Chaotic Billiards. Part I Introduction to Dynamical Billiards. 1 Review of the Dierent Billiard Systems. Ben Parker and Alex Riina.

Similar documents
1 The pendulum equation

Math 1270 Honors ODE I Fall, 2008 Class notes # 14. x 0 = F (x; y) y 0 = G (x; y) u 0 = au + bv = cu + dv

ENGI Parametric Vector Functions Page 5-01

Chaos in the Swinging Atwood Machine (SAM)

UNIVERSITY OF HOUSTON HIGH SCHOOL MATHEMATICS CONTEST Spring 2018 Calculus Test

Parametric Equations *

Chaotic motion. Phys 420/580 Lecture 10

Section Arclength and Curvature. (1) Arclength, (2) Parameterizing Curves by Arclength, (3) Curvature, (4) Osculating and Normal Planes.

PHY411 Lecture notes Part 5

1998 Harvard/MIT Math Tournament GEOMETRY Answer Sheet

4 The Trigonometric Functions

Point Vortex Dynamics in Two Dimensions

Created by T. Madas LINE INTEGRALS. Created by T. Madas

Lecture 6, September 1, 2017

No-Slip Billiards in Dimension Two

Minimum and maximum values *

GRE Math Subject Test #5 Solutions.

Is Quantum Mechanics Chaotic? Steven Anlage

Math 117: Calculus & Functions II

Contents. 1 Introduction to Dynamics. 1.1 Examples of Dynamical Systems

Chapter 9 Uniform Circular Motion

ENGI 4430 Parametric Vector Functions Page dt dt dt

MATH 162. FINAL EXAM ANSWERS December 17, 2006

Orbits, Integrals, and Chaos

= 3, the vertical velocty is. (x(t), y(t)) = (1 + 3t, 2 + 4t) for t [0, 1],

MAC2313 Final A. (5 pts) 1. How many of the following are necessarily true? i. The vector field F = 2x + 3y, 3x 5y is conservative.

a k 0, then k + 1 = 2 lim 1 + 1

Devil s Staircase Rotation Number of Outer Billiard with Polygonal Invariant Curves

MATH2111 Higher Several Variable Calculus Integration

MA 126 CALCULUS II Wednesday, December 14, 2016 FINAL EXAM. Closed book - Calculators and One Index Card are allowed! PART I

Solutions for B8b (Nonlinear Systems) Fake Past Exam (TT 10)

36. Double Integration over Non-Rectangular Regions of Type II

Integration of Constraint Equations in Problems of a Disc and a Ball Rolling on a Horizontal Plane

MA CALCULUS II Friday, December 09, 2011 FINAL EXAM. Closed Book - No calculators! PART I Each question is worth 4 points.

APPM 1360 Final Exam Spring 2016

The Liapunov Method for Determining Stability (DRAFT)

Circles Unit Test. Secondary Math II

MATH H53 : Final exam

Curves I: Curvature and Torsion. Table of contents

Practice Final Exam Solutions

Problem Set Number 01, MIT (Winter-Spring 2018)

Integrals. D. DeTurck. January 1, University of Pennsylvania. D. DeTurck Math A: Integrals 1 / 61

AP Calculus AB Winter Break Packet Happy Holidays!

WHAT IS A CHAOTIC ATTRACTOR?

DIFFERENTIATION RULES

Arc Length and Riemannian Metric Geometry

One Dimensional Dynamical Systems

Chapter 2 Polynomial and Rational Functions

MAC Calculus II Spring Homework #6 Some Solutions.

