Chaos in Hamiltonian systems

Similar documents
= 0. = q i., q i = E

Physics 106b: Lecture 7 25 January, 2018

Hamiltonian Dynamics

Lecture 11 : Overview

Physics 5153 Classical Mechanics. Canonical Transformations-1

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

Hamiltonian Lecture notes Part 3

Lecture 8 Phase Space, Part 2. 1 Surfaces of section. MATH-GA Mechanics

Hamiltonian Chaos and the standard map

Area-PReserving Dynamics

A Glance at the Standard Map

Chaotic transport through the solar system

ANALYTICAL MECHANICS. LOUIS N. HAND and JANET D. FINCH CAMBRIDGE UNIVERSITY PRESS

Chaotic motion. Phys 420/580 Lecture 10

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

Lectures on Dynamical Systems. Anatoly Neishtadt

Chapter 1. Principles of Motion in Invariantive Mechanics

15. Hamiltonian Mechanics

Aubry Mather Theory from a Topological Viewpoint

Orbits, Integrals, and Chaos

Sketchy Notes on Lagrangian and Hamiltonian Mechanics

Nonlinear Single-Particle Dynamics in High Energy Accelerators

Theory of Adiabatic Invariants A SOCRATES Lecture Course at the Physics Department, University of Marburg, Germany, February 2004

Chaotic motion. Phys 750 Lecture 9

GEOMETRIC QUANTIZATION

An introduction to Birkhoff normal form

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

HAMILTON S PRINCIPLE

Part II. Classical Dynamics. Year

Chaos in the Hénon-Heiles system

Hamilton-Jacobi theory

Oscillatory Motion. Simple pendulum: linear Hooke s Law restoring force for small angular deviations. small angle approximation. Oscillatory solution

PHY411 Lecture notes Part 2

Oscillatory Motion. Simple pendulum: linear Hooke s Law restoring force for small angular deviations. Oscillatory solution

Eva Miranda. UPC-Barcelona and BGSMath. XXV International Fall Workshop on Geometry and Physics Madrid

Introduction to Applied Nonlinear Dynamical Systems and Chaos

Homework 3. 1 Goldstein Part (a) Theoretical Dynamics September 24, The Hamiltonian is given by

Lecture 1: Oscillatory motions in the restricted three body problem

From quantum to classical statistical mechanics. Polyatomic ideal gas.

PHY411 Lecture notes Part 3

Canonical transformations (Lecture 4)

Symmetries. x = x + y k 2π sin(2πx), y = y k. 2π sin(2πx t). (3)

Hamiltonian. March 30, 2013

Lecture 5. Alexey Boyarsky. October 21, Legendre transformation and the Hamilton equations of motion

PHY411 Lecture notes Part 1

Hamiltonian aspects of fluid dynamics

M2A2 Problem Sheet 3 - Hamiltonian Mechanics

Generalized Coordinates, Lagrangians

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

M. van Berkel DCT

Linear and Nonlinear Oscillators (Lecture 2)

Variational principles and Hamiltonian Mechanics

Hamiltonian Chaos. Niraj Srivastava, Charles Kaufman, and Gerhard Müller. Department of Physics, University of Rhode Island, Kingston, RI

REVIEW. Hamilton s principle. based on FW-18. Variational statement of mechanics: (for conservative forces) action Equivalent to Newton s laws!

The Geometry of Euler s equation. Introduction

An Exactly Solvable 3 Body Problem

Physical Dynamics (SPA5304) Lecture Plan 2018

Action-Angle Variables and KAM-Theory in General Relativity

Physics 106a, Caltech 13 November, Lecture 13: Action, Hamilton-Jacobi Theory. Action-Angle Variables

Classical Statistical Mechanics: Part 1

Synchro-Betatron Motion in Circular Accelerators

L = 1 2 a(q) q2 V (q).

