Section 9 Variational Method. Page 492

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

Physics 221A Fall 2017 Notes 27 The Variational Method

PY 351 Modern Physics - Lecture notes, 3

( ) = 9φ 1, ( ) = 4φ 2.

Lecture 45: The Eigenvalue Problem of L z and L 2 in Three Dimensions, ct d: Operator Method Date Revised: 2009/02/17 Date Given: 2009/02/11

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

TP computing lab - Integrating the 1D stationary Schrödinger eq

Ch 125a Problem Set 1

Physics 342 Lecture 17. Midterm I Recap. Lecture 17. Physics 342 Quantum Mechanics I

Path integrals and the classical approximation 1 D. E. Soper 2 University of Oregon 14 November 2011

Physics 505 Homework No. 1 Solutions S1-1

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

Quantum Mechanics Solutions

Summary of free theory: one particle state: vacuum state is annihilated by all a s: then, one particle state has normalization:

Quantum Mechanics: Vibration and Rotation of Molecules

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

Project: Vibration of Diatomic Molecules

Physics 137A Quantum Mechanics Fall 2012 Midterm II - Solutions

CHAPTER 8 The Quantum Theory of Motion

Lecture 3 Dynamics 29

Lecture 5. Hartree-Fock Theory. WS2010/11: Introduction to Nuclear and Particle Physics

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

Generators for Continuous Coordinate Transformations

Quantum Mechanics Solutions. λ i λ j v j v j v i v i.

Simple Harmonic Oscillator

Sample Quantum Chemistry Exam 2 Solutions

Harmonic Oscillator I

The Sommerfeld Polynomial Method: Harmonic Oscillator Example

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

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

7. Numerical Solutions of the TISE

Lecture 10: Solving the Time-Independent Schrödinger Equation. 1 Stationary States 1. 2 Solving for Energy Eigenstates 3

Problems and Multiple Choice Questions

The Particle in a Box

REVIEW: The Matching Method Algorithm

Statistical Interpretation

8.04 Spring 2013 March 12, 2013 Problem 1. (10 points) The Probability Current

1 Mathematical preliminaries

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

5.4 Given the basis e 1, e 2 write the matrices that represent the unitary transformations corresponding to the following changes of basis:

Chem 3502/4502 Physical Chemistry II (Quantum Mechanics) 3 Credits Spring Semester 2006 Christopher J. Cramer. Lecture 9, February 8, 2006

df(x) = h(x) dx Chemistry 4531 Mathematical Preliminaries Spring 2009 I. A Primer on Differential Equations Order of differential equation

Chapter 1.6. Perform Operations with Complex Numbers

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

Reading: Mathchapters F and G, MQ - Ch. 7-8, Lecture notes on hydrogen atom.

Harmonic Oscillator with raising and lowering operators. We write the Schrödinger equation for the harmonic oscillator in one dimension as follows:

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

Introduction to Electronic Structure Theory

Lecture 9: Eigenvalues and Eigenvectors in Classical Mechanics (See Section 3.12 in Boas)

Lecture 12. The harmonic oscillator

Physics 741 Graduate Quantum Mechanics 1 Solutions to Midterm Exam, Fall x i x dx i x i x x i x dx

The quantum state as a vector

Continuous quantum states, Particle on a line and Uncertainty relations

CHAPTER 6 Quantum Mechanics II

Chem 3502/4502 Physical Chemistry II (Quantum Mechanics) 3 Credits Spring Semester 2006 Christopher J. Cramer. Lecture 7, February 1, 2006

Chemistry 432 Problem Set 4 Spring 2018 Solutions

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

For a system with more than one electron, we can t solve the Schrödinger Eq. exactly. We must develop methods of approximation, such as

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-XXVI. Time-Independent Schrodinger Equation

Relativistic Quantum Mechanics

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

Two and Three-Dimensional Systems

Quantum mechanics in one hour

Lecture 6 Quantum Mechanical Systems and Measurements

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

3 The Harmonic Oscillator

Mathematical Structures of Quantum Mechanics

PHY 407 QUANTUM MECHANICS Fall 05 Problem set 1 Due Sep

! 4 4! o! +! h 4 o=0! ±= ± p i And back-substituting into the linear equations gave us the ratios of the amplitudes of oscillation:.»» = A p e i! +t»»

Quantum Mechanics I Physics 5701

You may not start to read the questions printed on the subsequent pages until instructed to do so by the Invigilator.

in terms of the classical frequency, ω = , puts the classical Hamiltonian in the form H = p2 2m + mω2 x 2

Perturbation Theory 1

DS-GA 1002 Lecture notes 0 Fall Linear Algebra. These notes provide a review of basic concepts in linear algebra.