Unit IV Derivatives 20 Hours Finish by Christmas

Math 20C Homework 2 Partial Solutions

Trigonometry: Graphs of trig functions (Grade 10) *

Unit IV Derivatives 20 Hours Finish by Christmas

HOMEWORK 2 SOLUTIONS

Monte Carlo simulations of harmonic and anharmonic oscillators in discrete Euclidean time

Study Guide for Benchmark #1 Window of Opportunity: March 4-11

MTH301 Calculus II Glossary For Final Term Exam Preparation

Chaotic motion. Phys 750 Lecture 9

Bi. lkent Calculus II Exams

Engineering Mechanics Prof. U. S. Dixit Department of Mechanical Engineering Indian Institute of Technology, Guwahati Kinematics

Synchronization, Chaos, and the Dynamics of Coupled Oscillators. Supplemental 1. Winter Zachary Adams Undergraduate in Mathematics and Biology

One dimensional Maps

SOLUTIONS TO SECOND PRACTICE EXAM Math 21a, Spring 2003

PhysicsAndMathsTutor.com

Lecture 17 - Gyroscopes

BILLIARD DYNAMICS OF BOUNCING DUMBBELL

Exercises for Multivariable Differential Calculus XM521

Complex Analysis MATH 6300 Fall 2013 Homework 4

Practice Problems for Final Exam

MTH101 Calculus And Analytical Geometry Lecture Wise Questions and Answers For Final Term Exam Preparation

Growth of energy density for Klein-Gordon equation coupled to a chaotic oscillator

Secondary School Certificate Examination Syllabus MATHEMATICS. Class X examination in 2011 and onwards. SSC Part-II (Class X)

APPM 2350, Summer 2018: Exam 1 June 15, 2018

18.02 Multivariable Calculus Fall 2007

The Minimal Element Theorem

Math 112 (Calculus I) Final Exam

The Not-Formula Book for C2 Everything you need to know for Core 2 that won t be in the formula book Examination Board: AQA

Practice Final Exam Solutions

PRACTICE PROBLEMS. Please let me know if you find any mistakes in the text so that i can fix them. 1. Mixed partial derivatives.

MA 114 Worksheet Calendar Fall 2017

Chaos in the Hénon-Heiles system

Written test, 25 problems / 90 minutes

BC Exam 1 - Part I 28 questions No Calculator Allowed - Solutions C = 2. Which of the following must be true?

mathematical objects can be described via equations, functions, graphs, parameterization in R, R, and R.

Identifying Safe Zones for Planetary Satellite Orbiters

Department of Physics, Korea University Page 1 of 8

Advanced Higher Mathematics of Mechanics

In these notes we will outline a proof of Lobachevskiĭ s main equation for the angle of parallelism, namely

Introduction to Real Analysis

Vector Geometry Final Exam Review

Answer Key for AP Calculus AB Practice Exam, Section I

Math 11 Fall 2018 Practice Final Exam

LAMC Intermediate I March 8, Oleg Gleizer. Theorem 1 Any two inscribed angles subtending the same arc on a circle are congruent.

Exam 3 - Part I 28 questions No Calculator Allowed - Solutions. cos3x ( ) = 2 3. f x. du D. 4 u du E. u du x dx = 1

HUDSONVILLE HIGH SCHOOL COURSE FRAMEWORK

Integrated Math II. IM2.1.2 Interpret given situations as functions in graphs, formulas, and words.

MA 441 Advanced Engineering Mathematics I Assignments - Spring 2014

Our goal is to solve a general constant coecient linear second order. this way but that will not always happen). Once we have y 1, it will always

3 Stability and Lyapunov Functions

Flux Invariants for Shape

Transcription:

Chaotic Billiards Ben Parker and Alex Riina December 3, 2009 Part I Introduction to Dynamical Billiards 1 Review of the Dierent Billiard Systems In investigating dynamical billiard theory, we focus on two important examples that demonstrate a variety of behaviors and represent clear gradation in complexity. This paper mixes analytic and computational billiard theory to give a detailed picture of a circular billiards and billiards in a stadiuma two semi-circular regions separated by a rectangular region. We investigate sensitive dependence on initial coordinates and the structure of periodic orbits before considering whether each exhibits chaos. 1.1 Case Examples of the Billiards The circular billiard is one of the simplest nite billiards to analyze because the collision map can be described by a very simple system of equations. If θ is the angular representation of a collision point relative to the positive x axis and ψ is the angle of the trajectory against the tangent line to the circle, we can easily relate an element in the phase space by the system ( 1 2 0 1 ) ( Θn Ψ n ) = ( Θn+1 Ψ n+1 Since θ [0, 2π) and ψ [0, π] we can dene the phase space as C 1 [0, π]. We see that the map simply twists the θ coordinate by how high it is on the cylinder. By also calculating the determinant of this map, we see that the area of a shape in the phase space is preserved under the mapping. Computationally however, we represent the boundary as a single curve parameterized as (cos(t), sin(t)) for t [0, 2π). The stadium billiard however, represents an even more dicult problem to model mathematically but proves still tractable to computational analysis. The boundary of the stadium billiard can be thought of as a union of four curves. ) 1

