. Each experiment runs continuously for four hours. Fig. 4.5 shows the tracking results at the beginning of the experiment (I = 0 h), while Fig. 4.6 shows the results when approaching the end of the experiment (I = 4 h). To better compare the control schemes, two metrics are defined for the tracking error. Given a starting time to and a constant T > 0, we define the normalized average error ea t +T A 1,3 ly(t) -ym(t)|dt ea = to + T , (4.21) fro lymmid’ and the normalized maximum error em max 0’0) "Ymmi t e t ,t + T em 2 [0 0 ] (4.22) maxt e [t0,t0 + T] Irm(t)l Throughout this thesis, T is chosen to be 100 3. Under the robust adaptive scheme, ea drops from 11% at to = 0 h to 7% at t = 4 h, and em drops from 15% to 9% for the same period. In comparison, ea increases from 7% to 28% under the PID controller, from 7% to 50% under the fixed model-following controller, and em increases from 8% to 25% under the PID scheme, and from 10% to 48% under the model-following scheme. Fig. 4.7 shows the evolution of ea and em, measured and calculated every half a hour, under the three schemes. It is clear that the robust adaptive control scheme delivers consistent tracking performance during the four-hour continuous operation, while the tracking performance under the PID scheme or the fixed model-following scheme deteriorates over time. A second batch of experiments (Batch Two) is conducted to examine the effectiveness 114 Robust adaptive ___ Achieved (raj, T T ' — Desired tra]. .5 (BOO! OI .a O 15 20 25 P I 9 UI 9 01° 1 l i I 1 I l l l 1 ._ _- . e.-_L~ . 4_._—7._~— ,.__. .. . -_ l i 1 t 1 l 1 I l l . 1 , I I - l 1 1 L 1 . Bending disp. (mm) o r- '°'5o 5 1o 15 2o 25 Time (s) (a) Robust adaptive ’5‘ 5 § 0 20025 , , c '1 1 1 ‘ g 0 5 10 15 20 25 E o 5 10 15 20 25 Time (s) (b) Figure 4.5: Experimental results on trajectory tracking (Batch One), I = 0 h. (a) Achieved trajectories versus desired one under the three controllers; (b) instantaneous tracking errors under the three schemes. 115 ROW“ adaptive --.-Achleved tral. : _ I 1 — Desired tral. I P PID P 5 1‘0 1‘5 20 25 Model-following I P Bending dlsp. (mm) 01 001 O P o 5 10 115 210 25 Time (s) (a) Robust adaptive i l I P 0 UI Tracking error (mm) C) l .. b.--e-..- )— 0 '5 10 115 20 25 Time (s) (b) Figure 4.6: Experimental results on trajectory tracking (Batch One), I = 4 h. (a) Achieved trajectories versus desired one under the three controllers; (b) instantaneous tracking errors under the three schemes (note the different vertical-axis scales). 116 0.5 4 + Robust adaptive : .i 0.4-~"'F"D __,..w’. +ModeI-foliowing , .I" o“ 2 a 0-3' """" )2", """"" ’1} t z . ,I‘ [I at ' ’0’ . II < ; ,2 ‘ Ix .I" ,x - 0.1’ ":.'-’¢ ;' ’:-a’f ,. .‘ _ L‘:::——-0"'"" : N) 0O 2 3 4 Time (h) (a) 0.5 . . . )L + Robust adaptive 2 x. o4..-.-°-P|D __,:- ' +Model-followingJ ; ,’ GE 1 .1 § 3 x' a E 0O 2 4 Time (h) (b) Figure 4.7: Normalized average error ea and maximum error em under the three control schemes (Batch One experiments). (a) Evolution of ea; (b) evolution of em. 117 of the three schemes in tracking trajectories of much larger magnitude. With the reference input r(t) = 2.1 sin(0.47rt) + 2.25in(0.17rt)V, the desired trajectory ym(t) has peak-to-peak variation of 3 mm. All three schemes show good tracking performance at the beginning. The experiment under PID control has to be stopped after 3 hours since at that time the voltage input exceeds the limit (1.6 V). Fig. 4.8 shows the tracking results when t = 3 h. Fig. 4.9 shows the evolution of ea and em. The trend is consistent with that in Batch One experiments. It can be seen that the robust adaptive controller keeps ea under 3.5% and em under 4% throughout the four-hour experiment. In the mean time ea rises from 5.5% to 8.4%, em from 5.7% to 10%, under the PID controller (in three hours), and ea rises from 5% to 80%, em from 5.5% to 81%, under the fixed model-following controller (in four hours). It is also important to compare the control efforts required under the different control schemes. Low control effort is highly desirable since that leads to long working life for the conjugated polymer actuator. Fig. 4.10 shows the evolution of the magnitude of voltage input under each scheme, for both Batch One and Batch Two experiments. The required voltage increases over time under every scheme, which is due to the deteriorating actuation capability of the actuator as the solvent evaporates. However, it can be clearly seen that the voltage input under the adaptive scheme is much lower than that under the PID scheme, and also lower than that under the model-following scheme most of the time. In the Batch Two/PID experiment, the polymer actuator was actually damaged and stopped functioning after 3 hours due to continuous high-voltage (> 1.5 V) actuation. There is another interesting observation during the experiments that supports the valid- ity of the reduced model. The effect of solvent evaporation can be incorporated by taking the diffusion coefficient D —> 0. This leads to p1 -+ 0, p3 —> 0, and 22 —-> 0, where p1, p3, and 2:2 are the poles and the zero of the reduced model (5.2), as defined by (4.4), (4.6), and (4.8). Fig. 4.11 (a) shows the evolution of the poles and the zero during the Batch One experiment every half an hour when the adaptive control scheme is adopted. Fig. 4.11 (b) 118 2_ Robust adaptive L-Achleveditgj. — Desired tra]. O 20 25 Ni O” - < at _I o _fi 01 '9 a" 1'5 20 25 Bending disp. (mm) O 1‘0 1'5 20 ‘25 Time (s) (a) Robust adaptive 15 20 25 0 5 10 1‘5 20 25 Time (s) (b) Figure 4.8: Experimental results on trajectory tracking (Batch Two), I = 3 h. (a) Achieved trajectories versus desired one under the three controllers; (b) instantaneous tracking errors under the three schemes. 119 + Robust adaptive , -°- PID E 0.3” , ......... ,1. -- Model-followm 3 x’ a ; : . ’. d) : .5 a 0-6'.” — .-:-..’0’,___.- t E .1.“ Q I ’°’ ' f ,1 . gt 0 4 .9]... < I.’ : .l’ 0.2” ' ‘;’*’----- ......... : ~’ , E L‘J:rJ:L'{__—:———2-—--t-"t OY v Y Y Y V L : 0 1 2 3 4 Time (h) (a) 0.8 I , z .7 + Robust adaptive § x” -.- PID 3 X E 05"" 4* Model-following (4 - -- -- —- 0 I 3 ’3‘: b ‘ I O f 1 ”ya : E 0.4.. _ ' I.’°’E:“"“ -" I . ’0 . 1 2 ; ,- 0.2 1' """"""" i- - l.- - - ...... z! .’.I-.-‘ ‘0‘ _o- __.