“mom. In"; .3031 , cu v y uVJw 1.x ..w .x .‘r. J yx.'z 31:34, . t a 1: 1.1... 571.125.: In... hat... :5 Wu! MRI-n... :Ibit'éuu nun z¢iV mum: I!’ 41‘ a! a. .fhuii. uhhuf THESIS WNW““1111111“WW‘MHIHIUUI\IW :3 3 1293 014172 7“» This is to certify that the dissertation entitled A METHOD OF PARAMETRIC IDENTIFICATION FOR CHAOTIC SYSTEMS presented by Ching-Ming Yuan has been accepted towards fulfillment of the requirements for Ph.D. , Mechanical Engineering degree in M. r dajor professoj Date ‘61 IN}, 9( MS U is an Affirmariw Action/Equal Opportunity Institution 0- 12771 way wa— " m: LIBRARY Mlchlgan State Unlverslty PLACE N RETURN BOX to remove this Moot from your record. TO AVOID FINES Mum on or before date duo. DATE DUE DATE DUE DATE DUE I’T’I s MSU In An Affirmative Action/Equal Opportunity Institution W A METHOD OF PARAMETRIC IDENTIFICATION FOR CHAOTIC SYSTEMS By Ching-Ming Yuan A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCI' OR OF PHILOSOPHY Department of Mechanical Engineering 1995 ABSTRACT A METHOD OF PARAMETRIC IDENTIFICATION FOR CHAOTIC SYSTEMS By Citing-Ming Yuan Amethodforidenfifyingpmmeminamathemadealmodelofachaodcsystemis presented. It is an extension of an existing method for nonlinear systems with Stable periodic response. The method exploits the chaotic attractor, and extracts the unstable puiodicmbitsnomtheamactormteprecentthesystembehavim.fiachtaminthe mathematical model is expressed in a finite Fourier series using the extracted periodic- orbits,andthe harmonic-balancemethodisappliedtoformasetoflinearalgebraic equations in system parameters for least-squares estimation. This method has been successfully applied numerically to a forced Duffing oscillator, a smooth Coulomb friction system. a parametrically forced system. and a Lorenz oscillator, and experimentally to a forced oscillator with a two-well stifi'ness potential. The identified models have been verified by comparing the Lyapunov exponents. the suucuneoftheummblepaiodicorbimandmebifincadondiagramsofdwmiginal system and the identified model. To my parents ACKNOWLEDGEMENTS I am very grateful to my thesis chairman and advisor, Professor Brian Feeny, for his introduction to chaotic systems, for his understanding of my poor English, and his encouragement and guidance in the research. His promptness and attention to detail were extremely helpful. I thank Professor Alan Haddow for his excellent teaching in nonlinear mechanical vibrations, for his friendship, and for helping my family in adjusting to a foreign environment. I thank Professor Hassan Khalil, who also provided excellent teaching in linear and nonlinear control theory, and stimulated ideas in my research. I also thank professor Tien-Yien Li for participating in my thesis committee. My warm thanks are also due to Professor David S.-Y. Yen, who left my thesis committee due to a medical leave. I wish him the best for his health. A special appreciation goes to Professor Joe Cusumano and Bart Kimble at Penn State University for generously giving me the experimental data and detailed explanations of the experiment. Without this, I would probably be working on this research for another year. A warm thank you goes to my colleague Shyh-Leh Chen, who was always enthusiastic and eager to help in many ways. He provided many critical ideas and suggestions in my thesis. Also, thanks to my other colleagues, Matt Brach, C. P. Chao, J. W. Liang, and Ramana Kappagantu, for good discussions and companionship during my stay in MSU. iv Without my parents, I would not be here. My parents have never gone to school—not even a single day in their lives-~but have worked very hard to send me and my brothers to schoolforashighaneducadonuwecouldanaimmybelievethatacademic achievement is the highlight of the family history. Their spiritual support played a crucial roleinthecomplefionofthisthesialamalsoindebtedtomypuents-in-law, who provided extra financial support for my family, subsiding my worries about living in America. ’ My wife and my two daughters, Ting-fang and Yu-hua, came to this country shortly after me. Ting-fang and Yu-hua came here with no English background. and were my main oonoernotherthanmy study.'I'heyweottoSpartanVillageElementary SchooLwithgreat courage and determination. They slowly but steadily improved their English, made friends and enlarged their knowledge of American culture. Ting-fang went on to Hannah Middle SchoolasaéthgradaandtheanacDonaldMiddleSchooluahhgradadmmthe school district reorganization. Yu-hua continued through 4th grade in Spartan Village School. They now enjoy their school lives very much. I am grateful to the faculty and Stafi‘ of Spartan Village School for pmViding mold-national cultural environments, and excellent teaching for the international students. The same goes to Hannah Middle School and MacDonald Middle School too. OfcomsemywifeplayedanimportantroleregardinglifeinAmerica. Shetakescareof ourdaily lives. She isavery goodcookthhout her, livinginAmericawouldheoome more dimcult and less enjoyable. I deeply appreciated her moral support. This research was supported financially by a grant from CSIST, Taiwan. TABLE OF CONTENTS 1 Introduction 1 1.1 1.2 1.3 1.4 AnOverviewoftheParametricIdentificationMethods ............. 2 ChaoticMotion ............................................ 4 Motrvatron ................... 7 ThesisOverviews .......................................... 8 2 Methodology 10 2.1 2.2 2.3 2.4 2.5 2.6 Introduction .............................................. 10 Periodic Orbit Extraction .................................... 11 The Choice of a Mathematical Model .......................... l4 Algorithm ................................................ 17 2.4.1 Externally Excited Systems ............................ 17 2.4.2 Parametrically Excited Systems ......................... 21 2.4.3 Autonomous Systems ................................. 23 Strategy for Model Validation ................................ 24 Summary ................................................. 25 3 NumericalResults.........'............................. 26 3.1 The Forced Duffing Oscillator ................................ 27 3.1.1 Extraction of the Periodic Orbits ........................ 28 3.1.2 arousing a Mathematical Model ........................ 29 3.1.3 Identification Results ................................. 30 3.1.4 Model Verification ...... _ ............................. 31 3.1.5 Efl‘ect of Noise ...................................... 32 3.2 A Smooth Coulomb Friction System ........................... 3‘7 3.3 3.4 3.5 3.6 3.2.1 Effect of Noise ...................................... 41 A Parametrically Excited System ............................. 42 3.3.1 Effect of Noise ...................................... 46 An Autonomous System: the Lorenz Equations ................... 46 3.4.1 Effect of Noise ...................................... 51 A Case Study on Modeling the Nonlinearity with a Power Series . . . . 52 3.5.1 Using the Known Function in the Model .................. 54 3.5.2 Using the Power Series Approximation ................... 54 Conclusion ............................................... 56 4 ExperimentalResults................................... 58 4.1 4.2 4.3 4.4 4.5 4.6 4.7 4.8 4.9 Introduction ............................................... 58 Experimental Setup ......................................... 58 Phase-Space Reconstruction .................................. 60 Periodic-Orbit Extraction .................................... 64 Choosing a Mathematical Model .............................. 67 Data Processing Issues ...................................... 67 Identification Results and Model Verification .................... 69 Estimation of the Natural Frequency and the Damping Ratio ........ 74 Discussion ................................................ 76 5 ErrorsinParameterEstimates........................... 78 5.1 Introduction ............................................... 78 5.2 Errors Induced by Noise ..................................... 79 5.3 Errors Induced by the Periodic Orbit Extraction .................. 81 5.4 Sensitivity of the Parameter Estimates to Errors .................. 86 5.5 Using Several Periodic Orbits ................................. 88 6 ConclusionsandFutureWork............................ 90 6.1 Conclusions ............................................... 90 6.2 Future Work .............................................. 93 fi BIBLIOGRAPHY ........................................ 95 LIST OF TABLES Table 1 Identification results using individual periodic orbit for Duffing’s equation ....... 3 1 Table 2 Identification results using 4 periodic orbit data for Duffing’s equation .............. 32 Table 3 Identification results for Duffing’s equation using noisy data. ............................ 36 Table 4 Identification results using 4 periodic orbits for Coulomb friction system .......... 4 I Table 5 Noise effect on identification results for Coulomb friction system ...................... 42 Table 6 Identification results using 4 periodic orbits for a parametrically excited system 45 Table 7 Noise effect on identification results for a parametrically excited system ............ 4 7 Table 8 Identification results for the Lorenz equation (a) ................................................. 51 Table 9 Identification results for Lorenz equation (b) ...................................................... 52 Table 10 Identification results for Lorenz equation with 1% noise ................................... 52 Table 11 Force and response in model (3.13) .................................................................... 53 Table 12 Identification results using the exact function in Model (3.15) ........................... 54 Table 13 Identification results using power series in Model (3.16) ................................... 55 Table 14 Identification results for the experimental system. ............................................. 69 Table 15 Comparison of the natural frequency and the damping ratio ............................... 7o LIST OF FIGURES Figure 2.1 (a) A close recurrence of a chaotic trajectory, and (b) a precise recurrence after a gentle adjustment of the starting point of a chaotic trajectory ....... 12 Figure 3.1 Phase portrait of the Duffing oscillator ........................... 28 Figure 3.2 Some extracted periodic orbits of the Dufi‘rng oscillator .............. 29 Figure 3.3 The simulated chaotic attractor and some of the extracted periodic orbits of the identified model ........................................... 33 Figure 3.4 Bifurcation diagrams of the Duffing’s equation (a) using the original equation, and (b) using the identified model with the average values in Table 2 . 