Quantum Chaos: An Exploration of the Stadium Billiard Using Finite Differences

Similar documents
Is Quantum Mechanics Chaotic? Steven Anlage

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

Homoclinic and Heteroclinic Motions in Quantum Dynamics

Stadium Billiards. Charles Kankelborg. March 11, 2016

Chapter 29. Quantum Chaos

Lecture-XXVI. Time-Independent Schrodinger Equation

Chaos, Quantum Mechanics and Number Theory

Lecture 2: simple QM problems

Ergodicity of quantum eigenfunctions in classically chaotic systems

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

Quantum mechanics is a physical science dealing with the behavior of matter and energy on the scale of atoms and subatomic particles / waves.

TitleQuantum Chaos in Generic Systems.

Quantum Mechanics: Particles in Potentials

QUANTUM CHAOS IN NUCLEAR PHYSICS

If electrons moved in simple orbits, p and x could be determined, but this violates the Heisenberg Uncertainty Principle.

CHAPTER 8 The Quantum Theory of Motion

Appendix A. The Particle in a Box: A Demonstration of Quantum Mechanical Principles for a Simple, One-Dimensional, One-Electron Model System

Sample Quantum Chemistry Exam 2 Solutions

Lecture 6. Four postulates of quantum mechanics. The eigenvalue equation. Momentum and energy operators. Dirac delta function. Expectation values

Introduction to Theory of Mesoscopic Systems

2. As we shall see, we choose to write in terms of σ x because ( X ) 2 = σ 2 x.

1. For the case of the harmonic oscillator, the potential energy is quadratic and hence the total Hamiltonian looks like: d 2 H = h2

An operator is a transformation that takes a function as an input and produces another function (usually).

The Postulates of Quantum Mechanics Common operators in QM: Potential Energy. Often depends on position operator: Kinetic Energy 1-D case: 3-D case

Lecture #5: Begin Quantum Mechanics: Free Particle and Particle in a 1D Box

Particle in one-dimensional box

Opinions on quantum mechanics. CHAPTER 6 Quantum Mechanics II. 6.1: The Schrödinger Wave Equation. Normalization and Probability

The Schrodinger Equation and Postulates Common operators in QM: Potential Energy. Often depends on position operator: Kinetic Energy 1-D case:

Introduction to Quantum Mechanics (Prelude to Nuclear Shell Model) Heisenberg Uncertainty Principle In the microscopic world,

The Simple Harmonic Oscillator

The Schrodinger Wave Equation (Engel 2.4) In QM, the behavior of a particle is described by its wave function Ψ(x,t) which we get by solving:

= X = X ( ~) } ( ) ( ) On the other hand, when the Hamiltonian acts on ( ) one finds that

Physics 221A Fall 2017 Notes 27 The Variational Method

Spin Dynamics Basic Theory Operators. Richard Green SBD Research Group Department of Chemistry

Lecture #13 1. Incorporating a vector potential into the Hamiltonian 2. Spin postulates 3. Description of spin states 4. Identical particles in

The Transition to Chaos

Quantum Mechanics for Scientists and Engineers. David Miller

Experimental and theoretical aspects of quantum chaos

Potentially useful reading Sakurai and Napolitano, sections 1.5 (Rotation), Schumacher & Westmoreland chapter 2

Harmonic Oscillator I

Physics 221A Fall 1996 Notes 16 Bloch s Theorem and Band Structure in One Dimension

Chemistry 3502/4502. Final Exam Part I. May 14, 2005

C/CS/Phys C191 Particle-in-a-box, Spin 10/02/08 Fall 2008 Lecture 11

Section 9 Variational Method. Page 492

Mathematical Tripos Part IB Michaelmas Term Example Sheet 1. Values of some physical constants are given on the supplementary sheet

Quantum Mechanics in One Dimension. Solutions of Selected Problems

Chemistry 3502/4502. Final Exam Part I. May 14, 2005

The Sommerfeld Polynomial Method: Harmonic Oscillator Example

chmy361 Lec42 Tue 29nov16

Vibrational motion. Harmonic oscillator ( 諧諧諧 ) - A particle undergoes harmonic motion. Parabolic ( 拋物線 ) (8.21) d 2 (8.23)