———--O fur-1"- ;‘ - " 1 1 - off ' j Y Time (h) (b) Figure 4.9: Normalized average error ea and maximum error em under the three control schemes (Batch Two experiments). (a) Evolution of ea; (b) evolution of em. 120 _L O + Robust adaptive 1.4"” ‘9'PID ----------- 3 ------ ‘9 -I- Model-following 1 ’1’ 1.2- ----------- 2. 4- ————————— --; ....... ‘ .II 1 E ,’ a 1 ' ’ ‘ ‘IIII' .5 i 0 1’0 g 0.8 ~~~~~~~~~~~~~~~~~~~~~~ ~ ,4 --- '6 > (a) 2 7 7 v -o- Robust adaptive 1'8“-._P|D ““““““““ 1 —————————— --.4 In... . : 1.6-- Model following: _______ .75. __________ . ’(l I 2,9" 1.4 ------------------------- 3,4" ...................... 1 3 e" : g ",e’ . '3 1 2 ’3’, - - ~ -, j G ,I 2 2’ 5 ‘1 1 0" 1 0.8L """"""""" ‘ ------------- :r---—.-...).¢'.z | I ,o'1 ”L """ 5 M Em: *; e "'9 ’ 5 ' : 0. 1 Figure 4.10: Evolution of voltage input magnitude under the three schemes. (a) Batch One experiments; (b) Batch Two experiments. 121 shows this evolution during 100 seconds after the experiment started for half an hour. It can be seen that the poles and zero all tend to O, as predicted by the model. 1 -0—Pole1 -----Pole 2 Time (h) (a) ° : JyaI-aaam-émawflam 0 5 Check valves 9 o o O l O / Outlet ° °/ Inlet Section view of pump chamber Figure 5.2: The assembly schematic of micropump. 5.2 Physics-based, Control-oriented Model for the Pump Because PPy has sophisticated electrochemomechanical dynamics that can vary signifi- cantly over time [6], a model that captures the complicated dynamics yet amenable to con- 128 L 1L 7L . wilting l J lfi # ¥ 1 1 |—_ 1 11mm __ Figure 5.3: The mechanism of flap check valves. trol design is important for applying advanced control algorithms to deliver accurate and consistent flow rates for different applications. As shown in Fig. 5.7, the developed model in this thesis consists of three cascaded modules: 1) the electrical admittance module of PPy relating the current I (and thus the charge transferred) to the voltage input U; 2) the electromechanical coupling module of PPy expressing the generated strain 8m in terms of the transferred charge; and 3) the mechanical module connecting the generated strain to the diaphragm curvature K or flow rate <1) of the pump, which captures the dynamics of both the PPy membranes and the flap valves. In this thesis, the modules 1) and 2) follows those in Section 2 and therefore will only be briefly reviewed. Module 3) is derived based on an energy-based method, which will be elaborated in detail. 5.2.1 Electrical Admittance Module of PPy A diffusive-elastic-metal model was proposed for PPy, where it was assumed that the poly- mer matrix is perfectly conducting and the ion transport within the polymer is solely deter- mined by diffusion [16]. It was adapted to model the ions dynamics in the trilayer beam [6]. 129 -SU-8 PDMS 1:] Si 1 spin on SU-8 W lpattern SU-8 with photo mask {51 m m E] O O @ 0 sMnonPDMS lpeel off PDMS /97/ Figure 5.4: The microfabrication process to make a flap valve. The admittance model of a trilayer conjugated polymer was derived as [6] 1(5) ___ s[%—5tanh(h\/S/D)+\/E] U(S) §+Rs3/2+R3%§stanh(h\/S/D)’ (5.1) where U (s) and l (s) are the applied voltage and the resulting current in the Laplace domain, respectively, 3 is the Laplace variable, 6 is the double layer thickness, D is the diffusion coefficient, R is the resistance across the trilayer polymer, C is the double-layer capacitance, 130 Outlet Conjugated polymer 0’) Figure 5 .5: (a) The assembled micropump (top View); (b) The assembled micropump (bot- tom view). Alignment holes Check valves Electrode Figure 5 .6: A magnified view of the micropump to show the structure. 131 V Admittance I Electromechanical 5 Mechanical Y Module f Module Module Figure 5.7: The complete model structure for conjugated polymer actuators. and h is the thickness of the PPy layer. This infinite-dimensional system can be reduced to the following second—order transfer function [(3) ~ K~s(s+Z1) U(S) ~ (8+p1)(S+p2)’ (5.2) where the parameters are functions of physical parameters [6], which are either known or measurable. 5.2.2 Electromechanical Coupling of PPy The anions transferred to the polymer cause expansion of the polymer. It can be shown that the strain 8m introduced by the volumetric change is proportional to the density p of the transferred charges [16]: em = ap, (53) where a is the strain-to-charge ratio, which varies for different anions. PPy doped with TFSI' is used in this thesis, and the strain-to-charge ratio is estimated to be 7 x 10—10 m3 -C"1. Because the double-layer capacitance is much smaller than the bulk capacitance of the PPy polymer, the charges stored in the double layer at the steady state is negligible comparing with those in the bulk. Therefore one can obtain the density p(s) NS) = . (5.4) where A is the area of PPy. 132 5.2.3 Mechanical Module of the Micropump Given the actuation strain 8m in the PPy layers, an energy-based method is used to model the deformation of the pump diaphragm and consequently the flow rate. In this method the equilibrium of a mechanical structure is obtained by minimizing a properly defined total energy, and such as approach has been taken to predict the curvature of a composite plate generated by the strain mismatch in different layers [47, 84, 85]. In the following, we first discuss the common framework that applies to both a clamped whole diaphragm and a petal—shape PPy actuated diaphragm (simply called petal-shaped diaphragm hereafter), and then specialize the discussion to individual cases. There are two relevant energy terms: the elastic energy stored and the work done to the fluid. The elastic energy of actuation diaphragm will be calculated based on mechanics of constitutive modeling. The analysis holds for both the whole diaphragm and the petal- shaped diaphragm cases, because the infinitesimal elements in both cases experience radial and transverse strains as defined in Fig. 5.8. The diaphragm bends as a result of the de- formation of all these elements. The difference between the two cases is that the midplane strain 80 depends on the diaphragm curvature K in the whole diaphragm case because of the edge constrains of the edges, while in the petal-shaped diaphragm case the curvature is not related to the middle