and many of the diverse methods for producing approximate eigenfunctions have come to be labelled \quasimode" constructions. For a given potential V (

Similar documents
A Minimal Uncertainty Product for One Dimensional Semiclassical Wave Packets

The Time{Dependent Born{Oppenheimer Approximation and Non{Adiabatic Transitions

Molecular propagation through crossings and avoided crossings of electron energy levels

Magnetic wells in dimension three

Representations of Sp(6,R) and SU(3) carried by homogeneous polynomials

QUASI-UNIFORMLY POSITIVE OPERATORS IN KREIN SPACE. Denitizable operators in Krein spaces have spectral properties similar to those

Linear Algebra (part 1) : Vector Spaces (by Evan Dummit, 2017, v. 1.07) 1.1 The Formal Denition of a Vector Space

Generators for Continuous Coordinate Transformations

Congurations of periodic orbits for equations with delayed positive feedback

Ordinary Differential Equation Introduction and Preliminaries

Generating Function and a Rodrigues Formula for the Polynomials in d Dimensional Semiclassical Wave Packets

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

Counterexamples to witness conjectures. Notations Let E be the set of admissible constant expressions built up from Z; + ;?;;/; exp and log. Here a co

Non Adiabatic Transitions in a Simple Born Oppenheimer Scattering System

1 Lyapunov theory of stability

PARAMETER IDENTIFICATION IN THE FREQUENCY DOMAIN. H.T. Banks and Yun Wang. Center for Research in Scientic Computation

Bohr Sommerfeld Quantization Condition Derived by a Microlocal WKB Method

Applied Asymptotic Analysis

Semi-Classical Dynamics Using Hagedorn Wavepackets

Exponentially Accurate Semiclassical Tunneling Wave Functions in One Dimension

The Erwin Schrodinger International Boltzmanngasse 9. Institute for Mathematical Physics A-1090 Wien, Austria

Physics 70007, Fall 2009 Answers to Final Exam

Waves on 2 and 3 dimensional domains

Vector Space Basics. 1 Abstract Vector Spaces. 1. (commutativity of vector addition) u + v = v + u. 2. (associativity of vector addition)

Integral-free Wigner functions

ON TRIVIAL GRADIENT YOUNG MEASURES BAISHENG YAN Abstract. We give a condition on a closed set K of real nm matrices which ensures that any W 1 p -grad

An inverse scattering problem for short-range systems in a time-periodic electric field. François Nicoleau

Lecture 2: Review of Prerequisites. Table of contents

Introduction to Spectral Theory

THE PRINCIPLE OF LIMITING ABSORPTION FOR THE NON-SELFADJOINT. Hideo Nakazawa. Department of Mathematics, Tokyo Metropolitan University

Math Ordinary Differential Equations

4.1 Eigenvalues, Eigenvectors, and The Characteristic Polynomial

B. Physical Observables Physical observables are represented by linear, hermitian operators that act on the vectors of the Hilbert space. If A is such

Complex WKB analysis of energy-level degeneracies of non-hermitian Hamiltonians

The Erwin Schrodinger International Pasteurgasse 6/7. Institute for Mathematical Physics A-1090 Wien, Austria

Molecular Resonance Raman and Rayleigh Scattering Stimulated by a Short Laser Pulse

QM and Angular Momentum

Garrett: `Bernstein's analytic continuation of complex powers' 2 Let f be a polynomial in x 1 ; : : : ; x n with real coecients. For complex s, let f

1 Introduction It will be convenient to use the inx operators a b and a b to stand for maximum (least upper bound) and minimum (greatest lower bound)

2 Section 2 However, in order to apply the above idea, we will need to allow non standard intervals ('; ) in the proof. More precisely, ' and may gene

Solutions to the Hamilton-Jacobi equation as Lagrangian submanifolds

Representation theory and quantum mechanics tutorial Spin and the hydrogen atom

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

CANONICAL FORMS FOR LINEAR TRANSFORMATIONS AND MATRICES. D. Katz

The Sommerfeld Polynomial Method: Harmonic Oscillator Example

Computations of Critical Groups at a Degenerate Critical Point for Strongly Indefinite Functionals

2 Garrett: `A Good Spectral Theorem' 1. von Neumann algebras, density theorem The commutant of a subring S of a ring R is S 0 = fr 2 R : rs = sr; 8s 2

Simple Harmonic Oscillator

UNCERTAINTY PRINCIPLES FOR THE FOCK SPACE

2 EBERHARD BECKER ET AL. has a real root. Thus our problem can be reduced to the problem of deciding whether or not a polynomial in one more variable

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

view that the mathematics of quantum mechanics (in particular a positive denite scalar product) emerges only at an approximate level in a semiclassica

2 G. D. DASKALOPOULOS AND R. A. WENTWORTH general, is not true. Thus, unlike the case of divisors, there are situations where k?1 0 and W k?1 = ;. r;d

An introduction to Birkhoff normal form

Lecture 4: Numerical solution of ordinary differential equations

1 Mathematical preliminaries

Comm. Korean Math. Soc. 13(1998), No. 1, pp SEMI-QUASITRIANGULARITY OF TOEPLITZ OPERATORS WITH QUASICONTINUOUS SYMBOLS In Hyoun Kim and Woo You

Quantum mechanics. Chapter The quantum mechanical formalism

A Time Splitting for the Semiclassical Schrödinger Equation

BASIC MATRIX PERTURBATION THEORY

REMARKS ON THE TIME-OPTIMAL CONTROL OF A CLASS OF HAMILTONIAN SYSTEMS. Eduardo D. Sontag. SYCON - Rutgers Center for Systems and Control

On the Resolvent Estimates of some Evolution Equations and Applications

Taylor and Laurent Series

in order to insure that the Liouville equation for f(?; t) is still valid. These equations of motion will give rise to a distribution function f(?; t)

Abstract. Jacobi curves are far going generalizations of the spaces of \Jacobi

Wave equation on manifolds and finite speed of propagation

the set of critical points of is discrete and so is the set of critical values of. Suppose that a 1 < b 1 < c < a 2 < b 2 and let c be the only critic

The effect of Group Velocity in the numerical analysis of control problems for the wave equation