34 Figure 3.5 (a) A noise-free periodic orbit, (b) the noise-contaminated counterpart . . 35 Figure 3.6 The simulated chaotic attractor and some of the extracted periodic orbits of the identified model using the noise-contaminated periodic orbits ....... 37 Figure 3.7 Bifurcation diagrams of the Duffing’s equation (a) using the original equation, (b) using the identified model of the noisy case ................... 38 Figure 3.8 Phase portrait of a Coulomb friction system ....................... 39 Figure 3.9 Some extracted periodic orbits of a Coulomb friction system .......... 40 Figure 3.10 Phase portrait of the parametrically excited system .................. 43 Figure 3.11 Some extracted periodic orbits of the parametrically excited system . . . . 44 Figure 3.12 Phase portrait of a Lorenz system ............................... 48 Figure 3.13 Recurrence plot of the Lorenz system (period lengths are indicated in the num- bers of time steps used in the numerical integration) ............... 49 Figure 3.14 Two extracted periodic orbits of a Lorenz system: (a) period length of 110 time steps, (h) period length of 144 time steps .................... 50 Figtne 3.15 Nonlinear function in a power series ............................. 56 Figure 4.1 Sketch of the experimental setup ................................ 59 Figure 4.2 Autocorrelation function of the experimental data ................... 61 X Figure 4.3 Correlation function of the experimental data . . . . . l ................ 63 Figure 4.4 Singular system analysis of the experimental data ................... 64 Figure 4.5 Reconstructed phase space of the experimental system ............... 65 Figure 4.6 Reconstructed phase space in singular coordinates .................. 65 Figure 4.7 Some extracted periodic orbits from the reconstructed phase space ..... 66 Figure 4.8 Some extracted periodic orbits in singular coordinates ............... 66 Figure 4.9 Qualitative nonlinear function of the experimental system ............ 71 Figure 4.10 Bifurcation diagram of the identified model, (a) with the cubic nonlinear term in the model, and (b) with the fourth power term in the model ....... 72 Figure 4.11 Phase portrait of the identified model with cubic nonlinearity ......... 73 Figure 4.12 Some periodic orbits extracted from the phase portrait of Figure 4.11 . . . 73 Figure 4.13 Transfer function of the right well of the experimental system ......... 75 Figure 5.1 Close look of the periodic orbit extraction on the Poincare section ...... 81 Figure 5.2 A sketch of the construction of a linearized map .................... 84 CHAPTER 1 Introduction Parametric identification deals with the problem of determining the values of the parameters in a mathematical model that represents a dynamical system, based on the observed data from the system. It is a field of increasing interest, in part because of applications in prediction and control. Linear models have dominated in the description of the dynamical systems and control theoretic approaches. Very complex and randomlike behavior has been viewed from a statistical perspective in which many degrees of freedom were involved. More recently nonlinear models have emerged, capable of describing chaotic dynamics and other nonlinear behaviors. Such systems can exhibit extremely complex dynamical behavior, even though the underlying dynamics may be low dimensional. On the other hand, high- dimensional systems such as fluids and lasers can Show simple and low—dimensional dynamics, which may be described by a low dimensional model [43]. Chaotic motion features the sensitive dependence on initial conditions. Nearby orbits that cannot be distinguished will diverge exponentially and soon become uncorrelated. Along with new theoretical concepts have come practical techniques, such as Lyapunov 2 exponents and fractal dimension, for characterizing the dynamics of such systems. Yet, the techniques for identifying the parameters of a chaotic system are not as well developed as those for analyzing its dynamical behaviors. We briefly review parametric identification methods below. 1.1 An Overview of the Parametric Identification Methods Parametric identification work generally presupposes that a mathematical model has been chosen to represent a nonlinear system and that the goal is to identify the unknown parameters in the given model. The unknown parameters are determined by optimizing in some sense the fit of the chosen model to the available data. For linear systems, the superposition principle can be applied to the system response and the transfer function that characterizes the system behavior can be obtained by a variety of techniques, such as transient analysis, frequency analysis, correlation analysis, and spectral analysis [37]. The system parameters are estimated by a curve fitting of the transfer function. For a nonlinear system, the techniques for linear systems fail fundamentally because the superposition principle is no longer applicable. However, for small nonlinearities, perturbation techniques were widely used in analyzing the system response and in identifying the system parameters as well. For example, Hanagud et al. [32] used the method of multiple scales to determine the nominal system response, which was used iteratively to estimate parameters. Nayfeh [45] and Feeny er al. [23] used the method of multiple scales to exploit resonances and produce expressions relating the parameters to the experimentally measured nonlinear behavior such as jump phenomena and nonlinear drifi. The parameters could then be determined algebraically, or in a least-squares sense. 3 Ibanez [33] used a describing function method to construct an approximate transfer function of the nonlinear structural system and hence uncouple the original nonlinear equations. System parameters were obtained by iteratively minimizing the error function between the measured data and the theoretical solution. Gottlieh et al. [27] used the Hilbert Transform in their parametric identification of weakly nonlinear systems. Another approach proposed by Mook er al. [41, 42, 52], called the method of minimum model error (MME), combined the assumed model with the measurements to determine the correct form of model for the nonlinear system under investigation. A correction term which represented the model error was added to the assumed model and a cost function was formed. By minimizing the cost function, a two—point-boundary-value problem was formed and yielded the correction term, which was then fitted to an assumed polynomial form to obtain the correct model of the nonlinear system. Mohamrmd [40] used a direct approach by assuming a general form of the equation to represent the nonlinear system under investigation. By measuring all of the system responses, such as acceleration, velocity, and position, and directly introducing them into the assumed equation, a set of algebraic equations was formed by balancing these measurements and the input function. System parameters were then estimated by a singular-value decomposition method. In a similar direct approach, Yasuda et al. [59, 60, 61] represented the system nonlinearity as a sum of polynomials in the system equation, with unknown coeficients as the system parameters to be determined. Periodic responses under periodic excitation were measured and expressed in Fourier series. The harmonic balance method was used to balance the Fourier coemcients of each harmonic and a set of algebraic equations in the system 4 parameters was formed. The system parameters were then eStimated by a least squares fit to the algebraic equations. Recently, methods for modelling a chaotic system and identifying the system parameters based on experimental data have been developed. Abarbanel er a1. [1] proposed a method for constructing a parameterized map which evolved points in the phase space into the future. This map was regarded as a dynamic system. and the parameter values were chosen through a least-squares optimization procedure, constrained by the invariants of the system, such as the Lyapunov exponents. Eisenhammer et al. [20] proposed a trajectory method to extract ordinary difi'erential equations from an experimental time series. The experimental data were represented in a state space and the corresponding flow vectors were approximated by polynomials of the state vector components. Starting from appropriately chosen initial states, the model equation was used to obtain an estimation of the states for later times, and the coefficients were fitted by minimizing the distances between the states predicted by the model and the experimental states. Breeden and Hubler [6] proposed a flow method for reconstructing a set of coupled maps or ordinary differential equations from a trajectory of the system in state space. By choosing some trial coefficients for a series expansion in the state variable, the error in these parameters were computed by comparing the predicted values and the experimental values. The parameters of the model were obtained by solving a chi-squared minimization problem. 1.2 Chaotic Motion Chaos was known by Henri Poincare (1854-1912) about a century ago in the course of his investigations on the three-body problem. Through his discovery of homoclinic solutions (homoclinic intersection, or homoclinic tangles), Poincare showed that the three-body 5 problem has no solutions of the type envisioned by Jacobi or Hamilton, in the sense that a small error in the initial conditions produced an enormous error in the final response. In his book, New Methods of Celestial Mechanics, Poincare motel: When we try to represent the figure formed by these two curves and their intersections in a finite number, each of which corresponds to a doubly asymptotic solution, these intersections form a we of trellis, tissue, or grid with infinitely serrated mesh. Neither of the two curves must ever cut across itself again, but it must bend back upon itself in a very complex manner in order to cut across all of the meshes in the grid an infinite number of times. The complexity of this figure will be striking, and I shall not even try to draw it. Nothing is more suitable for providing us with an idea of the complex nature of the three-body problem, and of all the problems of dynamics in general, where there is no uniform integral and where the Bohlin series are divergent. In the words of modern dynamical systems dreary, the solution is sensitive to initial conditions due to the inherent stretching and folding process of the nonlinear dynamics. This sensitivity to initial conditions makes the nearby states on the attractor divergent exponentially on the average, and results in a long-term unpredictability emanating from a small amount of uncertainty in the initial conditions. Confronted with his discovery of the homoclinic solution, Poincare went on inventing several theories for new branches of mathematics, including topology, ergodic theory, homology theory, and the qualitative theory of difi‘erential equations. He also pointed out the possible uses of periodic orbits in characterizing his discoveryzz 1. “mummaanymuapawwmuwmommmcm 1992forrnore historicalcornmatts[53]. 