DYNAMICS OF BOUNCING RIGID BODIES AND BILLIARDS IN THE SPACES OF CONSTANT CURVATURE KI YEUN KIM DISSERTATION

Similar documents
BILLIARD DYNAMICS OF BOUNCING DUMBBELL

A FRANKS LEMMA FOR CONVEX PLANAR BILLIARDS

Periodic Sinks and Observable Chaos

HAMILTON S PRINCIPLE

1 Lyapunov theory of stability

Analysis II: The Implicit and Inverse Function Theorems

Hamiltonian Dynamics

April 13, We now extend the structure of the horseshoe to more general kinds of invariant. x (v) λ n v.

Lecture 1: Oscillatory motions in the restricted three body problem

Chapter 23. Predicting Chaos The Shift Map and Symbolic Dynamics

The Homoclinic Tangle and Smale s Horseshoe Map

One Dimensional Dynamical Systems

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

LECTURE Itineraries We start with a simple example of a dynamical system obtained by iterating the quadratic polynomial

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

Physics 5153 Classical Mechanics. Canonical Transformations-1

notes 5d. Details of the Construction of the Smale horseshoe map.

No-Slip Billiards in Dimension Two

11 Chaos in Continuous Dynamical Systems.

arxiv: v1 [math.ds] 6 Apr 2011

Physics 106b: Lecture 7 25 January, 2018

1 Differentiable manifolds and smooth maps

B5.6 Nonlinear Systems

HYPERBOLIC SETS WITH NONEMPTY INTERIOR

Metric Spaces and Topology

M3-4-5 A16 Notes for Geometric Mechanics: Oct Nov 2011

Noncollision Singularities in the n-body Problem

LECTURE 15: COMPLETENESS AND CONVEXITY

The Structure of Hyperbolic Sets

The Liapunov Method for Determining Stability (DRAFT)

Physics 106a, Caltech 4 December, Lecture 18: Examples on Rigid Body Dynamics. Rotating rectangle. Heavy symmetric top

PHY411 Lecture notes Part 5

arxiv:math/ v1 [math.ds] 8 Oct 2006

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

THE LAGRANGIAN AND HAMILTONIAN MECHANICAL SYSTEMS

CALCULUS ON MANIFOLDS

8.1 Bifurcations of Equilibria

Lebesgue Measure on R n

7 Pendulum. Part II: More complicated situations

Sketchy Notes on Lagrangian and Hamiltonian Mechanics

Lecture 5: Oscillatory motions for the RPE3BP

APPPHYS217 Tuesday 25 May 2010

S. Mrówka introduced a topological space ψ whose underlying set is the. natural numbers together with an infinite maximal almost disjoint family(madf)

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

Persistent Chaos in High-Dimensional Neural Networks

Stable periodic billiard paths in obtuse isosceles triangles

MATH 415, WEEK 11: Bifurcations in Multiple Dimensions, Hopf Bifurcation

INVERSE FUNCTION THEOREM and SURFACES IN R n

Smooth Structure. lies on the boundary, then it is determined up to the identifications it 1 2

CHAOTIC UNIMODAL AND BIMODAL MAPS

UNIFORM SUBHARMONIC ORBITS FOR SITNIKOV PROBLEM

Def. A topological space X is disconnected if it admits a non-trivial splitting: (We ll abbreviate disjoint union of two subsets A and B meaning A B =

COMPLETELY INVARIANT JULIA SETS OF POLYNOMIAL SEMIGROUPS

ẋ = f(x, y), ẏ = g(x, y), (x, y) D, can only have periodic solutions if (f,g) changes sign in D or if (f,g)=0in D.

2000 Mathematics Subject Classification. Primary: 37D25, 37C40. Abstract. This book provides a systematic introduction to smooth ergodic theory, inclu

On Periodic points of area preserving torus homeomorphisms

1 Hamiltonian formalism

CHAPTER 7. Connectedness

Luminy Lecture 2: Spectral rigidity of the ellipse joint work with Hamid Hezari. Luminy Lecture April 12, 2015

1 The Local-to-Global Lemma

On dynamical properties of multidimensional diffeomorphisms from Newhouse regions: I

Chapter 2 Metric Spaces

Curves in the configuration space Q or in the velocity phase space Ω satisfying the Euler-Lagrange (EL) equations,

An introduction to Birkhoff normal form

The Existence of Chaos in the Lorenz System

An Undergraduate s Guide to the Hartman-Grobman and Poincaré-Bendixon Theorems

MATH Final Project Mean Curvature Flows

Course Summary Math 211

Topological properties

arxiv: v3 [math.dg] 19 Jun 2017

Finding numerically Newhouse sinks near a homoclinic tangency and investigation of their chaotic transients. Takayuki Yamaguchi

1 The Derivative and Differrentiability

Singular Perturbations in the McMullen Domain

CP1 REVISION LECTURE 3 INTRODUCTION TO CLASSICAL MECHANICS. Prof. N. Harnew University of Oxford TT 2017

Rose-Hulman Undergraduate Mathematics Journal

for changing independent variables. Most simply for a function f(x) the Legendre transformation f(x) B(s) takes the form B(s) = xs f(x) with s = df

Introduction. Chapter 1. Contents. EECS 600 Function Space Methods in System Theory Lecture Notes J. Fessler 1.1

Stability of Feedback Solutions for Infinite Horizon Noncooperative Differential Games

Lozi-like maps. M. Misiurewicz and S. Štimac. May 13, 2017

Problem List MATH 5173 Spring, 2014

Aubry Mather Theory from a Topological Viewpoint

Hamiltonian systems with linear potential and elastic constraints

Hyperbolic Dynamics. Todd Fisher. Department of Mathematics University of Maryland, College Park. Hyperbolic Dynamics p.

Diffraction by Edges. András Vasy (with Richard Melrose and Jared Wunsch)

Introduction to Applied Nonlinear Dynamical Systems and Chaos

THE INVERSE FUNCTION THEOREM

TOPOLOGICAL EQUIVALENCE OF LINEAR ORDINARY DIFFERENTIAL EQUATIONS

Homoclinic Points and the Horseshoe Map - A Reprise

= w. These evolve with time yielding the

Symplectic maps. James D. Meiss. March 4, 2008

Good Problems. Math 641

MORSE HOMOLOGY. Contents. Manifolds are closed. Fields are Z/2.