MATHEMATICAL PHYSICS

INTERACTION OF TWO CHARGES IN A UNIFORM MAGNETIC FIELD I: PLANAR PROBLEM

Physical Dynamics (PHY-304)

Physics GRE: Classical Mechanics. G. J. Loges 1. University of Rochester Dept. of Physics & Astronomy. xkcd.com/815/

Liouville Equation. q s = H p s

Nonlinear Single-Particle Dynamics in High Energy Accelerators

HANDOUT #12: THE HAMILTONIAN APPROACH TO MECHANICS

The Principle of Least Action

Classical mechanics of particles and fields

THE LAGRANGIAN AND HAMILTONIAN MECHANICAL SYSTEMS

Gauge Fixing and Constrained Dynamics in Numerical Relativity

ON JUSTIFICATION OF GIBBS DISTRIBUTION

Perturbation theory, KAM theory and Celestial Mechanics 7. KAM theory

THE POINCARÉ RECURRENCE PROBLEM OF INVISCID INCOMPRESSIBLE FLUIDS

= w. These evolve with time yielding the

Nonlinear dynamics & chaos BECS

Modified Equations for Variational Integrators

11 Chaos in Continuous Dynamical Systems.

Pentahedral Volume, Chaos, and Quantum Gravity

Consider a particle in 1D at position x(t), subject to a force F (x), so that mẍ = F (x). Define the kinetic energy to be.

Two-Body Problem. Central Potential. 1D Motion

A forced spring-pendulum

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

The Toda Lattice. Chris Elliott. April 9 th, 2014

L(q, q) = m 2 q2 V (q) 2 m + V (q)

PHYS 705: Classical Mechanics. Hamiltonian Formulation & Canonical Transformation

Some Collision solutions of the rectilinear periodically forced Kepler problem

2.10 Saddles, Nodes, Foci and Centers

[#1] R 3 bracket for the spherical pendulum

Classical Mechanics Comprehensive Exam Solution

1 Hamiltonian formalism

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

Part II. Dynamical Systems. Year

Nonlinear Mechanics. A. W. Stetz

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

Lecture 4. Alexey Boyarsky. October 6, 2015

Lecture I: Constrained Hamiltonian systems

Classical Mechanics. Character: Optative Credits: 12. Type: Theoretical Hours by week: 6. Hours. Theory: 6 Practice: 0

Transcription:

Chaos in Hamiltonian systems Teemu Laakso April 26, 2013 Course material: Chapter 7 from Ott 1993/2002, Chaos in Dynamical Systems, Cambridge http://matriisi.ee.tut.fi/courses/mat-35006 Useful reading: Goldstein 2002, Classical Mechanics, Addison Wesley Hand & Finch 1998, Analytical Mechanics, Cambridge Advanced: Lichtenberg & Lieberman 1992, Regular and Chaotic Dynamics, Springer Arnold 1989, Mathematical Methods of Classical Mechanics, Springer Contents 1 Introduction: classical mechanics 2 2 Symplectic structure 4 3 Canonical changes of variables 6 4 Hamiltonian maps 7 5 Integrable systems 10 6 Perturbations and the KAM theorem 13 7 The fate of resonant tori 15 8 Transition to global chaos 17 1

1 Introduction: classical mechanics Newtonian mechanics Consider a particle of mass m, subject to a force field F, in a d-dimensional Euclidean space (one-body system). Newton s second law; m r = F = a system of 2d first order ODEs for position r R d and velocity ṙ R d. These are equations of motion in phase space (the space R d R d where (r, ṙ) are coordinates). The number of degrees of freedom of the mechanical system is N = d. Example (harmonic oscillator). Set d = 1, F = kr = r = kr/m [draw a picture]. dr dt = ṙ dṙ dt = k m r. The solution for initial values r(0) = r 0, ṙ(0) = 0 is r(t) = r 0 cos ( t ) k/m. [Characteristic equation z 2 + k/m = 0 = z 1,2 = ±i k/m = ±iω h = r(t) = a cos ω h t + b sin ω h t.] Lagrangian mechanics For an unconstrained system of n bodies: N = dn (2N ODEs). Under holonomic constraints f j (r 1, r 2,..., r n, t) = 0, j = 1,..., k we can define generalized coordinates q i, i = 1,..., N, where N = dn k, using transformation equations r 1 = r 1 (q 1, q 2,..., q N, t). (1) r n = r n (q 1, q 2,..., q N, t). Definition. The Lagrangian (function) is L = T V, where T = n i=1 m i(ṙ i ṙ i )/2 is the kinetic energy, and V = V (r 1, r 2,..., r n, t) is the potential energy. 2