6 there is a zero probability for the initial conditions of the motion to be precisely those corresponding to a periodic solution. However, it can happen that they difi’er very little from them, and this takes place precisely in the case where the old methods are no longer applicable. We can then advantageously take the periodic solution as first approximation, as intermediate orbit, to use Gylden’s language... Given equations of the form defined in art. I 3 and any particular solution of these equations, we can always find a periodic solution (whose period, it is true, is very long), such that the difi'erence between the two solutions is as small as we wish, during as long a time as we wish. In addition, these periodic solutions are so valuable for us because they are, so to say, the only breach by which we may attempt to enter an area heretofore deemed inaccessible. The periodic orbits are dense in the chaotic attractor, and all of them are unstable. This is a characteristic sign of chaos that only the presence of unstable periodic orbits but absence of the stable ones [2]. The periodic orbit theme has been pursued by many authors in modern dynamical system theory in characterizing a chaotic attractor [5, 19, 35, 49, 53], and in the course of controlling a chaotic system [16, 47, 50]. The unstable periodic orbits have also been used in system identification [31], and in recognizing parameter variations [35]. We use them as a major tool in our parametric identification scheme for a chaotic system. Chaotic signals have been discarded in the past as ‘noise’. But, as pointed out by Abarbanel [2], “chaos is not an aspect of physical systems which is to be located and discarded, but is an attribute of physical behavior which is quite common and whose 2. MncKly, R. and Main, J.,eds., Hamiltonian dynamical systems (Adam Hilger, Philadelphia, 1987), cited from ‘l‘ufil- m Abbott, an! Reilly inAnExperimental qpmaeh toNonlinearDynamicsandChaos.1992[53] 7 utilization for science and technology is just beginning’. It has been discovered in many nonlinear systems in the laboratory and in the mathematical models in the past two decades, and has become a well-known phenomenon and an important subject in modern dynamical system theory. Much of the work has concentrated on learning how to classify the nonlinear systems by analyzing the output from known systems. These efi‘orts have provided, and continue to provide, significant insights into the kinds of behavior which one might expect from nonlinear dynamical systems, and have led to an ability to evaluate now familiar quantities such as fractal dimension, Lyapunov exponents, and other invariants of the nonlinear systems [2]. Efl'orts have also been extended to predicting and controlling the chaotic behaviors. For example, Farmer and Sidorowich [20] proposed a local approximation approach for predicting a short-term chaotic time series using the nearby states. Du et al. [47], Ditto et al. [16], and Shinbrot et al. [50] tried to control the chaos by exploiting the periodic orbits embedded in a chaotic attractor and perturbing some parameters of the system, so as to stabilize one of the unstable periodic orbits, making the system become stable and more flexible under different operating conditions. Cusumano and Sharkady [l2] experimentally studied the bifurcation and dimensionality of a chaotic attractor occurred in a low dimensional parametric-excited system, and built a valid model for the physical system. This trend of study shows that the chaotic motion may be often regarded as an annoyance, yet it provides an extremely useful capability without counterpart in non-chaotic systems. 1.3 Motivation Chaos is inherent to nonlinear dynamical systems, and is rich in information content as compared to a periodic trajectory. This richness has been exploited in dimensionality 8 studies, nonlinear prediction, and control schemes as stated in the previous section. The potential usage of chaotic system in parametric identification has not been fully exploited, because of the sensitive dependence on initial conditions and the long-term unpredictability. It is well known that a chaotic attractor is the closure of the set of unstable periodic orbits [5]. They can be extracted and used to characterize the chaotic attractor [5, 19, 35, 49, 53], and hence can be used for identifying the system parameters, because they are the solution to the system equation. Meanwhile, Yasuda and co-workers [59, 60, 61] have demonstrated that the stable periodic solution to a nonlinear system can be used to identify the system parameters. This inspires us to explore the applicability of the unstable periodic orbits in a parametric identification scheme for a chaotic system. 1.4 Thesis Overview In Chapter Two, we describe the methodology for identifying the parameters of a chaotic system. A mathematical model is chosen to fit the characteristic of the original system from which the chaotic data are obtained. The unstable periodic orbits are extracted from a chaotic set for use in the identification algorithm. Then the harmonic-balance method is applied to form a set of linear algebraic equations in system parameters, which are then solved by a least-squares fit. This approach is applied to different kinds of nonlinear systems, such as externally excited, parametrically excited, and autonomous systems. Chapter Three contains the identification results for several numerical examples. Chaotic data are generated numerically from known governing equations. Mathematical models are chosen in polynomial form generally if no knowledge about the system nonlinearity is 9 available. Random noise is added to the periodic data to assess its efi‘ects on the identification results. The identified models are verified by comparing the Lyapunov exponents, the structure of the unstable periodic orbits, and the bifurcation diagrams of the original system and the identified one. In Chapter Four, we apply the method to a set of experimental data, taken from J. P. Cusumano and B. W. Kimble at Pennsylvania State University. The phase space is reconstructed by the method of delays, from which the unstable periodic orbits are extracted for use in the identification procedtu'e. A model is obtained and verified for the experimental system. Two sources of error, noise and the extraction of the periodic orbits, in the identification process are discussed in Chapter Five. We examine a bound on the error in the Fourier coefficients induced by the noise. We examine the extraction of the unstable periodic orbits closely, and establish a bound on the deviation of the extracted periodic orbit from the real one. We discuss the sensitivity of the errors induced by the noise and the extraction of the unstable periodic orbit to the identification results. Chapter Six contains some conclusions and future work. CHAPTER 2 Methodology 2.1 Introduction Parametric identification method is not well developed for a chaotic system, partly because in the past chaos has been treated as noise to be discarded, and partly because chaotic motion exhibits sensitive dependence on initial conditions and defies long-term predictability. Traditional usage of time series data in a parametric identification scheme for non-chaotic systems may not be appropriate for chaotic systems, because of sensitivity to initial conditions. However, a chaotic system features a chaotic attractorl, in which infinitely many unstable periodic orbits are present, but absent of stable ones [2]. These unstable periodic orbits can be extracted and used to characterize the chaofic attractor [5, 35]. They provide a skeleton of the chaotic set, which can be used in characterizing a chaotic system. Each unstable periodic orbit is a “solution” to the system which generated the chaofic set. Once a periodic orbit is extracted, each term in the mathematical model can be expressed in a Fourier series, and the harmonic-balance method can be applied to form a set of 1. Mmhmuficfingsetwhichconmadaueubinhisdiffiadtmshowmmpluthetaderrseorbit Mmdmfactmmy oftlrenmnericallyobeavedfittrectorfmaymtbetrueattncm butmerely attractingsets. Weusethistermlooselytodenoteasetofpoirrtsinplmespecetowudwhiehatirnehistoryapproechesaftertnn- mdcmkGuckenhermenndHohnee [301andMoorr[43] forstrictmathernaticaldefinitionandexamphs. 10 11 algebraic equations, in which the system parameters are estimated by a least-squares method. The harmonic-balance method has also been used as a parametric identification technique by Yasuda an coworkers [59, 60, 61] for some nonlinear systems that have stable periodic response. We extend this technique to chaotic systems, where unstable periodic orbits are abundant in the chaoric attractor. In this chapter, we will demonstrate the methodology in detail with three kinds of chaotic systems, categorized as externally excited, parametrically excited, and autonomous. They are treated difl'erently because the excitation can affect the formulation of the identification problem. We will discuss two important issues to our identification method. the extraction of the unstable periodic orbits and the choice of a valid mathematical model. We will also discuss the method for model verification. 2.2 Periodic Orbit Extraction The genesis of a chaotic trajectory can be visualized as a random walk on the union of infinitely many periodic orbits [14]. A physical trajectory approaches the saddle orbit along its stable manifold, and remains nearby for a time before it is thrown out along the unstable direction. It wanders around the union of periodic orbits, tracing out a strange1 attractor [14]. When the trajectory is near a periodic orbit, it approximately follows the motion of that periodic orbit for an interval of time. If this time interval exceeds the period of the reference orbit, the trajectory exhibits a recrurence. This property can be used to 1. mmnmmgemmfarmgmdBmmmmephnewmwhiehchmfieubhemehm withageornetricobjectcalledefncteleet,whileadraoticefiractorfienotingebomdedmfionthntiemitiveb changeehrhrifialeondifim,huuleastoneposidvel.yapmovexpmrau[43].Inouupecifictnrrpoee,weintendnot todistinguishdwdifiamhnmdunmuchmgablymmfawdnm-tambehavbrofdnnmlmm 12 approximate the positions of the unstable periodic orbits embedded within the attractor [5, 35]. Figure 2.1 is a sketch of a recurrent three-dimensional flow in the vicinity of a hyperbolic periodic orbit [53]. The chaotic trajectory has a recurrent segment, shown in Figure 2.1(a), which is very close to an unstable periodic orbit, shown in Figure 2.1(b). We can gently adjust the “starting point” of the trajectory segment so that the segment nearly coincides with the unstable periodic orbit and returns almost precisely to its starting point [14, 53]. This idea has been used in a control scheme to stabilize one of the unstable periodic orbits for a chaotic system [16, 47, 50]. In practice, we may have a sufficiently large chaotic data set {xi} , i = 1, ...N, in state space. We scan the data set for recurrences by seeking points that come within a specified spatial distance a of one another after a fixed elapsed time, such that [5, 35] Figure 2.1: (a) A close recurrence of a chaotic trajectory, and (b) a precise recurrence after a gentle adjustment of the starting point of a chaotic trajectory. 