Implementation of a Chaotic Electromechanical Oscillator Described by a Hybrid Differential Equation A Thesis Presented to The Division of Mathematics and Natural Sciences Reed College In Partial Fulfillment of the Requirements for the Degree Bachelor of Arts Xueping Long May 2013 Approved for the Division (Physics) Professor Lucas Illing Acknowledgments I would like to thank the following people: Professor Lucas Illing for being a great thesis advisor, Gregory Eibel, Robert Ormond, Jay Ewing, Cristian Panda, Yudan Guo and Kuai Yu for the help they gave me during my thesis project, and my parents for always supporting me. I also want to thank Reed College Science Research Fellowship for funding my experimental setup. Table of Contents Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Chapter 1: System Equation and Experimental Setup 1.1 System Equation . . . . . . . . . . . . . . . . . . . . 1.2 Magnetic Coil Design . . . . . . . . . . . . . . . . . . 1.3 Primary Coil Circuitry . . . . . . . . . . . . . . . . . 1.4 Secondary Coil Circuitry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 . . . . . . . . . . . . . . . . . . . . . . . . . 3 5 8 10 13 Chapter 2: Introduction to the Hybrid System . . . . . . . . . . 2.1 A Brief Introduction to Dynamics and Chaos . . . . . . . . . . 2.1.1 Dynamics and Dynamical Systems . . . . . . . . . . . 2.1.2 Differential Equations and Iterated Maps . . . . . . . . 2.1.3 Fixed Points, Periodic Orbits and Quasiperiodic Orbits 2.1.4 Chaos and Lyapunov Exponent . . . . . . . . . . . . . 2.1.5 Poincaré Section and Poincaré Return Map . . . . . . . 2.1.6 Symbol Sequences, Shift Maps and Symbolic Dynamics 2.2 Solution to the Hybrid Differential Equation . . . . . . . . . . 2.2.1 Discussion of the Solution for β = ln 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 17 17 19 21 22 24 25 28 32 Chapter 3: Experimental Data and Analysis . . . . . . . . . . . . . . . 35 Chapter 4: Further Discussion on the Hybrid System . . . . . . . . . 41 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47 Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49 List of Figures 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 Schematic setup of the mechanical oscillator . . . . . . . . . . . Picture of the mechanical oscillator . . . . . . . . . . . . . . . . Primary coil circuit . . . . . . . . . . . . . . . . . . . . . . . . . Calculation of emf using Amperian loops and a Gaussian surface Equivalent circuit for the primary coil circuit . . . . . . . . . . . First subcircuit of the secondary coil circuit . . . . . . . . . . . Second subcircuit of the secondary coil circuit . . . . . . . . . . Last subcircuit of the secondary coil circuit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 6 11 11 13 14 15 15 2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 Typical waveform for u(t) . . . . . . . . . . . . . . . . Phase space projection for β = ln 2 . . . . . . . . . . . Phase portrait of the phase space of 1-D spring . . . . Separation of nearby trajectories . . . . . . . . . . . . Example of a Poincaré section in 2-D phase space . . . Example of a Poincaré section in 3-D phase space . . . Determination of symbol sequence representation of 0.3 Bernoulli Shift Map . . . . . . . . . . . . . . . . . . . . un+1 versus un for β = ln 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18 18 19 23 25 26 26 28 33 3.1 3.2 3.3 3.4 3.5 Position v.s. time measurement . . . . . . . . . . . . . . . . . . Phase space diagram of experiment data . . . . . . . . . . . . . Poincaré return map for experiment data . . . . . . . . . . . . . Poincaré return map in the form of Bernoulli shift map . . . . . Comparison of analytic solution to measured experimental data . . . . . . . . . . . . . . . 36 36 37 37 38 4.1 Waveforms for numerical solution and its corresponding u, v and s . . 43 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Abstract This thesis describes an electromechanical oscillator whose governing equation of motion is an exactly solvable differential equation. The differential equation is a hybrid system that takes the form of a force driven harmonic oscillator with negative damping coefficient and a discrete switch condition to control and set two different equilibrium positions. Very strong agreement is found between the waveforms produced by the oscillator and the waveforms predicted by the analytic solutions. An extension to the theory of hybrid system is also proposed, which potentially allows the hybrid system to be perceived as the limiting case of an ordinary differential equation. Introduction The dynamics of chaotic systems is often described by nonlinear differential equations, which generally cannot be exactly solved using analytic methods. Recently a class of exactly solvable chaotic differential equations was discovered [1–4]. These differential equations are interesting not only because they are exactly solvable and may provide theoretical insights that would be otherwise unachievable, but also because they may allow us to investigate the properties of chaos and to explore possible applications of chaos. One potential application of these novel solutions would be matched filters for chaos communication [1]. It is known that matched filters are optimal receivers in the presence of noise in conventional communication protocols, where a small number of basis waveforms are used to transmit information. However, matched filtering for chaos communication is often thought to be impossible to design since chaotic waveforms, which are non-repeating, are used to transmit information in chaos communication. The class of novel solutions discussed in this thesis have enabled the design of matched filters [1], thus making chaos communication a viable option. Although research is in progress in order to realize these ideas at technological relevant microwave/radar frequencies, there remain many unresolved questions about these systems. An electromechanical oscillator that implements such an exactly solvable chaotic system is therefore of great interest. This thesis describes an electromechanical oscillator, whose governing equation of motion is a novel class of exactly solvable differential equations. The differential equation is a hybrid system which takes the form of a force driven harmonic oscillator with negative damping coefficient and a discrete switch condition to control the different oscillatory fixed point. Very strong agreement is found between waveforms produced by the oscillator and waveforms predicted by analytic solutions. An extension to the theory of hybrid system is also proposed, which potentially allows the hybrid system to be perceived as the limiting case of an ordinary differential equation. Chapter 1 System Equation and Experimental Setup Recently, an exactly solvable hybrid differential equation was discovered [1]. I am interested in implementing an electromechanical oscillator described by this differential equation, which is ü − 2β u̇ + (ω 2 + β 2 )(u − s) = 0, (1.1) where u(t) ∈ R is a continuous state, s(t) ∈ {±1} is a discrete state, and ω and β are fixed parameters satisfying ω = 2π and 0 < β ≤ ln 2. Transitions in s(t) are made when u̇ satisfies the guard condition u̇ = 0 ⇒ s(t) = sgn(u(t)), where the signum function is defined as +1, u ≥ 0 sgn(u) = −1, u < 0. (1.2) (1.3) This means that when the derivative of the continuous state hits zero, s(t) will be set to sgn(u(t)) and maintain this value until the same guard condition is met next. Note that sgn(0) = +1 is arbitrarily chosen for definiteness. Moving s to the right hand side in Eq. (1.1) yields ü − 2β u̇ + (ω 2 + β 2 )u = (ω 2 + β 2 )s. (1.4) For comparison, the equation of a damped driven harmonic oscillator is ẍ + 2ζω0 ẋ + ω02 x = F , m (1.5) where ω0 is the natural frequency, ζ is the damping ratio, m is mass, and F is the driving force. It is easy to see that Eq. (1.1) is the equation for a damped driven harmonic oscillator with negative damping. A setup of an electromechanical oscillator described by Eq. (1.1) has been proposed by Owens et al. [5]. Following their setup, an oscillator system composed of mechanical, electric and electromagnetic parts is constructed, and various improvements to 4 Chapter 1. System Equation and Experimental Setup Iron frame Metal bar Clamps Springs Secondary coil Magnets Primary coil Wooden board Position sensor Clamps Super pulleys Metal bar Clamps Figure 1.1: Schematic diagram of the mechanical oscillator. The coils are placed right below the spring sets. A set of cylindrical neodymium magnets is placed in the center of each coil, aligned along the vertical symmetry axis, and is attached to the spring set above with a fishing line. The magnet sets are also connected to one another using a fishing line that passes through the two super pulleys. A position sensor is placed below the primary coil to detect position of the magnets. 1.1. System Equation 5 the mechanical setup and circuit design are made. As schematically demonstrated in Figure 1.1, two 0.45 m metal bars are attached to a heavy, stable rectangular iron frame 0.56 m wide and 1.8 m high. A wooden board 0.45 m long and 0.36 m wide is placed horizontally in the middle of the metal frame. Two copper coils wound around PVC pipes are placed on top of the wooden board. Two sets of springs (McMaster-Carr Steel Extension Spring 9654K616) are hung down from the top bar, and two pulleys (Pasco Super Pulley ME09450A) are placed on top of the bottom bar. Each set of springs consists of four springs placed in parallel, and two aluminum set screws are used to hold the ends of the springs in place. Two sets of three cylindrical neodymium magnets (K&J Magnetics DX08B-N52), joined end to end, are placed in the hollow region inside the coils and are attached to the springs with fishing lines. The two sets of magnets are also connected to one another using a fishing line that passes through the two pulleys. The fishing lines are connected in specially designed knots such that the lengths of the fishing lines can be adjusted or fixed at will. A position sensor (Measurement Specialties DC-EC Series) is placed below the primary coil and is secured to the side of the metal frame using a combination of clamps. Figure 1.2 is a photo showing the real physical setup of the oscillator. 