Eigenvalues and eigenfunctions of the Laplacian. Andrew Hassell

THE INVERSE FUNCTION THEOREM FOR LIPSCHITZ MAPS

MATH 205C: STATIONARY PHASE LEMMA

Spurious Chaotic Solutions of Dierential. Equations. Sigitas Keras. September Department of Applied Mathematics and Theoretical Physics

Math 210B. Artin Rees and completions

Lecture 3 Dynamics 29

Dynamics and Quantum Channels

Problem Set No. 3: Canonical Quantization Due Date: Wednesday October 19, 2018, 5:00 pm. 1 Spin waves in a quantum Heisenberg antiferromagnet

only nite eigenvalues. This is an extension of earlier results from [2]. Then we concentrate on the Riccati equation appearing in H 2 and linear quadr

114 EUROPHYSICS LETTERS i) We put the question of the expansion over the set in connection with the Schrodinger operator under consideration (an accur

Non Adiabatic Transitions near Avoided Crossings: Theory and Numerics

C*-algebras, composition operators and dynamics

and the nite horizon cost index with the nite terminal weighting matrix F > : N?1 X J(z r ; u; w) = [z(n)? z r (N)] T F [z(n)? z r (N)] + t= [kz? z r

Least Sparsity of p-norm based Optimization Problems with p > 1

Entrance Exam, Differential Equations April, (Solve exactly 6 out of the 8 problems) y + 2y + y cos(x 2 y) = 0, y(0) = 2, y (0) = 4.

arxiv: v7 [quant-ph] 22 Aug 2017

EIGENVALUES AND EIGENVECTORS 3

A Time Dependent Born Oppenheimer Approximation with Exponentially Small Error Estimates

THE GUTZWILLER TRACE FORMULA TORONTO 1/14 1/17, 2008

Chapter 2 The Group U(1) and its Representations

Contents. 2.1 Vectors in R n. Linear Algebra (part 2) : Vector Spaces (by Evan Dummit, 2017, v. 2.50) 2 Vector Spaces

Computation Of Asymptotic Distribution. For Semiparametric GMM Estimators. Hidehiko Ichimura. Graduate School of Public Policy

1 Holomorphic functions

Dynamical Systems & Lyapunov Stability

We denote the space of distributions on Ω by D ( Ω) 2.

Notes on symplectic geometry and group actions

Special Functions of Mathematical Physics

E.G. KALNINS AND WILLARD MILLER, JR. The notation used for -series and -integrals in this paper follows that of Gasper and Rahman [3].. A generalizati

SPECTRAL PROPERTIES OF THE LAPLACIAN ON BOUNDED DOMAINS

Bohr-Sommerfeld rules to all orders

Project: Vibration of Diatomic Molecules

PY 351 Modern Physics - Lecture notes, 3

Transcription:

Bohr{Sommerfeld Quantization Rules in the Semiclassical Limit George A. Hagedorn Department of Mathematics and Center for Statistical Mechanics and Mathematical Physics Virginia Polytechnic Institute and State University Blacksburg, Virginia 2461-123 Sam L. Robinson Department of Mathematics The William Paterson University of New Jersey Wayne, New Jersey 747 Abstract We study one-dimensional quantum mechanical systems in the semiclassical limit. We construct a lowest order quasimode (~) for the Hamiltonian H(~) when the energy E and Planck's constant ~ satisfy the appropriate Bohr-Sommerfeld conditions. This means that (~) is an approximate solution of the Schrodinger equation in the sense that k[h(~) E] (~)k C~ 3=2 k (~)k. It follows that H(~) has some spectrum within a distance C~ 3=2 of E. Although the result has a long history, our time-dependent construction technique is novel and elementary. 1 Introduction In this paper we construct approximate eigenfunctions for the one{dimensional, time{independent Schrodinger equation in the semiclassical limit. The history of such constructions is as rich as that of the Schrodinger equation itself, 1

and many of the diverse methods for producing approximate eigenfunctions have come to be labelled \quasimode" constructions. For a given potential V (x) and the associated Schrodinger operator H(~) = ~2 + V (x), we 2 dx2 seek a quasi{energy E(~) in R and a quasimode (~; x) in L 2 (R; dx) that satisfy k (~; )k L 2 (R) = O(1) as ~ & and k[h(~) E(~)] (~; )k L 2 (R) C~ (1) for > 1 and some sequence of positive values of ~ converging to zero. This implies that H(~) has spectrum in the interval [E(~) C~ ; E(~) C~ ]. Under the hypotheses we assume, the spacing between eigenvalues of H(~) is O(~), so our construction yields non-trivial information. We present a novel quasimode construction based on time{dependent methods. The Bohr{Sommerfeld quantization rules arise as a sucient condition under which our approximation holds. Our techniques require little more than some functional analysis, a little ODE theory, and some L 2 estimates. The construction of quasimodes using coherent states has been studied by others (see [1], [3], [8], and the references therein). Some historical comments and multidimensional results can be found in [13] and [11]. We note that assumptions made in some papers, such as [13] and [11], are never satised by non{trivial one{dimensional systems. Near the completion of this work we learned of the unpublished doctoral thesis of Khuat-Duy [9] which contains ideas similar to those underlying our proof, though the techniques and details of the presentation are dierent. We thank Professors Paul and Uribe for bringing this result to our attention. The basic idea of this paper is to construct quasimodes of the form (~; x) = C(~) 1=4 (E) ~ it(e+ e (E) )=~ e is(t)=~ ' (A(t); B(t); ~; a(t); (t); x) ; (2) where E satises the Bohr{Sommerfeld conditions. The quantities A(t), B(t), a(t), (t), and S(t) are determined by classical mechanics, and e is(t)=~ ' (A(t); B(t); ~; a(t); (t); x) is an approximate solution to a time{dependent Schrodinger equation that is dened below. The Bohr{Sommerfeld conditions arise in a simple, intuitive fashion as conditions on the phase of the time{dependent wave function as the system propagates around a classical periodic orbit. 2 d 2