13 Putt-'19] $6, (2.1) where K is the number of points per period of the unstable periodic orbit. If x‘. is a recurrent point, then x‘. H, xi +2, , are likely to be “near” the unstable periodic orbit. Thus, the segment of data, {x‘.,x‘.+ v ...,x‘.+K_1}, is then taken as the ‘approximated’ unstable periodic orbit. The ‘true’ one is generally unobtainable. The value i is the starting point of the unstable periodic orbit, and is related to the phase angle associated with that periodic orbit relative to the forcing function. It is important to record the phase angle for later use in the calculation of the Fourier coefficients. This will become clear when we do the calculation. In a periodically forced system. all periodic orbits have a period that is an integer multiple of the forcing period such that K = no, 2no, 3n0..., where n0 is the number of points per forcing period [30]. However, in an autonomous system, such as in the Lorenz oscillator, there is no such forcing period. Instead there are infinitely many unstable periodic orbits with incommensurate periods. These incommensurate periods can be obtained using a recurrence plot, which can be constructed by varying the period length and counting the number of recurrent points found for each period length. The recurrence periods will be clustered around certain values, indicating the periodicity of the periodic orbits and hence the number of points in a period, which is then used as the fixed elapsed time K in Eq. (2.1) for locating the periodic orbits, and also used as the fundamental period in the calculation of their Fourier coefficients. This procedure is quite successful in finding the unstable periodic orbits in many chaotic 14 systems. However, if the positive Lyapunov exponent associated with the unstable orbit is large, then in one period the orbit will most likely have so departed from the unstable periodic structure in phase space that one will probably not be able to identify the unstable periodic orbit within the spatial criterion. In such case, the spatial distance criterion 8 can then be relaxed to include more points in the neighborhood. For finding low-order periodic orbits, say less than ten, it is adequate to set a to be 0.5% of the maximum extent of the chaotic attractor [5, 35]. The searching process for the periodic orbits may reveal several distinct unstable periodic orbits with the same period number. Nonetheless, all extracted periodic orbits can be used in the parametric identification algorithm. 2.3 The Choice of a Mathematical Model For the task of parametric identification, it is important to choose a valid mathematical model to represent the physical system from which the measurements are taken. To do this, we need to know the order of the system and the form of the system nonlinearity. For typical mechanical vibratory systems, the system order is twice the number of degrees of freedom. Also, for a nonlinear system to be deterministically chaotic, the system has to be three or more dimensional. For a forced single-degree-of-freedom system, a general mathematical model can be written as mx+f(1.x.t) = 0. (22) where the time variable is taken as an additional dimension. For an autonomous chaotic system, a general mathematical model can be written as Y = My) . (2.3) 15 forye R",n23. In the forced case, the general function f (x, x, t) can take some specific forms when the excitation and the nonlinearity of the system are known. For example, a system is externally excited if it is modeled with an inhomogeneous term in the governing equation, and parametrically excited if the system difi‘erential equations have time-varying coefficients. Also, the form of the system nonlinearity can be determined using the physical law that governs the system dynamics and the background knowledge about the physical system. For example, sin(x) is usually used to model a pendulum system. A power series can be used to model a system with an unknown smooth nonlinear function. Whenever possible, models based on the physical mechanism should be employed. Thus, for an externally excited single-degree-of-freedom nonlinear system, Eq. (2.2) can be recast more specifically as P m! + 2 Bi,- (x. x) = E (t) . (2.4) i I l where E(t) is a known external excitation, fi (x, 2) are some known functions of x and x , p is the number of nonlinear terms in the system model, and m and B,- are the unknown parameters to be determined. For a parametrically excited nonlinear system, Eq. (2.2) can be recast as r Pt mx+ 2,2,0) { 2 Myrna} = o. (25) 8'81 jlll where g ,- (t) are the known parametric excitation functions, fil- (x, x) are some known 16 functions of x and x , r is the number of excitation terms, p i are the numbers of the nonlinear terms associated with each excitation functions, m and Bi]. are the unknown parameters to be identified. And for an autonomous system, Eq. (2.3) can be written as Pt y,- = Z Bijhgj (”ti = 1, ..., n, (n 2 3) . (2.6) fall where y = [y], ..., yn] T, and hij (y) are the nonlinear functions of the state variable y, and p,- are the number of terms in each equation of the model. If the form of the system nonlinearity in Eq. (2.4) and (2.5) is unknown, but can be assumed as a smooth function, and the system is operated in the neighborhood of the equilibrium point, then the unknown function can be approximated by a truncated power series. This is reasonable, because any smooth function can be represented by a power series in some neighborhood of the origin (equilibrium). However, this approximation by a power series may be accompanied by issues such as convergence and optimal truncation. Ideas of convergence and divergence make sense when we consider infinite series. Since we are using a truncated series, these ideas are not critical. If a power series indeed converges to our function to be identified, then it is best to use as many terms as possible without introducing numerical problems associated with large exponents. If the underlying function has a divergent power series in the range of data, then there would be some optimal truncation which is unknown. Thus, an imperfect identification result seems to be the norm. 17 2.4 Algorithm 2.4.1 Externally Excited Systems The mathematical model for an externally excited single-degree—of-freedom system is chosen as Eq. (2.4) P mx+ 2‘, B,f,-(x.x) = Em. (2.7) i I 1 where the external exciting function, E(t), is considered to be periodic with single frequency to , such as E (t) a aeos (art) . (2.8) Upon extraction of the periodic orbits from the chaotic attractor, the nonlinear functions become periodic and can be expressed in Fourier series, such as xp (t) =a — 202;] {a cos(-k—‘) + b sing—i”) (2.9) 15,, (t) -.‘=. i (L?) {bjcos(‘Z-(£—t)—ajsin(J-—:”)} (2.10) j- 1 it"E‘§.(’#2-°)’ijs(’—1”)+b8141—?) fi(x,X)p2% 0‘r- 2 {cijcos owl?” til)+d sin(L(£—t)} (2.12) j at with the Fourier coefficients calculated as mttl balar, (2.13) (2.14) dij= cij =::%-.J: f(x,Jt) Pcos(— 1: )dt 2 7' where the subscript p denotes the function being evaluated using the period-k data, T is the period of the employed periodic orbit, and t) is the phase angle of the extracted periodic orbit relative to the forcing. Since the phase angle has been included implicitly in the periodic-orbit data and the nonlinear functions (the beginning of the periodic orbit is the index of the phase angle), the limits of integration in Eq. (2.13) and (2.14) are used in the numerical integration of the data. Ignoring the phase angle will cause an inconsistency in the Fourier series representation in Eq. (2.9) to (2.12), and consequently produce incorrect identification results. Substituting these Fourier series into the model equation (2.7), and balancing the Fourier coefficients of Eq. (2.13) through (2.14) of any set of harmonics, a set of linear algebraic equations in system parameters can be constructed. This usage of the harmonic balance method contrasts its usual usage for response analysis, where the ordinary difi‘erential equation is known, and the efi‘ort is to solve a set of nonlinear equations in Fourier coefficients. For systems forced with a single harmonic, and for autonomous systems, the method of harmonic balance requires nonlinearity so that several harmonics can be balanced. In this thesis, we typically use the multiples of the primary harmonic. Thus, for the wh COI' rcpn matn‘ H2n 19 example of k = l , the balance equations, in matrix form, are .