Through the transformation equations (1), we have L = L(q, q, t). The Hamilton s principle δi = 0, I = t2 t 1 L(q, q, t)dt is a variational equation for finding a path q(t) from t 1 to t 2 for which the line integral I (action) is stationary [draw a picture]. Solution [e.g., Goldstein] yields the (Euler-)Lagrange equations of motion: d L L = 0, i = 1,..., N. dt q i q i [For all possible paths, the system takes the one requiring the least action.] Example (harmonic oscillator). L = T V = mṙ 2 /2 kr 2 /2, d L dt ṙ = m r, Hamiltonian mechanics L r = kr = r = k m r. Definition. The conjugate momenta p i = L/ q i and Hamiltonian H = q i p i L (Einstein summation convention). We do a Legendre transformation; dl = L q i dq i + L q i d q i + L t dt = ṗ idq i + p i d q i + L t dt dh = q i dp i ṗ i dq i L t dt Since we wish H = H(q, p, t), dh = H q i dq i + H p i dp i + H t dt. Equating terms, we have the Hamiltons equations of motion: p i = H q i q i = H. p i The (q, p) R N R N are canonical phase-space variables. 3 (2)

Example (harmonic oscillator). p = L/ ṙ = mṙ, H = mṙ 2 L = mṙ 2 /2 + kr 2 /2 = T + V = E. In general, if the transformation equations (1) do not depend on t explicitly, and the forces are conservative (of the form F = Φ), Hamiltonian is the total energy. For a time-independent Hamiltonian H(p, q) the total energy is conserved; dh dt = H H q + q p ṗ = H H q p H H p q = 0. = H(p, q) = E is a constant of motion. 2 Symplectic structure The class of Hamiltonian systems is very special. Let x = (p, q) T and f(x, t) = Ω ( ) T H, where Ω = x [ ] 0 I. I 0 Now, Hamilton s equations are ẋ = f(x, t). Note that f(x, t) R 2N Hamiltonian vector field) is fully determined by H(x, t) R. (the Theorem (Liouville s theorem). Hamilton s equations preserve 2N-dimensional volumes. Proof. d d 2N x = dt V x f = p S dx dt ds = ( H q S ) + ( ) H q = 0, p ( f ds = x f V ) d 2N x = 0. [differentiation under integral sign, Hamilton s equations, divergence (Gauss ) theorem]. Corollary. There are no attractors in Hamiltonian systems. 4

Hamilton s equations are symplectic. Consider three orbits (p, q), (p + δp, q + δq), and (p + δp, q + δq ). The differential symplectic area is the sum of parallelogram areas: [Ott, fig. 7.1] N δp i i=1 δp i δq i N = (δp i δq i δq i δp i) = δp δq δq δp = δx T Ωδx. δq i i=1 d ( δx T Ωδx ) = dδxt Ωδx + δx T Ω dδx dt dt dt ( ) T f = x δx Ωδx + δx T Ω f x δx [ ( f ) ] T = δx T Ω + Ω f δx x x = δx T [ ( Ω 2 H x 2 = δx T [ ( 2 H x 2 = 0. ) ] T Ω + ΩΩ 2 H δx x 2 ) ] T Ω T Ω + ΩΩ 2 H δx x 2 [linearization of f(x + δx, t), ΩΩ = I, Ω T = Ω, 2 H/ x 2 symmetric]. Symplectic condition = conservation of volume. Definition. The differential symplectic area is a differential form of Poincaré s integral invariant N p dq = p i dq i, γ i=1 where γ is a closed path in (p, q) at constant t. A generalization in extended 2N + 1 phase space (p, q, t) [Ott, fig. 7.2]: Theorem (Poincaré-Cartan integral theorem). (p dq Hdt) = (p dq Hdt), Γ 1 Γ 2 where Γ 1 and Γ 2 are paths around the tube of trajectories. 5 γ

Γ 1 and Γ 2 at constant times (dt = 0) = Poincaré s integral invariant. If H = H(p, q) and Γ 1 and Γ 2 are on this surface, we have Hdt = 0 and p dq = p dq, Γ 1 Γ 2 where Γ 1 and Γ 2 does not need to be at constant t. 3 Canonical changes of variables Let us define a new set of phase-space variables X = (P, Q) T transformation g : R 2N R 2N, x X. by using a Definition. The transformation g is canonical, if it preserves the differential symplectic area; δp δq δq δp = δp δq δq δp. (3) Canonical transformations are typically performed by using a generating function, e.g., S = S(P, q, t); Q = S P, p = S q, which gives (P, Q) implicitly. The transformed Hamiltonian is K(P, Q, t) = H(p, q, t) + S/ t. [Can be derived using the Hamilton s principle; see Goldstein.] In order to check the canonicity of the above one can substitute δq = 2 S P 2 δp + 2 S P q δq, δp = 2 S P q δp + 2 S q δq, 2 into the condition (3) [homework]. The new variables satisfy Hamilton s equations: dx/dt = Ω( K/ X) T. Hence, the underlying symplectic structure of a given Hamiltonian system is invariant, and can be represented using any suitable choice of canonical phase-space variables. 6