1.1 Some notation and denitions We handle a rather large class of potentials. Our assumptions on the potential V (x) are: (V1) V 2 C 5 (R), (V2) V is bounded from below by a constant, (V3) jv (x)j Ce Mx2 for some constants C and M, (V4) V = lim V (x) 2 R [ f1g. x!1 Under assumptions (V1 - V2) the Hamiltonian H(~) is essentially selfadjoint on C 1 (R) L 2 (R). The degree of smoothness in (V1) and the growth condition in (V3) facilitate some estimates that arise in our proof. Assumption (V4) serves to simplify the spectral information we can extract from the quasimodes. We restrict ourselves to quasi{energies below E max = min fv ; V + g so our quasi{energies correspond to discrete eigenvalues of nite multiplicity. We assume (V1 - V4) for the remainder of this paper. These assumptions are not optimal: the degree of smoothness in (V1) can be relaxed to, say, V 2 C 4 (R) with V (4) uniformly Lipschitz; the growth condition in (V3) could be avoided by introducing \cutos"; and (V4) could at least be generalized (e.g., by using the limit superior or inferior). Crucial to our proof is the semiclassical time evolution of a class of Gaussian states [5] [6]. Given complex numbers A and B satisfying real numbers a and, and ~ > we dene ' (A; B; ~; a; ; x) = (~) 1=4 A 1=2 exp AB + BA = 2; (3) 1 2~ BA 1 (x a) 2 + i ~ (x a) : (4) We explicitly dene the branch of the square root in this denition when necessary. The function ' is normalized in the sense that it has L 2 norm (hereafter denoted by k k) equal to one. We write H(q; p) for the classical Hamiltonian H(q; p) = 1 2 p2 + V (q); 3

H(~) for the quantum Hamiltonian H(~) = ~2 2 d 2 dx 2 + V (x); and use the symbol H alone when the context is clear. For E < E max, let (E) denote a single connected component of the energy surface (actually a curve) (E) = f(q; p) 2 R 2 : H(q; p) = Eg: We call the trajectory (E) regular if its projection onto the spatial component q avoids the top of a potential barrier, i.e., if q = minfq : (q; p) 2 (E)g and q + = maxfq : (q; p) 2 (E)g are distinct adjacent roots of V (q) = E with V (q ) < and V (q + ) >. Under such conditions, it is well known (see, e.g., [4]) that the classical motion in phase space is periodic with a positive minimal period. The period depends on the initial conditions only through the energy. For a regular trajectory (E) we dene the \action function" I by I(E) = I (E) p dq: Geometrically, I(E) is the phase space area enclosed by the trajectory (E). This function is well-dened, and implicit function arguments show that, in a neighborhood of a regular trajectory, it is as smooth in E as V is in x. It is straightforward to show that the period of the motion conned to (E) is (E) = @ @E I(E): 1.2 The results We now state our main results. We assume V satises assumptions (V1 - V4) and that (E) is a regular trajectory. Next, we determine time-dependent quantities a(t), (t), A(t), B(t), and S(t) from classical mechanics. When the Bohr{Sommerfeld condition I(E) = I (E) p dq = 2~n; n 2 + 4

is satised, (~; x) dened by (2) is a quasimode in the sense that it satises H(~) E + ~ (~; ) (E) = O(~3=2 ): The precise statement is the following: Theorem 1 Suppose V satises assumptions (V1 - V4), and E < E max. Suppose (E) is a regular trajectory, and dene (E) = (E) 2(E). Let A and B be complex numbers that satisfy (3), (a ; ) 2 (E), and let a(t), (t), A(t), B(t), and S(t) be given by the unique solution of the system of ordinary dierential equations _a(t) = (t) (5) _(t) = V (a(t)) (6) A(t) _ = ib(t) + 2(E)(t) (V (a(t))a(t) + i(t)b(t)) (7) _B(t) = iv (a(t))a(t) + 2i(E)V (a(t)) (V (a(t))a(t) + i(t)b(t)) (8) _S(t) = 1 2 (t)2 V (a(t)) (9) subject to the initial conditions Then (a(); (); A(); B(); S()) = (a ; ; A ; B ; ) : I(E) = E (E) + S((E)): (1) We assume ~ and E satisfy the Bohr{Sommerfeld condition I(E) = 2~n; n 2 + (11) and dene s jj (~; x) = (~) 1=4 2(E) (E) ~ it(e+ e (E) )=~ e is(t)=~ ' (A(t); B(t); ~; a(t); (t); x) ; (12) where denotes the conserved quantity = V (a(t))a(t) + i(t)b(t); 5

and the branch of the square root in (4) is determined by continuity in t. Then, (~; x) and E(~) = E + ~=(E) are a quasimode/quasi-energy pair for H(~), i.e., k (~; )k = 1 + O(~ 1=2 ) and there is a constant C such that H(~) E + ~ (E) (~; ) C~3=2 k (~; )k : (13) In [1], Paul and Uribe use an integral containing a certain coherent state to produce similar results (in fact, the quasimode and quasi-energy are expanded to all orders in ~) for one-dimensional Schrodinger operators having polynomial symbols. Karasev [8] obtains results similar to ours and to the leading terms in [1] using coherent states integrated over arbitrary Lagrangian manifolds. De Bievre et al. [2], [3] use integration of a coherent state along a classical trajectory to obtain approximate eigenfunctions for n-dimensional harmonic oscillators. Our contribution lies in the ideas and methods used in the construction of the quasimode (~; x). The wave packets used in our construction are semiclassical approximations to the quantum evolution. The construction detailed in Theorem 1 is easily implemented numerically. Figure 1 compares the absolute square of an exact eigenfunction and our quasimode for the explicitly solvable Morse potential V (x) = (1 e x ) 2 [1]. To construct we took a =, = 1, A = B = 1, n = 1, and approximated all other necessary quantities using standard numerical schemes. The value of ~ is :413224 and the quasi-energy E + ~= is :52682. The tenth eigenvalue above the ground state for the Morse potential with this value of ~ is :519478. Before proceeding with the proof of the theorem, we present the motivation and intuition that led to our ideas. 1.3 A remark on a \natural" quasimode construction Integration of an approximate solution of the time{dependent Schrodinger equation over the corresponding classical trajectory is a clear, natural way to attempt to construct a quasimode. However, this naive construction based on the wave packets of [5], [6] fails to work, except in very special cases, such as 6