1.1 System Equation In order to derive the equation of motion for the system, a few assumptions are made to simplify the system. The pulleys are assumed to be both massless and frictionless, and the fishing lines connecting the magnets and springs are assumed to share equal and constant tension throughout the line and are always stiff. It follows from the assumptions that the magnets are always moving along the central vertical axis of the copper coil, and the two sets of magnets have the same oscillatory motion (although it would look like one set of magnets is moving up if the other set is moving down). Electric circuits are coupled to the mechanical oscillator via electromagnetic interaction of the magnetic masses and the coils. The primary coil is connected to a circuit that acts as a negative resistor and provides a positive feedback. The secondary coil is connected to a circuit that switches between constant offset states based on the position and velocity information it receives. This hybrid system can therefore be viewed as a damped harmonic oscillator system with a driving force that depends on the displacement and velocity of the test mass. After carefully tuning various parameters in the circuits, the predicted chaotic behavior, as will be derived in Chap. 2, can be obtained for the oscillator. The chaotic oscillator can thus be implemented successfully. In addition, it is assumed that the current applied to the coil varies sufficiently slowly in time such that the magnetic field Bcoil generated inside the coil can be treated as a steady state field. The coils are designed to display cylindrical symmetry such that when a current I is applied to the coil, the magnetic field generated by the coils, Bcoil , is in the vertical direction ŷ. Since each set of magnets can be treated as a single cylindrical magnet with three times the length of the original magnet, and the magnets only oscillate in the vertical ŷ direction along the central axis of the copper 6 Chapter 1. System Equation and Experimental Setup Figure 1.2: Picture of the mechanical oscillator. coil, when the magnets interact with the magnetic field Bcoil , the resultant force Fmag is vertical: Fmag = Fmag ŷ = µ ∂B ŷ, ∂y (1.6) where µ is the magnitude of the magnetic dipole moment generated by the cylindrical magnets. Since the system can be viewed as a force-driven damped harmonic system, one can write down the equation of motion describing the system: mÿ + cẏ + ky = Fmag , (1.7) where Fmag is the effective magnetic force, k is the effective spring constant of the system, and c is the lumped viscous damping coefficient. The magnetic force in this system can be written as Fmag = Fpc + Fsc , (1.8) where Fpc is the magnetic force due to the primary coil and Fsc is the magnetic force due to the secondary coil. As will be shown in Sec. (1.2) and Sec. (1.3), Fpc is directly proportional to the velocity of the oscillating magnets inside the primary coil and represents a negative viscous friction term, whereas Fsc represents an offset term that is switched between two constant values as a function of position and velocity of 1.1. System Equation 7 the magnets. It will be shown in Sec. (1.3) that Fpc can be written in the form Fpc = βc2 ẏ, Reff (1.9) where βc is a constant that relates the current running in the coil to the magnetic force on the cylindrical magnet due to the coil (an effective negative viscous friction coefficient), and Reff is the magnitude of the negative resistor that will be discussed in Sec. (1.3). Therefore Fmag can be written as Fmag = βc2 ẏ + Fsc (y, ẏ). Reff (1.10) Rewriting Eq. (1.7) using Eq. (1.10) gives βc2 mÿ + cẏ + k(y − y0 ) = ẏ + F̃sc (y, ẏ), Reff (1.11) where y0 is the position of the top of the coil such that it effectively allows the origin of the coordinate system to be chosen at will, and F̃sc = Fsc − ky0 is the effective magnetic force due to the secondary coil. To nondimensionalize Eq. (1.11), a coordinate transformation is performed: y = uY + Y0 , t = τ T, (1.12) where Y , Y0 and T are constants to be determined later, and u and τ are dimensionless. Then Eq. (1.11) can be written as 2 βc Y 0 Y0 y 0 mY 00 u − −c u + kY u + − = F̃sc , (1.13) T2 Reff T Y Y where u0 is the τ -derivative of u. To simplify Eq. (1.13), T is chosen such that β 2 + 4π 2 = where T β= 2m k 2 T , m βc2 −c . Reff (1.14) (1.15) Solving Eq. (1.14) and Eq. (1.15) yields 4πm T =p , 4mk − βc2 /Reff + c 2π(βc2 /Reff − c) β =p . 4mk − βc2 /Reff + c (1.16) 8 Chapter 1. System Equation and Experimental Setup Then Eq. (1.13) becomes Y0 y0 u − 2βu + (β + 4π ) u + − Y Y 00 0 2 2 = F̃sc T2 , mY or u00 − 2βu0 + (β 2 + ω 2 ) (u − s) = 0, (1.17) where ω = 2π, and s = = = = 1 Y 1 Y 1 Y 1 Y T2 y0 − Y0 + F̃sc 2 (β + 4π 2 )m T2 y0 − Y0 + F̃sc k/m · T 2 m y0 − Y0 + F̃sc /k Fsc − Y0 . k To find appropriate values for Y0 and Y , we demand ( 1 when Fsc = βc I max s= −1 when Fsc = βc I min . (1.18) (1.19) Solving Eq. (1.18) with the definitions given in Eq. (1.19) yields Y0 = βc max (I + I min ). 2k (1.20) Plugging Eq. (1.20) into Eq. (1.18) for s = 1 yields Y = βc max (I − I min ). 2k (1.21) One can check that Eq. (1.20) and Eq. (1.21) together satisfy Eq. (1.18) for s = −1. Therefore Eq. (1.17) is the nondimensionalized differential equation that governs the electromechanical system. Note that Eq. (1.17) is exactly the same as the hybrid differential equation I am interested in studying, Eq. (1.1). The conditions given in Eq. (1.19) are fulfilled by controlling the electromagnetic force in the secondary coil through the attached circuitry and position sensor, which recognizes the guard condition and outputs the desired voltage accordingly. The theory and development of the magnetic coils are described in the next section. 1.2 Magnetic Coil Design As has been discussed in Eq. (1.6), the magnetic force inside a coil is in the ŷ direction and is proportional to ∂B/∂y: Fmag = µ ∂B ŷ. ∂y (1.22) 1.2. Magnetic Coil Design 9 In order to achieve a constant offset for the secondary coil, a magnetic field that is linear in y for some region of the coil is needed. The copper coils were made in the machine shop in the subbasement of the Reed College Physics Department. Copper wires of American wire gauge 24 are wound onto two PVC pipes, each having a length of 40.6 cm, an outer diameter of 4.7 cm and an inner diameter of 4.0 cm. For each coil one long continuous wire is wound in layers; each successive layer is 12.7 cm shorter than the previous layer. The first layer has a length of 30.5 cm, and a total of 24 layers are on a coil. A theoretical calculation predicts the resistance of coils to be 103 Ω. The measured resistance for both coils is 104.5 Ω, in good agreement with the theoretical prediction. The measured inductance is 0.567 H for both coils. With such winding, the number of turns of copper wire per length, n(y), increases approximately linearly with distance from the top of the coil: n(y) = cn y, y ∈ [0, Lc ], (1.23) where the origin of the coordinate system is chosen to be the top of the coil and the y-axis to point downward. Lc is the length of the coil and cn = 2Nturns /L2c ≈ 14 cm−2 is a constant. It will be demonstrated next that this design indeed gives a region of constant magnetic field gradient to a good approximation. Because of its dense helical windings, the coil can be modeled as a continuum of current-carrying circular loops. By Newton’s third law, each loop exerts a vertical force on the magnet that is equal in magnitude but opposite in direction to the force the magnet exerts on the loop. Therefore, the total force the coil exerts on the magnet is equal in magnitude to the total force the magnet exerts on the coil, which is the sum of the forces the magnet exerts on all the loops. Assuming that each loop has the same radius (which is to say that the wire is thin), when the geometric center of the magnet is at height y, by the Lorentz force law, in the continuous limit, Z Fmag = − Jcoil × Bmag (rc , ȳ − y) dV Z Jcoil Bρ,mag (rc , ȳ − y)ŷ dV = Z Lc = Bρ,mag (rc , ȳ − y) × 2πrc n(ȳ)I dȳ ŷ 0 Z Lc = 2πrc I n(ȳ)Bρ,mag (rc , ȳ − y) dȳ ŷ, (1.24) 0 where Jcoil is the current density running in the coil, Bmag (rc , ȳ − y) is the magnetic field felt by the wire at height ȳ due to the magnet, Bρ,mag is the radial component of Bmag (rc , ȳ − y), rc is the coil radius, and I is the current running in the coil-wire. The direction of the current is chosen such that Bmag , Jcoil and ŷ form a right-handed system. The magnetic force Fmag depends on the geometry of the coil through n(ȳ) and the geometry of the magnet through Bρ,mag . These dependencies can be put into 10 Chapter 1. System Equation and Experimental Setup a single term βc , where Lc Z n(ȳ)Bρ,mag (rc , ȳ − y) dȳ, βc (y) = 2πrc (1.25) 0 which allows the magnetic force to be expressed as Fmag = βc (y)I. (1.26) Therefore, if one can show that βc (y) is independent of y, a constant current I would result in a constant force Fmag and subsequently a constant magnetic field gradient as desired. Evaluating the value of βc for a dipole magnet using Eq. (1.25) with the definition given in Eq. (1.24) yields ! y µµ0 cn (Lc − y)3 − yrc2 +p βc (y) = . (1.27) 2 y 2 + rc2 (rc2 + (Lc − y)2 )3/2 Assuming Lc rc (long coil) and Lc > y rc (the magnet is away from the coil edges), Eq. (1.27) can be simplified to βc (y) ≈ µµ0 cn , (1.28) which is independent of y as desired. 