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

Similar documents
TitleQuantum Chaos in Generic Systems.

Physics 106b: Lecture 7 25 January, 2018

= 0. = q i., q i = E

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

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

The Transition to Chaos

Quantum Billiards. Martin Sieber (Bristol) Postgraduate Research Conference: Mathematical Billiard and their Applications

Pentahedral Volume, Chaos, and Quantum Gravity

Lecture 11 : Overview

Averaging II: Adiabatic Invariance for Integrable Systems (argued via the Averaging Principle)

The wave function for a particle of energy E moving in a constant potential V

Chaos, Quantum Mechanics and Number Theory

GEOMETRIC QUANTIZATION

ON JUSTIFICATION OF GIBBS DISTRIBUTION

Topics for the Qualifying Examination

Hamiltonian Dynamics

Qualifying Exam. Aug Part II. Please use blank paper for your work do not write on problems sheets!

Path Integral for Spin

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

Linear and Nonlinear Oscillators (Lecture 2)

Resonance Capture. Alice Quillen. University of Rochester. Isaac Newton Institute Dec 2009

Section B. Electromagnetism

Bifurcations of phase portraits of pendulum with vibrating suspension point

Maslov indices and monodromy

Classical adiabatic angles and quantal adiabatic phase

On the Torus Quantization of Two Anyons with Coulomb Interaction in a Magnetic Field

Chaos in Hamiltonian systems

Semiconductor Physics and Devices

1. Electricity and Magnetism (Fall 1995, Part 1) A metal sphere has a radius R and a charge Q.

J10M.1 - Rod on a Rail (M93M.2)

(a) Write down the total Hamiltonian of this system, including the spin degree of freedom of the electron, but neglecting spin-orbit interactions.

List of Comprehensive Exams Topics

Ergodicity of quantum eigenfunctions in classically chaotic systems

Fluctuations for su(2) from first principles

Introduction to Applied Nonlinear Dynamical Systems and Chaos

Quantum Physics II (8.05) Fall 2002 Outline

FINAL EXAM GROUND RULES

Topology of the Fermi surface wavefunctions and magnetic oscillations in metals

From Last Time. Gravitational forces are apparent at a wide range of scales. Obeys

Chem 467 Supplement to Lecture 19 Hydrogen Atom, Atomic Orbitals

Chaos Indicators. C. Froeschlé, U. Parlitz, E. Lega, M. Guzzo, R. Barrio, P.M. Cincotta, C.M. Giordano, C. Skokos, T. Manos, Z. Sándor, N.

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

Review for Final. elementary mechanics. Lagrangian and Hamiltonian Dynamics. oscillations

Eigenvalues and eigenfunctions of the Laplacian. Andrew Hassell

Lecture 4. Alexey Boyarsky. October 6, 2015

M2A2 Problem Sheet 3 - Hamiltonian Mechanics

Chaotic motion. Phys 420/580 Lecture 10

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

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

Under evolution for a small time δt the area A(t) = q p evolves into an area

ORIGINS. E.P. Wigner, Conference on Neutron Physics by Time of Flight, November 1956

ψ s a ˆn a s b ˆn b ψ Hint: Because the state is spherically symmetric the answer can depend only on the angle between the two directions.

Theoretical physics. Deterministic chaos in classical physics. Martin Scholtz

ORTHOGONAL FUNCTIONS EXACT INVARIANT AND THE ADIABATIC LIMIT FOR TIME DEPENDENT HARMONIC OSCILLATORS

The Kepler Problem and the Isotropic Harmonic Oscillator. M. K. Fung

Single particle motion

arxiv:nlin/ v2 [nlin.cd] 8 Feb 2005

2 Canonical quantization

PHY 407 QUANTUM MECHANICS Fall 05 Problem set 1 Due Sep

Anatoli Polkovnikov Boston University

A Glance at the Standard Map

The general solution of Schrödinger equation in three dimensions (if V does not depend on time) are solutions of time-independent Schrödinger equation

Chapter 29. Quantum Chaos

The kinetic equation (Lecture 11)

Introduction to Theory of Mesoscopic Systems

Chaotic motion. Phys 750 Lecture 9

Hamiltonian Chaos and the standard map

Canonical transformations (Lecture 4)

Physics 139B Solutions to Homework Set 4 Fall 2009

General Exam Part II, Fall 1998 Quantum Mechanics Solutions

The Simple Harmonic Oscillator

QUANTUM MECHANICS. Franz Schwabl. Translated by Ronald Kates. ff Springer

Problems and Multiple Choice Questions

Quantum Physics III (8.06) Spring 2007 FINAL EXAMINATION Monday May 21, 9:00 am You have 3 hours.

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

Second quantization: where quantization and particles come from?