(x k ) sequence in F, lim x k = x x F. If F : R n R is a function, level sets and sublevel sets of F are any sets of the form (respectively);

Stabilization of a 3D Rigid Pendulum

Legendre Transforms, Calculus of Varations, and Mechanics Principles

1.1 Single Variable Calculus versus Multivariable Calculus Rectangular Coordinate Systems... 4

CALCULUS ON MANIFOLDS. 1. Riemannian manifolds Recall that for any smooth manifold M, dim M = n, the union T M =

Quick Tour of the Topology of R. Steven Hurder, Dave Marker, & John Wood 1

ASYMPTOTIC STRUCTURE FOR SOLUTIONS OF THE NAVIER STOKES EQUATIONS. Tian Ma. Shouhong Wang

The Symmetric Space for SL n (R)

Transcription:

016 Ki Yeun Kim

DYNAMICS OF BOUNCING RIGID BODIS AND BILLIARDS IN TH SPACS OF CONSTANT CURVATUR BY KI YUN KIM DISSRTATION Submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Mathematics in the Graduate College of the University of Illinois at Urbana-Champaign, 016 Urbana, Illinois Doctoral Committee: Professor Yuliy Baryshnikov, Chair Professor Vadim Zharnitsky, Director of Research Associate Professor Robert DeVille Associate Professor Zoi Rapti

Abstract Mathematical billiard is a dynamical system studying the motion of a mass point inside a domain. The point moves along a straight line in the domain and makes specular reflections at the boundary. The theory of billiards has developed extensively for itself and for further applications. For example, billiards serve as natural models to many systems involving elastic collisions. One notable example is the system of spherical gas particles, which can be described as a billiard on a higher dimensional space with a semi-dispersing boundary. In the first part of this dissertation, we study the collisions of a two-dimensional rigid body using billiard dynamics. We first define a dumbbell system, which consists of two point masses connected by a weightless rod. We assume the dumbbell moves freely in the air and makes elastic collisions at a flat boundary. For arbitrary mass choices, we use billiard techniques to find the sharp bound on the number of collisions of the dumbbell system in terms of the mass ratio. In the limiting case where the mass ratio is large and the dumbbell rotates fast, we prove that the system has an adiabatic invariant. In case the two masses of the dumbbell are equal, we assume gravity in the system and study its infinitely many collisions. In particular, we analytically verify that a Smale horseshoe structure is embedded in the billiard map arising from the equal-mass dumbbell system. The second part of this dissertation concerns the billiards in the spaces of constant curvature. We provide a unified proof to classify the sets of three-period orbits in billiards on the uclidean plane, the hyperbolic plane and on the two-dimensional sphere. We find that the set of three-period orbits in billiards on the hyperbolic plane has zero measure. For the sphere, three-period orbits can form a set of positive measure if and only if a certain natural condition on the orbit length is satisfied. ii

To my parents who believed in me and supported me all along. iii

Acknowledgments I would like to express my gratitude to my advisor Vadim Zharnitsky for his consistent guidance and encouragement. I would also like to thank the rest of my committee members, Yuliy Baryshnikov, Lee DeVille and Zoi Rapti, for their helpful comments and insights. This work was financially supported by Dr. Lois M. Lackner mathematics fellowship, the Hone-Nash award, and a gift to the Mathematics department at the University of Illinois from Gene H. Golub. I deeply appreciate the generosity of the donors. I am grateful to my family and friends who have supported me through my ups and downs during the past years. A special thanks goes to Mike for proof-reading my work as a fellow mathematician and as a native speaker, and for being a great part of my life. iv

Table of Contents CHAPTR 1 Introduction................................ 1 CHAPTR Preliminaries................................ 6.1 Lagrangian and Hamiltonian mechanics.................... 6.1.1 Lagrange s equations.......................... 6.1. Hamilton s equations.......................... 7.1.3 Adiabatic invariance.......................... 8. The Smale horseshoe map............................ 10..1 Deterministic chaos.......................... 10.. The Smale horseshoe map....................... 11..3 The Conley-Moser conditions..................... 13.3 Mathematical billiards.............................. 15 CHAPTR 3 The dumbbell system........................... 17 CHAPTR 4 stimate on the number of collisions of the dumbbell......... 0 CHAPTR 5 Adiabatic invariance in the dumbbell.................. 6 CHAPTR 6 Chaotic dynamics of the two-dimensional coin............. 34 6.1 The coin system as a gravitational billiard................... 34 6. The topological picture of the billiard map.................. 37 6..1 Construction of the domain D..................... 37 6.. The image of D under the billiard map................ 40 6.3 Construction of the strips satisfying the Conley-Moser conditions..... 45 CHAPTR 7 Three-period orbits in billiards on the surfaces of constant curvature 50 7.1 Billiard system on the surface of constant curvature............. 51 7.1.1 Jacobi fields............................... 51 7.1. volution and reflection matrices................... 5 7. Billiard on the hyperbolic plane........................ 54 7.3 Billiard on the -sphere............................. 56 References......................................... 60 v

CHAPTR 1 Introduction Mathematical billiards is the study of the free motion of a mass point in a domain: the point moves along a straight line inside the domain, and makes mirror-like reflections at the boundary. These billiard models are simple yet they illustrate various fundamental concepts in dynamical systems. Moreover, billiards have close connections to other fields of mathematics as well as to other sciences. In this dissertation, we focus on mathematical billiards and its applications to the motion of a simple rigid body. Along the way, we explore topics from classical mechanics and chaos theory. The plan of the paper is the following. In the remainder of this introduction, we briefly explain our main results and provide some motivation to the study. In Chapter, we discuss essential terminologies and other relevant background materials in detail. In Chapter 3, we introduce a dumbbell system consisting of two point masses. We present the main results regarding the dumbbell in Chapter 4 through Chapter 6. Chapter 4 concerns the dumbbell with two arbitrary masses, whereas Chapter 5 considers the case of a limiting mass ratio and Chapter 6 considers the case of equal masses with an extra assumption of the gravitational force. Chapter 7 contains independent topics and results about periodic orbits in billiards on the non-uclidean spaces. Let us begin with a motivation why we study the dumbbell. Many mechanical systems which involve elastic collisions are often easier to analyze when they are considered as billiard systems. In the 1960s, the billiard problem appeared in the context of statistical mechanics to study a system of spherical gas particles to verify the Boltzmann ergodic hypothesis [Si]. More recently, non-spherical particles, which have rotational in addition to translational velocities, were studied to build a more realistic gas model [Co]. In this direc- 1

tion, it is natural to consider the behavior of a billiard ball with two degrees of freedom moving inside a domain. ven in the case where the boundary is flat, the extra degree of freedom of a moving object can generate many interesting phenomena including deterministic chaos and adiabatic invariance. In Chapter 3, we introduce a dumbbell system, which consists of two point masses m L and m R connected by a weightless rigid rod. We study the elastic collisions of the dumbbell with a flat boundary. This is an important first step before understanding the full dynamics of a two-dimensional object in some general planar domains. If we assume no gravity, the dumbbell will leave the flat boundary after finitely many collisions. In Chapter 4, by first considering the dumbbell system as a single point billiard then applying the unfolding technique [Ta], we estimate the maximal number of collisions of the dumbbell in terms of the mass ratio m L m R [Ba]. One of the basic questions about an ideal gas model is to estimate the maximum number of collisions for a system of spherical gas particles moving freely in an open space [Bu]. Our collision estimate of the dumbbell can be useful to extend such study to a system of non-spherical gas particles. For the previous result, we did not impose any assumptions on the two masses m L and m R. We now consider the two natural special cases when the mass ratio m L m R m L m R 1 or when = 1. In the former case, we prove that there exists a function, which depends on the phase variables of the dumbbell system and a slowly varying parameter, such that the function value remains nearly constant along an orbit. We call this function an adiabatic invariant. One classical example of adiabatic invariance is in the Fermi-Ulam model. It is a onedimensional system where an elastic particle travels between two slowly moving walls. One can use the laws of physics to show that the product of the velocity of the particle and the distance between the walls is an adiabatic invariant. In Chapter 5, we show that the dumbbell system, which is a two-dimensional system, has a comparable adiabatic invariant. If we assume the dumbbell has a small mass ratio m L m R 1 and approaches the flat boundary very slowly, then the adiabatic invariant of the system is given by the product of the angular velocity of the dumbbell and the angular distance traveled by the lighter mass m L between