1.3 Primary Coil Circuitry The function of the primary coil circuit is to provide the negative damping term −2β u̇ in Eq. (1.1) to the hybrid system. As shown in Figure 1.3, the primary coil circuit has three substructures placed in series: the coil, the negative resistor, and the active inductor. From the perspective of circuit design, the primary coil can be treated as an in-series combination of a voltage source, a resistor, and an inductor. That the coil acts as a voltage source follows from Faraday’s law, which implies that the time varying displacement of the magnets inside the coil induces an emf. The goal is to provide positive feedback that is proportional to the emf, and hence a primary circuit that compensates for the coil’s self-inductance and overcompensates the coil’s internal resistance needs to be constructed. The negative damping force provided by the primary coil is the force exerted on the magnet due to the primary coil. As derived in Eq.(1.26), Fpc is Fpc = βc Ipc . (1.29) I will derive an expression for Ipc by analyzing the emf ε that the moving magnet induces across the coil. Consider a horizontal Amperian loop of radius rc located at a height of ȳ, as shown in Figure 1.4. By Faraday’s law, the induced emf ε is the negative time derivative of the magnetic flux through the loop 1.3. Primary Coil Circuitry 11 100 Ω +15 V 104 LM675 104 104 1 kΩ 1 kΩ 560 Ω Primary Coil -15 V -R Negative Resistance 1 kΩ L Active Inductance 1 kΩ 5.6 kΩ 105 1 kΩ Figure 1.3: The primary coil circuit. It is composed of the primary coil, the negative resistor and the active inductor placed in series. Amperian Loop Cylindrical Magnet Coil Figure 1.4: Calculation of emf using Amperian loops and a Gaussian surface. The dotted loops are the Amperian loops, and the cylinder formed by connecting the two Amperian loops provides a Gaussian surface one can work with. By Gauss’s law, the difference of the magnetic flux through the top and bottom Amperian loop is exactly canceled by the magnetic flux through the sidewall of the cylinder, as no magnetic monopole exists in nature. 12 Chapter 1. System Equation and Experimental Setup ε=− ∂Φ(ȳ − y) dΦ(ȳ − y) = ẏ, dt ∂ ȳ (1.30) where y is the position of the magnet. Consider a second Amperian loop that is dȳ below the first Amperian loop. Together the two Amperian loops form a cylinder whose surface can be considered as a Gaussian surface. By Gauss law the total magnetic flux through the Gaussian surface is equal to the magnetic “charge” enclosed by the Gaussian surface, and, as there is no magnetic monopole, the difference dΦ of the magnetic flux through the top and bottom Amperian loop is exactly canceled by the magnetic flux through the sidewall of the cylinder: 2πrc Bρ,mag (rc , ȳ − y)dȳ. Then ∂Φ(ȳ − y) = −2πrc Bρ,mag (rc , ȳ − y), ∂ ȳ (1.31) Since there are n(ȳ) loops of wire at position ȳ per unit length, the total induced emf is Z Lc ε = − 2πrc n(ȳ)Bρ,mag (rc , ȳ − y)dȳ ẏ 0 = −βc ẏ. (1.32) If the effective resistance of the attached circuit is a negative resistor −Reff , then the current in the primary coil is βc ẏ Reff (1.33) βc2 ẏ, Reff (1.34) Ipc = −ε/Reff = and the force due to the primary coil would be Fpc = βc Ipc = as has been claimed in Eq. (1.9). The effective resistance of the primary coil circuitry can be derived from its equivalent circuit, as shown in Figure 1.5 (the 104 capacitor used to eliminate high-frequency noise is ignored because it has a negligible effect on the circuit). The value of Lv is chosen deliberately such that R1 Lc = . (1.35) R2 Lv Since V1 =V2 , I1 V1 − V = I2 R1 V2 − V R2 = . R2 R1 (1.36) Applying Kirchhoff’s voltage law to the circuit in Figure 1.5 yields V 2 + I 2 Rv + dI2 Lv = 0 dt (1.37) 1.4. Secondary Coil Circuitry 13 R1 I1 V1 V Rc R2 I2 Lc V0 V2 Rv Lv Figure 1.5: Equivalent circuit for the primary coil circuit. V0 is the induced emf as a result of moving magnets inside the primary coil, Lc and Rc are the effective inductance and resistance of the primary coil respectively, Rv is a variable resistor, Lv is an active inductor whose value is controlled by choosing appropriate components, and R1 = 100Ω, R2 = 1kΩ. and dI1 Lc − I1 Rc = V1 . dt Solving Eq. (1.36), (1.37) and (1.38) for V0 yields R1 Rv V0 = −I1 − Rc , R2 V0 − or Reff V0 = =− I1 R1 Rv − Rc . R2 (1.38) (1.39) (1.40) The variable resistor is tuned to keep Reff negative ( (R1 /R2 )Rv −Rc = Rv /10−Rc > 0, or Rv > 10Rc ) while maximizing gain. As a consequence, the primary coil acts as a source that provides positive feedback (negative damping) proportional to the velocity of the oscillating magnet. In other words, the amplifier provides the power that is required to sustain the mechanical oscillations in the system. 1.4 Secondary Coil Circuitry The secondary circuit provides the constant offset force on the right hand side of Eq. (1.1). It shifts the oscillation fixed point between two levels based on the switch condition s defined by Eq. (1.2). It is clear from the switch condition that to produce the desired driving force, two pieces of information are required: u(t) and u̇(t). Therefore the secondary circuit needs to have three subcircuits: one that collects position 14 Chapter 1. System Equation and Experimental Setup Vin AD620 100 kΩ Vg 105 LM339 x 500 kΩ 10 kΩ 5.6kΩ v Figure 1.6: The first subcircuit of the secondary coil circuit. This circuit takes information from the position sensor and derives information on position and velocity, which are then used to output the desired offset voltage. and velocity information from the data gathered by the position sensor, one logic circuit that implements the desired switch condition, and one that outputs current to the secondary coil. The first subcircuit, which collects position and velocity information, is shown in Figure 1.6. Position is determined by a Measurement Specialties DC-EC Series position sensor, which relates position linearly to a voltage output. As the position sensor puts out voltage from -2V to 8V while the displacement of the system is relatively symmetric, a zero-displacement position corresponding to approximately 4V is set. However, it is necessary to acquire the exact zero-displacement position (voltage) with the help of a LabVIEW program before actually running the experiment, because the chaotic system is extremely sensitive to this parameter. I denote this zero-displacement voltage by Vg . In Figure 1.6, the voltage signal Vin collected by the position sensor is buffered, the zero-displacement voltage Vg is subtracted from Vin using an AD620 instrumental op-amp, and the difference is compared to ground using an LM339 comparator. The digital output of the LM339 comparator, called x, encodes the relative position (higher or lower) with respect to the zero-displacement point. Vin is also used to derive the velocity information. Since velocity is the time derivative of relative position, and Vin is directly proportional to relative position, Vin is converted to velocity information by being sent into an R-C differentiator (the 100kΩ variable resistor in front of the differentiator is used to suppress high frequency noise). Velocity information is then compared to ground, yielding a HIGH or LOW voltage (±14V depending on the sign of the input, and converted to a suitable amplitude for digital signals). The resulting signal encodes the sign of the velocity. A final buffer is used to isolate the voltage divider from the logic components in the second subcircuit. Signals encoding position and velocity information are next sent into the second 1.4. Secondary Coil Circuitry 15 Q1 A1 x 7475 D Q v C CLR1=(H) A2=(L) Q2 Comparator Low level OR gate 7432 Vdout LM339 2.7 kΩ B1=(H) Vcc=5V B2 CLR2=(H) DM74123N Figure 1.7: The second subcircuit of the secondary coil circuit. This circuit takes position and velocity information from the first subcircuit and outputs the correct waveform for the offset signal. Vdout +15 V 10 kΩ -15 V 104 Sec. Coil 10 kΩ 100 kΩ 2.2 kΩ +15 V LM675 104 Figure 1.8: The last subcircuit of the secondary coil circuit. This circuit scales down the signal from Vdout by a factor of 1/6, which is then suitable as an input to a LM675 high-output current precision amplifier that generates the current driving the secondary coil. subcircuit. As the offset voltage corresponds to position and updates its value only when velocity vanishes (in other words when velocity changes sign, since it is extremely unlikely for velocity to stay zero for a considerable period of time in a real experiment), a circuit that determines the position and sets the output in accordance to the position (HIGH or LOW) for every change of sign of the velocity can be built. In Figure 1.7, the DM74123N retriggerable one-shot circuit is connected in series with an OR gate such that it outputs a HIGH every time the velocity changes sign. This information, together with position information, is then sent into an SN7475 bistable latch, such that every time C is HIGH, Q will be set to the value of D (position). Finally a LM339 comparator is used to set a LOW value since a LOW 16 Chapter 1. System Equation and Experimental Setup output from SN7475 latch is normally greater than 0V. The comparator LOW level is set to 0.6V. The output signal from the digital circuit is then sent into the third subcircuit, which generates the desired signal to the coil. The third subcircuit is shown in Figure 1.8. The input signal to this circuit, i.e. Vdout , is scaled by a factor of 1/6 and is sent to a LM675 high-output-current precision amplifier that generates the current driving the secondary coil. Chapter 2 Introduction to the Hybrid System The governing equation for the hybrid system is ü − 2β u̇ + (ω 2 + β 2 )(u − s) = 0, (2.1) where u(t) ∈ R is a continuous state, s(t) ∈ {±1} is a discrete state, and ω and β are fixed parameters satisfying ω = 2π and 0 < β ≤ ln 2. Transitions in s(t) are made when u̇ satisfy the guard condition u̇ = 0 ⇒ s(t) = sgn(u(t)), where the signum function is defined as +1, u ≥ 0 sgn(u) = −1, u < 0. (2.2) (2.3) Using an adjustable step size Runge-Kutta integrator (MATLAB’s ODE45) to integrate the ordinary differential equation and implementing the switching condition as a detectable event in the integrator, a typical waveform for the hybrid system (2.1) is obtained using numerical integration for β = ln 2 and is shown in Figure 2.1 (source code for the program was provided by Professor Lucas Illing). The corresponding phase-space projection is shown in Figure 2.2. The solution obtained by numerical integration appears to be chaotic. A brief introduction to dynamics and a discussion on the meaning of chaos will be given in Sec. 2.1. An exact solution to Eq. (2.1) will also be derived, which demonstrates that the solution obtained is indeed chaotic. The textbooks consulted in Sec. (2.1) are Ott [6], Hilborn [7], Strogatz [8] and Zheng [9]. The method for solving Eq. (2.1) in Sec. (2.2) is from the paper by Corron, Blakely and Stahl [1]. 