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

Outline 1. Real and complex p orbitals (and for any l > 0 orbital) 2. Dirac Notation :Symbolic vs shorthand Hilbert Space Vectors,

Thermalization in Quantum Systems

Notes for Special Relativity, Quantum Mechanics, and Nuclear Physics

Particle in a 3 Dimensional Box just extending our model from 1D to 3D

PHYS 3313 Section 001 Lecture #20

Universality for random matrices and log-gases

COMP 558 lecture 18 Nov. 15, 2010

Problem Set 5: Solutions

Chem 452 Mega Practice Exam 1

The Quantum Theory of Atoms and Molecules

Atkins & de Paula: Atkins Physical Chemistry 9e Checklist of key ideas. Chapter 7: Quantum Theory: Introduction and Principles

Quantum Theory and Group Representations

Lecture 5: Harmonic oscillator, Morse Oscillator, 1D Rigid Rotor

Final Exam. Tuesday, May 8, Starting at 8:30 a.m., Hoyt Hall.

Dynamical Systems and Chaos Part I: Theoretical Techniques. Lecture 4: Discrete systems + Chaos. Ilya Potapov Mathematics Department, TUT Room TD325

The Birth of Quantum Mechanics. New Wave Rock Stars

(Refer Slide Time: 1:20) (Refer Slide Time: 1:24 min)

There is light at the end of the tunnel. -- proverb. The light at the end of the tunnel is just the light of an oncoming train. --R.

... it may happen that small differences in the initial conditions produce very great ones in the final phenomena. Henri Poincaré

Chem 467 Supplement to Lecture 19 Hydrogen Atom, Atomic Orbitals

Page 404. Lecture 22: Simple Harmonic Oscillator: Energy Basis Date Given: 2008/11/19 Date Revised: 2008/11/19

Eigenvalue statistics and lattice points

Spacing distributions for rhombus billiards

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

Chem 3502/4502 Physical Chemistry II (Quantum Mechanics) 3 Credits Fall Semester 2006 Christopher J. Cramer. Lecture 5, January 27, 2006

Page 684. Lecture 40: Coordinate Transformations: Time Transformations Date Revised: 2009/02/02 Date Given: 2009/02/02

Random Matrix: From Wigner to Quantum Chaos

Continuous quantum states, Particle on a line and Uncertainty relations

PHYSICS 721/821 - Spring Semester ODU. Graduate Quantum Mechanics II Midterm Exam - Solution

We also find the development of famous Schrodinger equation to describe the quantization of energy levels of atoms.

Chm 331 Fall 2015, Exercise Set 4 NMR Review Problems

Applied Nuclear Physics (Fall 2006) Lecture 3 (9/13/06) Bound States in One Dimensional Systems Particle in a Square Well

Hydrogen atom energies. Friday Honors lecture. Quantum Particle in a box. Classical vs Quantum. Quantum version

The Schrödinger Equation

Quantum mechanics (QM) deals with systems on atomic scale level, whose behaviours cannot be described by classical mechanics.

Homework Problem Set 3 Solutions

Lecture 4 (19/10/2012)

What is quantum unique ergodicity?

6. Qualitative Solutions of the TISE

Review of Quantum Mechanics, cont.

H ψ = E ψ. Introduction to Exact Diagonalization. Andreas Läuchli, New states of quantum matter MPI für Physik komplexer Systeme - Dresden

Arithmetic quantum chaos and random wave conjecture. 9th Mathematical Physics Meeting. Goran Djankovi

TP computing lab - Integrating the 1D stationary Schrödinger eq

Problems and Multiple Choice Questions

Quantum chaos analysis of the ideal interchange spectrum in a stellarator

Physical Chemistry Quantum Mechanics, Spectroscopy, and Molecular Interactions. Solutions Manual. by Andrew Cooksy

CHAPTER 6 Quantum Mechanics II

PHY 407 QUANTUM MECHANICS Fall 05 Problem set 1 Due Sep


Transcription:

Quantum Chaos: An Exploration of the Stadium Billiard Using Finite Differences Kyle Konrad & Dhrubo Jyoti Math 53: Chaos! Professor Alex Barnett Dartmouth College December 4, 2009 Abstract We investigate quantum chaos in chaotic billiards by modelling the square (non-chaotic) and the stadium (chaotic) billiards as 2D infinite square wells. We developed MATLAB code that uses grid points and the method of finite differences to numerically solve the Schrödinger equation for either case. We successfully obtained the scar structures in higher energy eigenfunctions for the stadium case, discovered by Joseph Heller in 1984. We also studied the eigenvalue spacings, and obtained the Poisson distribution for the square case and the GOE distribution for the chaotic case. Thus, we were able to demonstrate both features (scarring in eigenfunctions and GOE distribution in eigenvalue separation) that indicate the presence of chaos in the classical limit. Background For more than 80 years, predictions from quantum mechanics (QM) have been validated in laboratory experiments with infallible accuracy. QM is, to date, the most accurate description of nature in the limit where relativistic effects can be ignored. On the other hand, classical mechanics as formulated by Newton, Lagrange, Hamilton and others give accurate results for less than exotic, every-day length scales and energies (known as the classical limit), until we go much smaller and reach the molecular level where quantum effects become important. We expect the more general and accurate theory, QM, to reduce to the less general, classical mechanics, in the classical limit. This is Bohr s correspondence principle. The phenomenon of deterministic chaos was discovered by Henri Poincaré in the late 18 th century when he was studying the seemingly erratic orbit of the moon in the 3-body system of Sun-Earth-Moon. The classical context in which chaos was discovered and also continued to study until late into the 20 th century naturally meant the definition of chaos in a dynamical

system was cast exclusively in classical terms: chaotic systems are those that have orbits that (i) are not asymptotically periodic and (ii) have positive Lyapunov exponent(s). Note how chaos is formulated entirely in terms of orbits, i.e. the trajectory of a particle. But in QM, the particle does not have a definite position due to the Heisenberg Uncertainty Principle (in any finite potential, a particle has a finite likelihood to be anywhere at any moment in time due to quantum tunneling). Hence, the definition of chaos in classical mechanics does not carry over to QM. It is not even obvious what chaos would entail in the quantum world, if anything. What kind of behavior would a classically chaotic system produce if studied in the quantum regime? Also, from the other logical direction, would it be possible to look at quantum mechanical results (i.e. eigenfunctions and eigenvalues) of a system and deduce whether or not that system is classically chaotic? Intuitively, one might expect that since most orbits of a chaotic system are dense in the phase space, the quantum eigenfunction would also be broadly, thinly and randomly smeared out and show no significant localization. Surprisingly,however, in many of the eigenfunctions corresponding to large eigenvalues, we see pockets of high probability densities in the coordinate space, often forming ornate patterns with various symmetries but no discernible similarity or connection with other eigenfunctions. Even more curiously, for some of these eigenfunctions, we can identify the pockets as lying along some of the unstable periodic orbits of the corresponding classical system. For this last subset of eigenfunctions, the pockets around the unstable orbits are known as scars. One-Dimensional Square Infinte Potential Well We began by solving the Schrodinger equation for the one-dimensional squre infinite potential well defined by the potential function V x = 0 if 0 x 1 otherwise and compared our numerical results with the exact results. Doing so allowed us to develop and test our method of finite differences code in a case much simpler than the stadium. In computing solutions we began with the time-independent Schrödinger equation, ħ 2 2m 2 Ψ + VΨ = EΨ, where 2 = d2 dx 2. V = 0 everywhere inside the well, and we take ħ2 simplifies to the Helmholtz equation 2m to be unity so the Schrodinger equation

2 Ψ = EΨ We approximated 2 as a matrix A that acts on a set of grid points. Using secant line approximations to the derivative we obtain d 2 f dx 2 x n df df dx xn dx +1/2 xn 1/2 h u n+1 u n h u n u n 1 h h = u n+1 2u n + u n 1 h 2 where h = x n x n-1 is the lattice spacing and u i = Ψ(x i ) as illustrated in Figure 1. Figure 1: Illustration of quantities used in the method of finite differences. Hence, the matrix A takes the form 2 A = 2 0 1 2 1 0 2 Computed solutions match theoretical solutions Ψ n = sin nπx, E n = n 2, n Z + very well, with errors scaling as k 2, the scale factor depending on resolution. Deeper analysis of error scaling is reserved for the two-dimensional case.