4 Hamiltonian maps Consider the map M h : R 2N R 2N : M h (x(t), t) = x(t + h). Its differential variation [linearize M h (x + δx, t)] is Symplectic condition = M h δx(t) = δx(t + h). x δx T (t)ωδx (t) = δx T (t + h)ωδx (t + h) ( ) T ( ) Mh = x δx(t) Mh Ω x δx (t) ( ) T ( ) = δx T Mh Mh (t) Ω δx (t) x x ( ) T ( ) Mh Mh = Ω = Ω x x We say, M h / x is symplectic. Generally, matrix A is symplectic, if Ω = A T ΩA. The product of symplectic matrices A and B is symplectic; (AB) T Ω(AB) = B T (A T ΩA)B = B T ΩB = Ω. Theorem (Poincaré recurrence theorem). Examine a Hamiltonian H(p, q) with orbits bounded in a finite subset D R 2N. Let a ball R 0 D have radius ε > 0. In time, some of the orbits leaving R 0 will return to it. Proof. Evolve R 0 with the volume-preserving M h, obtaining subsequent regions R 1, R 2,.... Conclude that R r R s = R r s R 0, r > s [Arnold, fig. 51]. Poincaré maps In general, a Poincaré map gives an intersection of an orbit with a lowerdimensional subspace of the phase space, called Poincaré section (or surface of section, SOS). For periodic orbits, plotting subsequent intersections typically draws a figure which reflects deeper characteristics of the dynamical system. 7

Definition. Poincaré section for a Hamiltonian system H(p, q, t) where H is τ-periodic in time: extended (2N + 1) -dimensional phase space (p, q, ξ), where ξ = t, additional equation dξ/dt = 1, let ξ = ξ mod τ, set ξ = t 0 [0, τ) as the 2N-dimensional SOS. Because of the periodicity, M τ (x, t 0 ) = M τ (x, t 0 + kτ), k Z. surface of section map x n+1 = M(x n ) = M τ (x n, t 0 ) is symplectic. The Example (Standard map). The kicked rotor : bar of moment of inertia I and length l, frictionless pivot [Ott, fig. 7.3]. Vertical impulse of strength K/l at times t = 0, τ, 2τ,... Canonical variables (p θ, θ), Hamiltonian H(p θ, θ, t) = p2 θ 2I + K cos θ n δ(t nτ), where δ denotes the Dirac delta, and equations of motion (2) dp θ dt = K sin θ δ(t nτ), dθ dt = p θ I. SOS: (p θ, θ) after each kick. Integration over t (nτ, (n + 1)τ] yields p n+1 p n = K sin θ n+1 and θ n+1 θ n = p n τ/i. Setting τ/i = 1, we have p n+1 = p n + K sin θ n+1, θ n+1 = (θ n + p n ) mod 2π. The mapping (N = 1) preserves differential areas, if det(δx n, δx n) = det [( x n+1 / x n )(δx n, δx n)], where x n = (p n, θ n ) T. Symplecticity can hence be verified by calculating [ ] [ ] pn+1 / p det n p n+1 / θ n 1 + K cos θn+1 K cos θ = det n+1 = 1. θ n+1 / p n θ n+1 / θ n 1 1 8

Definition. Poincaré section for a Hamiltonian system H(p, q): motion in a (2N 1)-dimensional energy surface H(p, q) = E, by choosing, e.g., q 1 = 0, we determine a (2N 2)-dimensional SOS. For a unique SOS mapping x n+1 = M(x n ), we often need an additional constraint, e.g., p 1 > 0. Example (Logarithmic potential). The potential Φ near the centre of a flattened galaxy can be modelled as Φ(x, y) = 1 ( ) 2 ln x 2 + y2 a + 2 b2, (4) where a, b R are parameters [draw a picture]. For a star moving in this potential, the Hamiltonian per unit mass is H(p x, p y, x, y) = 1 [ )] p 2 x + p 2 y + ln (x 2 + y2 2 a + 2 b2. Equations of motion (2) become x ṗ x = x 2 + y 2 /a 2 + b, 2 y ṗ y = a 2 (x 2 + y 2 /a 2 + b 2 ), ẋ = p x, ẏ = p y. Choose y = 0 as the SOS. Now H(p x, p y, x, 0) = E = p 2 y = 2E p 2 x ln ( x 2 + b 2). Hence, there are two points ±p y corresponding to y = 0. We choose that p y > 0 on the SOS. In this case, we cannot solve the equations of motion analytically, and points (p x, x) on the SOS have to be determined by numerical integration. 9