two consecutive collisions. For the second case when the two masses are equal, we add gravity to the system and interpret this scenario of the equal-mass dumbbell with gravity as a special case of a coin toss with collisions. Previous work on the dynamics of a coin either assumes a tossed coin is caught in hand before a collision [Di, Ke] or investigates the collision effects of a coin using numerical simulations [Str, Vu]. Here, we present our analytic result on the collision dynamics of a two-dimensional coin: a thin disk whose motion is restricted in the twodimensional space. Note that our equal-mass dumbbell can be viewed as an equivalent model to the twodimensional coin. That is, despite having the different moments of inertia, the basic motions of the equal-mass dumbbell and the two-dimensional coin are the same. Since we already have established the billiard dynamics for the dumbbell, we continue to use the dumbbell to study the two-dimensional coin. Our results below, which are proven using the equal-mass dumbbell, remain true if we use the coin instead. The main results in Chapter 6 are on the chaotic behavior of the two-dimensional coin equal-mass dumbbell. Since the gravity pulls the coin back to the boundary and there is no energy loss, the coin will bounce infinitely many times. We prove that the two-dimensional bouncing coin can produce any infinite sequence of collisions: if a collision of m L or m R is labeled by L or R respectively, then any infinite collision sequence of L s and R s can be realized by choosing an appropriate initial condition. The proof begins by viewing the coin system as a single point gravitational billiard in a transformed domain. Due to gravity, the mass point moves along a parabola inside the domain. Then we show that the Smale horseshoe [Sm] is embedded in the billiard map. Motivated from the study of a nonlinear forced oscillator, the Smale horseshoe map is one of the famous examples of a chaotic map. The horseshoe map is defined topologically: in the standard setting, it transforms a square into a long and thin strip and places the strip over the square in the shape of a horseshoe. When iterated forward and backward infinitely many times, the horseshoe map generates an invariant Cantor set in the square. The horseshoe map acting on the invariant set is topologically conjugate to a shift map on 3

the set of infinite sequences of two symbols. Therefore, once we have a horseshoe, we can use the fact that there exists an one-to-one correspondence between an infinite symbolic sequence and a phase point to complete the proof. One way to verify that a certain map contains a horseshoe is by showing that the map satisfies the Conley-Moser conditions [Mo, Wi], which are a combination of geometric and analytic criteria. If met, these conditions guarantee the existence of a horseshoe. Although there are many systems in science and engineering that are proven to contain a horseshoe [Lv, Ho], to the best of our knowledge, all the already-known horseshoe maps are defined on one connected domain. For our coin system, we iterate the billiard map on the union of two disjoint rectangles. We prove that the billiard map on the two rectangles produces the horizontal and vertical strips satisfying the Conley-Moser conditions. The associated set of sequences contains combinatorially more complex sequences than the sequences for the standard horseshoe map. These results on the bouncing coin assert the existence of chaotic behavior in the system, which leaves the question of how the entire phase space of the coin system will evolve over time. Since the system is highly nonlinear, only partial numerical results are known [Mi, Stz]. The billiard boundary for the coin consists of piecewise-smooth dispersing curves forming convex wedges of angle π at the discontinuities. It is known that a gravitational billiard in a straight wedge of angle bigger than π is ergodic [Le, Wo]. Also, when a gravitational force does not exist, all dispersing billiards are ergodic [Si]. Since the billiard system for the coin has a boundary which is some sort of a combination of these, we expect that the coin system is ergodic. We plan to pursue this question as future work. In Chapter 7, we turn our attention to billiards on the hyperbolic plane and on the - sphere. One important question in billiards concerns the presence of large sets of periodic orbits. Ivrii conjectured [Iv] that, in every bounded planar billiard domain with a smooth boundary, the set of n-period orbits has measure zero for all n. The Weyl estimate formula, which describes eigenvalue asymptotics for the Dirichlet problem, has a certain special form if the set of periodic orbits in the associated billiard has measure zero. The conjecture has been proven for, 3 and 4-period orbits only. Since -period orbits 4

must be normal to the boundary at both ends, we can easily see that in any billiard the set of -period orbits has measure zero. The conjecture for the set of 3, 4-period orbits was proved in various literature [Gl, Gl, Ry, Sto, Wo]. There have been extensions to other types of billiard systems such as higher dimensional [Vo], outer [Ge, Tu], and spherical billiards [Ba3]. We present a unified approach to study the set of 3-period orbits for the spherical and the hyperbolic cases [Bl]. We extend the Jacobi fields approach from Wojtkowski s paper [Wo] to write a underlying geometry based relation which must be satisfied on a neighborhood filled with 3-period orbits. On the hyperbolic plane, we obtain a contradiction in the relation and conclude that there cannot be an open set of 3-period orbits. For the sphere, the relation remains true if and only if a certain natural condition on the orbit length is satisfied. Thus, under the condition, the set of 3-period orbit can have a positive measure. Otherwise, it has an empty interior and has measure zero. It is expected that understanding the structure of the sets of periodic orbits in non- uclidean billiards would help studying eigenvalue asymptotics in non-uclidean geometries. 5

CHAPTR Preliminaries Before we move on to the main results, we provide some necessary background materials. In Chapter.1, we exposit the basics of the Lagrangian and the Hamiltonian mechanics and explain the notion of adiabatic invariance. Many definitions and results in this chapter come from the book by Arnold [Ar]. In Chapter., we first discuss deterministic chaos, then give the definition of the Smale horseshoe map and the sufficient criteria for the existence of the map. The book by Wiggins is an excellent reference for this topic [Wi]. Finally in Chapter.3, we talk about mathematical billiards. For more details about billiards, we recommend the books by Tabachnikov [Ta, Ta] and the book by Cvitanovic [Cv]..1 Lagrangian and Hamiltonian mechanics.1.1 Lagrange s equations Newtonian mechanics is one branch of classical mechanics, and is indisputably a breakthrough development to study the motion of bodies. However, it mainly uses Cartesian coordinates which may lead to extremely complex calculations depending on the systems. Lagrangian mechanics is a reformulation of Newtonian mechanics, which is better adapted to the change of coordinates. The configuration space is the the collection of the parameters that defines the position of a system. These parameters which we denote by q = q 1,...q n are called the generalized coordinates. The time derivatives of the generalized coordinates, q, are called the generalized velocities. In general, a configuration space is a differential manifold. 6