Transient Phenomena in Quantum Bound States Subjected to a Sudden Perturbation

An Exactly Solvable 3 Body Problem

The Adiabatic Invariance of the Action Variable in Classical Dynamics

The general solution of Schrödinger equation in three dimensions (if V does not depend on time) are solutions of time-independent Schrödinger equation

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

Laplace s Method for Ordinary Differential Equations

QUANTUM CHAOS IN NUCLEAR PHYSICS

Complex Numbers. The set of complex numbers can be defined as the set of pairs of real numbers, {(x, y)}, with two operations: (i) addition,

Applied Asymptotic Analysis

Modified Schrödinger equation, its analysis and experimental verification

Physics Mechanics. Lecture 32 Oscillations II

3. Quantum Mechanics in 3D

arxiv: v1 [cond-mat.mes-hall] 26 Sep 2013

Non-linear dynamics Yannis PAPAPHILIPPOU CERN

Towards stability results for planetary problems with more than three bodies

Topological insulator part II: Berry Phase and Topological index

Is Quantum Mechanics Chaotic? Steven Anlage

C. Show your answer in part B agrees with your answer in part A in the limit that the constant c 0.

Hamiltonian aspects of fluid dynamics

Notes on x-ray scattering - M. Le Tacon, B. Keimer (06/2015)

The Harmonic Oscillator: Zero Point Energy and Tunneling

Chapter 2 Ensemble Theory in Statistical Physics: Free Energy Potential

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

Fundamentals of wave kinetic theory

Transcription:

Preprint CAMTP/03-8 August 2003 Theory of Adiabatic Invariants A SOCRATES Lecture Course at the Physics Department, University of Marburg, Germany, February 2004 Marko Robnik CAMTP - Center for Applied Mathematics and Theoretical Physics, University of Maribor, Krekova 2, SI-2000 Maribor, Slovenia E-mail: Robnik@uni-mb.si Abstract. We review the basic theory on adiabatic invariants in classical and quantum mechanics. PACS numbers: 45.30.+s,05.45.-a,03.65.-w, 03.65.Sq Extended abstract Adiabatic invariants, denoted by I, are approximate constants (invariants) of motion of a given dynamical system (not necessarily Hamiltonian), which are approximately preserved (constant) during a process of very slow change of the system s parameters (denoted by λ), on a time scale T, which is supposed to be much larger than any typical dynamical time scale such as traversal time or the period of the shortest periodic orbits etc. This is an asymptotic statement, in the sense that the adiabatic invariants are the better preserved the slower the driving of the system, that is the more slowly the switching function λ = λ(t) varies on the typical evolutionary time scale T, and the preservation is perfect in the limit T. The important point is that whilst the system s parameters λ(t) and their dynamical quantities like the total energy, angular momentum etc, can change by arbitrarily large amounts, their combination involved in the adiabatic invariant I is very well preserved to a very high degree of accuracy, and that allows us to calculate changes of important quantities in dynamical systems. Examples are in celestial mechanics, in other Hamiltonian systems, in motion of charged particles in magnetic and electric fields, etc. The accuracy of preservation can be calculated in one degree of freedom systems and is exponentially good with T if the switching function λ(t) is analytical (of class C ), that is to say the change of the adiabatic invariant I is of the form I = α exp( βt ), (1)

Adiabatic invariants 2 where α and β are certain constants. If, however, the switching function λ(t) is only of class C m (m-times continuously differentiable), then the change of the adiabatic invariant I during an adiabatic change over a time period of length T is algebraic only, namely I = αt (m+1). (2) In both cases I 0 as T. The fact that the evolutionary time scale T is very large compared to the typical shortest dynamical time scales (average return time etc), points to the intuitive idea of the averaging method or the so-called averaging principle. Here it is thought that the long-term evolution (adiabatic evolution) of the system can be calculated by replacing the actual dynamical system with its averaged correspondent, obtained by averaging it over the shortest dynamical time scales, i.e. over the fast variables. Such a procedure is very well known e.g. in celestial mechanics where the secular effects of the thirdbody perturbations of a planet are obtained by averaging the perturbations over one revolutionary period of the perturbers. This has been done already by Gauss in the context of studying the dynamics of planets. The adiabatic invariants can be easily calculated in one dimensional systems, and in completely integrable systems with N degrees of freedom, and something is known about the ergodic Hamiltonian systems, whilst little is known about the adiabatic invariants in mixed type Hamiltonian systems (with divided phase space), where for some initial conditions in the classical phase space we have regular motion on invariant tori and irregular (chaotic) motion for other (complementary) initial conditions. One elementary example is the simple (mathematical) pendulum, of point mass m and of length l with the declination angle ϕ, described by the Hamiltonian H = p2 ϕ mgl cos ϕ (3) 2ml2 where p ϕ = ml 2 ϕ is the angular momentum, and for small oscillations ϕ 1, around the stable equilibrium ϕ = 0, it is described by the harmonic Hamiltonian H = p2 ϕ 2ml + mgl 2 2 ϕ2. (4) Here the angular oscillation frequency is ω = 2πν = g/l, where ν is the frequency, and g is the gravitational acceleration. By E we denote the total energy of the Hamiltonian H. It has been discovered by Ehrenfest, that the quantity I = E/ω is the adiabatic invariant of the system, so the change of E(t) on adiabatic evolutionary very large time scale T 1/ν is such that I = E(t)/ω(t) remains constant. Therefore, if for example the length of the pendulum l = l(t) is slowly, adiabatically, changing, then the energy of the system will change according to the law E = E 0 l0 l, (5) where E 0 and l 0 are the initial and E and l the final values of the two variables. One can easily show that the oscillation amplitude ϕ 0 changes as l 3/4 as the length l