plane strain. The latter releases the constrains on the curvature K from the clamped edge. Figure 5.8: Definition of the principal strains. For small deflection, the principal strains defined in Fig. 5 .8 can be expressed as follows 133 [47]: 2 du d w €rr — d— —Z:l'—2 + Es, (55) r u zdw €99 —;—;—d-‘;+€3, (5.6) where z is the axis in the thickness direction, 8,- is the swelling strain in different layers (8m in the oxidized PPy layer, —8m in the reduced PPy layer, and O in the PVDF layer), u(r) and w(r) are the radial and transverse displacements of points in the midplane. Notice that for the case of a petal-shaped diaphragm, there is a very thin PDMS layer underneath the conjugated polymer to seal the chamber, but the Young’s modulus of PDMS used in this thesis is 0.5 MPa, which is much smaller than the Young’s modulus of PPy (60 MPa) and PVDF as (612 MPa) [1 1]. Therefore, the influence of the PDMS layer on the conjugated polymer deformation is ignored, and the midplane is taken to be the middle plane of the PVDF layer. The displacements u(r) and w(r) in the midplane can be expressed as u = £0r, (57) w = g, (5.8) where 80 represents the strain of the midplane, and K is the curvature of the plane as shown in Fig. 5.9. Substituting (5.7) and (5.8) into (5.5) and (5.6), one obtains the following expression errzeggzeo—z-K+£,. (5.9) 134 Figure 5.9: The coordinates and geometry definition. The corresponding strain energy density is: E. ‘P(r,z) :2 2—(—1Tl-—2—)-(E,2r + £39 + 2Vi€rr€96) 1+V in —V-2 l where E,- and v,- are Young’s modulus and Poisson ratio of the material in the i-th layer, which are notated as E pp), and VPPy for the PPy layers, EPVDF and VPVDF for the PVDF layer. Finally the total strain energy of the trilayer conjugated polymer is obtained by superimposing three energy densities and integrating over the volume —h1(1+Vp )EP 2 Welastic =_/7r ”/00 [-/h2 Py Py(£0 sin—1'70 dz ”1 (1+V:VDF)EPVDF(80 ___z M261,z “hl 141%va h 1+V E +/ 2( PPy) PPy(£0+£m —z K)2dz]rdrd1:, (5.11) where r0 is the undeformed diaphragm radius. Since the whole diaphragm and the petal- shaped diaphragm have the same radius under undeformed configuration, r0 is used in both cases. 135 In the whole diaphragm case, the curvature denoted as K is assumed to be uniform when the diaphragm displacement is small. The geometrical relationship is illustrated in Fig.5.10, where the solid curve represents the sectional view of the middle plane of the deformed diaphragm. The angle 9 is obtained as 0 = arcsin(x‘r0). (5.12) One can obtain the strain in the middle plane as r—ro 9 =———-=———1 5.13 80 r0 KTO ’ ( ) which relates 80 and K. ‘ \ --> C L-----__--_--_---_-_-_-__-__ J 1 Figure 5.10: Illustration of the geometrical relationship in (5. 12) and (5.13), R = E. Since the deformation of the diaphragm will generate the chamber pressure, the work done by the chamber pressure needs to be considered in the energy based method. Consid- ering an incompressible fluid, we have the following equation from the mass conservation 136 principle, I pv = p /0 (china) —ow)dr, (5.14) where p is the density of the fluid, (pin and (1)0“, are the flow rates at the inlet and outlet, respectively,with the unit of m3 /s. The relationship between the chamber pressure p and the flow rates is modeled as follows: dp 1 27 = 'C—CW’in —¢0ut)a (515) where CC is the chamber capacitance that can be determined experimentally with the unit of m3 /Pa [86]. Combining (5.14) and (5.15), one can obtain p = —, (5.16) where pl, : 0 is considered as 0, since there is no initial net pressure in the chamber. Therefore, one can express the work done by the diaphragm as ’0 W W = 275/ / p r drrdr p 0 0 ( ) 27: ’O/W = — th'rdr. 5.17 Cc/O O ( ) . . . . . dW . The total energy rs W = We + WP. The equrhbnum is reached when E(- = O m lastic the whole diaphragm case, since K is the only independent variable. In the petal-shaped (9W 8W diaphragm case, the equilibrium is reached when — = 0 and 5? = O. 980 In order to obtain the explicit expressions for both cases, firstly the volume change in the chamber is calculated for the whole diaphragm and petal-shaped diaphragm respectively. In the case of the whole diaphragm, the volume change can be characterized as a dome’s volume 7‘ 2 2 137 where ho is the height of the dome that equals R — ‘/R2 — r3 based on the geometry. Substituting this expression into Eq. (5.18), one has 7: _ _ _ 2_ 2 2 2_ 2_ 2 V _ 3(R j/R r0)(R +r0 R‘/R r0) ”3 5 19 z — K . 6 , ( ) r2 2 . . . . . 0 r where the approxrmatron 1S based on the Taylor series expansron at R—z— = 0 for 1 — 13%. since the small deflection of the diaphragm implies that R > r0. Substituting (5.19) into (5.17), one can obtain the expression of WI) and consequently dW W. The equilibrium is reached when :1? = 0 in the whole diaphragm case, since it“ is the dW only variable. The equation of —d—r_c_ = O can be rearranged as follows: 7 _7‘50_ 32cc +C1r0h%(h2 x — 38m) x + 3C1r0h%£mx° — 6C1r0h2]2 = 0, (5.20) —36C12r8h% + (1 — Kzrgfl x2 + (C2 — Cl )r0h315 x2 where C1 __ (I + VPPy)EPPy 02 _ (1 + VPVDZF)EPVDF _ 2 a — — 9 1 “ VPPy 1 VPVDF The curvature K can be obtained by numerically solving this polynomial and appropriately choosing the root. In the petal-shaped diaphragm case, the expressions of volume change V and work Wp are different. As illustrated in Fig. 5.10, the following geometric relationship holds for the petal-shaped diaphragm when the deflection is small, 1 2a—=m, an) K 138 which implies K . sine z a = 592—. (5.22) The volume change in the petal-shaped diaphragm case can be approximated by the shad- owed volume (a top-off cone) in Fig. 5.12 under the assumption of (5.22) 7: V 2 508+ (r0 —sin26or0)2) -23in29-r0 4 Mo K , (5.23) 22 Figure 5.11: Geometric relationship in the calculation of the volume V and the work done by the diaphragm. The displacement 5 defined in Fig. 5.11 is expressed as rgx .5: sin6 - r0 as —2—. (5-24) The displacement at each point of the diaphragm is defined as fix in Fig. 5.11, which can be expressed in terms of x and 5 as :: . (5.25) Deformed configuration l l Original configuration Figure 5.12: Illustration of the volume change under the actuation voltage for the petal- shaped diaphragm. By substituting (5.23) into (5.17), one