2.1 2.1.1 A Brief Introduction to Dynamics and Chaos Dynamics and Dynamical Systems Dynamics is a subject that studies how a given system evolves with time. In particular, people are interested in questions such as whether the system will eventually 18 Chapter 2. Introduction to the Hybrid System 2 u 1 0 −1 −2 0 5 10 15 20 25 Time 30 35 40 45 50 Figure 2.1: Typical waveform for u(t). The waveform is obtained from numerical integration of the hybrid system for β = ln 2. 10 udot 5 0 −5 −10 −2.5 −2 −1.5 −1 −0.5 0 u 0.5 1 1.5 2 2.5 Figure 2.2: Phase space projection from numerical integration of the hybrid system for β = ln 2. 2.1. A Brief Introduction to Dynamics and Chaos 19 Figure 2.3: Phase portrait of the phase space of 1-D spring. The spring oscillates harmonically about its equilibrium position. Every ellipse is a possible trajectory for a point on the spring. The origin is a special point, where the trajectory is a simple fixed point for all time. settle to an equilibrium state, or demonstrate periodic behavior, or do something more complicated. In order to discuss dynamics, it is convenient to introduce the concept of phase space. In general, the phase space of a system is a space in which all possible states of the system are represented. For this reason, the phase space is also called the state space. Every possible state of the given system corresponds to a point in the phase space uniquely. In the study of moving particles, the phase space often consists of all possible values of position and velocity variables. Figure 2.3 demonstrates the phase space of a 1-D spring oscillating about its equilibrium position: Each ellipse in Figure 2.3 represent a possible time revolution for a given initial state, and thus is a possible trajectory in the phase space. The origin alone constitute another kind of trajectory: it represents a spring in its equilibrium position where it is fixed for all future time. Such plot, which includes a collection of several different trajectories originating from different initial conditions is called a phase portrait for the system. As each point in phase space can be considered an initial condition, the phase space is completely filled with trajectories. Finally, a deterministic dynamical system consists of a phase space and a fixed rule that determines the time evolution of any possible starting state of the system. One key thing to note about deterministic systems is that for a fixed time interval, only one future state can result from a given current state. 2.1.2 Differential Equations and Iterated Maps When considering dynamical systems one can distinguish dynamical systems governed by differential equations and dynamical systems governed by iterated maps. The choice of the description of a dynamical system depends on the nature of time that describes the system: differential equations are used to describe dynamical systems in which time is a continuous variable, while iterated maps are used to describe dynamical systems in which time is discrete. If a system can be described solely by some variables and their time derivatives, it is possible to convert the system into a system of first-order, autonomous ordinary 20 Chapter 2. Introduction to the Hybrid System differential equations, ẋ(1) = f1 (x(1) , ..., x(N ) ), .. . (N ) ẋ = fN (x(1) , ..., x(N ) ), (2.4) where ẋ(i) ≡ dx(i) /dt, and the expressions for f1 , ..., fN will depend on the actual system one is trying to describe. Eq.(2.4) can also be written in vector form as ẋ(t) = f[x(t)], (2.5) where x is an N -dimensional vector. The differential equation system given in Eq. (2.4) is said to be autonomous because it has no explicit time dependence. In the case of non-autonomous systems, in which the differential equations explicitly depend on time, one more equation can be added to Eq. (2.4) by setting x(0) = t: ẋ(0) = 1 ẋ(1) = f1 (x(0) , ..., x(N ) ), .. . (N ) ẋ = fN (x(0) , ..., x(N ) ), (2.6) The advantage of writing the non-autonomous system in the autonomous form (2.6) is that non-autonomous systems and autonomous systems can be treated on an equal footing, but the advantage gained does come with a price: there is one more dimension in phase space. Sometimes information about a system can only be obtained at discrete, integervalued times. For example, in my experiment the position is recorded every 7 ms. In such cases, it is useful to describe the system using a mapping: xn+1 = M(xn ), (1) (2.7) (2) (N ) where xn is an N -dimensional vector xn = (xn , xn , ..., xn ) representing the state of the system after n time intervals, and M represents a fixed rule that governs how the current state maps to the next state. Again this is a dynamical system, because given any initial state x0 , the mapping M can be applied n times to get the uniquely determined state at t = n∆t, where ∆t is the time interval between adjacent measurements. Systems described by Eq. (2.7) are known as iterated maps. They are also referred to as difference equations, recursion relations or simply maps. It should be noted that continuous time systems can be turned into discrete time systems by sampling at regular time intervals or using the Poincaré section method. Sampling at a regular time interval T is also known as the time T map, in which a continuous time trajectory x(t) is evaluated at discrete times tn = t0 + nT (n = 0, 1, 2 . . .). In this way, a continuous time trajectory x(t) yields a discrete time orbit xn ≡ x(tn ). The Poincaré section method will be discussed in Sec. (2.1.5). Of course, there are other types of maps that are not derived from a discretization of continuoustime ordinary differential equations. For more information on iterated maps, refer to Ott [6] and Hilborn [7]. 2.1. A Brief Introduction to Dynamics and Chaos 2.1.3 21 Fixed Points, Periodic Orbits and Quasiperiodic Orbits Often people are interested in the time evolution of systems. While generally points in the phase space of a dynamical system follow predestined trajectories as dictated by the deterministic rule, some special points never change: the fixed points. If a dynamical system is described by a set of autonomous, first-order differential equations, as in Eq. (2.4), a point in the phase space of the system for which the time derivatives of the phase space variables are 0 is called a fixed point for the system. That is, a fixed point is a point for which dx(i) = 0, dt (2.8) where i = 1, ..., N , and N is the dimension of the system. Equivalently, in vector form fixed points must satisfy ẋ = 0. (2.9) Fixed points for systems of differential equations are also referred to as equilibrium points, or critical points, or singular points. If a dynamical system is described by an iterated map, as in Eq. (2.7), a fixed point would be one whose next iteration is the same as the current point: x∗ = M(x∗ ). (2.10) While the fixed points stay at their values and never change, some trajectories return to their previous value after some fixed time interval. These trajectories are called periodic orbits. For a dynamical system described by differential equations, such periodic behavior is characterized by x(t + T ) = x(t), (2.11) where T is the period, while for a dynamical system described by iterated maps, the periodic behavior is characterized by xn+P = xn , (2.12) xn+P = MP (xn ), (2.13) or equivalently, where P is the period. Periodic orbits and fixed points for a dynamical system described by differential equations are shown in Figure 2.3. The ellipses are periodic orbits while the origin is a fixed point. In addition, sometimes one encounters another type of trajectories called quasiperiodic orbits (or almost periodic orbits), which can be described by a quasiperiodic function F (t) = f (ω1 t, . . . , ωm t) (2.14) for some continuous function f (ϕ1 , . . . , ϕm ) of m variables (m ≥ 2), periodic in ϕ1 , . . . , ϕm with period 2π, and some set of positive frequencies ω1 , . . . , ωm . The 22 Chapter 2. Introduction to the Hybrid System frequencies are rationally linearly independent, meaning that m X ki ωi 6= 0 (2.15) i=1 for all non-zero integer valued vectors k = (k1 , . . . , km ) [10]. This thesis will not discuss quasiperiodic orbits in detail. 