- 0 COO Cm)1 ,. F 1 2 m.T 0 2 1 -0) b1 dll dpl B: ____ 0 (2.15) 2 _(nm) a. c1" c” LBP 3 2 -' l. .. :(nm) bu d1" dP'U 01'. At! = q, (2.16) where a is the parameter vector of the system model; A is a (2n+1) x (p+l) coemcient matrix, with each column containing the Fourier coefficients of the corresponding term in the system model; q is a (p + 1) vector, containing the Fourier coefficients of the external forcing function, which contains a non-zero element a in our periodic excitation case; and n is the number of terms retained in the Fourier series representation. For general values of k, the indices and frequencies in the elements of matrix A are scaled by k. If 2n = p and the matrix A is non-singular, the parameter vector u can be determined uniquely. In practice, it is statistically better if the algebraic equation of Eq. (2.16) is overdetermined, so that 2n > p . Consequently the exact solution will not generally exist, but a best solution can be obtained by a method such as a least-squares fit. We seek a solution that can minimize the average error in all of the equations. The error function is most conveniently chosen as the sum of squares, or defined as [4] 20 e :2 IAa—ql. (2-17) The solution that minimizes Eq. (2.17) is called the least-squares solution of the linear system. The minimization of the error function is performed by setting the partial derivatives of the squared error function with respect to the parameters equal to zero, i.e. an/aot = 0 , which leads to the so-called normal equations as ATMOI- q) = 0. (2.13) and the least-squares solution of the parameters vector u is —r a = (ATA) ATq. (2.19) Since the operation of a matrix inversion is less accruate and time consuming, the normal equation is often not recommended in the numerical implementation. The most general least-squares solution using the singular-value decomposition method is a = Vz’rUT q = Aiq, (2.20) where through the singular-value decomposition, A = UZVT , and AT is its pseudo- inverse; U and V are the orthogonal matrices with each column consisting of the left and the right singular vectors of matrix A respectively; and ET is the pseudo-inverse of 2 , which has the non-negative diagonal elements being the inverse of that of the corresponding terms in 2. (See, for examples, Atkinson [4] and Strang [51] for a geometric discussion). Here arises a question as to how many terms should be retained in the Fourier series representation of a periodic solution. Theoretically, the number of terms in the Fourier series should be infinite, but Mickens [39] has shown that the upper bounds of the absolute 21 magnitudes of the harmonic coefficients decrease exponentially, such that they become ineffective in the least-squares estimation procedures. We found that the rule of thumb for retaining the number of terms in the Fourier series is SSnSS. can where n is the number of harmonics of the primary (driving) frequency. This limits the number of unknown parameters in the model that can be estimated using a single periodic orbit. However, we can use several periodic orbits to form several sets of algebraic equations, thereby augmenting the matrix A to increase the redundancy of algebraic equations for the least-squares estimation. This treatment can improve estimation result even if the number of unknown parameters is not excessively large. Moreover, when the parameter set is small, each set of algebraic equations formed by individual periodic orbit can be used separately to obtain statistical information such as mean values and standard deviations. This availability of several extracted periodic orbits fiom a chaotic set increases the applicability beyond that of a simple periodic response, such as the case by Yasuda and coworkers [59, 60, 61]. 2.4.2 Parametrically Excited Systems A parametrically excited system has time-varying coefficients in the governing equations of motion. Examples 'of this kind of nonlinear system are a pendulum with a moving support [46], a column with an axial time-varying force [46], and a flexible beam under an electromagnetic force [12]. Previous studies have focused on dynamic stability, in which the introduction of a small vibrational loading can stabilize (destabilize) a system which was statically unstable (stable) [46]. Recent studies show that the system can exhibit !f char The cho Sid: {ac div aPP Usi COT. Wit Hen thg. 22 chaotic behavior in a large range of parameters [12, 38]. The mathematical model for a single-degree-of-fi'eedom parametrically excited system is chosen as Eq. (2.5), r Pr mx'+ 2 gi(t) {2 a... 0(a)} = o. (2.22) i-l j-l This model is a degenerate case of a parameter estimation problem because the right-hand side of the equation is zero. Also, the Fourier series representations must account for the fact that the system variables are coupled with a time-varying function. To proceed, we divide through by m. The 1: term is taken as a known quantity by the fact that the approximate periodic solution of the original system is known, and moved to the right- hand side of the equation. Using the extracted periodic orbits, the evaluated nonlinear functions in the model are periodic. The excitation functions in time and the nonlinear functions in x and x are combined together when they are to be expressed in Fourier series, such that ~ c.. " 8.30:.1. t) = g,o(t)f,-j (x. x) = 45+ 2 cijkcos(kmt) +dmsin (kart), (2.23) t-r with the Fourier coefficients calculated as 2 IT“- 2 +¢ (224) dijt = 7‘ ¢ it,- (1,1,!) sin(lttnt)dt Here, the phase angle is included in the combined nonlinear function g”. (x, x, I) through the variable x. The limits of integration are chosen to match with the phase angle of the CXt Sui COR 2.4 far: We Perle Cant pl’lasC 23 extracted periodic orbits. Substituting the Fourier series into the model equation (2.22), and balancing the Fourier coefficients of each harmonic, a set of algebraic equations such as Eq. (2.16) can be constructed, and the parameters can be estimated by a least-squares fit. 2.4.3 Autonomous Systems An autonomous system of dimension three or more can exhibit chaotic behavior [30]. The famous example is the Lorenz equation, given by Y1 = 0(Y2’yl) Y2 = pyl‘yz-ylygg (225) Y3 = "BY3+YIY2 There exists a ‘butterfiy shaped’ chaotic attractor, in some region of the parameter space, which consists densely of infinite many unstable periodic orbits whose periods are incommensurate. To extract the unstable periodic orbits, the incommensurate periods have to be determined by constructing a recrnrence plot, as stated in section 2.2. The mathematical model for an autonomous system is chosen as Eq. (2.6). It can be chosen more specifically if we know the type of the autonomous system under investigation. In an experiment, each state variable y‘. must be measured. Using the periodic orbits extracted from the chaotic attractor, each term in the model is periodic and can be expressed in a Fourier series with the fundamental frequency as the one obtained from the recurrence plot The Fourier coefficients are calculated as before, except the phase angle can be ignored since there is no forcing function involved. Treating the y‘. - the me anc‘ inv: [11] [27] We data Data The tech, knov 24 terms as known quantities, and balancing the Fourier coeflicients of each harmonic in each equation, a set of algebraic equations of the form of Eq. (2.16) is formed. The system parameters are then obtained using a least-squares fit. 2.5 Strategy for Model Validation The final step, and perhaps the most difficult step, in parametric identification procedure is the model validation. The objectives of validation are to seek answers to questions such as: Is the identified model adequate? Under what conditions is the model representative the system? Traditionally, the method of model validation is to simulate both system and model under similar conditions and compare the respective responses. This is subjective and lacks consistency for chaotic systems, due to the system’s sensitivity to the initial conditions [3, 57]. More sophisticated criteria based on geometrical and statistical invariants have been proposed, such as embedding trajectories [9, 43], Poincare sections [11], bifurcation diagrams [3], Lyapunov exponents [1, 58], and the correlation dimension [27], to characterize and compare reconstructed attractor and identified model. We will seek consistency of the identification results from using different periodic-orbit data sets. This is the most convenient way to check the quality of the identified parameters. The positive Lyapunov exponent is an invariant quantity of a chaotic system. Several techniques have been developed into algorithm for estimating Lyapunov exponents from a known dynamical system or from observable [5, 19, 35, 58]. We use the computer codes by Wolf et al. [58] to calculate the Lyapunov exponents, which will be used in verifying the identified model. We will also compare the structure of the unstable periodic orbits that are extracted from the l The chao bylk 2.6 VVbli unsui cone: physi 25 the original attractor and from the one generated from the identified model. This is reasonable because the unstable periodic orbits are the skeleton of the chaotic attractor. The structure of the periodic orbits would provide some geometric information of the chaotic system that is useful to assess the quality of the identified model. The criterion of bifurcation diagrams of the system and the identified model, as suggested by Aguirre and Billings [3], will also be used as a supplementary criterion when available. 2.6 Summary We have outlined a scheme for identifying the parameters of chaotic systems by using the unstable periodic orbits that are extracted from the chaotic attractor. The method is simple conceptually and easy to implement. Models are chosen based on the knowledge of the physical system, or on approximation by a power series. Bach term in the mathematical model is expressed in a Fourier series, and the Fourier coeficients of each harmonic are balanced to form a set of algebraic equations in system parameters, which are estimated by a least-squares method. Methods for model verification are proposed. This is an extension of an existing method, previously applied to systems with a stable periodic response [59, 60, 61], to chaotic systems. By using the unstable periodic orbits, the method exploits the structure of the chaotic set. Thus, we overcome issues such as sensitivity to initial conditions. Cl N1 In th math to ex cmbc find: set of esu'm ffiCtiu talter: Swen “first imm'a 53’s [cl-11 CHAPTER 3 Numerical Results In the previous chapter, we presented an approach for identifying parameters in a mathematical model of a nonlinear system that exhibits chaotic behavior. The strategy is to exploit the chaotic attractor of the system and extract the unstable periodic orbits embedded within it. The extracted periodic orbits are used to express each term in the model in a Fourier series, and their coefficients of each harmonic are balanced to form a set of algebraic equations in system parameters, which are then obtained by a least-square estimation. This approach can be applied to a general class of nonlinear systems with smooth nonlinearity. In this chapter, numerical studies on the forced Duffing oscillator, a smoorh Coulomb friction system, a nonlinear parametrically excited system, and a Lorenz equation, are taken to demonstrate the applicability of this approach. Numerical integration of the governing equations is carried out using a Stir-order Runge-Kutta method on a Sun workstation. Typically, 50000 chaotic data points are generated for with a time step interval of one-100th of the forcing period, or with 0.005 time step size in autonomous systems. 