ECE 487 Lecture 6 : Time-Dependent Quantum Mechanics I Class Outline:

Time part of the equation can be separated by substituting independent equation

Lecture Notes 2: Review of Quantum Mechanics

Quantum Mechanics II

Mathematical Physics Homework 10

One-dimensional Schrödinger equation

Page 712. Lecture 42: Rotations and Orbital Angular Momentum in Two Dimensions Date Revised: 2009/02/04 Date Given: 2009/02/04

We do not derive F = ma; we conclude F = ma by induction from. a large series of observations. We use it as long as its predictions agree

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

Introduction to Quantum Mechanics PVK - Solutions. Nicolas Lanzetti

Vector Spaces. Vector space, ν, over the field of complex numbers, C, is a set of elements a, b,..., satisfying the following axioms.

Quantum Physics 130A. April 1, 2006

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

Addition of Angular Momenta

5.1 Classical Harmonic Oscillator

Quantum Mechanics: Postulates

Maxwell s equations. electric field charge density. current density

Path integral in quantum mechanics based on S-6 Consider nonrelativistic quantum mechanics of one particle in one dimension with the hamiltonian:

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

Hilbert Spaces. Hilbert space is a vector space with some extra structure. We start with formal (axiomatic) definition of a vector space.

We start with some important background material in classical and quantum mechanics.

Waves and the Schroedinger Equation

P3317 HW from Lecture and Recitation 7

Total Angular Momentum for Hydrogen

1.3 Harmonic Oscillator

Transcription:

Section 9 Variational Method Page 492

Page 493 Lecture 27: The Variational Method Date Given: 2008/12/03 Date Revised: 2008/12/03

Derivation Section 9.1 Variational Method: Derivation Page 494 Motivation It is not always possible to find an analytic solution to the Schrödinger Equation. One can always solve the equation numerically, but this is not necessarily the best way to go; one may not be interested in the detailed eigenfunctions, but rather only in the energy levels and the qualitative features of the eigenfunctions. And numerical solutions are usually less intuitively understandable. Fortunately, one can show that the values of the energy levels are only mildly sensitive to the deviation of the wavefunction from its true form, and so the expectation value of the energy for an approximate wavefunction can be a very good estimate of the corresponding energy eigenvalue. By using an approximate wavefunction that depends on some small set of parameters and minimizing its energy with respect to the parameters, one makes such energy estimates. The technique is called the variational method because of this minimization process. This technique is most effective when trying to determine ground state energies, so it serves as a nice complement to the WKB approximation, which works best when one is interested in relatively highly excited states, ones whose debroglie wavelength is short compared to the distance scale on which the wavelength changes.

Section 9.1 Variational Method: Derivation Page 495 The Hamiltonian s Eigenstates and Eigenvalues are Stationary Points of the Energy Functional For any wavefunction ψ, we may calculate the expectation value of the energy, E[ψ] = ψ H ψ ψ ψ = R dx ψ (x)h ψ(x) R dx ψ (x)ψ(x) (9.1) (We will explain below why we need to explicitly include the normalizing denominator.) We call E a functional of the wavefunction ψ because the above expression maps a wavefunction ψ to a number E. We aim to show that we can obtain the energy eigenvalues by requiring that E[ψ] be stationary with respect to ψ. By this, we mean that, if there is a function ψ such that, for small variations δψ away from ψ, the corresponding variation δe in E[ψ] vanishes, then ψ is an eigenstate of the Hamiltonian with energy E[ψ]. This is the same kind of requirement one places on the classical action in Lagrangian mechanics to yield a differential equation for the classical path, so we are essentially just doing the calculus of variations with the E[ψ(x)] functional instead of the S[x(t)] functional.