can express the work done by the diaphragm as r0 5;; WI) 2 271’ O A; p(t')d’t'rdr 2 8 2 ___ 7t rOK IZCC . W W The equilibrium is reached when _8__ :: O and L = 0. Since 880 8K 8W 58—0— = 47tr(2)£0[C1 (112 — h1)+C2hl], a—K‘ = “475 0[C1€m(h2“h1)- 3 ’T1 2 8 +7! I‘OK’ 6CC ’ the equilibrium is: 80 = 0, 3C1 8,7101% — h?) m6 + 3 C, h2 — C1h% + C25? (5.26) (5.27) (5.28) (5.29) (5.30) Eq. (5.30) captures the static deformation of the diaphragm in terms of em. The material damping effect is introduced as (2.26) to incorporate the dynamics during the deformation process. The values of parameters in (2.26) are kept the same for both PPy and PVDF. 140 Other damping effects in the pump, such as the flap valve damping, are captured by using the equivalent chamber capacitance that includes a damping term Cc C' (s) = ———————, (5.31) C 1 + CCRCS where RC represents the equivalent damping resistance with the unit of Pa - s/m3. Thus one can write (5 .30) in the Laplace domain as K=H(S)'€m, (5.32) where 2 2 3C h —h 11(3) = l( 2 1) 3 3 3 33,5 , Clhz -C1h1+C2hl + mag and C1, C2, and CC! are the frequency—dependent terms, because the Young’s moduli in (5.30) are replaced by the ones in (2.26). 5.2.4 Complete Model One can obtain the complete model for the petal-shaped diaphragm in an analytical form by combining (5.2), (5.3) and (5.32), which is shown as follows: K(s) : aK- (s+zl )H(s) (1(8) 4012 —h1)-'S(S+P1)(S+P2)' (5.33) Note that h = h2 - hl in (5.2). However,the complete analytical model for the whole diaphragm is not easy to obtain due to the complexity of (5.20). Therefore, the model prediction in that case is obtained by numerically solving (5.20). 141 Considering (5.14) and (5.23), one gets 7LT Oh (bin — (bout = ' S ' K. (5.34) 5| Combining (5.33) and (5.34), one can furthermore obtain the transfer function model from the voltage input to the flow rate: (13mm —¢our(5) nrgaK-(s+zl)I-I(s) Um :4A(h2—hl)'(5+p1)(s+p2)' (5’35) 5.3 Experimental Results A trilayer PPy circular actuator is used in the experiments, with the radius r0 of 10 mm. The curvature is measured by a laser sensor (OADM 2OI6441/S14F, Baumer Electric Inc) with a resolution of 5 pm. The model parameter R is identified by applying a high-frequency (100 Hz) sinusoidal input U Sin(a)t). From (2.12), the impedance approaches R as a) —+ 0°. To measure the parameter C, a step voltage is applied and the transferred charge into the PPy layer is computed by integrating the charging current. The calculation details can be found. in [6]. Finally, the resistance and capacitance are identified to be: R = 27 Q, C = 8.69 x 10’5 F. The diffusion coefficient D is chosen to be 2 x 10’10 mz/s based on [16]. The double-layer thickness 5 is estimated to be 25 nm based on [18]. 5.3.1 Admittance Since cutting does not influence the electrical property of conjugated polymer, (2.12) is used to predict the admittance of both the whole diaphragm and the petal-shape diaphragm. The comparison is shown in Fig. 5.13, where a good match of admittance is seen between the model and the experimental results. 142 Magnitude (dB) 1: 1 .1 11111111 11111111 11111111 111111111 1111111 60 1o" 10'1 10° 101 102 103 100 i l 1 l v v vvvr r I v [V l[[ v y-vvv-v A l 0 Unit diaphragm ,_ 3 ; 6 Petal-shape diaphragm‘ 3 8 ; -—Model predictlon " 5° 7 tit: ' 1"":i - 7 m I l l 1 10 . . .c . IL I q 0-2 A Lil-1* LL..1.0.......1 1'2 ms 10 10 10 10 10 10 Frequency (rad/sec) Figure 5.13: Comparison of model prediction from (2.12) with the experimental results for the whole diaphragm and the petal-shaped diaphragm. 143 5.3.2 Displacement In experiments the displacement at the center of the diaphragm is measured by a laser distance sensor. The predicted curvature of the petal-shape diaphragm from (5.33) is used to calculate the displacement based on the following equation [6]: 2 K] where l is the distance between the clamped end and the laser incident point when the beam is at rest, which is r0 minus 1 mm in the setup. The predicted displacements for the petal-shape diaphragm are compared with the ex- perimental results when there is no water in the pump (Fig. 5.14) and when pumping water (Fig. 5.15). There the actuation voltage applied is sinusoidal with an amplitude of 4 V and with different frequencies. All the other parameters in (5.33) are given or identified in the previous admittance model experiments. CC and RC can be determined by fitting the experimental data for petal-shaped diaphragm. They are identified as CC = 1.02 x 10—8 m3 / Pa and RC = 3 x 108 Pa - s/m3 . To predict the results when there is no water in pump, the parameters CC and RC are set to be cc and 0 respectively, which sets the fluid energy to be zero to eliminate the fluid energy term. Thus the curvature is merely determined by the equilibrium of elastic energy. The experimental results for the whole diaphragm case are also compared with the numerical solution of (5.2), (5.3), and (5.20). The result when there is no water is shown in Fig. 5.16, while the result when pumping water is shown in Fig. 5.17. The actuation voltage is 4 V. It can be seen that the petal-shaped diaphragm can generate almost 10 times larger displacment than the whole diaphragm when there is no wa- ter in the chamber. When pumping water, the displacement of the petal-shaped diaphragm is almost 3 times larger than the whole diaphragm. These experimental results here proved the effectiveness of the petal-shaped diaphragm design in alleviating the constrains from the clamped edge. 144 Displacement (mm) [0 Viv 3: Experimental ‘ L: Simulated 5‘31 1'1: 0 1 Phase (deg) '5 _ ..;:..;.. 1. ;:.;; 150-2 -1 0 1 2 10 1o 10 10 10 Frequency (rad/sec) Figure 5.14: Comparison of model prediction from (5.33) with the experimental results for the petal-shaped diaphragm when there is no water (actuation voltage amplitude 4 V). 145 1 P a) l Displacement (mm) D .b 0.2—-- o L 10'2 o I ' fi' ViTTrT. ' ' ‘ 1 III I I I .11 A . . “ ‘ ; 1 T 3: Experimental 9 X j j , .. 1 1 . j , — Simulated 13 ; ' 1 1.2: . 1 ; 1 . ‘ . : . . V 1 1‘1 i 1 1 a, '50 ' ********* ‘ I i .- ge ; z .1: . 1 .x ; E ; 1 1 _100 i . L .. i1 1 1 1 . .1. 10' 10'1 10° 101 Frequency (rad/sec) Figure 5.15: Comparison of model prediction from (5.33) with the experimental results for the petal—shaped diaphragm when pumping water (actuation voltage amplitude 4 V). 