3.1 The It is 27 3.1 The Forced Duffing Oscillator The forced Dumng oscillator is given as mx' + cx + Bx + 1x3 =- acostnt. (3.1) It is a classic differential equation that has been used to model the nonlinear dynamics of mechanical and electrical systems. With B = 0, Eq. (3.1) is a model for a circuit with a nonlinear inductor [55, 56], and with B<0,y>0, it is a model for the postbuckling vibrations of an elastic column under compressive loads [44]. It can be written as a set of first-order differential equations 11 = xz/m xz = — cx2 - 311- 'yxl3 + acoscut (32) to fit the format of the computer integration routine in public libraries such as IMSL. This equation admits chaotic motions for a large range of parameters. We choose the parametervaluesas m =1, c = 0.2, B = y =1,andtheforcingtermasa = 27,and to = 1.33 [54]. These parameters are to be estimated by the present method. Using the numerical data generated from the governing equation, a phase portrait is constructed as shown in Figure 3.1. We see that the trajectory wanders around the phase space in the attracting set. Any initial condition within the basin of attraction leads to the same qualitative appearance in the phase space. This is the attracting set from which the unstable periodic orbits are to be extracted. Also, the Lyapunov exponents, indicating the average exponential rates of divergence or convergence of nearby orbits in phase space, are calculated using the computer code by Wolf at al. [58]. They converge to r, = 0.13, and x, = —0.468, indicating that the 28 15 7 Y T I T V T f 10 ‘2‘ “If“ —10 ‘P j, I XOP .e n.- al- (I Figure 3.1 Phase portrait of the Duffing oscillator system of Eq. (3.2) is chaotic, since there is one positive value and the sum of them is negative. (There is a zero exponent from the computer code, corresponding to the time variable of the vector field. We omit it for convenience). 3.1.1 Extraction of the Periodic Orbits The unstable periodic orbits can be extracted using the recurrence property of the chaotic attractor, as stated in previous chapter. We repeat the idea here to emphasize its importance. We scan the data set in state space forward to locate the recurrent points that are close within a spatial distance a , such that Pt + K—x‘l s c (3.3) for a periodic orbit with K data points in the orbit. Here the index i is taken as the phase angle of the periodic orbit relative to the forcing function, which is to be used in the mmmn woof dx/dt dx/dl 29 calculation of the Fourier coefficients. The spatial distance 8 is chosen to be 0.5% of the span of the data set [5,35]. Some of the extracted periodic orbits are shown in Figure 3.2. pd—1 pct—2 Figrue 3.2 Some extracted periodic orbits of the Duffing oscillator 3.1.2 Choosing 3 Mathematical Model To identify the system parameters, we need a mathematical model that can catch the essential feature of the original system. Some a priori knowledge about the original system will help choosing a valid model. In this case, we know that the system is an externally excited, Duffing-type nonlinear system. Hence we choose a model in polynomial form, which has been commonly used in modeling the Duffing type nonlinear systems. The model with viscous damping is written as 30 p t' mx'+a.x+ 2 Ba = acostut, (3.4) i-l where m, (x, and B5 are the parameters to be identified. 3.1.3 Identification Results Applying the extracted periodic-orbit data to the mathematical model of Eq. (3.4), each term in the model is expressed in a Fourier series. The phase angle associated with each periodic orbit is included in the calculation of their Fourier coefficients, as discussed in detail in Chapter Two. Then the principle of harmonic balance is applied to the primary harmonics of the Fourier series, resulting in a set of algebraic equations in system parameters to be estimated by a least-squares fit. We first apply four sets of the periodic-orbit data separately to the model of Eq. (3.4), with five terms retained in the polynomials, the identification results are shown Table l. ’ Also, we apply four sets of periodic-orbit data together to increase the redundancy of the least-squares fit with different number of terms included in the model. The identification results are shown in Table 2. The results are accurate compared to the actual values, and consistent with each other for using different set of periodic- orbit data. The non-zero parameter values are recovered within 1% of their nominal values, and the zero-valued parameters are close to zero, even when the mathematical model contains many unnecessary high-order nonlinear terms. The standard deviations are less than 1% of the non-zero parameter values, or close to the average values of the zero-valued parameters, indicating the consistency of the results. Combining individual sets of algebraic equations increases the redundancy in the least- square; include The id. “0111111. 3.1.4 From t the, id, “MCI “0313b 31 Table 1: Identification results using individual periodic orbit for Duffing’s equation periodic orbits (1 Bl 52 53 B4 as actual 1.0000 .2000 1.0000 0.0000 1.0000 0.0000 0.0000 pd-3 0.9999 .1997 0.9915 -.0023 1.0009 .0002 .0000 pd-4 1.0008 .2002 1.0498 -.0050 0.9803 .0007 .0010 pd-5 0.9997 .1998 0.9659 -.0020 1.0124 .0001 -.0006 pd-6 1.0001 .1999 1.0015 .0016 1.0009 -.0002 -.0061 Average 1.0001 .1999 1.0022 -.0019 0.9968 .0002 -.0014 std. dev. 0.0005 .0002 0.0351 .0027 0.0134 .0004 .0032 squares fit, and improves the accuracy of the identification results when the model includes many parameters. The identified results suggest that the model can be refined by removing the higher-order nonlinear terms whose parameter values are negligible. The reduction of the unnecessary terms in the model tends to yield higher accuracy in the identification results. 3.1.4 Model Verification From the numerical results, the model of Eq. (3.4) can be easily verified. However, we use the identified model with the average values in Table 2 to generate a set of data, and extract the unstable periodic orbits, for comparison with the original ones. The extracted unstable periodic orbits from the identified model are shown in Figure 3.3. They resemble 32 Table 2: Identification results using 4 periodic orbitsfor Duffing’s equation orders m (1 Bl 92 B3 94 95 55 97 actual 1.000 .21XX) 1.“)0 0.CXX) MIX) 0.000 0.000 0.1K” 0.(X)0 p84 1.1XX) 0200 1.005 -0.(X)l ”XXI 0.1K” p=5 1.000 0200 1.015 p.001 0.996 0.000 0.000 p=6 MIX) 0200 1.010 0.(X)4 0.998 -o.oor 0.(XX) 0.000 ps7 l.(X)0 0.200 1.(X)5 0.004 1.000 -0.(X)l 0.(X)0 01X” -0.(X)0 Avg. 1.(X)0 0.200 1.009 .0015 0.999 -.(X)1 0.1K” 0.(XX) 0.(X)0 std.dv 0.000 0.000 0.005 0.(X)3 0.002 0.1K“ 0.(XX) 0.(XX) 0.1K” closely their counterparts in Figure 3.1 and Figure 3.2. The bifurcation diagrams, as shown in Figure 3.4, are calculated using the original equation and the identified one, by slowly increasing the forcing amplitude and sampling the steady-state response at the same Poincare section. The resemblance of the original bifrncation diagram and the identified one can be clearly seen. The Lyapunov exponents of the identified model calculated by the computer code of Wolf et al. [58] are convergent to i, = 020, and 212 = —0.49, which are close to the original values of 2.1 = 0.18 and 12 = -0.468 , with deviations of 11.1% and 4.7% respectively. Thus our model is verified. 3.1.5 Effect of Noise Numerically generated data are considered to be essentially noise free. The excellent 33 phone space period—1 Figure 3.3 The simulated chaotic attractor and some of the extracted periodic orbits of the identified model identification results in the example above may deteriorate if noise is present in the periodic data. To assess the influence of noise on the identification results, a set of uniformly—distributed random noise is added to each periodic orbit for use in the identification algorithm to test sensitivity of A, q and At: = q . If the noise is added to the chaotic set before the extraction of the periodic orbits, the spatial criterion 8 may need an adjustment. The noise level is set by the ratio of its maximum amplitude to that of the employed periodic-orbit data. Figure 3.5(a) shows a period—3 noise-free periodic orbit and Figure 3.5(b) shows its 2% noise contaminated counterpart in phase space. We examine the noise as tore. unplltudo Figure 3.4 Bifurcation diagrams of the Duffing’s equation (a) using the original equation, and (b) using the identified model with the average values in Table 2 eflect by using the model (3.4) with varying nonlinear terms and varying noise levels. Four sets of noisy periodic-orbit data are used together in the identification algorithm. The identification results are shown in Table 3. Comparing with the previous results in Table 2 for the same model, we find that, (1) within 5% noise level, the noise effect is not significant for a model with three or four terms in the polynomial. As the nonlinear terms increase beyond five, the noise effect increases. The errors are within 2.3% of the non-zero parameters. The zero-valued parameters have larger deviations than in previous case; (2) For the nonlinear terms in the model are greater than five, the noise efiect increase rapidly, resulting in less accurate identification results. The issue of noise will be 35 ,2 :0) h tel Figure 3.5 (a) A noise-free periodic orbit, (b) the noise-contaminated counterpart discussed in Chapter Five; (3) For a small amount of noise, the effect of noise is not catastrophic to our method, but in a robust way. With the model identified using the noisy periodic data, we proceed to verify the model by comparing the Lyapunov exponents, the structure of the unstable orbits, and the bifurcation diagrams as before. Using a model with the parameter values as in the last second row of Table 3, the Lyapunov exponents calculated by the computer code of Wolf et al. [58] are convergent to £1 = 0.21 , and i2 = —0.5 , which are close to the original values of )‘r = 0.18 and 3.2 = —0.468, with deviations of 16.67% and 6.84% respectively. The simulated chaotic attractor and the extracted periodic orbits from the identified model Table 3: Identification results for Duffing’s equation using noisy data 36 noise orders m (1 Bl 52 B3 94 55 actual 1.000 .200 1.000 0.000 LNG 0.000 0.“ 1% p=3 1.(X)0 .2“) 0.999 -.(X)6 MIX) 1% p=4 1.000 .2“) MIX) -.(X)1 1.000 .(XX) 1% p=5 1.000 .2“) 1.058 .016 0.983 .001 .001 2% p=3 1.1K” .2“) 0.999 -.013 1.(X12 2% p=4 1.000 .200 1104 -.029 1.001 .001 2% p=5 1.001 .201 0.947 .016 1.036 -.009 -.(X)5 3% p83 1.000 .2“) 0.995 -.020 1.(X)2 3% p=4 1.000 .2“) 1.002 -.042 1.1K” .001 3% p=5 1.(X)1 .201 1.136 -.044 0.958 .002 .002 5% p=3 MIX) .2“) 0.977 -.032 1.002 5% p=4 0.999 .200 0.988 -.069 1.001 .002 5% p=5 1.(X)1 .201 1.202 -.07 3 0.931 .(XB .