For a conservative mechanical system on its configuration space, the Lagrangian is defined by Lq, q, t = K U where K is the kinetic energy and U is the potential energy of the system. Lagrangian mechanics is based on the calculus of variation which concerns the extremals of functionals: functions from the space of curves to the reals. Given a Lagrangian Lq, q, t, we call the functional Φγ = t 1 t 0 Lq, q, tdt the action of the system. The evolution of q is governed by the following variational condition on the action functional. Theorem.1. Hamilton s principle of least action states that between two points qt 0 and qt 1, the motion of the system follows the curve γ such that the action Φγ = t 1 t 0 is an extremum. Theorem.. The derivative of the functional Φγ = t 1 t 0 Lq, q, tdt is given by Lq, q, tdt F h = t1 where h is the variation of the curve γ. t 0 [ L q d dt L ] L t 1 hdt + q q h t 0 An extremal of the action Φγ is a curve such that F h is equal to zero. From there, we get Lagrange s equations, which determine the trajectory of q on the configuration space. Definition. Lagrange s equations for the functional Φ = t 1 t 0 d dt L L q q = 0. Lq, q, tdt are given by.1. Hamilton s equations Hamiltonian mechanics is another formulation of classical mechanics, where a system is expressed in terms of a set of canonical coordinates q, p. In a mechanical system, q are the generalized coordinates and p = L are the generalized momenta. q We can convert a Lagrangian system to a Hamiltonian system via the Legendre transformation. 7

Definition. Let fx be a convex real-valued function. Given a number p, let x p = max{px fx}. x Such x p can be found by solving f x = p. The Legendre transform of fx is a new function gp = px p fx p. Theorem.3. Suppose Lagrange s equations describe the evolution of q on an n-dimensional configuration space. The system of Lagrange s equations is equivalent to the system of n first order differential equations, called Hamilton s equations ṗ = H q q = H p, where Hp, q, t = p q Lq, q, t is the Legendre transform of the Lagrangian function viewed as a function of q. We call H the Hamiltonian. In a mechanical system, the Hamiltonian coincides with the total energy of the system. The evolution of the system is given by the solution to Hamilton s equations. When interpreted geometrically, it means that the evolution is described as the flow on a vector field on the phase space, a n-dimensional symplectic manifold. The Hamiltonian formalism allows us to canonically transform one set of canonical coordinates p, q to another set P, Q. That is, the transformation conserves the -form, Σ dp i dq i = Σ dp i dq i. In the next section, we give an example of such transformation..1.3 Adiabatic invariance Consider a Hamiltonian system with one degree of freedom, which exhibits periodic motion. Suppose the system depends on a slowly varying external parameter λ and Hp, q; λ 8

is a twice differentiable in λ. H = Hp, q; λ, ṗ = H q q = H p..1 Then for the system.1, there exists a function of the phase variables and the parameter λ, called an adiabatic invariant Ip, q, λ, such that the value of I remains nearly constant along an orbit. An adiabatic invariant can be intuitively found as follows. For a fixed λ, the periodicity in motion implies that an orbit on the phase space must be a closed curve. We denote the area bounded by such a curve the action. A uniformly varying angular variable on this phase curve is called the angle and is usually denoted by ϕ. As λ slowly changes, the closed phase curve evolves slowly. However, the enclosed area stays nearly constant and it is the adiabatic invariant. A formal definition of an adiabatic invariant is given below. Definition. The quantity Ip, q, λ is called an adiabatic invariant if for every κ > 0 there exists an ɛ 0 > 0 such that if 0 < ɛ < ɛ 0 and 0 < t < 1/ɛ, then Ipt, qt; ɛt Ip0, q0; 0 < κ. Although we did not formally prove here, the transformation of the phase variables p, q to the action-angle variables I, ϕ is canonical. It can be done using an indirect method involving a generating function. If we drop the smoothness condition from the system.1, then it does not follow immediately, even in the case of one-dimensional systems, that the action is an adiabatic invariant. In Chapter 5, we consider a non-smooth system with two degrees of freedom and prove that the system has an adiabatic invariant. 9

. The Smale horseshoe map..1 Deterministic chaos In dynamical systems, chaos refers to high sensitivity to initial states in a deterministic system. Although no universal definition of chaos exists, we often determine the chaoticity of a system using Lyapunov exponents. Below is a definition of the Lyapunov exponents in a two-dimensional discrete map. Definition. Let f be a smooth map on R. Consider Bx 0, δ, a disk of small radius δ centered at x 0. Then Df n x 0 Bx 0, δ is an ellipse where the two axes have the lengths r1 n r n. These two numbers measure the separation rates of the points near x 0 after n- iterations of f. The Lyapunov numbers of x 0 are defined by r n 1/n L 1 = lim 1 r n 1/n L = lim. n δ n δ The Lyapunov exponents of x 0 are given by ln L 1 and ln L. When the maximal Lyapunov exponent is greater than 0, we generally say the system is chaotic. There are some counter-examples such as the Perron effects where the positive maximal Lyapunov exponent does not indicate chaos. The history of chaos theory goes back to the time of Poincaré, when he studied the stability of the three-body problem in celestial mechanics. During his study, he introduced a Poincaré section P, a transversal hyperplane to the flow of the system. Then he considered the Poincaré map f defined on P, which takes the intersection of the flow with the section P to the next intersection. If the flow is periodic, then there exists a fixed point p P such that fp = p. When Dfp has two eigenvalues where one is smaller than 1 and one is bigger than 1, we say p is hyperbolic and of a saddle type. Then by the stable manifold theorem, we can find the stable manifold W S and unstable manifold W U on which the hyperbolic fixed point p lies. Since by definition p W S W U implies f n p W S W U, one intersection of the two manifolds imply infinitely many intersections of the two. This 10

creates a complicated structure which we call the homoclinic tangle. The discovery of a homoclinic tangle is considered as the first observation of chaos. Due to its structure, the iteration of the map separates nearby trajectories arbitrarily far. In the next section, we introduce the Smale horseshoe map, a prototypical map possessing homoclinic invariant points... The Smale horseshoe map H T f H BT H TT H B f f H TB H BB f 1 V B V T Figure.1: A topological picture of the Smale horseshoe map A horseshoe map f is a diffeomorphism, defined topologically: the map take a square D and stretch it in one direction and squeeze it in another direction to make a long thin strip. Then the strip is placed over the square in the shape of a horseshoe Figure.1. We study set of the points which remain in the square after some forward and backward 11

iterations of f. We will call the invariant set N Λ N = {p p f n D}. n= N When n is sent to, we define Λ = lim n Λ n. From Figure.1, we see that after one forward iteration of the map, fd D consists of two horizontal strips H T and H B. If we iterate f once more, all but the four thinner horizontal strips stay inside the square. Observe that two thinner strips are nested in H T and H B respectively. We define each thinner nested strip fh s H s 1 = H s s 1 where s i {T, B}. Following this way, we label the four horizontal strips H BT, H T T, H T B, H BB. If we iterate f forward for n times, then only n horizontal strips remain in D. We inductively define the n horizontal strips, fh s n...s H s 1 = H s n...s 1 where s i {T, B}. Now we look at the backward iteration of f. We denote the preimages of the horizontal strips H T and H B as V T = f 1 H T and V B = f 1 H B. When we iterate f backward for n times, then f n D D is the union of n vertical strips. The nested vertical strips are defined f 1 V s1...s n V s0 = V s0...s n where s i {T, B}. We see that Λ n is the intersection of n horizontal strips and n vertical strips. When n, Λ consists of the intersection points of infinite number of horizontal curves and vertical curves, resulting a Cantor set structure. We assign a sequence s =...s s 1 s 0 s 1... to each point p Λ based on to which horizontal curve H...s s 1 and vertical curve V s0 s 1... the point p belongs. From the construction, this assignment is clearly one-to-one. 1