Adiabatic invariants 3 changes. This is an elementary example of a dynamically driven system, in which the change of energy E can be very large, as is the change of ω, but I = E/ω is a well preserved adiabatic invariant, in fact it is exponentially well preserved if the switching function ω(t) is analytic. More generally, for Hamiltonian systems H(q, p, λ), with one degree of freedom, whose state is described by the coordinate q and canonically conjugate momentum p in the phase space (q, p), and λ = λ(t) is the system s parameter, in general slowly changing on time scale T, one can show that the action integral I(E, λ) = I(E(t), λ(t)) = 1 2π pdq (6) is the adiabatic invariant of the system, where the contour integral is taken at fixed total energy E and fixed value of λ. In this case 2πI is interpreted as the area inside the curve E = const. in the phase plane (q, p). The accuracy is exponentially good if λ(t) is an analytic function and algebraic if it is of class C m. And the theorem holds true only if the frequency ω is nonzero. This implies that a passage through a separatrix (in the phase space of a one dimensional system) is excluded, because ω = 0 there, and indeed in such a case of crossing a separatrix a different approach is necessary with highly nontrivial result. When crossing a separatrix of a one-dimensional double potential well from outside in an adiabatic way going inside, a bifurcation takes place, and the capture of the trajectory in either of the two wells is possible with some probabilities. These probabilities can be calculated quite easily, and also the spread of the adiabatic invariant I after such a passage can be calculated, but this is more difficult. The important applications are in celestial mechanics, where an adiabatic capture of a small body near a resonance with a planet can take place, in plasma physics, in quantum mechanics of states close to the separatrix (in the semiclassical limit), etc. This is an interesting result, because I is precisely that quantity which according to the old quantum mechanics of Bohr and Sommerfeld has to be quantized, i.e. put equal to an integer multiple of Planck s constant. Of course, the old quantum mechanics is generally wrong, but it can be a good approximation to the solution of the Schrödinger equation. Even then, strictly speaking, the quantization condition in the sense of EBK or Maslov quantization, must be written in the form I = 1 2π pdq = (n + α ) (7) 4 where n = 0, 1, 2,... is the quantum number and α is the Maslov index, i.e. number of caustics (projection singularities) round the cycle E = const. in the phase plane. For smooth systems with quadratic kinetic energy it is typically α = 2. Thus at this semiclassical level we have the semiclassical adiabatic invariant, saying that in one dimensional systems under an adiabatic change the quantum number (and thus the eigenstate) is preserved. This agrees with the exact result in the theory of the Schrödinger equation in quantum mechanics. Round a closed loop in a parameter space,

Adiabatic invariants 4 a quantum system returns to its original state, except for the phase! (This closed-loop phase change is essentially the so-called Berry s phase.) The method of averaging can be used also in N dimensional Hamiltonians H = H(q, p), where q and p are N-dimensional vectors, but it works only in two extreme cases, the integrable case and the ergodic case. In a classical integrable Hamiltonian system we have N analytic, global and functionally independent constants of motion A i = A i (q, p), i = 1, 2,..., N, pairwise in involution, i.e. all Poisson brackets {A i, A j } vanish identically everywhere in phase space. The orbits in phase space are then confined to an invariant N-dimensional surface, and according to the Liouville-Arnold theorem the topology of these surfaces must be the topology of an N dimensional torus. Then an action integral I = 1 2π p.dq along a closed loop on a torus will be zero if the loop can be continuously shrunk to a point on the torus. But there are loops which cannot be shrunk to a point due to the topology of the torus. Then the integral I is different from zero, but otherwise its value does not depend on the particular loop, so in a sense it is a topological invariant of the torus. On an N dimensional torus there are N such independent elementary closed loops C i, i = 1, 2,..., N. The integrals which we call simply actions or action variables I i = 1 2π C i p.dq (8) are then the most natural momentum variables on the torus, whilst angle variables Θ specifying the position on the torus labeled by I can be generated from the transformation Θ = S(I, q), (9) I where S = p.dq is an action integral on the torus. Applying the averaging principle (the method of averaging) one readily shows that for an integrable system the actions I are N adiabatic invariants, provided the system is nondegenerate, which means that the frequencies ω = H I on the given torus are not rationally connected, that is to say there is no such integer vector k that ω.k = 0. The problem is that during an adiabatic process the frequencies ω will change and therefore strictly speaking there will be infinitely many points of λ = λ(t), where ω.k = 0, which will, strictly speaking, invalidate the theorem. However, it is thought that if the degree of resonances or rationality conditions ω.k = 0 is of very high order, meaning that all components of k are very large, that then the adiabatic invariants I i will be quite well preserved. But low order resonances (rationality conditions) must be excluded. The details of such a process still call for further investigation. When the N actions I i of an integrable system are quantized in the sense of Maslov, as explained before in the one-dimensional case, we again find the agreement, at this (10)