1-1 1 2 Figure 1: The absolute square of a quasimode (solid line) and the corresponding exact eigenfunction (dashed line) for the Morse potential. the harmonic oscillator. Understanding this failure provides the motivation for the construction we use. The wave packets of [5], [6] are the same as those in the theorem, except that A(t) and B(t) satisfy A(t) _ = ib(t) (14) _B(t) = iv (a(t))a(t) (15) instead of (7) and (8). We adopt the shorthand notation '(~; x; t) = ' (A(t); B(t); ~; a(t); (t); x); (16) with A(t) and B(t) satisfying (14) and (15). The function e is(t)=~ '(~; x; t) is an asymptotic solution of the time-dependent Schrodinger equation in the sense that e ith(~)=~ '(~; ; ) e is(t)=~ '(~; ; t) = O(~ 1=2 ) and i~ @@t H(~) e is(t)=~ '(~; ; t) = O(~3=2 ) for t. The rst of these properties is contained in the conclusion of Theorem 1.1 of [6], the second is imbedded in its proof. The branches of the square roots appearing in ' are determined by continuity in t. 7

The naive approach attempts to construct a quasimode 2 L 2 (R; dx) by (~; x) = ~ 1=4 (E) ~ it(e+ e (E) )=~ e is(t)=~ '(~; x; t); (17) where the factor of ~ 1=4 is inserted for purposes of normalization. This satises H(~) E ~ (E) = ~ 1=4 (E) = i~ 3=4 (E) (~; x) H(~) E ~ (E) ~ @ it(e+ e (E) )=~ e is(t)=~ '(~; x; t) @t ~ it(e+ e (E) )=~ e is(t)=~ '(~; x; t) + O(~ 5=4 ) = i~ 3=4 e i( E+~+S( ))=~ '(~; x; ) '(~; x; ) + O(~ 5=4 ); (18) where we have written F (~; x) = G(~; x) + O(~ p ) when kf (~; ) G(~; )k = O(~ p ). If we could arrange for '(~; x; ) = e i(~; )=~ '(~; x; ) for some real (~; ), then the imposition of the \quantization condition" E + S() + ~ + (~; ) = 2n~, n 2 or, its equivalent I (E) p dq + ~ + (~; ) = 2n~, n 2 would lead to being a quasimode for H(~) because the rst term in the last line of (18) would vanish. However, we can use elementary ODE theory (see the proof of the theorem) to show that B(t + )A 1 (t + ) 6= B(t)A 1 (t) for any t except in the non{generic special case when @ (E) =. Thus, @E the lack of periodicity of the solutions to (14) and (15) causes the naive construction to fail. Undaunted, we adopt an apparently less natural construction. We seek quasimodes for an operator f(h) rather than H itself. This is not as outlandish as one might think, because spectral mapping arguments relate the 8

spectra of H and f(h), and the classical trajectories are the same for the classical Hamiltonians H and f(h). Furthermore, there is a well-known canonical transformation on phase space (the action-angle variables) that produces dynamics with the period independent of E near a regular trajectory. Although this function of H may be a complicated object with which to work, we can approximate it by its Taylor series to obtain a simpler Hamiltonian. These ideas lead us to consider a Hamiltonian of the form f E (H) = H + 1 (E) (H E)2 2 (E) which is (up to an additive, E-dependent, constant and a scaling by (E)) the second order Taylor expansion about H = E of the action variable in the well known action-angle formalism of classical mechanics. Applying the techniques of [5] and [6] to this Hamiltonian, we obtain equations (7-8) instead of (14-15). The solutions to (7-8) have the periodicity to make the naive construction work for f E (H), i.e., (~; x) given by (12) satises f E (H(~)) E ~ (E) (~; ) = O(~3=2 ); as long as the Bohr-Sommerfeld quantization condition is satised. We then use spectral mapping arguments to prove the theorem. 2 The Proof of the Theorem The proof of the theorem follows the following outline: First, we develop the necessary classical mechanics. Next, we prove some results on the semiclassical evolution of our Gaussian wave packets. Then, we construct a quasimode for the Hamiltonian f E (H(~)). Finally, we argue that the construction actually yields quasimodes for H(~). 2.1 Some classical quantities In this section, we establish and collect some facts about classical quantities that arise from two Hamiltonian systems. The rst is the standard Newtonian system with Hamiltonian H(q; p) = 1 2 p2 + V (q): (19) 9

It is well known (see, e.g., [4]) that for any initial conditions (q ; p ) 2 R 2 the system _q = @H (q; p) = p (2) @p _p = @H @q (q; p) = V (q) (21) has a unique solution (q(q ; p ; t); p(q ; p ; t)) that satises (q(q ; p ; ); p(q ; p ; )) = (q ; p ): We often drop the explicit dependence on the initial conditions for convenience, even though much of this section is devoted to studying quantities generated by dierentiation with respect to initial conditions. We restrict attention to initial conditions (q ; p ) that are contained in a regular trajectory (H(q ; p )). Our study of this system mainly concerns the relations between the derivatives of the solution of (2-21) with respect to initial conditions evaluated at the initial time and at the period of the motion. We are also interested in the system with Hamiltonian f E (H(q; p)) where f E (H) = H + 1 (E) 2 (E) (H E)2 : (22) Here, E is considered a parameter and (E) denotes the period of the solution of (2-21) with initial conditions (q ; p ) satisfying H(q ; p ) = E. We introduce this Hamiltonian because derivatives with respect to initial conditions of certain solutions of the resulting Hamiltonian system are periodic with the same period as the orbit. This forces periodicity on certain quasiclassical quantities (namely, A(t) and B(t)) that arise in our construction. For the purpose of distinguishing the classical motions generated by the two Hamiltonians, we denote by ((a(a ; ; t); (a ; ; t)) the solution of the Hamiltonian system arising from (22): _a = @ @ f E(H(a; )) = f E(H(a; )) (23) _ = @ @a f E(H(a; )) = f E(H(a; ))V (a): (24) We note that, if H(a ; ) = E, then ((a(a ; ; t); (a ; ; t)) = (q(a ; ; t); p(a ; ; t)); 1