Moreover, there is a relation between the dynamics of the points in the invariant Cantor set and a shift map of the sequences. The horseshoe map acting on the invariant set is topologically conjugate to a shift map on the set of infinite sequences of two symbols. The sequence space provides a convenient setting to examine the periodicity of the orbits. It follows that the horseshoe map contains: 1. Countably many periodic orbits of arbitrarily long period,. Uncountably many non-periodic orbits, and 3. A dense orbit...3 The Conley-Moser conditions In general, the image of the horseshoe map needs not be in the shape of a horseshoe. The key features of the horseshoe map are the following. 1. The map has stretching and contracting aspects in two different directions.. Disjoint regions get mapped to disjoint regions in the domain. There is a formal criteria called the Conley-Moser conditions, which guarantee the existence of a horseshoe if the conditions are satisfied. To state the conditions, we first need some definitions. Consider a rectangular domain D = [X 0, X 1 ] [Y 0, Y 1 ] R. Definition. In D, a µ h -horizontal curve is the graph of a function Y = hx such that 1 Y 0 hx Y 1, and for every pair X i, X j [X 0, X 1 ], we have hx i hx j µ h X i X j for some 0 µ h. Definition. In D, a µ v -vertical curve is the graph of a function X = vy such that 1 X 0 vy X 1, and 13

for every pair Y i, Y j [Y 0, Y 1 ], we have vy i vy j µ v Y i Y j for some 0 µ v. Definition. Given two non-intersecting horizontal curves h 1 X < h X in D, we define a µ h -horizontal strip as H = {X, Y : Y [h 1 X, h X] for X [X 0, X 1 ]}. Given two non-intersecting vertical curves v 1 Y < v Y in S, we define a µ v -vertical strip as V = {X, Y : Y [v 1 Y, v Y ] for Y [Y 0, Y 1 ]}. Definition. The width of horizontal strips is defined as dh = max h X h 1 X, and the width of vertical strips is defined as dv = max v Y v 1 Y. Consider a diffeomorphism f : D R and let S = {1,,... N} be an index set. Let s S H s be a set of disjoint µ h -horizontal strips, and let s S V s be a set of disjoint µ v -vertical strips. The Conley-Moser conditions on f are: CM1 0 µ h µ v < 1 CM f maps V s homeomorphically onto H s. Also, the horizontal vertical boundaries of V s get mapped to the horizontal vertical boundaries of H s. CM3 Suppose H is a µ h -horizontal strip contained in s S H s, then fh H s is a µ h - horizontal strip and dfh H s < dh. Similarly, if V is a µ v -vertical strip contained in s S V i, then f 1 V V s is a µ v -vertical strip and df 1 V V s < dv. Theorem.4 Moser. Suppose f satisfies the Conley-Moser conditions. Then f has an invariant Cantor set Λ on which it is topologically conjugate to a full shift on N symbols, i.e. the 14

diagram below commutes, where φ : Λ Σ N is a homeomorphism and σ : Σ N Σ N is the shift map on the space of sequences of N symbols. φ Λ f Λ φ Σ N σ Σ N.3 Mathematical billiards A billiard domain Q is a Riemannian manifold with a piecewise smooth boundary Q. A billiard is the geodesic flow of a mass point in Q. From the view of classical mechanics, geodesic is a curve satisfying the Hamilton s principle of least action. At the boundary Q, the mass point reflects such that the normal component of the velocity to the boundary reverses its sign and the tangential component of the velocity stays the same. π Q α 1 q 1 f q 0 v + α 0 q 0,α 0 l Q q 1,α 1 0 a b Figure.: a A planar billiard and the billiard map. b The billiard map on its phase space. Commonly considered billiard domains are in the plane, since we can explore various mathematical questions even in this simple setting. For planar billiards, a mass point moves in a straight line inside a domain Q and makes mirror-like reflections at the boundary Q. It is sufficient to study a billiard flow by observing it at the moments of impact. A billiard flow has a natural Poincaré section, which has coordinates q, α, where q is the arc 15

length parameter along Q and α is the angle between the outgoing velocity vector v + at the footpoint q and the boundary Q Figure.a. The Poincaré section is topologically equivalent to a cylinder S 1 [0, π]. The Poincaré return map, which we call the billiard map f, sends a collision q 0, α 0 to the next collision q 1, α 1 Figure.b. If q n, α n := f n q 0, α 0 = q 0, α 0, we call the set {q 0, α 0, q 1, α 1,, q n, α n } an n-period orbit. The shape of a billiard domain influences the qualitative behavior of the billiard trajectories in it. We can divide the domains into roughly three different categories. The first type is of smooth convex billiards, called Birkhoff billiards, which are associated to stable dynamics. The common topics in Birkhoff billiards are periodic orbits and the integrability. The second type of billiards are hyperbolic billiards, which often consist of piecewise dispersing boundaries. The most famous hyperbolic billiard is probably the Sinai billiard whose domain is the complement of an array of circles in the plane. As we have seen in the horseshoe map, the hyperbolicity produces chaotic dynamics. The last type of billiards are the ones with polygonal domains. We will not say much about these, except we mention the unfolding technique which is at times used to study the billiard trajectories in polygons. Here, we explain the technique using a straight wedge instead. γ Figure.3: An illustration of the unfolding technique in a straight wedge of angle γ. Suppose a billiard trajectory enters a straight wedge of angle γ and hits one side of the wedge. Instead of reflecting the trajectory at the collision, we reflect the wedge across the side and unfold the trajectory into a straight line Figure.3. If we perform this procedure at every collision, then we will get a straight trajectory passing through π γ wedges. Therefore, the maximal number of collisions inside the wedge of angle γ is π. γ 16