(X13 37 “A an —-v ‘v --5 0 5 —5 0 I! X X Figure 3.6 The simulated chaotic attractor and some of the extracted periodic orbits of the identified model using the noise—contaminated periodic orbits are shown in Figure 3.6. The qualitative resemblance with the original ones of Figure 3.1 and Figure 3.2 is clearly seen. A bifurcation diagram is constructed for the identified model, by slowly increasing the force amplitude as the control parameter, and sampling the steady-state response at the same time interval, as shown in Figure 3.7(b). It closely resembles the original one in Figure 3.7(a). Thus the model is verified. 3.2 A Smooth Coulomb Friction System A smooth Coulomb friction system is given as x+cx+x+ (1+kx)tanh(ax) =fcos(tnt), (3.5) “(111011 is one of the models of a dry-friction system, studied extensively by Feeny and 38 35 6 r r 1 E5 - . M .. 0» 15K - g3 / .. as 210 l 3‘0 35 25 tomenplltudo Figure 3.7 Bifurcation diagrams of the Duffing’s equation (it) using the original equation, (b) using the identified model of the noisy case Moon [24]. The dry-friction is often modeled as a multivalued, discontinuous nonlinear force, which causes a “stick-slip” chaotic motion in a large parameter space. Here, the smooth function, tanh (ax) , is used to approximate the Coulomb fiiction model. This system exhibits ‘almost sticking’ motions, featuring a funnel-like structure in the phase space under the harmonic excitation [24]. We choose the parameter values as c = 0.03 , k = 1.5, a = 50, and the forcing term as 1.9cos(1.3t) , for numerical simulation. Numerical integration is carried out by a 5th—order Runge—Kutta method as before. A two- dimensional phase portrait is shown in Figure 3.8, where a funnel-like structure is clearly $6611. The Lyapunov exponents are calculated from the known equation using the computer code 39 1.5 ~ - 1 t. .. g 0.5 - , - o - ‘ -301- 10~ r 4—340 -20 0 2‘0 40 340 -2.0 40 20 40 x1 x1 Figure 3.12 Phase portrait of a Lorenz system shown in Figure 3.12, in which infinitely many unstable periodic orbits are dense and have an incommensurate period associated with each periodic orbit. Since the system is unforced, there is no fundamental period to be used as a guide for finding the unstable periodic orbits for use in our parameu'ic identification scheme. We use the recurrence property of the chaotic attractor to construct a recurrence plot to determine the period length of the periodic orbits, as stated in detail in Chapter Two. The recurrence plot is shown in Figure 3.13, in which the recurrent points that are clustered around certain values can be clearly seen. These values indicate the incommensurate periods of the periodic orbits. Using the values, the corresponding periodic orbits can be located within the Lorenz attractor. Some of the extracted periodic orbits are shown in Figure 3.14, which 49 8 385' § T 144 . 220 259 * 330 352 168 158 101. 129 110 5- 56 l .1 . . . . . o 250 0 50 100 150 200 300 350 400 450 500 length of delay momma“: 3 3 3 .5 0| T Figure 3.13 Recurrence plot of the Lorenz system (period lengths are indicated in the numbers of time steps used in the numerical integration) are to be used in our identification algorithm. Knowing that the system is a Lorenz type autonomous system, a mathematical model is chosen such that the linear terms and the quadratic nonlinear terms are included as Jt1 = i (“1111+ ibif‘ixj] (3.10) 1.1 j21' 1-1 jzi 12 = i (“2131* £61}:ij (3.11) 3 3 x3 = (03.x, + Zdyxixj] , (3.12) . 1 i jzi 50 Figure 3.14 Two extracted periodic orbits of a Lorenz system: (a) period length of 110 time steps, (h) period length of 144 time steps where aij, by, cij , and d". are the parameters to be determined. Using the periodic orbits, each term in the model is expressed in a Fourier series with the fundamental period obtained from the recurrence plot. The Fourier coefficients are calculated as before, except the phase angle is ignored, due to the fact that there is no forcing function involved. By balancing the Fourier coefficients of each harmonic in each equation, and ueafin g the It i -terms as known quantities, a set of algebraic equations in system parameters is constructed for the least-squares estimation. We use two periodic orbits with period lengths of 110 time steps and 144 time steps respectively, as shown in Figure 3.14, in the identification algorithm. The estimated parameter values are shown in Table 8. 51 Table 8: Identification results for the Lorenz equation‘| (a) 1‘1 ‘2 J"3 11x1 xrxz 1113 J‘2“2 12x3 x3x3 x -15.951 15.971 -.098 -.030 .019 -.003 .000 .003 .003 1 (-16.0) (16.0) 45.748 -.927 .230 .039 -.028 -.995 .003 -.004 -.006 (45.92) (- 1.0) (- 1.0) -.089 .041 -3.877 .031 .980 .003 .000 -0.003 -.004 (-4.0) (1.0) a. Wvdmmtltcpar’arrtetervaluesofeachequationinfiq.(3.12)indicamdbythefiratcohnnn. The actual parameters presented in the original system are identified accurately as highlighted in bold-face in the table, although some of the zero-valued parameters are not close to zero, such as the third term in the second equation. The model equation of (3. 12) can be refined by knowing that there is no ‘square’ term in the Lorenz equation. This refinement improves the accuracy of the identification results significantly, not only the non-zero parameters are closer to the real values, but also the zero-valued parameters are close to zero, as shown in Table 9. 3.4.1 Effect of Noise To assess the influence of noise on the identification results, we add a set of uniformly- distributed random noise to the periodic orbits as before. With 1% noise added to the extracted periodic orbits, the identification results of the model without square-terms are not significantly affected, although some of the zero-value terms have non-zero values, as shown in Table 10. With higher-level noise added to the periodic orbits, the identification 52 results deteriorate rapidly. This case shows that noise is influential to the parametric identification results for the autonomous system. Table 9: identification results for Lorenz equation‘ (b) Jr1 Jr2 Jr3 1132 11x3 Jr2"3 -16.011 15.953 -0.006 0.001 0.0002 0.001 1 x, (-l6.0) (16.0) x 45.900 4.0228 -0.0127 0.0020 0.9996 0.0021 2 (45.92) (-1.0) (-1.0) 13 0.0041 0.007 -3.9991 0.9999 0.0001 0.0002 (-4.0) (1.0) a. 111evalueaaretheparametervalueaofeachequationinfiq.(3.12)aaindieatedbythefiratcolunmwithoutthe aquaretermainthemodel. Table 10: Identification results for Lorenz equation' with 1% noise x, J‘2 x3 x1x2 1113 Jr2"3 x -15.5813 15.9582 0.0757 0.0006 -0.0087 0.0011 1 (-16.0) (16.0) 1 47.4206 -1.8632 0.0256 0.0034 -1.0252 0.0112 2 (45.92) (-1.0) (-1.0) x -0.9178 0.5717 -3.9468 0.9902 0.0200 -0.0129 3 (.40) (1.0) a. Thevalueaaretheparametervaluesofeachequationinfiq. (3.12)asindicatedbythefirsteohum1withoutthe square turns in the model. 3.5 A Case Study on Modeling the Nonlinearity with a Power Series We have confronted a problem in modeling a hyperbolic-tangent function with a power 53 series in Coulomb—friction system. We postulate that the power series representation may not be valid in such case. We examine this problem by a similar example, written as 1’ + cx + x + ktanh (x) = fcos (tut) (3.13) We want to show that, if the nonlinear function is known, our method is capable of identifying the parameters accurately, as shown in section 3.2; if the nonlinear function is unknown, and the power series is used to approximate it, then the radius of convergence of the power series and the truncated series representation are the factors influential to the identification results. The parameter values in Eq. (3.13) are chosen as c = 0.3, k = 0.5, and to = 1.3. Numerical data are generated using the Runge-Kutta method with several forcing amplitudes. The maximum periodic responses under difi'erent forcing amplitudes are listed in Table 11. Table 11: Force and response in model (3.13) case force, f max. x a 0.1 0.25 b 0.5 1.0 c 1.0 2.0 d 2.0 3.0 Note that, by Taylor series expansion, tanh(x) can be represented by _ _13 2 s_1_7 7 3 tanh(x) -x 3x +15x 3151c +...,Ix1$2. (3.14) 54 3.5.1 Using the Known Function in the Model We choose a mathematical model in a polynomial form as that of Eq. (3.4), with the addition of the known function of tanh(x) in the model, such that p . mx+ax+ 2 Bix'rytanh (x) = fcos (cot), i-l (3.15) where the parameters m, a, B: and ‘y are to be determined using the periodic data. The identification results are very accurate, as shown in Table 12, even when the mathematical model includes many unnecessary terms. Table 12: Identification results using the exact function in Model (3.15) cases 77: a 151 152 53 B4 ' BS 7 actual 1.0000 0.3000 1.0000 0.0000 0.0000 0.0000 0.0000 0.5000 a~b 1.0006 0.3000 1.0300 -.0009 -.0000 0.0000 0.0021 0.4700 a~c 0.9999 0.3000 1.0008 -.0000 -.0002 0.0000 0.0000 0.5000 a~d 1.0000 0.3000 1.0000 0000 .0000 0.0000 0.0000 0.5001 3.5.2 Using the Power Series Approximation Assuming that the nonlinear function of the system is unknown, our first choice is using a power series to approximate it. Part of the reason is that it is “easier” to fit the nonlinear function with a polynomial, and “possible” when data is within the radius of convergence. A model is chosen in a polynomial form as 55 P . mx+wz+ 2 Bax‘ = fcos(tot). (3.16) i=1 Applying the periodic data to this model, the identification results are liable to errors, depending on the amplitude of the response and the nonlinear terms retained in the model, as shown in Table 13. Table 13: Identification results using power series‘ in Model (3.16) cases 771 a 151 B; 133 B4 35 136 B7 138 B9 a~b 1.01 .300 1.52 .000 -.155 -.000 .036 a~d .951 .300 1.35 .000 -.062 -.000 .003 a~b .999 .300 1.50 -.000 -.166 .000 .060 -.000 -.014 a~d .991 .300 1.46 -.000 -.103 .000 .013 -.000 -.001 a~b 1.00 .300 1.50 .000 -. 167 -.000 .067 .000 -.025 -.000 005 a-d 1.11 .300 1.51 .000 -.139 -.000 .031 .000 -.003 -.000 .000 000 actual 1.00 .300 1.50 .000 -.l67 .000 .067 .000 -.027 a. Themnsretamedhrtheserieaiahtdicatedbythelstcolmnnmmber. In each case, better identification results are obtained using the smaller response data (cases a and b), which are within the radius of convergence of the series. The best result is obtained in the last case, in which the smaller response data are used in the model with nine terms included, which almost fits the power series in Eq. (3.14). Although the nonlinear function of the system is unknown, we may obtain a qualitative feature of the nonlinear function from the identification results. The nonlinear function is plotted using the identified values, as shown in Figure 3.15, in which the qualitative 56 0.5 I I V I r T T 0.4 - ad-5 is the curve using the values in the case of a~d with p=5. 0,3 . Analogous to other curves. 