5 Integrable systems Consider a Hamiltonian H(p, q). A constant of motion is a quantity that does not change when the system evolves in time, e.g., H(p, q) = E. In general, for a function f(p(t), q(t)), we have df dt = f p ṗ + f q f H q = q p f H p q. If f(p, q) is a constant of motion, we can write [f, H] = 0, where the Poisson bracket for functions f and g is defined as [f, g] := f g q p f g p q. Poisson bracket is anticommutative: [f, g] = [g, f]. Theorem (Liouville s theorem on integrable systems). The Hamiltonian system H(p, q) is integrable, if it has N independent constants of motion f i (p, q), i = 1,..., N which are in involution, i.e., [f i, f j ] = 0, i, j. [Proof: see, e.g., Arnold.] The independent constants of motion restrict an integrable system on an N-dimensional surface in the phase space. Because the constants are in involution, any bounded orbit on this surface is topologically equivalent to an N-dimensional torus [Ott, fig. 7.4a]. Action-angle variables The action-angle variables provide an explicit transformation between points (p, q) and points on the N-torus. Through a canonical change of variables, it is possible to transform to coordinates (P, Q) in which the new momenta P i (p, q), i = 1,..., N are constants of motion. In such a case dp i /dt = H/ Q i = 0 = H = H(P ). A particularly convenient choice is P i = J i := 1 2π γ i p dq, where the paths γ i wrap around each possible angle direction i = 1,..., N on the N-torus [Ott, fig. 7.4b]. The new variables J i are actions. They are constants of motion by the Poincaré-Cartan integral theorem. 10

The corresponding conjugate coordinates Q i = θ i are angles. Suppose that action-angle variables (J, θ) are obtained by the generating function S(J, q); θ = S J, p = S q By integrating the latter equation around γ i, we examine the change in S; i S = p dq = 2πJ i. γ i Then, or i θ = J is = 2π J J i, i θ j = 2πδ ij, where δ ij = 1, if i = j, and zero otherwise. Therefore, after one circuit around γ i, the angle θ i increases by 2π, and the other angles return to their original values. The new Hamiltonian is H = H(J), and dj dt = 0, dθ dt = H J =: ω(j). A solution from initial values (J 0, θ 0 ) is J(t) = J 0 and θ(t) = θ 0 + ωt. The constants of motion ω are frequencies which can be interpreted as angular velocities on the torus. Definition. An orbit on the torus is quasiperiodic if there is no vector of integers m Z n such that m ω = 0. A quasiperiodic orbit fills the surface of the torus as t. On the other hand, if ω i /ω j Q, i, j, the orbit is periodic, and closes on itself. The set of tori with periodic orbits is dense in the phase space, but has a zero Lebesgue measure. Hence, orbits on a randomly picked torus are quasiperiodic with probability of one. 11

Example (harmonic oscillator). In general, for N = 1, we have H(p, q) = p2 2m + V (q), where p, q R, and V (q) is a potential energy of a general form [Ott, fig. 7.5]. We have J = 1 π q2 q 1 {2m[E V (q)]} 1/2 dq. For the harmonic oscillator, V (q) = mωh 2q2 /2, where ω h = k/m is the angular velocity of the oscillation, and k is the spring constant. From V (q) = E we have q 2 = q 1 = [2E/(mωh 2)]1/2, and integration 1 yields J = E/ω h. Thus, H(J) = ω h J, and ω(j) = H/ J = ω h, and is (in this rare case) independent of J. The angle θ(t) = θ 0 + ω h t. In order to use the generating function S(J, q), we write H(p, q) = H(J), and substitute p = S/ q, obtaining S q = 2m(ω h J mωh 2q2 /2) Solving for θ, and integrating 2, gives θ = S J = mω [2m(ωh h J mωhq 2 2 /2) ] 1/2 dq ( ) 1/2 2J = q 2 dq mω h ( ) mωh = arcsin q, 2J and we have (p, q) as a function of (J, θ); 2J q = sin θ, mω h p = 2Jmω h cos θ. 1 (a 2 x 2 ) 1/2 dx = x 2 (a2 x 2 ) 1/2 + a2 2 arcsin x a 2 dx (a 2 x 2 ) 1/2 = arcsin x a 12