i.e., the two motions are identical for all time. As mentioned before, the distinction between these two systems is in the behavior of the derivatives of the motions with respect to initial conditions. It is this behavior we now begin to document. The existence and time dierentiability of the rst order partial derivatives of q, p, a, and with respect to the initial conditions follows from standard ODE theory (see, e.g., [12]). We rst concentrate on the system with Hamiltonian (19). Dierentiating (2-21) with respect to r 2 fq ; p g, we see that the rst order derivatives of q and p satisfy the system d @q @p where M(t) denotes the matrix M(t) = = M(t) @q @p 1 V (q(t)) : (25) A fundamental matrix for (25) is U(t) = " @q @q @p @q # @q @p @p @p ; i.e., _U(t) = M(t)U(t) and U() = I. We let = (E) (where E = H(q ; p )) denote the period of (q(t); p(t)), and dierentiate both sides of q(q ; p ; t + ) p(q ; p ; t + ) with respect to q and p to obtain: = q(q ; p ; t) p(q ; p ; t) U(t + ) = U(t) + (E) V (q )p(t) p(t)p V (q(t))v (q ) V : (26) (q(t))p We now turn our attention to the system generated by the Hamiltonian (22). Dierentiating (23-24) with respect to r 2 fa ; g and restricting to E = H(a ; ), we obtain d @a @ = M(t) @a @ 11 + (E) @E (E) V (a) : (27)

Using variation of parameters (and the fact that the nonhomogeneous term actually satises the homogeneous part of the equation), we see that @a @ @a @ = U(t) @a @ @q @p + t (E) @E (E) + t (E) @E (E) V (a) p V (q) : (28) This establishes a formula for comparing the solutions of (25) and (27): = : Evaluating (28) at time t + (E) and using (26), we see that @a (t + (E)) @a @ (t + (E)) = (t) @ (t) (29) if H(a ; ) = E. We now list some simple facts that we use in the next section. We rst note that equation (28) implies (t) = @a @a @a @ V (a ); (3) V (a(t)) = @ @ V (a ) @ @a : (31) Next, we dierentiate H(a ; ) = H(a; ) with respect to a and to obtain = @ @ (t) + @a @ V (a(t)); V (a ) = @ @a (t) + @a @a V (a(t)): (32) 2.2 The semiclassical evolution and a quasimode for fe(h(~)) We now turn our attention to the quantities A, and B that arise from the solution of (7-8) with E = H(a ; ). We rst note that the quantity (t) = V (a(t))a(t) + i(t)b(t) 12

is conserved by the motion generated by (5-8), i.e., It is easy to see that the two vectors A(t) and ib(t) both satisfy with ~x() = and (t) = V (a )A + i B : @a @a @ @a d ~x(t) = M(t)~x(t) + (E) (E). So, we conclude that ib A A + i A(t) = @a @a A + i @a @ B B(t) = @ @ B i @ @a A : V (a) " @a # @ @ B @ From these relations and (29), we see that A(t) and B(t) are periodic with period (E). One can easily check that A(t)B(t) + A(t)B(t) = A B + A B ; so, if A and B satisfy (3), then so do A(t) and B(t) for all t >. To determine the phase of (A()) 1=2 we need to show that the trajectory fa(t) : t g C has winding number about the origin equal to one. To prove this, we rst note that since @a ; @ A(t) = A @a @a Im( B A ) @a @ + i ja j 2 it suces to prove the trajectory (ja j 2 ( @a Im( B ) @a ); @a ) in R 2 winds @a A @ @ the origin exactly once. We rst consider the special case where the classical ; 13

motion originates at a turning point, say, (a ; ) = (q ; ). In this case, equation (3) implies (t) = @a @ V (q ) From this, we deduce that @a vanishes exactly twice, namely at t = and @ t = =2. At t =, and, at t = =2, ja j 2 @a @a Im( B A ) @a @ t= = ja j 2 > ja j 2 @a @a Im( B A ) @a @ t==2 = ja j 2 since equation (32), evaluated at t = =2, states @a @a t= =2 V (q + ) = V (q ): From (27) and (31) evaluated at t = =2 we nd so @a @ d @a = @ t= @ = =2 @ t==2 V (q + ) @a @a t==2 < ; V (q ) < ; changes sign at t = =2 and the claim follows. In the general case, for arbitrary (a ; ) we let t denote the rst positive time at which the trajectory (a(t); (t)) reaches (q ; ). By existence and uniqueness, we then have A(A ; B ; a ; ; t + t ) = @a @a (q ; ; t)a(t ) + i @a @ (q ; ; t)b(t ); (a ; ; t + t ) = @a @ (q ; ; t)v (q ); and the result follows by using the argument for the special case. Hence, if we determine the branch of the square root along the trajectory fa(t) : t g by continuity and use p to denote any xed branch, we have [A ] 1=2 = p A implies [A((E))] 1=2 = e ip A((E)): 14

We note that the phase e i that occurs here is directly related to the Maslov index of the orbit. We are now nearly ready to develop the semiclassical evolution of the state ' (A; B; ~; a; ; x) determined by the Hamiltonian f E (H). We use the following result to control the semiclassical errors (see Lemma 3.3 of [7]). Proposition 2 Let H(~) be a family of self-adjoint operators for ~ >. Suppose (~; x; t) is continuously dierentiable in t and belongs to the domain of H(~) for ~ >. Suppose further that (~; x; t) satises: i~ @@t H(~) (~; ; t) = O(~ ) for t 2 [; T ]. Then e ith(~)=~ (~; ; ) (~; ; t) = O(~ 1 ) for t 2 [; T ]. To estimate norms that arise in our proof, we rely on the following fact that is easily established by explicit calculation or a scaling argument. Proposition 3 If F (~; x) is such that jf (~; x)j C~ k (x a) m for some constants C, k, and non-negative integer m, then kf (~; )' n (A; B; ~; a; ; )k = O(~ k+m=2 ): (33) Moreover, the estimate (33) is uniform for a,, A, and B in compact sets. Now, with all quantities dened as in Theorem 1, we dene '(~; x; t) = ' (A(t); B(t); ~; a(t); (t); x) with the branches of the square roots appearing in the denition of ' determined by continuity in t starting with a given branch at t =. Explicit calculation and Proposition 3 show that i~ @ @t f E(H(~)) e is(t)=~ '(~; ; t) = O(~3=2 ): (34) 15