Section 9.1 Variational Method: Derivation Page 496 Let s explicitly insert a variation δψ into our equation for E[ψ] and determine the resulting δe: E + δe = R dx [ψ(x) + δψ(x)] H [ψ(x) + δψ(x)] R dx [ψ(x) + δψ(x)] [ψ(x) + δψ(x)] (9.2) It is now clear why we had to keep the denominator explicit: since we are varying ψ, the normalization of ψ will change. We thus must explicitly include the normalization correction in the denominator to get the correct energy functional. Since having the variations in the denominator on the right side is hard to deal with, but we may multiply through by the denominator to obtain» [E + δe] =» [E + δe] = dx [ψ(x) + δψ(x)] [ψ(x) + δψ(x)] [ψ(x) + δψ(x)] H [ψ(x) + δψ(x)] dx ψ (x)ψ(x) + dx ψ (x)h ψ(x) + dx δψ (x) ψ(x) + dx δψ (x) H ψ(x) + dx ψ (x) δψ(x) dx ψ (x)h δψ(x) (9.3) (9.4) where we have kept only terms to first order in δψ.

Section 9.1 Variational Method: Derivation Page 497 Next, we use the unvaried version of the equation to eliminate the terms that include no δψ factors, and we set δe = 0 to impose the stationarity condition.» E dx δψ (x) ψ(x) + dx ψ (x) δψ(x) = δψ (x)h ψ(x) + ψ (x)h δψ(x) dx δψ (x) (H E) ψ(x) + (9.5) dx ψ (x) (H E) δψ(x) = 0 (9.6) Next, we can show that the two terms must vanish independently by considering two special cases for the variation δψ. Suppose δψ(x) = χ(x) where χ(x) is purely real. Then we have dx χ(x) (H E) ψ(x) = dx ψ (x) (H E) χ(x) (9.7) Next, suppose δψ is completely imaginary, so δψ(x) = i χ(x) where χ(x) is again real. This gives dx χ(x) (H E) ψ(x) = dx ψ (x) (H E) χ(x) (9.8) where we have divided out by i.

Section 9.1 Variational Method: Derivation Page 498 Since δψ is completely arbitrary, it is necessary for both equations to hold simultaneously for any real function χ. That is only possible if both terms in the original equation vanish independently. So we have dx δψ (x) (H E) ψ(x) = 0 (9.9) dx ψ (x) (H E) δψ(x) = 0 (9.10) Since δψ is arbitrary in each equation, the integrands must vanish. The first equation yields (H E) ψ(x) = 0 (9.11) One must use the Hermiticity of H to transform the second equation so that H acts on ψ and not δψ, which will yield the complex conjugate of the above equation. We recover the eigenvector-eigenvalue equation for H.

Section 9.1 Variational Method: Derivation Page 499 Thus, we have proven that, if ψ(x) is an eigenfunction of H, then a small variation δψ(x) results in no change in E to first order in δψ(x). This proof is interesting in its own right we had no reason to expect that the Hamiltonian s eigenfunctions would result in an extremum of the energy functional in much the same way that there is no reason to expect ahead of time that the classical path that solves Newton s equations would correspond to an extremum of the action functional. We note that the we could have dealt with the normalization of ψ differently. We could have imposed a normalization requirement via a Lagrange multiplier by considering instead the alternate functional E [ψ] =» dx ψ (x) H ψ(x) λ dx ψ (x) ψ(x) 1 (9.12) and requiring δe = 0 under a variation δψ. The two functionals E[ψ] and E [ψ] are the same when one requires that ψ be normalized, so the requirement δe = 0 is equivalent to the requirement δe = 0 under that condition. The result would have been the same had we gone this route.

Section 9.1 Variational Method: Derivation Page 500 Is the converse of what we have proven true? That is, if ψ m is an eigenfunction of H with eigenvalue E m, does it hold that the energy functional E[ψ] is stationary with respect to variations away from ψ m? The answer is yes, and this is relatively easy to show. Let us allow for a variation δψ, and let s expand δψ in terms of the eigenstates: δψ = X n c n ψ n (9.13) Now, let s calculate the energy of the wavefunction with the variation: E[ψ m + δψ] = ( ψm + δψ ) H ( ψm + δψ ) ( ψ m + δψ ) ( ψ m + δψ ) Em + Em ( ψm δψ + δψ ψm ) = 1 + ( ψ m δψ + δψ ψ m ) = E m 1 + cm + c m 1 + c m + c m (9.14) (9.15) «= E m (9.16) Hence, δe = E[ψ m + δψ] E[ψ m] = 0. So, indeed, if ψ m is an eigenstate of H, then E[ψ] is stationary at ψ = ψ m: a variation δψ in ψ m results in no variation in E to first order in δψ.