0.2 - 0.1 - ' if i 3 o - . -o.1 « -o.2 .1 -o.3 - -0.4 ab-S, a“ ab—7, ab—9 , 0.5mm - "0'31 —o:a -o:e —o:4 —o:2 1:: 0:2 0:4 0:3 ate 1 Figure 3.15 Nonlinear function in a power series feature of the nonlinear function is clearly seen. 3.6 Conclusion Numerical examples taken from the Duffing’s equation, a Coulomb friction system, the nonlinear Mathieu equation, and the Lorenz equation, show that the present method can accurately identify the parameters in a mathematical model that has been well-chosen to match the characteristic of the original chaotic system. The mathematical model can be refined by removing the unnecessary terms that have negligible values. Consistent identification results are remained for the valid models, implying that the suspicious terms are indeed unnecessary. Models are verified by comparing the structure of the unstable periodic orbits, the Lyapunov exponents, and the bifurcation diagram. The usage of many periodic orbits in the identification scheme 57 improves the accuracy of the least-squares estimation and provides the statistical information of the identified results. This is suitable for systems with many parameters to be identified. Random noise added to the periodic orbits can deteriorate the accuracy of the identification results, but in a robust way. This efi'ect worsens when the mathematical models are not well-chosen, for example with many unnecessary terms. When the precise form of the nonlinearity is unknown, yet smooth, the accuracy of identified truncated power series coefiicients deteriorates. However, the truncated power series may be applicable for qualitative modeling. CHAPTER 4 Experimental Results 4.1 Introduction In this chapter, we investigate a chaotic data set taken from a periodically driven magneto- oscillator by J. P. Cusumano and B. W. Kimble at Pennsylvania State University. The experiment was designed for observing the global phase—space structure of basins of attraction and homoclinic bifurcation using the stochastic interrogation method [12]. The experimental system was known to be similar to a two-well potential system. The techniques developed in the previous chapters are to be applied to the given set of chaotic data, in effort to identify the parameters of this experimental system. The chaotic attractor is reconstructed using the method of delays [26, 53], from which the unstable periodic orbits are extracted for use in the identification algorithm. A mathematical model is chosen in polynomial form by knowing that the experimental system has a smooth two- well stiffness potential. The method of harmonic balance is used to form a set of algebraic equations in system parameters, which are estimated by a least-squares fit. 4.2 Experimental Setup The experiment conducted by Cusumano and Kimble consisted of a stiffened beam 58 59 buckled by two magnets. The beam had extra rigidity in the form of steel bars epoxied and bolted along the length away from the clamped end. This additional rigidity was included in effort to make the system behave as a single degree of freedom. The uncovered portion of the beam near the clamped end acted as an elastic hinge from which the position of the beam was measured by a strain gauge. Two rare-earth permanent magnets were placed on the base of the frame holding the beam to create the two—well potential. The frame was then fixed through a rigid mount to an electromagnetic shaker. A periodic driving signal was fed through a power amplifier to the shaker to provide the external forcing function. The experimental set-up is shown in Figure 4.1. [12] A .1. _ «MM 0-‘9 sauna-n. '1 T‘— I: 3 Ho E 5 7-“ JOF E Mao- O 3/ 3,. _l 2.11% J .1 I... |._ *|°-’|'— Figure 4.1 Sketch of the experimental setup. Data from the strain gauge was acquired using a 12-bit data-acquisition (AID) board, with the digital values from -2408 to 2407 corresponding to -5V to 5V. With no forcing, three equilibria exist; two are stable at digital values of -495 and 315, and one (saddle) is 60 unstable at approximately zero. When forcing is added, periodic orbits exist instead of equilibrium points. The driving frequency was set at 7.5 Hz with 1.5V of the function generator output, by which the chaotic data were generated and collected at the sampling frequency of 187.5 Hz for 7000 periods of excitation. 4.3 Phase-Space Reconstruction Since there is only one observable in the data set, denoted by {xi} , j = 1, ...N, with 1]. = xUAr) , A: is the sampling time interval, the phase space of the experimental system is to be reconstructed. The most common method of phase space reconstruction is the method of delays [26, 53]. It is used to construct a d—dimensional pseudo-vector with its elements being the single observable separated by a constant delay time, such that yj = (xi’xjn’ ""xj+t(d-1))’ (4.1) where 1: is the delay time, and d is embedding dimension. Both of which are to be determined. The pseudo-vector represents a data point in the embedding space. In theory, for any sufficiently large dimension d and almost any choice of delay time t , an embedding of the original attractor can be obtained, and the geometrical invariants such as dimension and positive Lyapunov exponents can be preserved. In practice, the delay time I should be chosen so that the elements of yj are uncorrelated. If t is too small, then the coordinates at X]. and xj +1 represent almost the same information. If t is too large, then xj and xi +1 represent distinctly unrelated components of the embedding space. If the embedding dimension dis too small, the trajectory may cross itself. The requirement of a 61 sufficiently large embedding dimension prevents such ambiguity and ensures that the reconstruction is difl‘erentiable and invertible [2]. But an excessively large embedding dimension may lead to excessive computation and corrupt data, since noise will dominate the additional dimensions of the embedding space where no dynamics is operating [26]. There are several methods that have been proposed to determine the suitable delay time and the embedding dimension [1, 5, 9, 25]. We use the criterion proposed by Abarbanel [1] to determine the delay time 1: to be approximately 1/10~ 1/20 of the time associated with the first local minimum of the autocorrelation function of the measurement data {xi}. The autocorrelation function is defined as 1 N R“) = [(7211431) (43) ill and is shown in Figure 4.2 for the chaotic two-well data. O 20 4o 60 80 100 120 140 160 180 200 time stop. Figtu'e 4.2 Autocorrelation function of the experimental data. The pro; the sing 015011» whcr then in t‘ [2. (1%) cl: 62 The proper embedding dimension is estimated by the correlation function method [27] and the singular system analysis method [9]. The correlation function method calculates the distribution of points within a small region for a large data set, such that N N C(r) = P30 5. 2 2410-14-81) <44) 3 i where H(z)=l if z is positive; and H(z)=0 otherwise; y is a pseudo-vector constructed as Eq. (4.1). If the attractor is properly unfolded by choosing a sufficiently large dimension, then any property associated with the attractor which depends on distances between points in the phase space would become independent of the value of the embedding dimension [2, 15, 27]. In a regime that C(r) becomes independent of d, and exhibits a power law dependence on r as r -> 0 , that is limoC (r) = ard, the correlation dimension could be r —9 obtained by measuring the slope of the plot of logC(r) versus logr, such as d = lim M (4.4) r -4 0 logr Figure 4.3 shows the plot of logC(r) versus log(r) for several values of the dimension d. The slopes are about 2.5, which becomes independent of the dimension as d 2 3 . Ding er al. [15] reported that the plateau begins when the embedding dimension first exceeds the correlation dimension. Thus, this criterion should produce a lower bound to our required embedding dimension. The singular-system analysis method involves constructing a covariance matrix C = YT Y and decomposing it into two unitary matrices U and V and a diagonal matrix 2 , such that $08. Whe Sin E She 63 Correlation function mothod Figure 4.3 Correlation function of the experimental data. C = UZVT , (4.5) where Y is the mauix with each column containing the pseudo-vector y as constructed in Eq. (4.1). Varying the dimension in constructing the pseudo-vectors, and conducting the singular values analysis, a plot of the singular values versus the embedding dimension is shown in Figure 4.4. By comparing the singular values with the values induced by ‘noise’, which is assumed uniformly distributed in the extra dimensions and will be nearly equal, the singular values become flat when d 2 4 . Thus we determine the suitable embedding dimensions to be four. A two dimensional projection of the reconstructed phase space is shown in Figure 4.5, from which the unstable periodic orbits are to be extracted for use in our identification -. 829 3:9... algo 4.4 Fro des W1". Ex 64 Singular value. vorouo ombodding dimension. éau: d=6 12 400.. ............... § .............................. § ................ , ............................... _ 35° .............................. Hd= .................................... - 30°. ...... oood=4 ................................... .1 325°). ........................................... §.m..®3 ................................... .1 3,..- ........................................... ................................................ _ “,0, ............................................ ............................................... . m. .......................................... ................................................ .. ,0. .......................................... ............................................... - oo 1 2L 4:_ 5 6 dimonoiono Figure 4.4 Singular system analysis of the experimental data. algorithm. Also, a uansformation of the reconstructed phase space into the singular coordinates is performed by using the singular vectors, as shown in Figtue 4.6. 4.4 Periodic-Orbit Extraction From the reconstructed chaotic attractor, the unstable periodic orbits can be extracted as described in Chapter two. In the pseudo phase space, we seek recurrent points such that lynr’yi $8 (4.6) where e is set as 0.5% of the maximum extent of the chaotic set as before. Some of the extracted periodic orbits are shown in Figure 4.7. The corresponding periodic orbits in the singular coordinate are shown in Figure 4.8. 65 Reconetructed pheee epece of the experimental eyetem 2 f T Y Y 1 .5 - ~ 1 r- .. 0.5 - ~ 4 .— . S’ —0.5 r .. _1 I- d -1 .5 " -4 —2 —1 5 —1 -0.5 0 0 5 1 1 5 2 X“) Figure 4.5 Reconsu-ucted phase space of the experimental system Projection of the reconetructed pheee epeoe on einguier ooordinetee ‘ .5 U V I I 1 V 1 . .,v ’ ;~.’.-~-‘...\ r.‘\~ . g 0.5 _ / I. _ — x ‘ " ,”l ’5 7//, .‘.. / [7({1'T4g‘pti W\\ \\ ‘\ .. ' «((10.01 . ' \ \ t “ n 1/ .‘ g ‘ ‘- \\.:\‘\ \\’;.‘:.?.’/,4/,/ {/11 ' \ ::\\