There exists an upper horizontal portion, a right semi-circular portion, a lower horizontal portion, and a left semi-circular portion. The left and right curves are parameterized by the vectors ( cos(t) a, sin(t)) and (cos(t) + a, sin(t)), respectively, for π/2 t < π/2. The other two horizontal curves which complete the billiard vary from a to a, a 0, at heights of y = 1 and y = 1. For depiction below, a is equal to 1. Part II Numerical Calculations with Circular and Stadium Billiards 2 Numerical Calculations 2.1 Sensitive Dependence on Initial Conditions For both of the circular and the stadium billiards, we can demonstrate sensitive dependence on initial conditions by simply varying any on of the parameters of one orbit and let the others remain constant. As is the case for both the stadium and circular billiards, we start with a known initial condition for two given orbits of each billiards. We then create another orbit in each billiard, which starts at a small distance from the rst initial condition in each billiard. We start by observing the case of the circular billiard. In this particular example, we start with an initial point (0.5, 0), and set the initial x velocity 2

to 0 and y velocity to 1. For the other initial condition in the billiard, each initial condition is the same except the starting x-coordinate of the second initial condition. In this case, we retain the same initial x and y velocities, but instead start at the point (0.5 + 1 10 3, 0). Thus, we seek to nd the rst collision at which two collision points are at least 0.05 units apart. The graphs on the next page show that this separation is reached by the twenty-fourth bounce along the circular billiard. The right-hand plot measures the Euclidean distance between each collision point on the vertical axis the nth collision along the boundary on the horizontal axis. Sensitive dependence on initial conditions for the stadium billiard follows the same premise as the circular billiard example, but with slight modications. That is, we start with an initial point (0, 0) in the stadium billiard, and let the initial x and y velocities equal 1 and 0.2, respectively: 3

In this particular plot, the blue trajectory is the is the initial condition starting at (0, 0), whereas the green trajectory starts at (1 10 10, 0). To show sensitive dependence, we specify that the trajectories must separate by a distance d of at least 0.25 after a certain number of collisions. The graph on the right indicates that this separation occurs at the twentieth collision along the billiard. As with the circular billiard, the horizontal axis species the nth collision along the stadium billiard, and the vertical axis is a measure of the Euclidean distance between each collision point. 2.2 Periodic Orbits Certain initial conditions in the circle and stadium billiard give rise to periodic orbits. For the circular billiard, there are many ways to create periodic orbits, one of which are shown below. In this case, we have a period-four orbit generated by the initial condition (0, 1/ 2). The stadium billiard exhibits similar properties to that of the circular billiards, particularly with period-four orbits. In this case, we observe that 4

starting at an initial condition of (0, 1/ 2) with initial x velocity 1 and y velocity 0 result in a similar period 4 orbit: For other periodic orbits of the circle and the stadium, they both share similarities in the generation of period-two orbits. We can generate a period two orbit in the circular case by starting at the point (0, 0) and choosing any x or y velocity, so long as they are not both zero. For period-two orbits in the stadium billiard, the only dierence lies in the fact that there are multiple period-two orbits that oscillate between y = 1 and y = 1, whereas the circular billiard has only one such orbit. 2.3 Lyapunov Exponents 2.3.1 Numerical Computation of Lyapunov Exponents Calculating the Lyapunov exponents of each billiard system in question will help to explain the behavior of the system, and consequently whether the given system is chaotic. To calculate a lower bound on the largest Lyapunov exponent, we measure the log of the growing dierence as we apply the map thousands of times to close by initial points. This estimation gives us a lower bound for the largest Lyapunov exponent since the maximum log of the growing distance occurs in the eigen-direction for the eigen-value corresponding to the leading Lyapunov number. Given that two points will not necessarily preserve their orientation under the mapping, the traditional method for calculating the Lyapunov exponent relies on constant orientation and normalization by the Gram-Schmidt process to calculate a Lyapunov number. Similarly, by not adjusting the direction, we cannot guarantee that the orientation of the points' orbits lie in the eigendirection, 5