The trajectory (p(θ), q(θ)) is an ellipse [Ott, fig. 7.6]. In general, the mapping (p, q) (J, θ) for N = 1 transforms a closed curve into a one-dimensional torus (a circle). For N > 1, a general procedure for finding the action-angle variables (J, θ) involves solving the Hamilton-Jacobi equation ( ) S H q, q = E. If we are lucky, the equation is solvable by separation of variables, and we have N separation constants which are also constants of motion. The motion is typically quasiperiodic. 6 Perturbations and the KAM theorem The phase space of integrable Hamiltonian systems is foliated by tori. These appear as nested closed curves on the SOS. The existence of N integrals of motion allows us to access the tori via action-angle variables. We know how to solve the Hamilton-Jacobi equation (N > 1) for only a handful of systems, e.g., harmonic oscillator, isochrone, and general Stäckel potentials. On the other hand, the logarithmic potential, for example, numerically demonstrates the existence of an additional constant of motion (besides E) almost everywhere on the SOS [computer demo, logarithmic potential]. Motion on a torus is regular, and chaotic orbits are possible only in regions of phase space where invariant tori do not exist. In order to see how the integrable tori change under perturbation, we study H(p, q) = H 0 (p, q) + εh 1 (p, q), where H 0 is an integrable Hamiltonian, H 1 is non-integrable, and ε R is small. We know that under small perturbations some Hamiltonian systems stay close to integrable (Solar System), but on the other hand, some are globally chaotic (statistical mechanics). Kolmogorov (1954), Arnold (1963), and Moser (1973) proved a result that for sufficiently small ε, most of the tori survive the perturbation (the KAM theorem). 13

The actual proof is difficult [see, e.g., Arnold]. Some considerations follow. In action-angle variables of H 0 ; H(J, θ) = H 0 (J) + εh 1 (J, θ). If there are tori in H, we have a new set of action-angle variables (J, θ ) for which H(J, θ) = H (J ), and which are obtained by a generating function S(J, θ) such that J = S θ, θ = S J. The corresponding Hamilton-Jacobi equation (HJE) is ( ) S H θ, θ = H (J ). We look for a solution in the form of a power series: S = S 0 + εs 1 + ε 2 S 2 +.... We set S 0 = J θ, because ε = 0 then corresponds to J = J, θ = θ. Substitution to HJE gives ( H 0 J + ε S 1 θ + S ) ( ε2 2 θ +... + εh 1 J + ε S ) 1 θ +..., θ = H (J ). By differentiating at J with respect to the first order ε-terms we have H 0 (J ) + ε H 0 J S 1 θ + εh 1(J, θ) = H (J ). (5) Next, we write θ-dependent terms as Fourier series; H 1 (J, θ) = m S 1 (J, θ) = m H 1,m (J ) exp(im θ), S 1,m (J ) exp(im θ), 14

where H 1, S 1 are real-valued, and the coefficients H 1,m, S 1,m C, m Z N. By substituting into (5), and requiring that H is a function of J only, we have S 1 = i H 1,m (J ) exp(im θ), m ω m 0 (J ) where ω 0 (J) = H 0 (J)/ J are the unperturbed frequencies. Similar terms S 2, S 3,... follow, if this technique is applied to higher orders of ε. Clearly, the resonant tori for which m ω 0 = 0 play a special role ( the problem of small denominators ). The KAM theorem essentially proves that the series for S converges for very nonresonant tori with a frequency vector ω satisfying m ω > K(ω) m (N+1), m Z N \{0}, where K(ω) > 0, and m = m 1 + m 2 +... + m N. The Lebesgue measure of the complement of this set, around resonant tori, goes to zero, as ε 0. However, since the unperturbed resonant tori are dense in the phase space, whenever ε > 0, arbitrarily near each surviving torus there is a region of destroyed resonant tori which can host irregular motion. 7 The fate of resonant tori Twist map (r n+1, φ n+1 ) = M 0 (r n, φ n ) is a model for the SOS map of an integrable Hamiltonian H 0 (J) with N = 2 [Ott, fig. 7.7]; r n+1 = r n, φ n+1 = [φ n + 2πR(r n )] mod 2π, where R(r) = ω 1 /ω 2 is the ratio of frequencies for the torus r, and the SOS is drawn at θ 2 = constant. On a resonant torus at r = ˆr, R(ˆr) = j/k, j, k Z such that kω 1 jω 2 = 0, and k applications of the map return to the original point; M0 k (r, φ) = (r, (φ + 2πj) mod 2π) = (r, φ). Hence, each point on r = ˆr is a fixed point of M k 0. We assume that R is smooth and increasing. In such a case there are circles r < ˆr < r + in such a way such that under M k 0, the points on r and r + travel clockwise and counterclockwise, respectively [Ott, fig. 7.8a]. 15