146 P a: .° C, 1 Displacement (mm) D b 0.2 9 10'2 0 ' I ' i ‘ #_1__._.___‘ A {’3} Experimental 9 50 — Simulated E ' . 8 (U -100~ ‘~ g 1 o. y : _150 1 1 1.11111 1 1111111 1 QLJ‘IA11 . . 1 1 1 10'2 10'1 ° 01 1o2 10 1 Frequency (rad/sec) Figure 5.16: Comparison of model prediction with the experimental results for the whole diaphragm when there is no water by numerically solving (5.20) (actuation voltage ampli- tude 4 V). 147 Displacement (mm) .0 1 1 1 '20 1 ' . 1 ' '11 1 1 1 1 ' 1F_1_1 . 1 1 '_'_>: T 8 ‘ ‘ i 1 5 1 ‘ I 1 3: Experimental _ ,1 1 —Simulated . . ‘ ‘ ‘ r ,7 ‘ ‘ ‘ J Phase (deg) 10'2 10'1 10 Frequency (rad/sec) Figure 5.17: Comparison of model prediction with the experimental results for the whole diaphragm when pumping water by numerically solving (5.20) (actuation voltage ampli- tude 4 V). 148 5.3.3 Flow Rate The flow rates of the petal-shaped diaphragm pump are measured and compared with the model prediction (5.35) for an actuation voltages of 3 V (Fig. 5.18) and 4 V (Fig. 5.19). It can be seen that the model can predict the experimental results well under different input voltages. The frequency to achieve highest flow rate is 0.5 Hz, which is also predicted well by the model. When the frequency of the input voltage is low, the flow rate is small due to the slow movement of the pump. The flow rate will increase as operating frequency increases. However, the flow rate will decrease as frequency becomes higher, because of the damping effect of the flap valves and the declining response of conjugated polymer actuator at high frequencies. Thus the model (5.35) can facilitate the design optimization and the feedback control of the flow rate. we have also tested the pumping performance of the whole diaphragm. However, the flow rate is barely observable in that case. 5.4 Chapter Summary In this thesis, circular conjugated polymer actuators are investigated for potential microp- ump applications. A petal-shaped diaphragm design is proposed to alleviate the constrains from the clamped edge. An analytical model is proposed that captures the relationship between the actuation voltage and the diaphragm deformation/flow rate. Experiments are conducted to verify the model and identify the parameters. The largest flow rate achieved in the current experimental setup is 1260 uL/min, when the operating frequency is 0.53 Hz. For comparison, modeling analysis and experiments are also conducted on the whole di- aphragm case. The model predicted much smaller diaphragm deformation compared with the petal-shaped diaphragm, which is also verified in experiments. 149 I _' )7. . 1109 ‘ MI E 1 A ‘ 1; 1 ‘ 1 : ; .11 ‘ rnfl- Experimental 1000~ 7—1‘ ‘9; ~ 931' j — Slmulated 800— 700L- 600~ 500 ~ Flow rate (pL/min) 400*”'i'1: 1 Frequency (Hz) Figure 5.18: Flow rate of the micropump at different frequencies when input voltage mag- nitude is 3 V. 150 14cc , _ f 1 :1‘1' 1 1 1 1‘1: : 1 fl Experimental 3 . l‘ I — Simulated 1200”” 3 ’1 ?‘1'”"1 ’ : ” f"? 1‘11’ A 51000 ' ‘ E ’ ’ \ 1 J l 1 1 1 1 v 1 1 q) 800” * 'E q..- . a h 11. 1 I 11 5 600” ”7””7' 3' WWW"? 1‘1 11*"; (”1’115’" 4oo~ »-1 4 209 . ...:;‘:1 L ..;;;.:. 1 ..;;.... 1 :..;;.i 10'2 10'1 ° 10‘ 1o2 10 Frequency (Hz) Figure 5.19: Flow rate of the micropump at different frequencies when input voltage mag- nitude is 4 V. 151 Chapter 6 Conclusion In this dissertation, we have focused on modeling, control, and application of conjugated polymer actuators and sensors. Experiments have been conducted to validate the modeling approach. Model-based control strategies are then proposed to control conjugated polymer actuators. The application of conjugated polymer in micropump is investigated from both modeling and experiment perspectives. The contributions of this dissertation on the modeling aspect include two main parts. The first part is the development of a scalable, low-order transfer function model for con- jugated polymer actuators that captures the major sophisticated electrochemomechanical dynamics, but is still control-oriented and therefore suitable for real-time control. This model is obtained by ignoring high-frequency dynamics in an infinite-dimensional model. The proposed model provides the theoretical basis for the adaptive control scheme. The second part is modeling of the nonlinearities existing in electrical and mechanical domains. The coupling of redox level change with ions migration in conjugated polymer is considered as the main source of nonlinearity in electrical dynamics. A partial differential equation (PDE) that governs its electrochemical dynamics is proposed and solved to ad- dress the nonlinear electrical property of conjugated polymer. The mechanical nonlinearity of conjugated polymer actuator is also investigated. Instead of using the elastic modulus 152 as in the linear elasticity theory, we use a nonlinear strain energy function to capture the stored elastic energy under actuation-induced swelling, which further allows us to compute the induced stress even under large actuation voltage. The proposed framework can also be applied to the analysis of large deformations in some other electroactive polymers. We have presented a fiber-directed conjugated polymer actuator that can generate tor- sional motion, which is due to the anisotropy associated with the interaction between the fiber and the material matrix. The nonlinear elasticity-based model is utilized to capture the actuator performance. The effectiveness of the model is verified through comparison with experimental results. Besides modeling on actuation performance of conjugated polymer, the electromechan- ical sensing behavior of conjugated polymer is experimentally characterized and mathemat- ically modeled as well. A theory for the sensing mechanism is proposed by postulating that, through its influence on the pore structure, mechanical deformation correlates directly to the concentration of ions at the PPy/PVDF interface. This provides a key boundary con- dition for the partial differential equation (PDE) governing the ion diffusion and migration dynamics. An analytical model is then obtained in the form of a transfer function that relates the open-circuit sensing