Section 9.1 Variational Method: Derivation Page 501 For our purposes, the practical implication of the relationship between ψ being an eigenstate and the energy functional being stationary with respect to variations in ψ is that the fractional error in the energy estimate obtained from the trial wavefunction will be much smaller than that fractional error in the wavefunction itself.

Section 9.1 Variational Method: Derivation Page 502 Ground State Energy Upper Bounds It is easy to show that whatever estimate we make for the ground state energy using this technique, it is always an upper bound. Suppose ψ(x) is our trial wavefunction, the energy eigenfunctions are {φ n(x)} with eigenvalues E n. We may expand ψ, ψ(x) = P n cn φn(x). Let us then calculate the energy: E[ψ] = P n cn 2 E n Pn cn 2 (9.17) Subtract off the ground state energy E 0 E[ψ] E 0 = P n cn 2 (E n E 0 ) Pn cn 2 (9.18) Because E 0 is the ground state energy, E n E 0 > 0 for all n. c n 2 0, so the right side is nonnegative. Therefore E[ψ] E 0 (9.19)

Section 9.1 Variational Method: Derivation Page 503 Excited State Energy Estimates If one knows the eigenfunctions for n < m and one wants to estimate E m, then one is assured of an upper bound by requiring that the trial wavefunction be orthogonal to the φ n for n < m. Explicitly, we consider a trial wavefunction ψ and require φ n ψ = 0 for n < m (9.20) This condition can be met for any trial wavefunction ψ(x) via Gram-Schmidt orthogonalization (Section 3.3). We then calculate the energy functional E[ψ] = P n=m cn 2 E n P n=0 cn 2 (9.21) because the terms with n < m vanish due to c n = φ n ψ = 0. We then subtract E m: E[ψ] E m = P n=m cn 2 (E n E m) P n=0 cn 2 (9.22) E n E m for n > m, so we are assured the right side is nonnegative, yielding E[ψ] E m when φ n ψ = 0 for n < m (9.23)

Section 9.1 Variational Method: Derivation Page 504 Of course, the above is somewhat useless because, if one is able to exactly solve the eigenvector-eigenvalue equation for some n, then one is usually able to solve it for all n and the above technique is unnecessary. The idea about Gram-Schmidt orthogonalization is good, though: even if one only has approximations to the lower-energy state eigenfunctions, one should still construct trial wavefunctions for the higher states using Gram-Schmidt orthogonalization. One can show, for example, that the error in estimating E m is related to the mismatch between one s estimates ψ n for the actual eigenfunctions φ n for n < m. In particular, if δ 0 = 1 ψ 0 φ 0 2 (9.24) expresses the fractional deviation of the approximate ground state wavefunction ψ 0 from the true one φ 0, then one can show that a trial wavefunction ψ 1 that has been constructed using Gram-Schmidt orthogonalization with respect to ψ 0 (not φ 0!) yields E[ψ 1 ] E 1 δ 0 (E 1 E 0 ) (9.25) That is, E[ψ 1 ] is no longer an upper bound on E 1, but the amount by which it underestimates E 1 is proportional to δ 0. Because δ 0 is quadratic in ψ 0 φ 0, the fractional error in estimating E 1 is much smaller than the error in estimating the wavefunction, as long as E 1 E 0 is of order E 1.