Figure 2: Calculated (green) and theoretical (red) wavefunctions of one-dimensional square infinite potential well for N=2000 degrees of freedom. Figure 3: Eigenvalue errors for one-dimensional square infinte potential well for N=2000 degrees of freedom. Two-Dimensional Square Infinite Potential Well Having successfully computed eigenfunctions and eigenvalues of the one-dimensional square infinite potential well we move on to the two-dimensional case with potential given by V x = 0 if 0 x 1 and 0 y 1 otherwise The two-dimensional Laplacian 2 = 2 x 2 + 2 y 2 is simply the sum of two one dimensional Laplacians. The matrix approximation A is therefore the sum of two matrices, one identical to the one-dimensional matrix and one with off diagonal elements a distance M+2 (rather than one) from the diagonal. 2 A = M + 1 0 1 0 1 4 1 0 1 0 M + 1 Using this matrix to find eigenfunctions requires us to represent two-dimensional wavefunctions as vectors. To accomplish this we discretize the two-dimensional square in the same fashion as the one-dimensional case but transform the two-dimensional grid into a vector

by stacking columns of the grid. This is the reason for the M+2 spacing (M+2 for M degrees of freedom plus boundary points on either side) of off-diagonal entries from the 2 y 2 term. Wavefuntions of the two-dimensional square infinite potential well can be found using separation of variables and are therefore products of wavefunctions of the one-dimensional square infinite potential well. Ψ nx,n y = sin n x πx sin n y πy, n x, n y Z + Probability distributions of these wavefunctions are periodic and symmetric, hence relatively simple and show little localization a feature common to all integrable systems. Figure 2: Probability distributions of two-dimensional square infinte potential well. Energy levels of the two-dimensional square infinite potential well are sums of the onedimensional values. E = n x 2 + n y 2 Knowing these exact values we can see how the computed solutions compare. A plot of energy level errors is shown in Figure 3.

Figure 3: Energy level errors for two-dimensional square infinite potential well with 90,000 degrees of freedom. Discarding the spike at the end (which is due to inaccuracies in the Matlab function eigs) we see that error scales linearly with energy, with the specific scale factor depending on resolution. This makes sense because in an area with zero potential the energy operator is equal to the kinetic energy operator which is proportional to the second derivative of the wave function i.e., the curvature. High curvature means the wavefunction is changing sharply over space and the discretization will fail to accurately capture these fine variations. We might expect the error scaling factor to scale linearly with degrees of freedom i.e., as N 2 where N is the degrees of freedom in either the x- or y-direction. Running a numerical analysis (using the Matlab function polyfit) we see we actually do slightly better: errors scales approximately as N 1.9. Based on these results we can expect highly accurate results for the stadium billiard for several hundred energy levels using NM 100,000 degrees of freedom. Indexing Stadium Lattice Nodes The key numerical difference between the two-dimensional square infinite potential well and the stadium billiard is in the entires of the matrix A, which effectively encodes boundary

conditions constraining solutions of the Helmholtz equation. To encode the boundary condition we begin by traversing lattice points of the circumscribed rectangle of the stadium giving each point that falls within the stadium an index number which is stored in a map. We also create an inverse map from index numbers to points. Now we can create A by traversing points in the index map and setting diagonals to -4 and entries corresponding to neighboring points within the stadium to 1. The neighboring points we test for the point (x k, y j ) are: east, (x k+1, y j ); west, (x k-1, y j ); north, (x k, y j+1 ); and south, (x k, y j- 1) again using our map to find the appropriate index for these points in rows of A. Stadium Enery Levels Transitioning from an integrable to nonintegrable system, we expect to see distinct changes in energy level distribution. Stadium energy levels should not follow a neat distribution but rather should reflect the classical chaos present in the nonintegrable system by following a pseudorandom distribution. Literature on the subject has established that energy level spacings follow the Wigner distribution, the same distribution that is obeyed by spacings between eigenvalues of matrices in the Gaussian Orthogonal Ensemble (GOE) [3]. A GOE matrix consists of Gaussian elements with standard deviation 1 off the diagonal and 2 on the diagonal and mean 0 in both cases. A GOE matrix can be formed by taking an n-by-n matrix R with all elements Gaussians of mean 0 and standard deviation 1 and forming W given by W = (R + R T )/ 2n [4] As n approaches infinity the distriubtion of spacings between eigenvalues of W approach the Wigner-Dyson distribution W s = π 2 se π 4 s2 which falls to 0 at s = 0. In the context of energy levels of the stadium billiard this means that there are very few energy levels with close to the same energy. This contrasts with the integrable case where many energy levels are close together and spacings follow a Poisson distribution. [3]