voltage to the mechanical input. We have proposed an effective adaptive control strategy for conjugated polymer actu- ators based on our control-oriented model. A recursive. least—squares algorithm is used to identify online the parameters of the transfer function model, which captures the essential actuation dynamics that varies significantly with time. A self-tuning regulator is designed based on the identified parameters to form the closed-loop control system. A parameter projection step ensures that the parameter estimates stay within the physically-meaningful region, and thus makes the system robust against noises and unmodeled dynamics. The application of conjugated polymer as a micropump is investigated. We have pro- posed a diaphragm micropump actuated by conjugated polymer in a petal-shape design to alleviate the edge constrains. Transfer function models from the actuation voltage to the 153 diaphragm deflection and flow rate are obtained. The pump body is fabricated through MEMS fabrication process to miniaturize the pump size. Experiments are conducted to evaluate the micropump performance. The effectiveness of the transfer function models are also verified by the experiments, which will facilitate design optimization and the use of feedback control tools in dealing with the complicated behavior of conjugated polymer actuators. There are several interesting directions to expand the work in this dissertation. Firstly, although adaptive control scheme we proposed has shown its superiority in overcoming the time-varying effects, its application is constrained in the linear range. The actuation voltage magnitude has been tuned small so that the change of redox level with actuation voltage can be ignored. The nonlinearity will impair the performance of the con- trol system based on the linear model. We have proposed nonlinear models to capture the significant nonlinearities in conjugated polymers in both electrical and mechanical do- mains. An interesting research direction is to apply nonlinear control schemes based on the nonlinear models of conjugated polymer actuator to achieve better performance, when the nonlinearities become significant, i.e., large deformation under large actuation voltage. Another direction is to further pursue the application of conjugated polymer in mi- crofluid medical devices that are experiencing significant development recently. A microp- ump based on conjugated polymer actuator has been investigated in this dissertation. We have built and tested the prototype of micropump. The model relating the flow rate to the actuation voltage is also proposed. It is meaningful to design feedback control system based on the proposed model to overcome the performance decaying of conjugated poly- mer, which would allow the conjugated polymer micropump to deliver desired and accurate fluid for a long period of time. 154 BIBLIOGRAPHY [l] E. Smela, O. Inganas, and I. Lundstrom, “Controlled folding of micro-size structures,” Science, vol. 268, pp. 1735-1738, 1995. [2] E. W. H. lager, O. Inganas, and I. Lunstrom, “Microrobots for micrometer-size objects in aqueous media: Potential tools for single cell manipulation,” Science, vol. 288, pp. 2335-2338, 2000. [3] J. Zhou, H. Chan, T. To, K. Lai, and W. J. Li, “Polymer MEMS actuators for under- water micromanipulation,” IEEE/ASME Transactions on Mechatronics, vol. 9, no. 2, pp. 334—342, 2004. [4] A. Mazzoldi and D. De Rossi, "Conductive-polymer-based structures for a steerable catheter,” in Smart Structures and Materials 2000: Electroactive Polymer Actuators and Devices, Y. Bar-Cohen, Ed. Bellingham, WA: SPIE - The International Society for Optical Engineering, 2000, pp. 273-280. [5] Z. Chen and X. Tan, “A dynamic model for ionic polymer-metal composite sensors,” Smart Materials and Structures, vol. 16, pp. 1477-1488, 2007. [6] Y. Fang, X. Tan, and G. Alici, “Robust adaptive control of conjugated polymer ac- tuators,” IEEE Transactions on Control Systems Technology, vol. 16, pp. 600—612, 2008. [7] R. Baughman, “Conducting polymer artificial muscles.” Synthetic Metals, vol. 78, pp. 339—353, 1996. [8] E. Smela, “Conjugated polymer actuators for biomedical applications,” Journal of Advanced Materials, vol. 15, no. 6, pp. 481—494, 2003. [9] G. G. Wallace, G. M. Spinks, L. Kane-Maguire, and P. R. Teasdale, Conductive Elec- troactive Polymers: Intelligent Materials Systems (2nd Edition). Boca Raton, FL: CRC Press LLC, 2003. ' [10] A. Della Santa, D. De Rossi, and A. Mazzoldi, “Characterization and modelling of a conducting polymer muscle-like linear actuator,” Smart Materials and Structures, vol. 6, pp. 23-34, 1997. 155 [l l] G. Alici, B. Mui, and C. Cook, “Bending modeling and its experimental verification for conducting polymer actuators dedicated to manipulation applications,” Sensors and Actuators A, vol. 126, pp. 396—404, 2006. [12] J. Janata and M. Josowicz, “Conducting polymers in electronic chemical sensors,” NATURE MATERIALS, vol. 2, pp. 19—24, 2003. [13] G. Alici, P. Metz, and G. M. Spinks, “A methodology towards geometry optimization of high performance polypyrrole (PPy) actuators,” Smart Materials and Structures, vol. 15, pp. 243—252, 2006. [14] G. Alici and N. N. Huynh, “Predicting force output of trilayer polymer actuators,” Sensors and Actuators A, vol. 132, no. 2, pp. 616—625, 2006. [15] M. Christophersen, B. Shapiro, and E. Smela, “Characterization and modeling of PPy bilayer microactuators. Part 1. curvature,” Sensors and Actuators B, vol. 115, pp. 596—609, 2006. [16] J. D. W. Madden, “Conducting polymer actuators,” PhD thesis, MIT, 2000. [17] P. G. A. Madden, “Development and modeling of conducting polymer actuators and the fabrication of a conducting polymer based feedback loop,” PhD thesis, MIT, 2003. [18] Y. Fang, X. Tan, Y. Shen, N. Xi, and G. Alici, “A scalable model for trilayer conju- gated polymer actuators and its experimental validation,” Materials Science and En- gineering C: Biomimetic and Supramolecular Systems, vol. 28, pp. 421—428, 2008. [19] G. Alici, B. Mui, and C. Cook, “Characterisation and bending modeling of conducting polymer actuators for use in micro/nano manipulation,” in Proceedings of the IEEE International Conference on Robotics and Biomimetics, 2005, pp. 560—565. [20] Y. Wu, G. Alici, J. D. Madden, G. M. Spinks, and G. G. Wallace, “Soft mechanical sensors through reverse actuation in polypyrrole,” Advanced Functional Materialss, vol. 17. PP. 3216—3222, 2007. [21] W. Takashima, K.Hayasi, and K. Kaneto, “Force detection with Donnan equilibrium in polypyrrole film,” Electrochemistry Communications, vol. 9, pp. 2056—2061, 2007. [22] T. Skotheim and J. Reynolds, Conjugated Polymers: Processing and Applications, 3rd ed. CRC Press, 2006. [23] D. L. Boxall and R. A. Osteryoung, “Switching potentials and conductivity of polypyrrole films prepared in the ionic liquid 1-butyl-3-methylimidazolium hexafluo- rophosphate,” Journal of The Electrochemical Society, vol. 151, no. 2, pp. E41-E45, 2004. 