Applications Section 9.2 Variational Method: Applications Page 505 How it Works in Practice How do we construct trial wavefunctions for the purpose of making these variational estimates of energy eigenvalues? One cannot usually guess the form for the correct wavefunction exactly. But one usually knows the general features of the wavefunction one wants. So one constructs a trial wavefunction that depends on some small set of parameters, calculates the energy functional for the trial wavefunction as a function of this small set of parameters, and then requires E be stationary with respect to those parameters i.e., that all the partial derivatives of E with respect to the parameters vanish. In effect, we are explicitly applying the stationarity condition to some subset of all possible wavefunctions under the expectation that we can get a very good approximation to the energy with a good approximation of the correct wavefunction thanks to the stationarity of E with respect to variations in ψ. An important guide is to require that the trial wavefunctions be eigenfunctions of any other Hermitian operators A that commute with H. These are usually related to symmetries of the Hamiltonian. For example, a Hamiltonian in which the potential is an even function of posiition commutes with the parity opertator and hence any eigenfunctions of H must have definite parity. The generators of continuous symmetry transformations are also good examples; rotation symmetry, for example, implies that angular momentum commutes with H. If one makes these requirements on the trial wavefunctions, one is assured that they are at least members of the same subspace as the true eigenfunctions.

Applications (cont.) Section 9.2 Variational Method: Applications Page 506 Example 9.1: Particle in a Box A good first example is to do a variational-method estimate for a problem whose exact solution we already know the particle in a box. The Hamiltonian guides our choice of trial wavefunction in two ways. First, we know the trial wavefunction should vanish at the box edge and outside. If the wavefunction s derivative is large at any point, the kinetic energy will be big, so the wavefunction should rise smoothly and slowly away from zero at the box edges. Second, the Hamiltonian commutes with the parity operator, so the trial wavefunction should be even or odd. We try an even function because it can have no zeroes and thus also have the smallest possible derivative and hence kinetic energy. We try 8 < 2 L ψ(x; c) = 2 x 2«`1 + c x 2 x L 2 : 0 x > L 2 (9.26) This is the simplest polynomial trial function we can use. Evenness requires that we only include even powers of x. A simple quadratic could satisfy the requirement of going to zero at the box edges, but admits no free parameters: it must be (L/2) 2 x 2 (an overall constant multiplier just sets the normalization and cancels out between the numerator and denominator of the energy functional). So the next possibility is a polynomial containing x 2 and x 4, and the above is just one convenient way to parameterize it (again, neglecting any overall scale factor).

Applications (cont.) Section 9.2 Variational Method: Applications Page 507 Calculating the integrals for the energy functional is a somewhat nontrivial algebraic exercise that can be done correctly in Mathematica, yielding 4 2 E[ψ] E[c] = 3 2 11 L 2 c 2 + 14 L 2 c + 35 m L 2 4 2 (9.27) L 2 c 2 + 6 L 2 c + 21 We then find the extrema of E(c) with respect to c, de/dc = 0, yielding with solutions 26 «L 4 c 2 + 196 2 «L 2 c + 42 = 0 (9.28) 2 c 1 = 0.221075 2 c 2 = 7.31771 2 (9.29) L L 2 2

Applications (cont.) Section 9.2 Variational Method: Applications Page 508 We may then calculate E for these values, giving E(c 1 ) = 4.93488 2 m L 2 E(c 2 ) = 51.0652 2 m L 2 (9.30) The reason that the second solution has so much higher energy is that it has zeros at x = ±c 1/2 2 0.185 L, hence its derivative and its kinetic energy is much larger than the c 1 state, which has no zeros. The true ground state energy is E n=1 = 2 π 2 2 = 4.93480 2 m L2 m L 2 (9.31) which is shockingly close to E(c 1 ). Plotting the trial function shows that it is a very good match to the true ground state wavefunction.

Applications (cont.) Section 9.2 Variational Method: Applications Page 509 Another interesting fact is that the second solution, using c 2, is a decent approximation to the second excited state, n = 3: E n=3 = 9 E n=1 = 44.4132 2 m L 2 (9.32) There is a general theorem on this point, the Hylleraas-Undheim Theorem, which essentially states that, if the trial wavefunction depends on a set of parameters, the alternate solutions for the parameters giving non-minimal extrema of E yield estimates for the energies of the excited states of the system.

Applications (cont.) Section 9.2 Variational Method: Applications Page 510 Example 9.2: Other Examples A few other examples of simple analytically soluble problems that can be approximated quite well or perfecty by the variational method are: The simple harmonic oscillator a Gaussian trial function will recover the ground state energy and wavefunction exactly because the ground state wavefunction is a Gaussian. A Gaussian times a polynomial will cover the excited states exactly, also, via the Gram-Schmidt orthogonalization procedure. Bound state of a δ-function potential well. The trial function should be a decaying exponential.