2.1.4 Chaos and Lyapunov Exponent Besides fixed points, periodic orbits and quasiperiodic orbits, there is another possible type of trajectories: trajectories that are said to be chaotic. A chaotic system must satisfy the following criteria: 1. the system demonstrates aperiodic long-term behavior on a nontrivial open set in phase space, 2. the system must be deterministic, 3. the system shows sensitivity to initial conditions. Aperiodic long-term behavior means that there are trajectories in the system that do not eventually settle to fixed points, periodic orbits or quasiperiodic orbits as t → ∞. Nontrivial open set ensures that there is a reasonable number of such trajectories. Deterministic system means that with knowledge of initial states, one can have complete knowledge of all future states. In other words, the system is not subject to random or noisy inputs or parameters. Sensitivity to initial conditions roughly means that neighboring trajectories separate exponentially fast. Here a more quantitative description of sensitivity to initial conditions is developed. For a dynamical system described by differential equations, suppose two nearby points start off with a separation vector d0 , and after time t the separation vector becomes dt , as shown in Figure 2.4. If neighboring trajectories separate exponentially fast, then |dt | ≈ |d0 |eλt , (2.16) where |d| is the length of the vector d, and λ is a positive number. This idea can be generalized by introducing the Lyapunov exponent, which characterizes the rate of separation between infinitesimally close trajectories. Suppose two trajectories with an initially infinitesimally small separation δx(0) diverge such that at time t the separation δx(t) satisfies |δx(t)| ≈ |δx(0)|eλt . (2.17) λ is called the Lyapunov exponent. If the Lyapunov exponent is positive, then separation between nearby trajectories grows exponentially fast. The system is extremely sensitive to small changes in the initial conditions. 2.1. A Brief Introduction to Dynamics and Chaos 23 Figure 2.4: Separation of nearby trajectories. Initially two points, x0 and x0 + d0 , are separated by a vector d0 . After time t, x0 travels to xt while x0 + d0 travels to xt + dt , and the separation becomes dt . For the system to be “sensitive to initial conditions”, the separation needs to grow exponentially fast: |dt | = eλt |d0 | for some positive coefficient λ. The definition of Lyapunov exponent can be generalized to iterated maps. Given the initial infinitesimal separation δx0 and the separation after n iterations δxn , if |δxn | ≈ |δx0 |enλ , (2.18) then λ is called the Lyapunov exponent. Note that if the phase space has N dimensions, N Lyapunov exponents can be defined, corresponding to the N dimensions. A dynamical system may be chaotic as long as the largest Lyapunov exponent is positive. For a one-dimensional iterated map xn+1 = M (xn ), (2.19) an explicit formula for its Lyapunov exponent can be found. Suppose two neighboring trajectories start off at x0 and x0 + δ0 , where δ0 → 0. Then the separation after n iteration is δn = M n (x0 + δ0 ) − M n (x0 ). By the definition of Lyapunov exponents, |δn | ≈ |δ0 |enλ . Dividing both sides by |δ0 | and taking the logarithm δn ln ≈ nλ δ0 (2.20) (2.21) yields 1 δn ln λ ≈ n δ0 1 M n (x0 + δ0 ) − M n (x0 ) = ln n δ0 1 = ln |(M n )0 (x0 )| . n (2.22) 24 Chapter 2. Introduction to the Hybrid System The last equality follows from the definition of the derivative. Eq. (2.22) can be simplified further by applying the chain rule: 0 (M n )0 (x0 ) = M n−1 ◦ M (x0 ) 0 = M n−1 (M (x0 )) · (M (x0 ))0 0 = M n−1 (x1 ) · M 0 (x0 ) .. . = M 0 (xn−1 ) · M 0 (xn−2 ) · ... · M 0 (x0 ) n−1 Y = M 0 (xi ). (2.23) i=0 Then Eq. (2.22) becomes n−1 1 Y 0 ln λ ≈ M (xi ) n i=0 n−1 1X ln |M 0 (xi )| . = n i=0 (2.24) More formally, for a trajectory starting at x0 in a one-dimension iterated map, the Lyapunov exponent can be calculated using n−1 1X λ = lim ln |M 0 (xi )| . n→∞ n i=0 (2.25) If an iterated map is obtained from a time T map of a well-behaved continuous time system, the Lyapunov exponent can be used to determine the nature of the continuous time dynamics. Suppose the original continuous time system is sampled at a regular time interval ∆t such that an iterated map with a positive largest Lyapunov exponent λ is obtained, and within each time interval the system does not behave crazily (such as performing wild oscillations). Consider two nearby trajectories of initial separation |δ0 | ≡ |δ(0)|. The separation after n iterations is |δn | ≡ |δ(n∆t)| ≈ |δ(0)|eλn∆t = |δ0 |en(λ∆t) . (2.26) Thus, as long as λ > 0, nearby trajectories separate exponentially fast at the sampling points. Since the continuous time system is well-behaved, the trajectories separate exponentially fast. Therefore a continuous time system is chaotic if the iterated map obtained by the time T map is chaotic. This fact will be used in Sec. (2.2) to verify that the continuous time hybrid differential equation is chaotic. 2.1.5 Poincaré Section and Poincaré Return Map As has been mentioned in Sec. (2.1.2), a continuous time system can be turned into a discrete time system via the Poincaré section method. A Poincaré section of 2.1. A Brief Introduction to Dynamics and Chaos 25 Figure 2.5: An example of a Poincaré section in 2-D phase space. In the x-v phase plane, a Poincaré section could be a straight line passing through the origin, such that the trajectories intersect the Poincaré section transversely, as shown in the figure. Given a particular trajectory and a point x0 where the trajectory crosses the Poincaré line, the next crossing point x1 can be found by following the trajectory, and similarly all the future crossings x2 , x3 , etc. can be found. The mapping P that carries one crossing point to the next P (xn ) = xn+1 is known as the Poincaré map, and x1 , x2 , etc. are the returns to the Poincaré map. an N -dimensional phase space is an (N − 1)-dimensional subspace of the original phase space such that the trajectories of the dynamical system intersect the plane transversely, which means that trajectories are not parallel to the Poincaré section. The Poincaré section of a 2-D system may be a straight line, while it may be a plane for a 3-D system. Examples of Poincaré sections are shown in Figure 2.5 and 2.6. Now given a particular trajectory of the continuous time system and a cross point x0 on the Poincaré section, one can treat the cross point as an initial condition and integrate over time, until the trajectory next crosses the Poincaré section at x1 . This process can be continued and a series of cross points x2 , x3 , etc. can be obtained. The Poincaré map P is the mapping that maps a cross point xn to the next cross point xn+1 : P (xn ) = xn+1 . (2.27) The crossing points xn are sometimes called the returns of the Poincaré map. As the returns of the Poincaré map are confined to an (N − 1)-dimensional Poincaré section, it simplifies the geometric description of the dynamics by removing one of the phase space dimensions. Nevertheless, it still keeps the essential information of the dynamical system such as periodicity, quasi-periodicity and chaoticity. For more discussion, see Ott [6] and Hilborn [7]. 2.1.6 Symbol Sequences, Shift Maps and Symbolic Dynamics Suppose the phase space of a system is the interval between 0 and 1 (including 0 but excluding 1) on the real axis. Every point in the interval represents a possible state of the system. Then the system has infinitely many possible states, corresponding to the uncountably many points in the interval [0, 1). The goal is to use some sequence of symbols to represent each state. 26 Chapter 2. Introduction to the Hybrid System Figure 2.6: An example of a Poincaré section in 3-D phase space. In the xyz phase space, a Poincaré section could be a plane passing through the origin, such that the trajectories intersect the Poincaré section transversely, as shown in the figure. x1 , x2 and x3 are the returns to the Poincaré map defined by the shown trajectory and initial crossing point x0 . 0 0.3 1/2 1 0 1/4 0.3 1/2 Figure 2.7: Determination of symbol sequence representation of 0.3. As 0.3 is in the left half interval of [0, 1), the first symbol for 0.3 is L. Next as 0.3 is in the right half interval of [0, 1/2) (0.3 ∈ [1/4, 1/2)), its second symbol is R. Hence the first two symbols of the symbol sequence representation of 0.3 is LR. It is clear that every point in [0, 1) must be either in [0, 1/2) or [1/2, 1). If the point is in [0, 1/2), write down the first symbol as L, otherwise write down R. Visually it makes sense: if the point is in the left half of the interval it can be represented as LEFT, or L, and if it is in the right half of the interval it can be represented as RIGHT, or R. Once the first symbol is determined, a similar procedure can be carried out to determine the second symbol. For instance, if the point is in [0, 1/2), then it must be in either [0, 1/4) or [1/4, 1/2). If it is in [0, 1/4), write down the second symbol L, corresponding to it being in the left half interval of [0, 1/2). Otherwise write down the second symbol as R. If the point is in [1/2, 1) instead, the interval can be partitioned in halves and the second symbol can be written down. By this procedure, 0.3 would have a LR representation for the first two symbols, as shown in Figure 2.7. This procedure can be continued infinitely many times for any point in [0, 1), and a infinite sequence of symbols for each point can be obtained. For example, 0 will be represented as LLL . . ., 1/3 will be represented as LRLR . . ., and 1/2 will be 2.1. A Brief Introduction to Dynamics and Chaos 27 represented as RLLL . . .. Two different points will always have different symbolic representations: it can be easily shown that two points that are δ apart will have different symbols in their sequences after at most b− log2 δc symbols, where bxc picks up the integer part of x. It can also be shown that each sequence will uniquely determine a point in the interval (the sequence RRR . . . will not be a valid representation under the definition above, since it will represent the point 1, but 1 is not in the interval). This means that there is a bijective mapping between each point in the interval and each different sequence of symbols. As each point corresponds to a possible state of the system, there is a bijective relation between states of the system and symbol sequences. Next the shift map operation S that acts on the symbol sequences will be defined. When S acts on a particular symbol sequence, it removes the first symbol in the sequence and shifts all other symbols one position to the left. For example, when the shift map is applied to RLLL . . ., LLLL . . . is obtained: S(RLLL . . .) = LLLL . . .. As RLLL . . . = 0.5 and LLL . . . = 0, the shift map has effectively mapped the symbol sequence for 0.5 to the symbol sequence for 0. This particular shift map is known as the Bernoulli shift map. It can be proved mathematically that the shift map is equivalent to the mapping S(x) ≡ 2x (mod 1), (2.28) which multiplies the original number by two and only keeps the decimal part of the product. As the shift map (in this case, the Bernoulli shift map) maps a point in the phase space to another point in the phase space, the system is closed, and by tracing out the points determined by the shift map, a trajectory for any initial point x ∈ [0, 1) can be obtained. Given the phase space ([0, 1)) and time evolution rule (the shift map), the system is effectively a dynamical system. The dynamics of such dynamical systems are known as symbolic dynamics. It should be noted that the dynamics of symbolic dynamics is completely deterministic, as the symbol sequence is completely fixed once the point is specified. I will demonstrate how the symbolic dynamics works with a few points. Firstly, starting with 0 (or LLL . . .), it is clear that no matter how many times the Bernoulli shift map is applied, LLL . . . (or 0) is always obtained. This means that it is a fixed point for the system. Next, starting with 0.25 (or LRLLL . . .), S(LRLLL . . .) = RLLL . . . (or S(0.25) = 0.5), S(RLLL . . .) = LLL . . . (or S(0.5) = 0). This means that it is a trajectory that eventually settles down to a fixed point. More complicated trajectories can be obtained as well. Starting with 2/7, or LRLLRLLRLLR . . ., it is easy to see that after applying S 3 times the original point is obtained. This means that it is a periodic orbit of period 3. It is possible that a symbol sequence does not demonstrate any periodicity. In fact, any symbol sequence obtained from an irrational number is aperiodic, and the resulting trajectory demonstrates aperiodicity. It should be noted that L and R are purely symbols, and hence they can be replaced with any other symbols. If instead of L and R, “0” and “1” are used for symbols, then interestingly, the symbol sequence of a point in the interval [0, 1) will be exactly the same as the digits after the decimal point in its binary representation (of 28 Chapter 2. Introduction to the Hybrid System 1 0.9 0.8 0.7 xn+1 0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.2 0.4 0.6 0.8 1 xn Figure 2.8: Bernoulli Shift Map. course infinitely many 0’s have to be inserted at the end of the binary representation). For example, in binary representation 0.5(10) = 0.1000 . . .(2) , and 0.5 in symbolic representation is 1000 . . .. The decimal representation of a number 0 ≤ x < 1 is related to its binary representation 0.s1 s2 ... via: x= ∞ X si i=1 2i . (2.29) A plot of the Bernoulli shift map is shown in Figure 2.8. It is composed of two straight lines of slope 2. It is clear from the plot that the Bernoulli shift map is not invertible, as for each horizontal section there are two intersections. 2.2 Solution to the Hybrid Differential Equation In this section a solution to the hybrid differential equation Eq. (2.1) will be derived. Consider the initial conditions u(0) = u0 , u̇(0) = 0, and s(0) = s0 , where |u0 | ≤ 1. For |u0 | = 1, it is not hard to see that u(t) = u0 , s(t) = s0 is a solution to Eq. (2.1) for all t > 0. In this case, the solution is a fixed point. For |u0 | < 1, let ũ ≡ u − s0 . Then the differential equation is equivalent to ũ¨ − 2β ũ˙ + (ω 2 + β 2 )ũ = 0, (2.30) which has the solution βt ũ(t) = ũ(0)e β cos ωt − sin ωt , ω (2.31) 2.2. Solution to the Hybrid Differential Equation or β cos ωt − sin ωt . u(t) = s0 + (u0 − s0 )e ω βt 29 (2.32) Note that the solution given in Eq. (2.32) subsumes the fixed point solution u0 = ±1, and hence Eq. (2.32) is valid for all |u0 | ≤ 1. The next step is to find when the guard condition is met, or when is u̇ = 0. To do this, differentiate Eq. (2.32) with respect to t: β βt cos ωt − sin ωt + (u0 − s0 )eβt (−ω sin ωt − β cos ωt) u̇(t) = β(u0 − s0 )e ω 2 2 ω +β = − (u0 − s0 )eβt sin ωt. (2.33) ω Since ω = 2π, the guard condition u̇ = 0 is first met when t = 1/2. Thus the next initial condition is 1 u = s0 −(u0 −s0 )eβ/2 = s0 [1 + (1 − u0 /s0 )] eβ/2 = s0 [1 + (1 − |u0 |)] eβ/2 . (2.34) 2 The last equality follows from the fact that |s0 | = 1 and u0 and s0 have the same sign. Because |u0 | ≤ 1, it follows that u(1/2) and s0 have the same sign as well: 1 = s0 . (2.35) sgn u 2 Therefore the discrete state s(t) remains unchanged by this initial trigger of the guard condition and the solution given by Eq.(2.32) is valid until at least the second trigger of the guard condition. The same process is repeated to find the time the guard condition is next met: t = 1. At this time, u1 ≡ u(1) = eβ u0 − (eβ − 1)s0 . (2.36) Now I will show |u1 | ≤ 1. Since 0 < β ≤ ln 2, 1 < eβ ≤ 2. Assume that 0 ≤ u0 ≤ s0 = 1. Then u1 = u0 eβ − (eβ − 1)s0 = s0 − (s0 − u0 )eβ ≤ s0 − (s0 − u0 ) = u0 ≤ 1, (2.37) u1 = u0 eβ − (eβ − 1)s0 = s0 − (s0 − u0 )eβ ≥ s0 − 2(s0 − u0 ) = 2u0 − s0 ≥ −s0 = −1. (2.38) and Thus under the assumption that 0 ≤ u0 ≤ s0 = 1, by Eq. (2.37) and Eq. (2.38), |u1 | ≤ 1. Similarly the same result is obtained for −1 = s0 ≤ u0 < 0. Therefore |u1 | ≤ 1. Note that s1 = sgn(u1 ) = sgn eβ u0 − (eβ − 1)s0 , (2.39) 30 Chapter 2. Introduction to the Hybrid System explicitly depends on the value of u0 , thus a transition in the discrete state s can only occur at t = 1. In the case that the value of s does change at t = 1, Eq. (2.30) is no longer valid for t > 1, and thus the solution given in Eq. (2.32) is not valid for t ≥ 1. It is also helpful to note that a Poincaré section defined by u̇(t) = 0 can be chosen. Then transitions can only occur at the returns of this Poincaré map, the points in phase space where the trajectory crosses the Poincaré section. To continue to solve Eq. (2.1) for t > 1, note that the initial condition now becomes u(1) = u1 , u̇(1) = 0 and s(1) = s1 , where |u1 | ≤ 1 and s1 = sgn(u1 ). Comparing this set of initial conditions to the set of initial conditions specified at the beginning of this section, it is easy to see that this problem is equivalent to the original initial value problem if a unit time translation and an increment of the subscripts are applied, which allows the solution for 1 ≤ t < 2 to be written as β β(t−1) (2.40) cos ωt − sin ωt . u(t) = s1 + (u1 − s1 )e ω Repeating the process above extends the solution to a general one that is valid for n ≤ t < n + 1, where n ∈ Z is a non-negative integer: β β(t−n) u(t) = sn + (un − sn )e cos ωt − sin ωt , (2.41) ω where the returns at the transition times satisfy the following recurrence relation: un+1 ≡ u(n + 1) = eβ un − (eβ − 1)sn , (2.42) sn+1 = sgn(un+1 ). (2.43) and sn+1 is defined as In principle, the hybrid system can be solved exactly using this method for all t ≥ 0, given the initial condition (u0 , s0 ). Defining a mapping M as M(un ) = eβ un − (eβ − 1)sn , (2.44) it is easy to see that Eq. (2.42) is an iterated map. As M 0 (un ) = eβ (2.45) is a constant for all un , use Eq. (2.25) to obtain the Lyapunov exponent λ of the iterated map: m−1 m−1 1 X β 1 X 0 ln |M (un )| = lim ln e = β. λ = lim m→∞ m m→∞ m n=0 n=0 (2.46) Since β > 0, the iterated map is chaotic. As the iterated map comprises returns at regular time intervals of a continuous time system, by the discussion in the last paragraph of Sec. (2.1.4), the continuous time system is also chaotic. 2.2. Solution to the Hybrid Differential Equation 31 Next an expression of u(t) using only si ’s will be derived. Changing the indices in Eq. (2.42) and multiplying by a suitable factor of eβ the following set of equations is obtained un = eβ un−1 − (eβ − 1)sn−1 , eβ un−1 = e2β un−2 − eβ (eβ − 1)sn−2 , .. . e(n−1)β u1 = enβ u0 − e(n−1)β (eβ − 1)s0 , (2.47) then adding up all the equations in Eq. (2.48) and canceling repeated terms yields nβ β un = e u0 − (e − 1) n−1 X e(n−i)β si i=0 ( = enβ u0 − (1 − e−β ) n−1 X ) si e−iβ . (2.48) i=0 Rearranging Eq. (2.48) yields −nβ u0 = e −β un + (1 − e ) n−1 X si e−iβ . (2.49) i=0 Since |un | ≤ 1 is bounded and e−nβ decays exponentially with increasing n, taking the limit of Eq. (2.49) as n → ∞ yields ( ) n−1 X −nβ −β −iβ u0 = lim e un + (1 − e ) si e n→∞ n→∞ i=0 = (1 − e−β ) ∞ X si e−iβ (2.50) i=0 In Eq. (2.50), the initial condition u0 has been expressed exclusively in terms of current and future si ’s for 0 ≤ i < ∞. Thus the si ’s can be “read” by resolving the initial condition u0 . Changing the indices in Eq. (2.50) yields −β un = (1 − e ) ∞ X si+n e−iβ , (2.51) i=0 which represents future returns purely in terms of current and future si ’s. Note that si takes value of only +1 or −1. If one views the possible values for si as symbols (as has been discussed in Sec. (2.1.6)), and forms symbol sequences Si as Si = si si+1 si+2 . . . , (2.52) then it is easy to see that each ui corresponds to a symbol sequence Si . The next return ui+1 is related to the current return ui via the shift map, as the shift map 32 Chapter 2. Introduction to the Hybrid System shifts the symbol sequence Si to Si+1 . Thus the symbols and the shift map form a symbolic dynamics for the chaotic iterated map. Now u(t) can be written purely in terms of current and future symbols. Plugging Eq. (2.51) into Eq. (2.32) gives ( ) ∞ X β −β −iβ β(t−n) cos ωt − sin ωt , (2.53) u(t) = sn + −sn + (1 − e ) si+n e e ω i=0 where n = btc is the largest integer that is smaller than t. As one can get the current and future symbols by resolving the initial condition u0 , the hybrid differential equation (2.1) has been solved exactly. 