156 ' [24] [25] [26] [27] [28] [29] [30] [31] [32] [33] [34] [35] H. Mao and P. Pickup, “In situ measurement of the conductivity of polypyrrole and poly[1-methyl-3-(pyrrol-l-ylmethyl)pyridinium]+ as a function of potential by medi- ated voltammetry. redox conduction or electronic conduction?” Journal of American Chemical Society, vol. 112, pp. 1776-1782, 1990. G. Spinks, L. Liu, G. Wallace, and D. Zhou, “Strain response from polypyrrole actu- ators under load,” Advanced Functional Materials, vol. 12, pp. 437 - 440, 2002. D. D. Rossia, A. D. Santaa, and A. Mazzoldia, “Performance and work capacity of a polppyrrole conducting polymer linear actuator,” Synthetic Metals, vol. 90, pp. 93— 100, 1997. Q. Pei and O. Inganas, “Electrochemlcal appllcatlons of the bending beam method. 1. mass transport and volume changes in polypyrrole during redox,” Journal of Physical Chemistry, vol. 96, pp. 10 507—10 514, 1992. P. Metza, G. Alici, and G. Spinks, “A finite element model for bending behaviour of conducting polymer electromechanical actuators,” Sensors and Actuators A, vol. 130-131, pp. 1—1 1, 2006. J. D. Madden, D. Rinderknecht, P. A. Anquetil, and I. W. Hunter, “Creep and cycle life in polypyrrole actuators,” Sensors and Actuators: A. Physical, vol. 133, pp. 210—217, 2007. T. F. Otero, J. J. L. Cascales, and G. V. Arenas, “Mechanical characterization of free- standing polypyrrole film,” Materials Science and Engineering: C, vol. 27, pp. 18-22, 2007. H. Tsai, T. J. Pence, and E. Kirkinis, “Swelling induced finite strain flexure in a rectangular block of an isotropic elastic material,” Journal of Elasticity, vol. 75, pp. 69—89, 2004. Y. Fang, X. Tan, and G. Alici, “Robust adaptive control of conjugated polymer actua- tors,” IEEE Transactions on Control Systems Technology, vol. 14, no. 6, pp. 600—612, 2008. H. Demirkoparan and T. J. Pence, “Torsional swelling of a hyperelastic tube with helically wound reinforcement,” Journal of Elasticity, vol. 92, pp. 61—90, 2007. P. G. A. Madden, J. D. W. Madden, P. A. Anquetil, N. A. Vandesteeg, and I. W. Hunter, “The relation of conducting polymer actuator material properties to performance,” IEEE Journal of Oceanic Engineering, vol. 29, pp. 696-705, 2004. B. Qi, W. Lu, and B. R. Mattes, “Control system for conducting polymer actuators,” in Smart Structures and Materials 2002: Electroactive Polymer Actuators and Devices (EAPAD), Y. Bar-Cohen, Ed. Bellingham, WA: SPIE - The International Society for Optical Engineering, 2002, pp. 359—366. 157 [36] T. A. Bowers, “Modeling, simulation, and control of a polypyrrole-based conducting polymer actuator,” Master’s thesis, Massachusetts Institute of Technology, 2004. [37] K. J. Astrom and B. Wittenmark, Adaptive Control, 2nd ed. Addison-Wesley, 1995. [38] T. S. Hansen, K. West, 0. Hassager, and N. B. Larsen, “An all-polymer mi- cropump based on the conductive polymer poly(3,4-ethylenedioxythiophene) and a polyurethane channel system,” Journal Of Micromechanics And Microengineering, vol. 17, pp. 860—866, 2007. [39] D. J. Laser and J. G. Santiago, “A review of micropumps,” Journal of Micromechanics and Microengineering, vol. 14, pp. R35—R64, 2004. [40] A. Wego, H.-W. Glock, L. Page], and S. Richter, “Investigations on thermo-pneumatic volume actuators based on pcb technology,” Sensors and Actuators A: Physical, vol. 93, pp. 95—102, 2001. [41] R. Zengerle, M. Richter, F. Brosinger, A. Richter, and H. Sandmaier, “Performance simulation of microminiaturized membrane pumps,” 7th lntemational Conference on Solid-State Sensors and Actuators, pp. 106—109, 1993. [42] J. G. Smits, “Piezoelectric micropump with three valves working peristaltically,” Sen- sors and Actuators A: Physical, vol. 21, pp. 203—206, 1990. 1 [43] TB. Xu and J. Su, “Development, characterization, and theoretical evaluation of electroactive polymer-based micropump diaphragm,” Sensors and Actuators A: Phys- ical, vol. 121, pp. 267—274, 2004. [44] W. Benard, H. Kahn, and A. Heuer, “Thin-film shape-memory alloy actuated microp- umps,” Journal of Microelectromechanical Systems, vol. 7, pp. 245-251, 1998. [45] J. H. Kim, K. T. Lau, R. Shepherd, Y. Wu, G. Wallace, and D. Diamond, “Perfor- mance characteristics of a polypyrrole modified polydimethylsiloxane (pdms) mem- brane based microfluidic pump,” Sensors and Actuators A: Physical, vol. 148, pp. 239—244, 2008. [46] T. T. N guyena, N. S. Goob, V. K. Nguyenc, Y. Y00, and S. Park, “Design, fabrication, and experimental characterization of a flap valve ipmc micropump with a flexibly supported diaphragm,” Sensors and Actuators A: Physical, vol. 141, pp. 640—648, 2008. ‘ [47] L. B. Freund, J. A. Floro, and E. Chason, “Extensions of the stoney. formula for sub- strate curvature to configurations with thin. substrate or large deformations,” Applied Physics Letters, vol. 74, pp. 1987—1989, 1999. [48] Y. Fang, X. Tan, and G. Alici, “Robust adaptive control of conjugated polymer ac- tuators,” IEEE Transactions on Control Systems Technology, vol. 16, pp. 600—612, 2008. 