When H 0 is perturbed by the term εh 1, we have a slightly changed map M ε ; r n+1 = r n + εg(r n, φ n ), φ n+1 = [φ n + 2πR(r n ) + εh(r n, φ n )] mod 2π. If ε is small enough, we still have the (distorted) circles r and r + where the points travel to opposite directions under Mε k. Thus, there must be a curve ˆr ε (φ) in between where Mε k moves points only in radial direction [Ott, fig. 7.8b]. When Mε k is applied to the points on ˆr ε (φ), we obtain another curve ˆr ε(φ) [Ott, fig. 7.9]. Since Mε k is volume-preserving, the two curves intersect at a finite and even number of points which are fixed points of Mε k. Since we know the directions where the points travel under Mε k with respect to the curves, we can identify half of the fixed points as elliptic and the other half as hyperbolic. [Ott, fig. 7.10] This is the Poincaré-Birkhoff theorem. The number of elliptic (and hyperbolic) points is equal to k. An elliptic orbit on the SOS circles around the elliptic points. [computer demo, logarithmic potential] The neighbourhood of an elliptic point can be further modelled by another perturbed twist mapping, resulting in a new set of elliptic and hyperbolic points in a smaller scale. If this process is continued ad infinitum, a fractal-like structure of the phase space is revealed, with subsequently smaller tori winding on bigger ones [Lichtenberg & Lieberman, fig. 3.5]. A hyperbolic point lies at an intersection of an unstable and stable manifold, W u and W s, respectively. Consider a repeated SOS mapping (Mε k ) n. A point on W s approaches the hyperbolic point, as n. The same happens to a point on W u as n. The two manifolds W u and W s connect to another hyperbolic point in a way that resembles the phase portrait of a simple pendulum. [Lichtenberg & Lieberman, fig. 3.3a] For the pendulum, W u and W s between the points coincide, forming a smooth separatrix. However, in the perturbed system, the W u and W s are generally different, and intersect at a homoclinic point. This point, mapped by (Mε k ) n must also lie on both W u and W s. Hence, we conclude that there must be an infinite number of homoclinic points as n ±. Since Mε k is area-preserving, W u (W s ) must oscillate wildly as n (n ). Near a hyperbolic point, these oscillations overlap, and form a 16

region where points from W u and W s are mixed (homoclinic tangle). [Lichtenberg & Lieberman, fig. 3.4b] Orbits in this region show extreme sensitivity to initial conditions, and hence, are chaotic [computer demo, logarithmic potential]. 8 Transition to global chaos In two degrees of freedom, the chain of resonant islands and chaotic regions surrounding them are confined between surviving KAM-tori. As the perturbation strength ε increases, chaotic regions grow, and more KAM-tori are destroyed. When all KAM-tori are destroyed, an can wander anywhere in the energy surface, and the motion becomes ergodic [computer demo, standard map]. For the standard map, the rotation number of the last surviving KAMtorus is the golden ratio which is the farthest away from any rational number in the sense that its representation as a continued fraction is 1 + 5 2 = 1 + 1 1 1 + 1 1 + For N = 2 the 2-tori can enclose volumes in the three dimensional energy surface H(p, q) = E. For N > 2 the situation is qualitatively different; e.g., for N = 3 the 3-tori cannot enclose 5-dimensional volumes much like lines cannot enclose 3-dimensional volumes. For any ε > 0, stochastic regions form a web where orbits can, in principle, reach any part of the energy surface through Arnold diffusion. However, Nekhoroshev theorem tells us that the Arnold diffusion happens very slowly; at time scales exp(1/ε) or even slower.. 17