2.2.1 Discussion of the Solution for β = ln 2 Plugging β = ln 2 into the solution for the returns to the Poincaré section Eq. (2.51) yields ∞ ∞ 1 X sn+i X sn+i = , (2.54) un = i 2 i=0 2i 2 i=1 and the relation between successive returns is un+1 = 2un − sn , (2.55) which is found by plugging β = ln 2 into Eq. (2.42). Figure 2.9 shows the plot of un+1 versus un . It is not hard to see that Figure 2.9 resembles Figure 2.8, except that the domain and range are slightly different: the domain and range are [−1, 1) instead of [0, 1). It is not surprising, because the solution Eq.(2.54) has the same form as Eq. (2.29), which is the binary representation of a real number between 0 and 1, except that in Eq.(2.54) the symbols are −1 and +1 instead of 0 and 1. In fact, making the substitution of xn = un2+1 yields xn un + 1 1 = = 2 2 1 = 2 = = ∞ X sn+i i=1 i=1 ∞ X + i=1 2i 1 2i ! +1 ! ∞ 1 X sn+i + 1 2 i=1 ∞ X s0n+i , 2i i=1 where s0n+i 2i ∞ X sn+i sn+i + 1 = = 2 2i (2.56) +1, sn+i = +1 0, sn+i = −1, (2.57) 2.2. Solution to the Hybrid Differential Equation 33 1 0.8 0.6 0.4 un+1 0.2 0 −0.2 −0.4 −0.6 −0.8 −1 −1 −0.5 0 un 0.5 1 Figure 2.9: un+1 versus un for β = ln 2. This mapping looks just like the Bernoulli shift map (see Figure 2.8), except that the domain and range of the mapping are [−1, 1) instead of [0, 1). which is exactly the same as Eq. (2.29). This means that the solution to the returns of the Poincaré map for β = ln 2 is equivalent to a Bernoulli shift map under a suitable change of coordinates. Since the Bernoulli shift map is known to be chaotic, the discrete mapping for β = ln 2 is chaotic, and since the points in the mapping are sampled at regular intervals, the continuous time system is also chaotic for β = ln 2. ˙ = 0, |u(t)| < 1 will be In Chap. 3, the returns to the Poincaré section u(t) determined and the returns un+1 versus un will be plotted, like what has been done in this section. The data points will then be fitted using linear regression method and the slope of mapping, k, will be determined using the slope of the fitted lines. Subsequently the slope can be used to determine β = ln k. Chapter 3 Experimental Data and Analysis Experiments using the setup described in Chap. 1 are carried out to test the theoretical solutions to the hybrid differential equation Eq.(2.1). The position of cylindrical magnets is measured every 7 ms and is recorded as voltage. Figure 3.1 shows a typical waveform obtained, which demonstrates the aperiodic behavior of the oscillator. Next, the position x is normalized to set the two offset states to be +1 and −1, and is then Fourier transformed and filtered to eliminate high frequency noise. The velocity v is calculated as the time derivative of position x. The resulting phase space plot is shown in Figure 3.2. A Poincaré section, defined as the points where v = 0 and |x| ≤ 1, is then taken and the returns to this Poincaré section are picked out and plotted, giving rise to the Poincaré map shown in Figure 3.3. Figure 3.3 has two notable features. Firstly, the end points (points for which un < −0.8 or un > 0.8) seem to behave differently from points in the middle region, possibly as a result of increasing friction that is inherent in the mechanical system. Therefore they are ruled out from the data analysis. Secondly, recognizing that Figure 3.3 resembles a Bernoulli shift map, the region [−0.8, 0.8] × [−0.8, 0.8] can be mapped to [0, 1] × [0, 1] in a way that linearity is preserved. Figure 3.4 is then obtained. The left and right stripes are then separately fitted to linear functions. The left stripe yields a linear fit of u0n+1 = 1.9736u0n + 0.0057, (3.1) while the right stripe yields a linear fit of u0n+1 = 2.0027u0n − 0.9603. (3.2) Therefore the Poincaré returns form a Bernoulli shift map to a good approximation under a suitable change of coordinates, and the system is indeed chaotic. Since the slopes of the linear fits approximate 2 to a good degree, the slope in Eq. (2.44) is eβ = 2. Therefore β = ln 2. All the symbols in the symbol sequence representation of the solution can be obtained by reading the sign for each u(t) where u̇(t) = 0, and the solutions to the hybrid differential equation Eq. (2.1) can be ob- 36 Chapter 3. Experimental Data and Analysis 4.5 zsmoothed (V) 4 3.5 3 2.5 2 1.5 6.5 7 7.5 8 8.5 9 9.5 10 Time (s) Figure 3.1: Position measured in Volts v.s. time measured in seconds. The diagram shows a typical waveform obtained from a 7501-point measurement of the position of the cylindrical magnets. Measurements are taken every 7ms. 80 60 40 20 v 0 −20 −40 −60 −80 −100 −3 −2 −1 0 u 1 Figure 3.2: Phase space diagram of experiment data. 2 3 37 1 0.8 0.6 0.4 un+1 0.2 0 −0.2 −0.4 −0.6 −0.8 −1 −1 −0.5 0 un 0.5 1 Figure 3.3: Poincaré return map for experiment data. 1 0.9 0.8 0.7 un+1 0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.2 0.4 un 0.6 0.8 1 Figure 3.4: Poincaré return map in the form of Bernoulli shift map. The Poincaré returns in the middle region [−0.8, 0.8] × [−0.8, 0.8] are selected and mapped to [0, 1] × [0, 1]. It is clear from the plot that they resemble a Bernoulli shift map. The left and right stripes are then separately fitted to linear functions. The slope for the left partition is calculated as 1.9736, while for the right partition is 2.0027. 38 Chapter 3. Experimental Data and Analysis 2 usmoothed 1 0 −1 −2 21.5 22 22.5 23 Time (s) 23.5 24 Figure 3.5: Comparison of analytic solution to measured experimental data. The analytic solution (black), calculated from Eq. (2.53) with β = ln 2, demonstrates a strong correlation with the measured data (blue). Also shown in the figure is the symbol sequence as obtained from experimental data (red) and Poincaré returns (square). tained from Eq. (2.53): u(t) = sn + 2 t−n −sn + ∞ X si+n i=1 2i ! β cos ωt − sin ωt , ω where n = btc. Then the waveform produced by the electromechanical oscillator can be “predicted” using the symbols as initial condition by plotting the analytic solution given in Eq.(3.3). A typical waveform in the resultant analytic solution is shown in black in Figure 3.5. Also shown in the same plot is the measured experimental data, which is painted in blue. As one can tell from Figure 3.5, the agreement is not perfect, which is possibly due to the following factors: 1. Amplitude noise. The mechanical system suffers friction: in addition to the friction forces due to the rotation of pulleys and movement of springs, the magnets also experience horizontal disturbances as the dipole moment of magnets are not perfectly aligned in the vertical direction. 2. Delays in the offset voltage signal. In the real experiment, transitions in the discrete state s in Eq. (2.1) cannot occur instantaneously. In addition, the RC differentiator used in the secondary coil circuit, as discussed in Sec. (1.4), inevitably introduces another delay in the form of a small phase shift. The outcome of this delay effect is obvious in Figure 3.2, where “overshooting” at v = 0 occurs. 3. Timing jitter. As a result of the aforementioned “overshooting” problem, the time intervals in between returns to the Poincaré section are not regular. These irregular time intervals make numerical solution hard, and for this thesis I take the average of the time intervals and use this averaged value as the unit time 24.5 39 standard. The time unit used for this thesis is 204 data points, which corresponds to 1.428 sec. Despite all these difficulties, the experimental data and analytic solution show a very strong correlation. Considering that it is a chaotic system, which is extremely sensitive to any small difference in initial condition, the agreement is really amazing. The agreement between the experiment and analytic solution demonstrates that it is possible to reproduce a chaotic wave signal from its symbol sequence representations. This may have potential application in data storage, since to recover a chaotic wave signal, it suffices to store only its symbol sequence representation, which hugely compresses the information needed to store the wave signal. Chapter 4 Further Discussion on the Hybrid System Despite the fact that the hybrid differential equation has been solved in Chap. 2 and the implementation of an electromechanical oscillator described by this equation has been discussed in Chap. 1 and Chap. 3, there are still many open questions. A few noticeable ones include: 1. The hybrid system consists of discrete states which are discontinuous. Is there a limit where an ordinary differential equation reduces to the hybrid system? 2. Hybrid systems are not ordinary differential equations, and hence the notion of phase space is complicated. There are basically two independent differential equations, and the two corresponding phase spaces, each being a 2D sheet, are combined to form a 2 × 2D phase space. Is there a well-defined embedding of the 2 × 2D sheets in a higher dimensional phase space? 3. Currently the electric circuit that outputs offset states to the mechanical oscillator calculates velocity using an RC differentiator, which inevitably introduces a small delay. While this delay is necessary for the design of the circuit1 , it changes the hybrid differential equation. Is it possible to reduce or even eliminate the delay while still implementing the discrete state signal? Whether these are answerable questions is not certain. Here, to finish up this thesis, the first question is explored. The goal is to write the discrete state s as a limit of continuous functions. One way to do this requires restating the guard condition for transition in s, given by Eq. (2.2) and Eq. (2.3): u̇ = 0 ⇒ s(t) = sgn(u(t)), where sgn(u) = 1 +1, u ≥ 0 −1, u < 0. (4.1) (4.2) In Figure 1.7, a DM74123N latch is used to detect a change in sign in v. The delay allows the oscillator to overshoot and causes a sign change in v, and without it a change in the discrete state will not be possible. In the setup proposed by Owen et al.[5], a similar mechanism, which requires a small delay, is also used to output the discrete state signal. 