By Proposition 2, this implies that e itf E (H(~))=~ '(~; ; ) e is(t)=~ '(~; ; t) = O(~ 1=2 ): The argument from Section 1.3 then shows that (~; x) = ~ 1=4 (E) satises f E (H(~)) E ~ (E) (~; x) ~ it(e+ e (E) )=~ e is(t)=~ '(~; x; t) = i~ 3=4 e i( E+~+S( ))=~ '(~; x; ) '(~; x; ) + O(~ 5=4 ): (35) Using the facts in the beginning of this section we conclude and therefore, if '(~; x; ) = e i '(~; x; ); e i( E+~+S( ))=~ '(~; x; ) '(~; x; ) = E + S() = 2n~ (36) for some integer n. This is precisely the Bohr{Sommerfeld condition (11). To see this, note that by using time to parametrize the integral and using @q @t I(E) = = = p, we have (E) (E) p(t) 2 I(E) = I (E) p dq; (E) p(t) 2 =2 + V (q(t) + p(t) 2 =2 V (q(t) = E (E) + S((E)): 16

Thus, if we restrict the values of ~ to the sequence I(E) ~ 2 2n we have f E (H(~)) E ~ (E) n2 + (~; x) = O(~ 5=4 ): (37) In the next section, we prove that the norm of is of order 1. Thus, (37) shows that and E + ~=(E) are a quasimode/quasi-energy pair for the Hamiltonian f E (H(~)). We also show in the next section that the power of ~ on the right side of equation (37) can be improved to 3=2. 2.3 The normalization of and an improved error estimate In this section, we prove some estimates which allow us to show that our quasimode is properly normalized and that the error in equation (37) is actually O(h 3=2 ). We need two preliminary results. We rst obtain a formula for the inner product of two Gaussians of the form (4), that is proved by explicit integration. Proposition 4 Suppose the pair A 1 and B 1 satisfy (3) and the pair A 2 and B 2 satisfy (3). Suppose a 1, 1, a 2, and 2 are real, and let ~ be positive. Then h' (A 1 ; B 1 ; ~; a 1 ; 1 ; ); ' (A 2 ; B 2 ; ~; a 2 ; 2 ; )i = s 2 exp A 2A 1 ( 2 1 ) 2 + B 2 B 1 (a 2 a 1 ) 2 B 2 A 1 + A 2 B 1 2~(B 2 A 1 + A 2 B 1 ) i (a 2 a 1 )(B 2 A 1 1 + A 2 B 1 2 ) ~(B 2 A 1 + A 2 B 1 ) Next, we prove an estimate that concerns integrals of a type we encounter. Proposition 5 Suppose f(t; s) is a complex C 2 function and g(t; s) is a complex C 3 function, for t 2 [; T ] and s 2 [ T=2; T=2]. Suppose there exists >, such that Re(g(t; s)) s 2, for t 2 [; T ] and s 2 [ T=2; T=2]; 17 :

g(t; ) = ; @g @s (t; ) = ; and @2 g (t; ) = (t) is real and positive. Then for @s2 any non-negative integer n, we have T T=2 T=2 ds f(t; s)s 2n e g(t;s)=~ = 1 3 j2n 1j p T 2 ~ n+1=2 f(t; )(t) n 1=2 + O(~ n+1 ): Proof: On the domain of interest, jf(t; s)j is uniformly bounded by some number C 1. Because Re g(t; s) s 2, we have je g(t;s)=~ j e s2 =~. Thus, jsj > ~ implies f(t; s)s 2n e g(t;s)=~ C1 (T=2) 2n e =~1 2 : Therefore, for any < 1=2, T ~ <jsjt=2 for any p. So, the integral in question equals T ~ ds f(t; s)s 2n e g(t;s)=~ = O(~ p ); ~ ds f(t; s)s 2n e g(t;s)=~ + O(~ p ): Since e g(t;s)=~ 1, standard error estimates for rst order Taylor series now show that the integral equals T ~ ds(f(t; ) + @f ~ @s (t; )s)s2n e g(t;s)=~ + T ~ ~ ds 1 2 for some (t; s) with values between and s. @ 2 f @s 2 (t; (t; s))s2n+2 e g(t;s)=~ ; (38) @ 2 f @s 2 (t; (t; s))e g(t;s)=~ is bounded by some C 2, the second term Since 1 2 in (38) is bounded by C 2 T ~ (2n+3) =(2n + 3). By choosing suciently close to 1=2, this can be made O(~ n+1 ). 18

Next, by standard Taylor series error estimates and our hypotheses on g, we have g(t; s) = (t)s 2 =2 + O(s 3 ), uniformly for t 2 [; T ]. Thus, jsj ~ implies Since T e g(t;s)=~ e (t)s2 =(2~) ~ ~ ds f(t; ) + @f the rst term in (38) equals T ~ e C 3 jsj 3 (t) s2 2 g(t;s) =~ 1 ~ e (t)s2 =(2~) : @s (t; )s jsj 2n+3 ~ e (t)s2 =(2~) e (t)s2 =(2~) C 4 ~ n+1 ; ~ ds(f(t; ) + @f @s (t; )s)s2n e (t)s2 =(2~) + O(~ n+1 ): (39) We make an exponentially small error by extending the s integration in (39) to the whole real line. We then explicitly compute the resulting s integral to obtain T ~ ds f(t; s)s 2n e (t)s2 =(2~) = ~ 1 3 j2n 1j p T 2 ~ n+1=2 f(t; )(t) n 1=2 + O(~ n+1 ): This implies the proposition. Now, let (~; x) be dened as in Theorem 1: s jj (E) (~; x) = (~) 1=4 ~ it(e+ e (E) )=~ e is(t)=~ 2(E) ' (A(t); B(t); ~; a(t); (t); x) : Proposition 6 The norm of the quasimode (~; ) satises k (~; )k = 1 + O(~ 1=2 ): 19