Our results showed a clear Poisson distribution for the eigenvalues spacings of the twodimensional square infinite potential well. To achieve an accurate Wigner-Dyson distrubtion for stadium eigenvalue spacings we plot only eigenvalues of wavefuntions that have odd-odd parity (i.e., antisymmetric through both axes), a technique known as desymmetrization originally utilized by McDonald and Kaufman [5]. To circumvent memory and time contraints, rather than finding many eigenvalues of a single billiard we parameterize the billiard by length in the x- direction and sum eigenvalue spacing distributions over many values of this parameter. Figure 4: Eigenvalue spacings for two-dimensional square infinite potential well (left) and stadium (right). Stadium Wave Functions Wavefunctions of the stadium billiard exhibit a surprising amount of localization. Looking at trajectories of the classical case, one might expect wavefunctions to be relatively unlocalized, as classical trajectories appear to fill the stadium relatively uniformly with no discernable patterns. There are of course certain classical trajectories in the stadium that stand out, namely low-period periodic orbits. These orbits are of course unstable as the stadium is a chaotic billiard, so in general they appear very infrequently for arbitrary starting conditions. Nonetheless these unstable periodic orbits are manifest in the quantum wavefunctions of the stadium billiard as scars, a term coined by Eric Heller who has shown the presence of these scars to follow from basic principles.[6] It should be made clear that scars corresponding to unstable period orbits are not visible in all wavefunctions but do appear fairly regularly (Heller claims about half the states have one or more recognizable scars, though we saw a proportion closer to a quarter) in high energy solutions. Furthermore, even in those wavefunctions that do not display scars corresponding to unstable period orbits, there is almost always a relatively high degree of localization. Typical scar patterns are shown in Figure 5.

0.5 0-0.5-1 -0.5 0 0.5 1 0.5 0-0.5-1 -0.5 0 0.5 1 Figure 5: Scars in the stadium billiard corresponding to classical period orbits of period 6 (top) and period 3 (bottom). Conclusion We have obtained results in strong agreement with previous research in the field, demonstrating the connection between classical and quantum chaos. Primary manifestations of this connection are seen in energy level spacing distributions and scarring in wave functions. In contrasting integrable and nonintegrable billiards, we have demonstrated the specific routes by which chaos is manifest is quantum systems. In a future project, we could explore how the eigenfunctions and eigenvalues change as we slowly vary the stadium parameter, changing the shape from disk to elongated stadium. Furthermore, it would be interesting to explore other chaotic (e.g. Sinai billiards) and nonchaotic (e.g. disk) billiard shapes in order to establish the generality of the features of the representative systems that have been analyzed here.

References [1] Gutzweiller, Martin Quantum Chaos. Scientific American. Retrieved Dec 4, 09. http://www.scientificamerican.com/article.cfm?id=quantum-chaos-subatomic-worlds [2] Ott, Edward. Chaos in Dynamical Systems. 2 nd Ed. Cambride Univ. Press, 2002. [3] Rudnick, Ze ev. What is Quantum Chaos? AMS Notices. Jan. 2008, Vol. 55, Issue 1. 32-34. [4] Edelman, Alan. The Eigenvalues of Random Matrices: Experiments with the Classical Ensembles MIT. Sept, 2005. [5] McDonald, Steven and Allan Kaufman. Spectrum and Eigenfunction for a Hamiltonian with Stochastic Trajectories. Phys. Rev. Let. Apr. 1979, Vol. 42, Issue 18. 1189-1191. [6] Heller, Eric. Bound-State Eigenfunctions of Classically Chaotic Hamiltonian Systems: Scars of Periodic Orbits. Phys. Rev. Let. Oct. 1984, Vol. 63, Issue 16. 1515-1518.