158 [49] P. Magnusson, G. Alexander, V. Tripathi, and A. Weisshaar, Transmission Lines and Wave Propagation, 4th ed. CRC Press, 2001. [50] G. F. Franklin, J. D. Powell, and A. Emami-Naeini, Feedback Control of Dynamic Systems, 5th ed. Upper Saddle River, NJ: Pearson Education, Inc., 2006. [51] S. Satyanarayana, “Surface stress and capacitive MEMS sensor arrays for chemical and biological sensing,” PhD thesis, University of California, Berkeley, 2005. [52] C. 0. Yoon, M. Reghu, D. Moses, and A. J. Heeger, “Transport near the metal- insulator transition: Polypyrrole doped with PF 6,” Physical Review B (Condensed Matter), vol. 49, pp. 10 851—10 863, 1994. [53] T. F. Otero and J. M. Sansinena, “Bilayer dimensions and movement in artificial mus- cles,” Bioelectrochemistry and Bioenergetics, vol. 42, no. 2, pp. 117-122, 1997. [54] 1. Ward, Mechanical Properties of Solid Polymers. Wiley, 1979. [55] Y. Shen, N. Xi, K. W. C. Lai, and W. J. Li, “A novel PVDF micro force/force rate sensor for practical application in micromanipulation,” Sensor Review, vol. 24, pp. 274-283, 2006. [56] N. Lakshminarayanaiah, Transport Phenomena in Membranes. Academic Press, 1969. [57] H. Mao, J. Ochmanska, C. Paulse, and P. Pickup, “Ion transport in pyrrole-based polymer films,” Faraday Discussions of the Chemical Society, vol. 88, pp. 165—176, 1989. [58] Y. Wu, G. Alici, G. M. Spinks, and G. G. Wallace, “Fast trilayer polypyrrole bending actuators for high speed applications,” Synthetic Metals, vol. 156, pp. 1017-1022, 2006. [59] G. Phillips, R. Suresh, J. Waldman, J. Kumar, J. I-Chen, S. Tripathy, and J. Huang, “Dielectric property of polypyrrole doped with tosylate anion in the far infrared and microwave,” Journal of Applied Physics, vol. 69, pp. 899—905, 1991. [60] F. Legros and A. Fourrier-Lamer, “Dielectric property of doped polypyrrole in the 5 hz - 1 ghz frequency range,” Materials Research Bulletin, vol. 19, pp. 1109—1117, 1984. [61] S. Saafan, M. El-Nimr, and E. El-Ghazzawy, “Study of dielectric properties of polypyrrole prepared using two different oxidizing agents,” Journal of Applied Poly- mer Science, vol. 99, pp. 3370—3379, 2006. [62] B. Chapman, R. Buckley, N. Kemp, A. Kaiser, D. Beaglehole, and H. Trodahl, “Low- energy conductivity of PF6 —doped polypyrrole, ” Physical Review B (Condensed Mat- ter), vol. 60, pp. 13479—134831999. 159 [63] D. W. Wang, H. S. Tzou, and H.-J. Lee, “Control of nonlinear electro/elastic beam and plate systems (finite element formulation and analysis),” Journal of Vibration and Acoustics, vol. 126, pp. 63—70, 2004. [64] S. Sathiyanarayan, S. M. Sivakumar, and C. L. Rao, “Nonlinear and time-dependent electromechanical behavior of polyvinylidene fluoride,” Smart Materials And Struc- tures, vol. 15, pp. 767—781, 2006. [65] J. Bonet and R. Wood, Nonlinear continuum mechanics for finite element analysis. Cambridge University Press, 1997. [66] L. Treloar, The Physics of Rubber Elasticity. Oxford Press, 1975. [67] R. C. Batra and T. S. Geng, “Enhancement of the dynamic buckling load for a plate by using piezoceramic actuators,” Smart Materials And Structures, vol. 10, pp. 925—933, 2001. [68] J. S. Yang and R. C. Batra, “A second-order theory for piezoelectric materials,” Jour- nal of Acoustical Society of America, vol. 97, pp. 280—288, 1995. [69] S. Nemat-Nasser and Y. Wu, “Tailoring the actuation of ionic polymercmetal com- posites,” Smart Materials And Structures, vol. 15, pp. 909—923, 2006. [70] T. J. Pence and H. Tsai, “Swelling induced cavitation of elastic spheres,” Mathematics and Mechanics of Solids, vol. 11, pp. 527—551, 2006. [71] J. Merodio and R. W. Ogden, “Mechanical response of fiber-reinforced incompress- ible non-linearly elastic solids,” International Journal of Non-Linear Mechanics, vol. 40, pp. 213—227, 2005. [72] Z. Guo, X. Peng, and B. Moran, “Mechanical response of neo—hookean fiber rein- forced incompressible nonlinearly elastic solids,” International Journal of Solids and Structures, vol. 44, pp. 1949—1969, 2006. [73] J. Ding, L. Liu, G. M. Spinks, D. Zhou, G. G. Wallace, and J. Gillespie, “High per- formance conducting polymer actuators utilising a tubular geometry and helical wire interconnects,” Synthetic Metals, vol. 138, pp. 391—398, 2003. [74] S. Therkelsen, “Constitutive modeling of active polymers,” Master thesis, MIT, 2005. [75] G. Qiu and T. Pence, “Remarks on the behavior of simple directionally reinforced incompressible nonlinearly elastic solids authors,” Journal of Elasticity, vol. 49, pp. 1-30, 1997. [76] K. Chen, L. Zhao, J. Tse, and J. Rodgers, “Elastic properties of platinum rh and rh3x compounds,” Physics Letter A, vol. 331, pp. 400—403, 2004. 160 [77] [78] 179] [80] [8‘1] [82] [83] [84] [85] [86] C. Ho, 1. Raistrick, and R. Huggins, “Application of a-c techniques to the study of lithium diffusion in tungsten trioxide thin films,” Journal of The Electrochemical So- ciety, vol. 127, pp. 343—350, 1980. ~ Y. Bar-Cohen, “Electroactive polymers as artificial muscles — reality and challenges,” in Proceedings of the 42ndAmerican Institute of Aeronautics and Astronautics, 2001, pp. 2001—1492. W. Dunham, Journey Through Genius: The Great Theorems of Mathematics. John Wiley & Sons Inc., 1990. S. M. Naik, P. R. Kumar, and B. E. Ydstie, “Robust continuous-time adaptive control by parameter projection,” IEEE Transaction on Automatic Control, vol. 37, no. 2, pp. 182—197, 1992. S. M. Naik and P. R. Kumar, “Robust indirect adaptive control of time-varying plants with unmodeled dynamics and disturbances,” SIAM Journal on Control and Opti- mization, vol. 32, no. 6, pp. 1696-1725, 1994. G. Alici, P. Metz, and G. M. Spinks, “A methodology towards geometry optimisation of high performance polypyrrole (ppy) actuators,” Smart Materials and Structures, vol. 15. PP. 243—252, 2006. D. Armani, C. Liu, and N. Aluru, “Re-configurable fluid circuits by pdms elastomer micromachining,” the 12th. IEEE International Conference on Micro Electro Me- chanical Systems, pp. 222—227, 1999. S. Huang and X. Zhang, “Extension of the stoney formula for film-substrate systems with gradient stress for mems applications,” Journal OfMicromechanics And Micro- engineering, vol. 16, pp. 382—389, 2006. V. T. Srikar and S. M. Spearing, “A critical review of microscale mechanical test- ing methods used in the design of microelectromechanical systems,” Experimental Mechanics, vol. 43, pp. 238—247, 2003. ~ R. Bardell, N. Sharma, F. Forster, M. Afromowitz, and R. Penney, “Designing high- performance micro-pumps based on no-moving-parts valves,” Microelectromechani- cal Systems (MEMS), ASME, vol. 62, pp. 47—53, 1997. 161 MICHIGAN STATE UNIVERSlTV LIBRARIES IIIII1111111111111111111||111111l1illllllHlIlllIlHlHllll1 3 1293 03063 2842