42 Chapter 4. Further Discussion on the Hybrid System In Figure 4.1, the waveform for a numerical solution to Eq. (2.1) as well as its corresponding waveforms for sgn(u), sgn(v) (v is the time derivative of u) and s are plotted. The following observation is made: s follows the shape of u but is delayed, the transition in s occurs when v = 0 is next met. Based on this observation, it is proposed that the modified continuous “discrete” state s should account for the following two events: 1. A transition in value of sgn(u). 2. The next occurrence for v = 0. In order to account for the occurrence of the above two events, intuitively two Dirac delta functions are needed. After trying out a few different combinations, the following expression is obtained: Z t Z t 0 0 0 0 0 sgn u(t ) δ u(t ) dt s(u, v, t) = sgn u(t) − 2 1− δ v(t ) dt , t−∆t t−∆t (4.3) where ∆t is a maximum time window between the transition in u and the occurrence of v = 0.2 In the expression, the first term demonstrates that s basically follows the waveform of sgn(u), the term Z t sgn u(t0 ) δ u(t0 ) dt0 t−∆t accounts for the occurrence and direction of transition in u, and the term Z t 1− δ v(t0 ) dt0 t−∆t accounts for the occurrence of v = 0. At this point, the signum function can be approximated using an error function, and the Dirac delta function using a normal distribution: sgn(x) = lim erf(kx), k→∞ 2 δ(x) = lim N (k 0 x), 0 k →∞ (4.4) One might doubt whether such a ∆t can be found at all. After all, if the ∆t is too big, then multiple events might be included in the integrals and the evaluation of s will be inaccurate. However, Figure 4.1 suggests that the time interval between a transition in u and a v = 0 event is less than 0.25. This should be obvious: taking the transition from positive to negative, for instance. The oscillator travels from the previous positive v = 0 point, where u > 1, to u = 0, then to the negative v = 0 point, where 0 > u > −1. The length of travel from the positive v = 0 point to u = 0 is larger than 1, while the length of travel from u = 0 to the negative v = 0 point is less than 1. It is then expected that the time interval between the transition in u and the next v = 0 event is less than the time interval between the previous v = 0 event and the transition in u. Since the total time interval between successive v = 0 events is 0.5, it is expected that the time interval between a transition in u and a v = 0 event is less than 0.25. Choosing ∆t = 0.25 should suffice for the purpose of this thesis, although a more careful proof should be done. 43 Figure 4.1: Waveforms for numerical solution and its corresponding u, v and s. 44 Chapter 4. Further Discussion on the Hybrid System and substitute v with u̇. Then the continuous form for the state function s is obtained: Z t Z t 0 0 0 0 0 0 0 N k u̇(t ) dt , 1− s(u, u̇, t) = erf ku(t) −2 erf ku(t ) N k u(t ) dt t−∆t t−∆t (4.5) where k and k 0 are some large numbers one can choose to approximate the Dirac delta function and the signum function. Comparing the continuous definition of s (Eq. (4.5)) to its original discrete definition (Eq. (4.1)), it is noted that s is now explicitly dependent on u̇. This piece of information, however, is not new, since the direction of transition in sgn(u) in the discrete definition (from −1 to +1 or from +1 to −1) implicitly tells the signum function of u̇. Eq. (4.5) allows the hybrid differential equation Eq. (2.1) to be written in terms of continuous functions of u and the time derivatives of u. Nevertheless, the continuous functions include integrals, which are not convenient for numerical integrations. The next goal is to convert Eq. (2.1) into a system of first-order, autonomous ordinary differential equations that are free from integrals, as in Eq. (2.4). In order to do so, denote 1 σ ≡ − s − erf ku(t) , 2 Z t x≡ erf ku(t0 ) N k 0 u(t0 ) dt0 , (4.6) t−∆t Z t y ≡1− N k 0 u̇(t0 ) dt0 , t−∆t then σ = xy. By the fundamental theorem of calculus, ẋ = erf ku(t) N k 0 u(t) − erf ku(t − ∆t) N k 0 u(t − ∆t) and ẏ = −N k 0 u̇(t) + N k 0 u̇(t − ∆t) , which are free from integrals. If one can represent some time derivatives of σ or their linear combinations as some function of the time derivatives of x and y, but not directly in terms of x and y, then a system of differential equations that are free from integrals can be obtained. Start from σ = xy. Differentiating both sides with respect to time yields σ̇ = ẋy + xẏ. (4.7) Rearranging Eq. (4.7) to solve for x, 1 x = (σ̇ − ẋy). ẏ (4.8) Differentiating both sides of Eq. (4.7) with respect to time yields σ̈ = ẍy + 2ẋẏ + xÿ. (4.9) 45 Plugging Eq. (4.8) into Eq. (4.9) gives 1 σ̇ ẋ σ̈ = ẍy + 2ẋẏ + (σ̇ − ẋy)ÿ = 2ẋẏ + ÿ + ẍ − ÿ y. ẏ ẏ ẏ (4.10) Rearranging Eq. (4.10) to solve for y, y= σ̈ − 2ẋẏ − σ̇ẏ ÿ ẍ − ẋẏ ÿ = σ̈ ẏ − 2ẋẏ 2 − σ̇ ÿ . ẍẏ − ẋÿ (4.11) Differentiating both sides of Eq. (4.9) with respect to time and plugging in Eq. (4.8) and Eq. (4.11) yields ... ... ... σ = x y + ẍẏ + 2ẍẏ + 2ẋÿ + ẋÿ + x y .. . after some algebra ... ... ... ... ... ... x ẏ − ẋ y ẍ y − x ÿ 2ẋẏ( x ẏ − ẋ y ) = 3(ẍẏ + ẋÿ) + σ̈ + σ̇ + . (4.12) ẍẏ − ẋÿ ẍẏ − ẋÿ ẍẏ − ẋÿ At this point, linear combination of time derivatives of σ has been successfully written ... as functions of time derivatives of x and y. Note that σ is symmetric under the exchange of x and y, which is expected since σ is symmetric under the exchange of x and y. Define a ≡ −2β and b ≡ ω 2 + β 2 . Then the hybrid differential equation Eq. (2.1) can be written as ü + au̇ + bu = bs = b[erf(ku) − 2σ] = −2bσ + b erf(ku). (4.13) Denote α ≡ −2b, β ≡ b erf(ku). Eq. (4.13) can be written as ü + au̇ + bu = ασ + β. (4.14) Differentiating both sides of Eq. (4.14) multiple times yields: ... u + aü + bu̇ = ασ̇ + β̇, ... (4) u + a u + bü = ασ̈ + β̈, ... ... ... u(5) + au(4) + b u = α σ + β . (4.15) ... Rearranging Eq. (4.15) to solve for σ̇, σ̈ and σ , 1 ... [ u + aü + bu̇ − β̇], α 1 ... σ̈ = [u(4) + a u + bü − β̈], α ... 1 (5) ... ... σ = [u + au(4) + b u − β ], α σ̇ = (4.16) 46 Chapter 4. Further Discussion on the Hybrid System Plugging Eq. (4.16) into Eq. (4.12) yields ... ... 1 (5) 1 x ẏ − ẋ y (4) ... ... ... (4) [u + au + b u − β ] = [u + a u + bü − β̈] α α ẍẏ − ẋÿ ... ... 1 ẍ y − x ÿ ... [ u + aü + bu̇ − β̇] + α ẍẏ − ẋÿ ... ... 2ẋẏ( x ẏ − ẋ y ) + + 3(ẍẏ + ẋÿ). ẍẏ − ẋÿ Let ... ... x ẏ − ẋ y X1 = ẍẏ − ẋÿ and ... ... ẍ y − x ÿ X2 = , ẍẏ − ẋÿ (4.17) then rearranging Eq. (4.17) to solve for u(5) yields ... u(5) = (X1 − a)u(4) + (aX1 + X2 − b) u + (bX1 + aX2 )ü + bX2 u̇ + ... ... ... 2ẋẏ( x ẏ − ẋ y ) α + 3(ẍẏ + ẋÿ)α . β − X1 β̈ − X2 β̇ + ẍẏ − ẋÿ (4.18) Now, Eq. (2.1) can be converted into a system of first-order, autonomous ordinary differential equations that are free from integrals: u̇(0) u̇(1) u̇(2) u̇(3) u̇(4) = = = = = u̇, ü, ... u, u(4) , u(5) , (4.19) ... where u(5) , given by Eq. (4.18), is a function of u̇, ü, u and u(4) . Numerical integration can then be carried out for Eq. (4.19), the result of which can be compared to the numerical solution for the original hybrid differential equation. If the two numerical integrations agrees, it may suggest that the original hybrid differential equation with the discrete offset states is in fact a limiting case for a continuous differential equation, and more interesting theoretical problems can be explored for this differential equation. Conclusion This thesis has demonstrated the implementation of a chaotic electromechanical oscillator described by a hybrid differential equation. The hybrid differential equation is ü − 2β u̇ + (ω 2 + β 2 )(u − s) = 0, (5.20) where u(t) ∈ R is a continuous state, ω and β are fixed parameters, s(t) ∈ {±1} is a discrete state, and transitions in s(t) are made when u̇ satisfy the guard condition u̇ = 0 ⇒ s(t) = sgn(u(t)). (5.21) Following the paper by Corron[1], an analytic solution to the hybrid differential equation is obtained. An electromechanical oscillator that can be described by the hybrid differential equation is built based on a setup proposed by Owen et al.[5], with multiple improvements on mechanical and circuit design. The waveform produced by the electromechanical oscillator closely matches the analytic solution (as shown in Figure 3.5). In light of all the difficulties of implementation (as discussed in Chap. 3), as well as the fact that the system is chaotic, the agreement between the two curves is amazing. An extension to the theory of hybrid system is also proposed, which potentially allows the hybrid system to be written as a limiting case to an ordinary differential equation. Therefore I conclude that the chaotic electromechanical oscillator, described by a hybrid differential equation, has been successfully implemented. The success of recovering the original waveform using only symbol sequence demonstrates that it is a powerful and promising technique for data storage. Bibliography [1] N. J. Corron, J. N. Blakely, and M. T. Stahl, “A matched filter for chaos,” Chaos 20, 023123 (pages 10) (2010). [2] N. J. Corron, J. N. Blakely, S. T. Hayes, and S. D. Pethel, “Determinism in synthesized chaotic waveforms,” Phys. Rev. E 77, 037201 (2008). [3] N. J. Corron, S. T. Hayes, S. D. Pethel, and J. N. Blakely, “Chaos without Nonlinear Dynamics,” Phys. Rev. Lett. 97, 024101 (2006). [4] N. J. Corron, S. T. Hayes, S. D. Pethel, and J. N. Blakely, “Synthesizing folded band chaos,” Phys. Rev. E 75, 045201 (2007). [5] A. M. B. Owen, N. J. Corron, M. T. Stahl, J. N. Blakely, and L. Illing, “Exactly solvable chaos in an electromechanical oscillator,” (submitted). [6] E. Ott, Chaos in dynamical systems (Cambridge University Press, New York, NY, 2002). [7] R. Hilborn, Chaos and nonlinear dynamics: an introduction for scientists and engineers (Oxford University Press, New York, NY, 1994). [8] S. H. Strogatz., Nonlinear dynamics and chaos: with application to physics, biology, chemistry and engineering (Addison-Wesley Publishing Company, Reading, MA, 1994). [9] W. Zheng and B. Hao, Practical symbolic dynamics (Shanghai Technology and Education Press, Shanghai, China, 1994). [10] A. M. Samoilenko, “Quasiperiodic oscillations,” Scholarpedia 2, 1783 (2007).