Adiabatic invariants 5 semiclassical level, with the quantum mechanics: In a family of integrable systems all N quantum numbers and the corresponding eigenstates are preserved under an adiabatic change. Another extreme of classically ergodic and thus fully chaotic systems has been considered already by Hertz. He found that in such ergodic Hamiltonian systems the phase space volume enclosed by the energy surface H(q, p) = E = constant, is the adiabatic invariant, denoted by Ω(E) = d N qd N p. (11) H(q,p) E Of course, here it is required that while system s parameter λ(t) is slowly changing the system itself must be ergodic for all λ(t). Sometimes this condition is difficult to satisfy, but sometimes it is easily fulfilled. Examples are the stadium of Bunimovich with varying length of the straight line between the two semicircles, or Sinai billiard with varying radius of the circle inside a square. For an ergodic two dimensional billiard of area A and point mass m, we have Ω(E) = 2πmAE. (12) Therefore when A is adiabatically changing, the energy E of the billiard particle is changing reciprocally with A. Diminishing A implies increasing E, and this can be interpreted as work being done against the pressure of only one particle, if we define the pressure as the time average of the momentum transfer at collisions with the boundary of our ergodic billiard. In fact, there is a formalism to proceed with this analysis close to the thermodynamic formalism, as derived from the statistical mechanics, except that here we are talking about the time averages rather than phase averages of classical variables. Again, this general result for ergodic systems is interesting from the quantal point of view, because N = Ω(E)/(2π ) N is precisely the number of energy levels below the energy E, in the semiclassical limit of very large N, which is well known as the Thomas-Fermi rule. It is the number of elementary quantal Planck cells inside the volume element H(q, p) E. Indeed, quantum mechanically, the eigenstate and the (energy counting sequential) main quantum number N is preserved under an adiabatic change. In case of a mixed type Hamiltonian system, which is a typical case in nature, the adiabatic theory is in its infancy. Moreover, in three or higher degrees of freedom, we have the universal diffusion on Arnold web, which is dense on the energy surface, even for KAM type Hamiltonian systems which are very close to the integrability, like our solar planetary system. On the Arnold web we have a diffusional chaotic motion, and there is a rigorous theory by Nekhoroshev giving a rigorous upper bound to the diffusion rate in such case. However, when compared with the numerical calculations, it is found that the difussion rate is many orders of magnitude smaller than the Nekhoroshev bound. In other words, the actual diffusion time is much longer than estimated by Nekhoroshev,

Adiabatic invariants 6 implying that there we have some approximate adiabatic invariant for long times, but not too long times. Further Reading Arnold, V.I. 1980. Mathematical Methods of Classical Mechanics, New York and Heidelberg: Springer-Verlag Cary, J.R. & Rusu, P. 1992. Separatrix Eigenfunctions Physical Review A 45: 8501-8512 Cary, J.R. & Rusu, P. 1993. Quantum Dyanmics near a Classical Separatrix Physical Review A 47: 2496-2505 Landau, L.D. & Lifshitz, E.M. 1996 Mechanics - Course of Theoretical Physics Volume 1 Exeter, Great Britain: Butterworth-Heinemann Landau, L.D. & Lifshitz, E.M. 1997 Quantum Mechanics (Non-Relativistic Theory)- Course of Theoretical Physics Volume 3 Exeter, Great Britain: Butterworth-Heinemann Lichtenberg, A.J. & Lieberman, M.A. 1992. Regular and Chaotic Dynamics - Second Edition New York and Heidelberg: Springer-Verlag Lochak, P. & Meunier, C. 1988. Multiphase Averaging for Classical Systems New York and Heidelberg: Springer-Verlag Reinhardt, W.P. 1994. Regular and Irregular Correspondences Progress of Theoretical Physics Supplement 116: 179-205