CHAPTR 3 The dumbbell system In the two-dimensional space with coordinates X, Y, let us consider a dumbbell system, which consists of two point masses m L, m R, connected by weightless rigid rod of length l. The coordinates of m L, m R, and the center of mass of the dumbbell are denoted by X L, Y L, X R, Y R, and X, Y, respectively. Let θ be the angular position of the vector X L X R measured counterclockwise from the positive X-direction Figure 3.1a. With the total mass of the dumbbell m = m L + m R, we define β L = m L m distance from the center of mass to m L and to m R are given by lβ R and lβ L. and β R = m R m. Then the The dumbbell moves freely in the half space Y > 0 until it reflects elastically on the floor Y = 0. The velocity of the center of mass in the X-direction is constant since there is no force acting on the system in that direction. Thus, we may assume that the center of mass does not move in the X-direction. With this reduction, the dumbbell configuration space is a two-dimensional cylinder with the natural choice of coordinates θ, Y. On the θy plane, the dumbbell is a mass point in a transformed domain. Note that on the XY plane, the dumbbell hits the floor when Y L or Y R becomes zero. Therefore, using the relations Y L = Y lβ R sin θ 0 Y R = Y + lβ L sin θ 0, we find the domain on the θy plane {θ, Y : Y max{lβ R sin θ, lβ L sin θ} for θ [0, π}. Now we examine the motion of the dumbbell. We assume no gravity in the system. The 17

Y y dumbbell θ X,Y m R boundary m L boundary X θ a b Figure 3.1: a The dumbbell model with the flat boundary. b The billiard system for the dumbbell. moment of inertia of the dumbbell is given by I = m L lβ R + m R lβ L = ml β L β R. It is convenient to rescale Y = y since in this way the Lagrangian is equal to the that of the free particle I m L = I θ + mẏ = I θ + Iẏ. 3.1 By Hamilton s principle of least action, true orbits extremize t 1 t 0 billiard trajectories on the θy plane are straight lines between two collisions. We also rescale the billiard domain accordingly and get Q = { θ, y : y max Lẏ, θdt. We see that the { } } β R β L sin θ, sin θ for θ [0, π. 3. β L β R The boundary Q has non-smooth corners at θ = 0, π. This is the case when the dumbbell s two masses hit the floor at the same time. We will not consider this degenerate case in our paper. At the boundary Q, there will be a mirror-like reflection. We use the subscript to 18

denote the values right before a collision and + to denote the values right after a collision. Given the incoming velocity vector v and normal vector n to the boundary Q, the outgoing velocity vector v + can be found using In the case when m L hits the boundary, we have [ v + = v n n n n + v. 3.3 v = [ θ, ẏ ] n = ] β R cos θ, 1. β L Then, using 3.3, we compute v + = [ θ +, ẏ + ]. Changing back to the original coordinates θ, Y, we express the translational and the angular velocities of the dumbbell after the collision in terms of the velocities before the collision. θ + = Ẏ + θ 1 β R cos θ β L + β R cos θ θ lβl β R cos θ β L + β R cos θ + Ẏ /l cos θ β L + β R cos θ β L. 3.4 1 β L + β R cos θ + Ẏ Remark 3.1. The post-collision velocities for the other case, when m R hits the boundary, can be obtained in a similar manner: we switch β L and β R, replace cos θ and sin θ with cos θ and sin θ, and replace Y L with Y R. 19

CHAPTR 4 stimate on the number of collisions of the dumbbell Consider the dumbbell with the arbitrary masses m L and m R moving in the XY plane where no gravity force is present. We estimate the maximal number of collisions of the dumbbell with the floor as a function of the mass ratios. As we have seen in Chapter 3, on the θy plane, the dumbbell reduces to a mass point that has unit velocity in a billiard domain on a cylinder and makes elastic reflections at the boundary. First, we discuss the properties of the billiard boundary for the dumbbell system on the θy plane. Recall from 3. that the boundary is given by Q = { θ, y : y max { } } β R β L sin θ, sin θ for θ [0, π. 4.1 β L β R When m L = m R, we have β L = β R = 1 and the boundary is y = sin θ. The angle between the two sine waves is π. When m L m R, it follows from 4.1 that the boundary consists of two sine waves with different heights. We will assume m L < m R, since the case m R > m L is symmetric. To visualize the billiard boundary more conveniently, we measure θ in the range [π/, 5π/ instead of [0, π Figure 4.1. It is easy to see that generically in the limit m L m R boundary. 0 most of repeated collisions will occur between the two tall parts of the Let us consider the straight wedge formed by the tangent lines to y = β R β L sin θ at θ = π and π. We call these tangent lines l π and l π respectively. We denote the angle of the straight wedge by γ. We define the hybrid wedge as the union of the sine waves when y > 0 and the tangent lines l π and l π when y 0. See Figure 4.1 for the construction of the wedges. 0

y π π θ γ Straight Wedge Hybrid Wedge With Bridge True Boundary Figure 4.1: The construction of the three different types of boundaries. σ s 1 v h 1 σ s 1 = σ h 1 v s 1 v s 1 = v h 1 a b Figure 4.: Two different base cases for Lemma 4.1. We introduce some notations. We denote a billiard trajectory interacting with the hybrid wedge by v h. Similarly, a trajectory bouncing inside the straight wedge is called v s. When v h or v s is written with the subscript i, it denotes the segment of the corresponding trajectory between the i-th collision and the i+1-st collision. Let σi h be the angle from the straight wedge to vi h, and σi s denote the angle from the straight wedge to vi s after the i-th collision. Define ρ h i as the angle deviation of the hybrid wedge from the straight wedge at the i-th collision of vi h. The trajectory v h or v s will terminate when the sequence of angles {σi h } or {σi s } termi- 1

nates due to the absence of the next bounce, or when there will be no more intersections with the straight wedge. This will happen when the angle of intersection, σi h or σi s, with the straight wedge is less than or equal to γ. Lemma 4.1. Consider the hybrid wedge and the straight wedge described above. The sequence of angles σi h, 1 i will terminate after or at the same index as the sequence of angles σi s, 1 i. Proof. Suppose that the initial segment v0 s = v0 h of the dumbbell trajectory crosses the straight wedge before it hits the hybrid wedge, as shown in Figure 4.a. If the incoming angle is α, then σ h 1 = π α ρ h 1 < π α = σ s 1. When the initial segment hits the hybrid wedge before crossing the straight wedge, as shown in Figure 4.b, then set σ h 1 = σ s 1. Now we can proceed by induction if σ h i, σ s i > γ and if the sequence σ h i has not terminated. σ h i+1 = σ h i γ ρ h i+1 σ s i+1 = σ s i γ which implies that σi+1 h σi+1. s Since σi h σi s, then v h will terminate at the same time or before v s. β Define the bridge as the smaller sine wave created by y = L β R sin θ for θ [π, π. The union of the bridge with the hybrid wedge is the true boundary defined by the dumbbell dynamics Figure 4.1. Lemma 4.. The presence of the bridge in the hybrid wedge will increase the number of collisions of the dumbbell by at most one from the number of collisions of the dumbbell to the hybrid wedge. Proof. Consider the true trajectory v that sees the bridge. Recall the definitions of σ h i