Proof: The square of the norm of the quasimode is jj 1=2 h (~; ); (~; )i = (~) 2 jj 1=2 = (~) 2 e it 2(E+ ~ )=~ e is(t 2)=~ ' (A(t 2 ); B(t 2 ); ~; a(t 2 ); (t 2 ); x) 2 ; )=~ e is(t 1)=~ ' (A(t 1 ); B(t 1 ); ~; a(t 1 ); (t 1 ); x) 1 e it 1(E+ ~ exp =2 i S(t 1 ) S(t 2 ) + (t 1 t 2 ) E + ~ h' (A(t 2 ); B(t 2 ); ~; a(t 2 ); (t 2 ); ); ' (A(t 1 ); B(t 1 ); ~; a(t 1 ); (t 1 ); )i 1 2 jj =2 1=2 = (~) ds exp i S(t) S(t + s) s E + ~ =~ 2 s exp exp 2 B(t)A(t + s) + B(t + s)a(t) ( A(t)A(t + s)((t + s) (t))2 + B(t)B(t + s)(a(t + s) a(t)) 2 2~(B(t)A(t + s) + B(t + s)a(t)) B(t)A(t + s)(t + s) + B(t + s)a(t)(t) 8 < : i (a(t) a(t + s)) ~ (B(t)A(t + s) + B(t + s)a(t)) =~ ) 9 = ; : In the last step we have changed the limits of integration by exploiting periodicity in t 2 and changed variables by t = t 1 and s = t 2 t 1. We have also used Proposition 4 to evaluate the inner product in the integrand. We rewrite the integrand in the form f(t; s) e g(t;s)=~, where s f(t; s) = e is= 2 B(t)A(t + s) + B(t + s)a(t) : We note that Re g(t; s) and g(t; ) =. Also, formula (3) implies f(t; ) = 1. Making use of equations (5){(9), we compute the second order Taylor series expansion of g(t; s) in the variable s. This is a lengthy calculation, but 2

the result is s 2 1 B(t)A(t) 2 ja(t)j 2 V (a(t)) 2 + jb(t)j 2 (t) 2 s 2 g(t; s) = i(t)v (a(t)) By using formula (3), we can rewrite this as g(t; s) = ja(t)v (a(t)) + ib(t)(t)j 2 s 2 4 + O(s3 ): 4 + O(s3 ) jj2 s 2 4 + O(s 3 ): It now follows that the hypotheses of Proposition 5 are satised. Thus, Proposition 5 and an explicit integration show k (~; )k 2 = 1 + O(~ 1=2 ): The proposition follows by taking square roots. We close this section with an outline of the proof that the error in the estimate (37) is actually of order ~ 3=2. The idea is to use a variant of Proposition 4 and Proposition 5 to estimate the right side of (34) rather than bringing the norm inside the integral and using Proposition 3. Explicit calculation shows that the quantity inside the norm in equation (34) is actually where i~ @ @t f E(H(~)) e is(t)=~ '(~; x; t) = ~ 3=2 e is(t)=~ (f 1 (t)' 1 (A(t); B(t); ~; a(t); (t); x) + f 3 (t)' 3 (A(t); B(t); ~; a(t); (t); x)) + O(~ 2 ) ' n (A; B; ~; a; ; x) = 2 n=2 (~) n=4 A (n+1)=2 A n=2 H n (~ 1=2 jaj 1 (x a)) exp 2~BA 1 1 (x a) 2 + i (x a) ~ with H n denoting the n th Hermite polynomial and f 1 and f 3 given by rather complicated expressions in A(t), B(t), a(t), (t), V (a(t)), and V (n) (a(t)) for n = 1, 2, 3 which we do not display here. This implies that the O(~ 5=4 ) term in equation (35) is ~ 5=4 ~ it(e+ e )=~ e is(t)=~ (f 1 ' 1 + f 3 ' 3 )(t) + O(~ 7=4 ): 21

We then estimate the L 2 norm of the integral above using the same trick as in the proof of Proposition 6, i.e., ~ it(e+ e )=~ e is(t)=~ (f 1 ' 1 + f 3 ' 3 )(t) 2 e it 1(E+ ~ )=~ e is(t 1)=~ (f 1 ' 1 + f 3 ' 3 )(t 2 ) 2 ; e it 2(E+ ~ )=~ e is(t 2)=~ (f 1 ' 1 + f 3 ' 3 )(t 1 ) 1 : (4) We break this into four integrals in x, s = t 2 t 1, and t = t 1 and bring the inner products \inside the integrals". We then use the following extension of Proposition 4 to evaluate the inner products: Proposition 7 Suppose the pair A 1 and B 1 satisfy (3) and the pair A 2 and B 2 satisfy (3). Suppose a 1, 1, a 2, and 2 are real, and let ~ be positive. Then h' l (A 1 ; B 1 ; ~; a 1 ; 1 ; ); ' k (A 2 ; B 2 ; ~; a 2 ; 2 ; )i = 1 p 2 l+k 2 h' (A 1 ; B 1 ; ~; a 1 ; 1 ; ); ' (A 2 ; B 2 ; ~; a 2 ; 2 ; )i l!k! (B 1 A 2 + A 1 B 2 ) l+k 2 min(l;k) X j= l j = k j!4 j (A 2 B 1 A 1 B 2 ) k j 2 j p p ) A2 B 1 A 1 B 2 B1 A 2 + A 1 B # 2 (A 1 B 2 A 2 B 1 ) l j 2 H k j (~ 1=2 B 1 (a 1 a 2 ) ia 1 ( 1 2 ) H l j ( ~ 1=2 B 2 (a 1 a 2 ) + ia 2 ( 1 2 ) p A1 B 2 A 2 B 1 p B1 A 2 + A 1 B 2 ) Proof: Induction and the properties of the Hermite polynomials establish the integral formula 1 1 e x2 H m (ax + b)h n (cx + d) dx = p 1 2 (m+n+1) min(m;n) X k= m n H m k (b k r k k!( a 2 ) m k 2 ( c 2 ) n k a 2 )H n k(d 22 r c ) 2 : 2 (2ac) k