but we do know that they cannot separate faster than they would in the eigendirection. So this calculation reveals a lower bound on the Lyapunov exponent. For the circular billiard, a lower bound on the largest Lyapunov exponent was obtained over thousands of iterations of the map between two nearby initial conditions. With many pairs of randomly chosen initial conditions from within the circle, we take the maximum value of the lower bounds Lyapunov exponents for many pairs of initial coordinates to nd an approximation of the largest Lyapunov exponent. We were able to calculate a lowever bound of the Lyapunov Exponent approximately equal to zero for the circular billiard. Which does not reveal anything in particular about the behavior of the circular billiard. The stadium billiard presents a more interesting case than the circular billiard because the horizontal side-length a is just a variable in our calculation. That is, by letting a vary for a 0, we obtain a range of Lyapunov exponents. Below, we plot Lyapunov exponents against values of a. 2.3.2 Alternative Calculations of the Lyapunov Exponents Alternate Method of Calculating the Lyapunov Exponents of the Circular Billiard Instead of chosing points close by and measuring their growing separation as we apply the map, we can calculate the Lyapunov exponent using the matrix relation. In terms of Θ and Ψ we have the relation ( 1 2 0 1 ) ( Θn ) = Ψ n ( ) Θn+1 Ψ n+1 To calculate the Lyapunov numbers of this map, we calculate the eigen-values of Φ n (Φ n ) τ which individually represent the squares of the nth powers of the 6

Lyapunov numbers of the map. Thus, we can calculate the Lyapunov exponents 1 using h k = lim n 2n log(rn k ) [3]. Using Matlab, this is a simple calculation, revealing that our approximation of each the Lyapunov exponent approaches zero, so the circular billiard is non-chaotic. Part III Invariant Measure of the Map Through the investigation of the circular billiard we found that a volume element in the phase space was preserved in the ow, but under what conditions does this invariance hold? To investigate whether a volume element in the phase space is preserved under the ow, we rst answer the same question regarding a time-t map for arbitrarily small time. The evaluation of a time-t map, Φ t (x, y, ẋ, ẏ), can be reduced to two cases by making t to be arbitrarily small. In the rst case no collision occurs during a time interval and in the second case, exactly one collision occurs. Φ t can be reduced to these two cases since given any starting point in the phase space since the ow can be broken into several segments, each of which fall into one of the two previous cases. If a particular ow has more than one collision in a time, t, then either we can break the time interval into many subintervals where only one collision occurs. If we cannot break it into subintervals where only one collision occurs, we have reached an accumulation point where the trajectory tends towards a point as t grows large. Instead of handling these cases, we restrict the following theorem to billiards where there are no accumulating points and where the boundary curves are well behaved through a series of assumptions made about the system. 7

3 Theorem The ow Φ τ preserves the volume element dx dy dω 3.1 Assumptions[1] 1. The boundary of the billiard is a nite union of smooth compact curves with continuous third derivatives. If the boundary were composed of an innite union of curves, we could easily violate the requirement that the third derivative be continuous by creating an innitely ne set of points connected by the rst order taylor series approximation of a function with a discontinuous third derivative. The other requirements restrict the boundary to parameterizable curves an important step in analytically calculating the aect of the ow on a volume element. 2. The individual curves that compose the boundary only intersect eachother at their end points. Restricting this helps us to classify the intersection and corners. 3. The second derivative of each curve in the boundary is either identically zero or is never zero. Which helps to classify the boundaries as either focusing, diverging, or at. 4. A billiard contains no cusps made by a focusing wall and a dispersing wall. According to Halpern's Theorem, Assumption 4 removes the remaining cases where a trajectory can accumulate. Such a trajectory has a limit as we allow the particle to travel and theoretically approaches the corner. 3.1.1 Proving the Theorem To describe the two cases rigorously, we introduce variables to represent the state of a particle before and after the map is applied. x ± y ± ω ± refers to the x coordinates of a point before and after the map is applied refers to the y coordinates of a point before and after the map is applied refers to the angles of the velocity against the x axis before and after the map is applied. Since the speed can be set to one, we can treat it similarly to the velocity of the particle. 8