and ρ h i. Similarly, we let σ i be the angle to v i and let ρ i be the angle deviation of the true boundary from the straight wedge at the i-th collision of v i. Before v intersects the bridge, by Lemma 4.1 we have v i = vi h, σ i = σi h, and σi h = σi 1 h γ ρ h i. Now define τ to be the signed angle measured from the horizontal line to the tangent line of the bridge curve where v hits the bridge Figure 4.3. Note that τ takes a positive value if v hits the left half of the bridge, and τ takes a negative value if v hits the right half of the bridge. We express σ i+1 after the bounce from the bridge in terms of σ i. By this convention, the bounce from the bridge does not increase the index count. We will adjust the index count by adding +1 at the end. σ i+1 σ i τ Figure 4.3: The real billiard trajectory v that sees the bridge. Then, we have σ i+1 = π σ i τ ρ i+1 4. σ h i+1 = σ h i γ ρ h i+1 We may assume that v hits the bridge with non-positive velocity in y. If the dumbbell hits the bridge with positive velocity in y, it will continue to move in the positive y direction after reflection from the bridge. Then, we consider the reverse trajectory to bound the 3

number of collisions. This allows us to restrict σ i. Moreover, v i naturally hits the upper part of hybrid wedge than vi h. We also assume that v hits the left half of the bridge. Otherwise, we can reflect the orbit around the vertical line passing through the middle point of the bridge. Utilizing the above arguments, we have the inequalities π + γ < σ i, 0 < τ < γ, and ρh i+1 < ρ i+1. 4.3 It is straightforward to verify that 4. and 4.3 imply σ i+1 σi+1. h From the i+-nd bounce, if σ i has not terminated, we can apply the induction argument similar to the proof of Lemma 4.1. We have the base case σ i+1 σi+1, h and ρ i+1 ρ h i+1. Note that ρ s indicate the relative position of a collision point in the hybrid wedge. That is, if ρ i+1 ρ h i+1, then the starting point of v i+1 is located at or above that of vi+1. h Since σ i+1 σi+1 h and v i+1 starts above vi+1, h we know v i+ will start on the hybrid wedge higher than vi+. h This implies ρ i+ ρ h i+. Then using the recursive relationship, σ i+ = σ i+1 γ ρ i+ σ h i+ = σ h i+1 γ ρ h i+, we obtain θ i+ θ h i+. By induction θ i θ h i for all i. Taking into account the bounce on the bridge, we conclude that the number of bounces of v will increase at most by one relative to that of v h. Remark 4.3. In most cases, the number of bounces of v will be less than the number of bounces of v h Figure 4.4. Now we are ready to prove the main theorem of the chapter. 4

Figure 4.4: a The bridge prevents the trajectory to travel deeper. b The bridge increases the number of collisions by 1. Theorem 4.4. The number of collisions of the dumbbell is bounded above by N γ = mr where γ = π arctan. m L π γ + 1, Proof. When m L = m R, the boundary consists of two equal sine waves which form hybrid wedges of angle π at θ = π, π. Using Lemma 4.1, we conclude that the upper bound for the number of collisions is π γ =, which is less than N γ = 3. When m L < m R, we consider the true boundary which consists of a hybrid wedge with the bridge. By Lemma 4.1 and Lemma 4., the the maximal number of collisions to the true boundary is bounded above by N γ = π γ + 1. Since γ = π arctan β R β L = π arctan mr m L, this completes the proof. 5

CHAPTR 5 Adiabatic invariance in the dumbbell m R Ẏ θ m L X Figure 5.1: The light mass bounces many times off the floor while the large mass slowly approaches the floor. We consider the case when the dumbbell has m L m R and m L rotates around m R with a high angular velocity θ. For easy computations, we set the length of the rod l = 1. We further assume that the center of mass has a slow downward velocity compared to θ. Since multiplying the velocities θ, Ẏ by a constant does not change the orbit, we normalize θ to be of order 1, then Ẏ is small. slowly approaching the floor, rotating with angular velocity of order 1. At some moment the small mass m L will hit the floor. If the angle θ = π or sufficiently close to it, then the dumbbell will bounce away without experiencing any more collisions. This situation is rather exceptional. A simple calculation shows that θ Ẏ 3π will be generically of order for our limit Ẏ 0. In this chapter we assume this favorable scenario. For the corresponding set of 6

initial conditions, we obtain an adiabatic invariant. We start by deriving approximate map between two consecutive bounces. To avoid cumbersome use of implicit constants, we use some standard notation when dealing with asymptotic expansions: Notation: f g f = Og f Cg for some C > 0 f g g = Of f g f g and f g. Lemma 5.1. Let β L = ɛ 1, θ 0 and assume m L bounces off the floor and hits the floor next before m R does. Then there exist δ = δɛ such that δ/ɛ and δ 0 as ɛ 0, and if δ < Ẏ < 0 and φ 3π δ, the collision map is given by θ + = θ ɛ Ẏ 1 Y + O δ 5.1 Y + = Y π arccos Y Ẏ + O δ θ 3/ ɛ + O. 5. δ Proof. We prove 5.1 in two steps. We first show that using the expression for θ + in 3.4. We have, θ + + θ + cos θ Ẏ = 1 1 β L cos θ β L + 1 β L cos θ θ + = θ ɛ + O cos φẏ δ = β L 1 β L cos θ θ cos θẏ β L sin + θ + cos θ θ + 1 + Ẏ sin θ = β cos θ L β L sin. θ + cos θ cos θ θ + β L + 1 β L cos θ θ + cos θ Ẏ Ẏ + θ + cos θ Ẏ 7

For sufficiently small δ, θ 3π δ implies cos θ δ. It follows that θ + + θ + cos θ Ẏ ɛ Observe from the Figure 5.1 that θ = 3π arccos θ + + θ + cos θ Ẏ = θ + + θ + = θ + + θ + 1 Y 1 β L 1 + δ δ δ Y 1 β L ɛ = O δ. Thus, 3π cos arccos Ẏ = θ + + θ +. Y 1 β L Ẏ 1 Y Ẏ + R 1, where Ẏ Y R 1 β L. 1 βl Y 3 Using that 1 β L Y = 1 β L cos θ, we obtain Combining the results, we have θ + + θ + Ẏ 1 Y = θ + + R 1 ɛ δ 1 β L δ 3. ɛ θ + cos θ Ẏ + R 1 1 + δ δ δ + δ 1 β L 3 δ 3/ ɛ = O. δ This completes the proof for 5.1. Let t be the time between the two consecutive collisions of m L. Then Y + = Y Ẏ t. 8