for Re() > from which the proposition follows easily by completion of the square in the exponent and a change of variable of integration. This implies that the right side of (4) is a sum of terms of the form where ~ (m+n)=2 =2 =2 ds f(t; s) z 1 (t; s) m z 2 (t; s) n e g(t;s)=~ (41) z 1 (t; s) = B(t)(a(t) a(s + t)) ia(t)((t) (s + t)); z 2 (t; s) = B(s + t)(a(t) a(s + t)) + ia(s + t)((t) (s + t)); m + n is an even integer, f(t; s) is a complex C 2 function, and g(t; s) is as in Proposition 6. For any < 1=2, (41) is equal to ~ (m+n)=2 ~ for any p. We make an error of order ~ ds f(t; s) z 1 (t; s) m z 2 (t; s) n e g(t;s)=~ + O(~ p ) O(~ (m+n)=2 ~ (m+n+2) ) = O(~ 2 (1 2)(m+n)=2 ) by replacing z 1 (t; s) and z 2 (t; s) by their Taylor series in s: z 1 (t; s) = is + O(s 2 ); z 2 (t; s) = is + O(s 2 ); and an exponentially small error in extending the s integration back to the interval [ =2; =2] to obtain: ~ (m+n)=2 =2 =2 =2 ~ (m+n)=2 ds f(t; s) z 1 (t; s) m z 2 (t; s) n e g(t;s)=~ = =2 Proposition 5 now applies with the result that ~ it(e+ e ds f(t; s) s m+n e g(t;s)=~ + O(~ 2 (1 2)(m+n)=2 ) )=~ e is(t)=~ (f 1 ' 1 + f 3 ' 3 ) = O(h1=4 ) for suciently large, and hence the right side of equation (35) is actually i~ 3=4 e i( E+~+S( ))=~ '(~; x; ) '(~; x; ) + O(~ 3=2 ); thus providing us with the estimate necessary to prove (13). 23

2.4 A quasimode for H In this section we complete the proof of the theorem. We must establish the connection between quasimodes for the Hamiltonian f E (H(~)) and quasimodes for the Hamiltonian H(~). We cannot establish this by use of spectral mapping arguments alone, because the map x 7! f E (x) is not invertible. This complication is easily overcome, however, with use of the following observations: 1. Our quasimode for f E (H) is also a quasimode for the Hamiltonian g E (H) = H + 1 (E) 2 (E) (H E)2 + (H E) 3 for arbitrary 2 R. This follows because the crucial estimate (34) (and hence, every result in the preceding two sections) holds with f E (H(~)) replaced with g E (H(~)) for arbitrary 2 R. 2. If > 2 =3, the polynomial function p(x) = x + x 2 + x 3 on R has an inverse. This follows by showing that, for such values of, p (x) > for all x 2 R. Armed with these two simple observations, we now settle the question of relating quasimodes of f E (H(~)) (or g E (H(~))) to quasimodes of H(~). Proposition 8 Suppose H(~) is self-adjoint on a Hilbert space H, and E 2 R. Suppose g(z) = z +(z E) 2 +(z E) 3, where is chosen large enough so that g(z + E) E is invertible. Suppose there exists a vector (~) 2 H with k (~)k = 1, such that k[g(h(~)) E] (~) k C ~ for some >. Then there exists C, such that k[h(~) E] (~)k C ~ : Proof: We dene g 1 : R! R by g 1 (z) = g(z + E) E = z + z 2 + z 3 ; and let h be its inverse function. Clearly, h is dierentiable; h() = ; and jh(z)j grows at most linearly for large jzj. Thus, there exists C, such that jh(z)j C jzj for all z 2 R. 24

By the spectral L theorem, there exists a unitary operator U ~ from H to the direct sum j L2 (R; d ~;j (x)), such that U ~ H(~)U 1 ~ is multiplication by x. Letting ~;j () denote the j th component of U ~ (~), we have X k[h(~) E] (~)k 2 = X j = j X j R R R jx Ej 2 j ~;j (x)j 2 d ~;j (x) jh(g 1 (x E))j 2 j ~;j (x)j 2 d ~;j (x) C 2 jg 1 (x E)j 2 j ~;j (x)j 2 d ~;j (x) = C 2 k g 1 (H(~) E) (~) k 2 = C 2 k [g(h(~)) E] (~) k 2 C 2 C 2 ~ 2 : This implies the proposition with C = C C. References [1] J. P. Dahl and M. Springborg, \The Morse oscillator in position space, momentum space, and phase space", J. Chem. Phys., 88 (7), p. 4535-4547, 1988. [2] S. De Bievre, \Oscillator eigenstates concentrated on classical trajectories," J. Phys. A, 25, p. 339{3418, 1992. [3] S. De Bievre, J. C. Hourd, and M. Irac-Astaud, \Wave packets localized on closed classical trajectories," in Dierential Equations with Applications to Mathematical Physics, Mathematics in Science and Engineering, Volume 192, edited by W. F. Ames, E. M. Harrell II, and J. V. Herod, Academic Press, San Diego, 1993. [4] G. Gallavoti, The Elements of Mechanics, Springer-Verlag, New York, 1983. [5] G. Hagedorn, \Semiclassical quantum mechanics I: the ~! limit for coherent states," Commun. Math. Phys., 71, p. 77-93, 198. 25

[6] G. Hagedorn, \Semiclassical quantum mechanics III: the large order asymptotics and more general states," Ann. Phys, 135, p. 58-7, 1981. [7] G. Hagedorn, Molecular Propagation through Electron Energy Level Crossings, Memoirs of the American Mathematical Society, Volume 111, Number 536, American Mathematical Society, Providence, Rhode Island, 1994. [8] M. V. Karasev, \Simple quantization formula," in Symplectic Geometry and Mathematical Physics (Aix-en-Provence 199), Progress in Mathematics, Volume 99, Birkhauser, Boston, 1991. [9] M. D. Khaut-Duy, \Formule des traces semi-classique pour une energie critique et construction de quasi-modes a l'aide d'etats coherents", doctoral thesis, Universite Paris IX Dauphine, 1996. [1] T. Paul and A. Uribe, \A construction of quasi-modes using coherent states," Ann. Inst. Henri Poincare, Vol. 59, n o 4, p. 357-381, 1993. [11] T. Paul and A. Uribe, \On the pointwise behavior of semi{classical measures," Commun. Math. Phys., 175, p. 229{258, 1996. [12] L. C. Piccini, G. Stampacchia, and G. Vidossich, Ordinary Dierential Equations in R n, Applied Mathematical Sciences 39, Springer-Verlag, New York, 1984. [13] J. V. Ralston, \On the construction of quasimodes associated with stable periodic orbits," Commun. Math. Phys., 51, p. 219-242, 1976. 26