Case: No Collisions In the case where there are no collisions, the map can be expressed as the simple system of equations relating the input coordinates and velocity and the output coordinates and velocity. x + = x + cos(ω) y + = y + sin(ω) Dierentiating reveals ω + = ω (= ω) dx + = dx sin(ω)tdω dy + = dy + cos(ω)tdω dω + = dω To compute the change in volume of an element dx dy dω as it is passed through the map Φ t, we relate the volumes by the equation dx + dy + dω + = det(j )dx dy dω where J is the Jacobian matrix of the mapping. dx + dy + dω + = 1 0 sin(ω) 0 1 cos(ω) 0 0 1 dx dy dω J = 1 0 sin(ω) 0 1 cos(ω) 0 0 1 dx + dy + dω + = det(j )dx dy dω = dx dy dω Thus, when there are no collisions, the time-t map preserves the volume of an element in the phase space. Case: Single Collision In the case where there is a single collision, we introduce `intermediate' variables to simplify the analysis. Let Γ i ( x, ȳ) be the segment of the boundary on which the collision occurs be the location of the collision 9

γ be the angle of the line tangent to Γ i at ( x, ȳ) o of the x axis ψ be the angle of the velocity o of the line tangent to Γ i at ( x, ȳ) r parametrize the arc length of Γ i s + s + = t represent the time interval before the collision and the time interval afterward. Using the system that these variables describe, we will rst model the map Φ t : (r, s, ψ) (x, y, ω ) and compare the volume elements drds dψ and dx dy dω using the Jacobian of Φ t, J Secondly, we will model the map Φ + t : (r, s +, ψ) (x +, y +, ω + ) and compare the volume elements drds + dψ and dx + dy + dω + using the Jacobian of Φ + t, J +. Backwards Map Φ t : (r, s, ψ) (x, y, ω ) x = x s cos(ω ) Since s cos(ω ) is the projection of the precollisional velocity on the x axis scaled to equal the distance traveled precollision y = ȳ s sin(ω ) Since s sin(ω ) is the projection of the precollisional velocity on the y axis scaled to equal the distance traveled precollision ω = γ ψ Dierentiating reveals dx = cos(γ)dr cos(ω )ds + s sin(ω )dω dy = sin(γ)dr sin(ω )ds s sin(ω )dω dω = Kdr dψ dx = cos(γ)dr dy = sin(γ)dr dγ = Kdr Substituting the denition of dω in resolves the system to dx = [cos(γ) Ks sin(γ ψ)]dr + [ cos(γ ψ)]ds + [ s sin(γ ψ)]dψ dy = [sin(γ) Ks cos(γ ψ)]dr + [sin(γ ψ)]ds + [s cos(γ ψ)]dψ dω = Kdr dψ 10