The angular distance that m L traveled is given by Y ψ = π arccos arccosy + 1 β L Y = π arccos arccos 1 β L Y Ẏ t 1 β L Y = π arccos + R = π arccos Y + R 3 + R, 1 β L where R and R 3 are the error estimates for the Taylor series expansion and are given explicitly by Therefore, we have Ẏ t R 1 βl Y β L Y R 3 1 β L 1 β L Y. Y + = Y Ẏ t = Y Ẏ ψ θ = Y Ẏ θ π arccos Y + R + R 3 = Y π arccos Y Ẏ + Ẏ R + R 3. θ θ Since m L can travel at most π between two collisions, t is bounded by t < π/ θ. Also note that R and R 3 contain the factor Ẏ and β L respectively. We finish the proof for 5. 9

by computing, Y + Y + π arccos Y Ẏ Ẏ R + R 3 θ θ Ẏ πy θ 1 βl Y + β πy L θ 1 β L 1 β L Y δ π + ɛ π δ 1 β L δ = O δ 3/ ɛ + O. δ Corollary 5.. Under the same assumptions as in Lemma 5.1 with the exception φ 3π δ k for 0 k 1 and ɛ δ k, the variables after the collision are given by the similar equations to 5.1 and 5. but with different error terms. θ + = θ ɛ Ẏ 1 Y + O δ k Y + = Y π arccos Y Ẏ + O θ Proof. When computing the error terms, use cos θ δ k. δ δ k 5.3 ɛ + O. 5.4 δ k Now, we can state the adiabatic invariance theorem for the special case when the light mass hits the floor and the dumbbell is far away from the vertical position: θ = 3π. Theorem 5.3. Suppose right before the collision θ 0 and θ 3π 0. Then there is δ > 0 such that if 0 < β L = ɛ = δ and δ < Ẏ0 < 0, then there exists an adiabatic invariant of the dumbbell system, given by I = θ fy, where fy = π arccos Y. In other words, θ n fy n θ 0 fy 0 = Oδ after N = Oδ 1 collisions. Proof. We prove this by finding fy that satisfies θ + fy + θ fy = Oδ. 30

When ɛ = δ and δ is sufficiently small, it follows from 5.4 that, fy + = fy π arccos Y Ẏ + Oδ θ f Y. Then, we have θ + fy + = θ + Ẏ 1 Y + Oδ π arccos Y fy Ẏ + Oδ θ f Y = θ fy π arccos Y Ẏ f Y + 1 Y Ẏ fy + Oδ. Therefore, fy satisfies θ + fy + θ fy = Oδ provided π arccos Y Ẏ f Y + Ẏ fy = 0. 1 Y The solution of the above equation is given by fy = π arccos Y. Let N = Oδ 1 and let θ N and Y N be the angular velocity and the distance after N th collision. Then, we have θ N fy N θ 0 fy 0 = N k=1 θ k fy k θ k 1 fy k 1 N δ δ. Remark 5.4. Adiabatic invariant has a natural geometric meaning: angular velocity times the distance traveled by the light mass between two consecutive collisions. Now, we state the theorem for a realistic scenario when a rapidly rotating dumbbell scatters off the floor. Theorem 5.5. Let the dumbbell approach the floor from infinity with θ 0. There exists δ > 0 such that if 0 < β L = ɛ = δ, δ < Ẏ < 0, θ 0 3π δ then, after N = Oδ 1 31

bounces the dumbbell will leave the floor after the final bounce by m L with I N = I 0 + O δ. The adiabatic invariant is defined as above I = θ fy. Remark 5.6. The condition on the angle θ 0 3π δ comes naturally from the following argument. If Ẏ = δ, the dumbbell approaching from infinity will naturally hit the floor when y 1 β L δ. Since δ is small, this implies θ 0 3π δ. If θ 0 happens to be too close to 3π, then there is no hope to obtain adiabatic invariant and we exclude such set of initial conditions. In the limit δ 0 the relative measure of the set where θ 0 3π = o δ tends to zero. Proof. We will split the iterations bounces into two parts: before the n th iteration and after it, where n = µ δ and µ is sufficiently small to be defined later. We claim that after n bounces, θ n 3π 4 δ. To prove this claim, we use energy conservation of the dumbbell system 3.1, and 5.3. We have ɛ1 ɛ θ + ɛ Ẏ 1 Y + O + δ Ẏ + = ɛ1 ɛ θ + Ẏ. 5.5 Next, Ẏ + Ẏ = ɛ1 ɛ θ ɛ ɛ1 ɛ θ + Ẏ 1 Y + O δ ɛ 4Ẏ 1 Y + 4 θ Ẏ + θ ɛ 1 Y O + 4Ẏ ɛ ɛ O + O. δ 1 Y δ δ By our assumptions, 1 Y δ so it follows that δ Ẏ + Ẏ δ δ + δ + δ + δ + δ δ 5/, δ δ which implies Ẏ+ Ẏ δ 3/. After n = µ δ bounces, Ẏn Ẏ0 δ if µ is sufficiently small and we still have the vertical velocity of same order, i.e. Ẏ n Ẏ0 δ. Then, at the n th collision, the center of 3

mass will be located at y n 1 δ, which will imply θ n 3π 4 δ. Now using Lemma 5.1, Corollary 5., and Theorem 5.3, we compute the error term of the adiabatic invariant under the assumption that the total number of collisions is bounded by N δ 1 and the heavy mass does not hit the floor. θ N fy N θ 0 fy 0 = n k=1 θ k fy k θ k 1 fy k 1 + = µ δ O δ + N C O δ 3/ = O δ. δ n θ k fy k θ k 1 fy k 1 By Theorem 4.4 proved in the previous chapter, there is indeed a uniform bound on the number of bounces. If the heavy mass does hit the floor it can do so only once as shown in the next section. We claim that the corresponding change in the adiabatic invariant will be only of order δ. Using 3.4 and Remark 3.1 from Chapter 3, we obtain Ẏ + = Ẏ + Oɛ θ + = θ + Oɛ + Oδ. Let the pairs θ m, Y m θ m+1, Y m+1 denote the corresponding values of θ, Y when the light mass hits the floor right before and after the large mass hits the floor. Then, since Ẏ = Oδ, we find that Y m+1 Y m = Oδ and θ m+1 θ m = Oδ. As a consequence, θ m+1 fy m+1 θ m fy m = Oδ and the change in adiabatic invariant due to large mass hitting the floor is sufficiently small I = Oδ. 33

CHAPTR 6 Chaotic dynamics of the two-dimensional coin Using the equal-mass dumbbell model, we study the dynamics of a bouncing coin whose motion is restricted to the two-dimensional plane. As we mentioned in the introduction, the choice of the two-dimensional coin models does not affect our results. To simplify the terminologies, we will refer to the equal-mass dumbbell as a coin. We first describe the coin system as a point billiard in a transformed domain. Although the majority of the steps overlaps with our discussion from Chapter 3, there are some additional details we need to note. 6.1 The coin system as a gravitational billiard Consider the coin interacting with the flat boundary under the influence of gravity. Let all other variables be the same as in Chapter 3, except the θ variable. We still define θ to be the angular position of the vector X L X R measured counterclockwise from the positive X-direction. However, we measure θ in R keeping track of the rotation number. We will identify θ and θ + πz as the same later. The moment of inertia of the coin is given by I = m L l + m R l = 1 4 ml. After rescaling, the energy of the system is given by = I θ + mẏ + mgy = I θ + Iẏ + g miy. 6.1 By considering the corresponding Lagrangian of the system and applying the Hamilton s principle of least action, we know that the motion of the coin between two collisions on 34