The dierential equation that follows is thus dx cos(γ) Ks sin(γ ψ) cos(γ ψ) s sin(γ ψ) dy = sin(γ) Ks cos(γ ψ) sin(γ ψ) s cos(γ ψ) dω K 0 1 dr ds dψ And the determinant dx dy dω = det(j )dr ds dψ = sin(ψ)dr ds dψ Forwards Map Φ + t : (r, s +, ψ) (x +, y +, ω + ) x + = x + s + cos(ω + ) Since s + cos(ω + ) is the projection of the precollisional velocity on the x axis scaled to equal the distance traveled postcollision y + = ȳ + s + sin(ω + )Since s + sin(ω + ) is the projection of the precollisional velocity on the y axis scaled to equal the distance traveled postcollision ω + = γ + ψ Dierentiating reveals dx + = cos(γ)dr + cos(ω + )ds + s + sin(ω + )dω + dy + = sin(γ)dr + sin(ω + )ds + + s + sin(ω + )dω + dω + = Kdr + dψ dx = cos(γ)dr dy = sin(γ)dr dγ = Kdr Substituting the denition of dω + in resolves the system to dx + = [cos(γ) + Ks + sin(γ + ψ)]dr + [cos(γ + ψ)]ds + + [ s + sin(γ + ψ)]dψ dy + = [sin(γ) Ks + cos(γ + ψ)]dr + [sin(γ + ψ)]ds + + [s + cos(γ + ψ)]dψ dω + = Kdr + dψ The dierential equation that follows is thus dx + cos(γ) + Ks + sin(γ + ψ) cos(γ + ψ) s + sin(γ + ψ) dy + = sin(γ) Ks + cos(γ + ψ) sin(γ + ψ) s + cos(γ + ψ) dω + K 0 1 dr ds + dψ And the determinant dx + dy + dω + = det(j + )dr ds + dψ = sin(ψ)dr ds + dψ 11

Now we can relate the volumes by the original denition s +s + = t, leading to ds + = ds. Thus dx dy dω = sin(ψ)dr ds dψ = sin(ψ)dr ds + dψ = dx + dy + dω + So for any billiard that satises our assumptions, we can take a volume element in the phase space and trace its ow as a series of time-t maps each of which preserves the volume. Following from this, we see that the volume is preserved over the general ow. Part IV The Eect of Changing a Billiard on the Flow In this investigation, we limitted analysis to only considering the possible eects of changing size of the circular billiard durring a ow to change the behavior. First we consider whether it is possible to change the size of the a billiard to change a non-periodic ow to a periodic ow. Consider a quasi-periodic orbit of a particle in the circular billiard. Where ψ represents the angle between the trajectory and the tangent to the circle and θ represents the angle of a collision point measured from the positive x axis, we nd the relations θ n = θ n 1 + 2ψ (mod 2π) ψ n = ψ n 1 So we can draw a triangle with vertices at θ n, θ n 1, and the origin and use the law of sines, to nd sin π 2 = sin(π/2 ψ) x where x is length of the perpendicular to the chord connecting θ n and θ n 1. Thus, we nd x = sin( π 2 ψ) = cos(ψ). So the inner caustic circle that each chord lies tangent to is distance cos(ψ) from the origin. To create a familiar periodic orbit, we extend the circle to have some radius r at which ψ, the angle between the tragectory and the tangent to the new circle will be π/3 to create an equilateral triangle. Thus, using the same relations in the new circle, we construct a similar right triangle with hypotenuse of length r and vertices at the origin, a point on the caustic circle, and a point on the circumference of the new circle. Thus, using the law of sines, we have sin( π 2 ψ ) cos(ψ) = 1/r which shows r = cos(ψ)/cos(ψ ) = 2cos(ψ). We are allowed to change the size of the billiard since for at least one point in its trajectory, the particle is cos(ψ) distance from the origin (and cos(ψ) 0 since the only quasi-periodic orbits have ψirrational with respect to π). 12

To convert from a periodic orbit to a quasi-periodic orbit, we need only preform these operations in reverse, allowing a period-3 orbit in a billiard of radius 2cos(ψ) to devolve into a non-periodic orbit as we change the radius back to 1 while it passes through a point distance cos(ψ) from the origin. References [1] Chernov, Nikolai, Markarian Roberto. Chaotic Billiards. Providence: American Mathematical Society, 2006. Print, 19-25 [2] Tabachnikov, Serge. Geometry and Billiards. Providence: American Mathematical Society, 2005. Print. [3] Alligood, Kathleen, Sauer, Tim, Yorke, James. Chaos: An Introduction to Dynamical Systems. New York: Textbooks in Mathematical Sciences, 1996. Print [4] "Billiards and Chaos Theory." Serendip's Exchange. Web. 27 Nov. 2009. <http://serendip.brynmawr.edu/chaos/home.html>. 13