Lecture 3 Dynamics 29

Similar documents
Quantum Mechanics. L. Del Debbio, University of Edinburgh Version 0.9

Lecture 7. More dimensions

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

Lecture 12. The harmonic oscillator

Lecture 6 Quantum Mechanical Systems and Measurements

5.1 Classical Harmonic Oscillator

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

Simple Harmonic Oscillator

Quantum Mechanics. p " The Uncertainty Principle places fundamental limits on our measurements :

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

PY 351 Modern Physics - Lecture notes, 3

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

Physics 215 Quantum Mechanics 1 Assignment 5

A few principles of classical and quantum mechanics

2. The Schrödinger equation for one-particle problems. 5. Atoms and the periodic table of chemical elements

The Simple Harmonic Oscillator

Angular Momentum in Quantum Mechanics.

Waves and the Schroedinger Equation

Columbia University Department of Physics QUALIFYING EXAMINATION

3 The Harmonic Oscillator

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

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

Quantum Mechanics: Vibration and Rotation of Molecules

Lecture 4: Equations of motion and canonical quantization Read Sakurai Chapter 1.6 and 1.7

Quantum Mechanics-I Prof. Dr. S. Lakshmi Bala Department of Physics Indian Institute of Technology, Madras. Lecture - 21 Square-Integrable Functions

Section 9 Variational Method. Page 492

SECOND QUANTIZATION PART I

15 Functional Derivatives

Time evolution of states in quantum mechanics 1

P3317 HW from Lecture and Recitation 7

1 Mathematical preliminaries

Lecture Notes 2: Review of Quantum Mechanics

The Harmonic Oscillator: Zero Point Energy and Tunneling

-state problems and an application to the free particle

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

Introduction to Path Integrals

SIMPLE QUANTUM SYSTEMS

221A Lecture Notes Path Integral

Statistical Interpretation

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

PLEASE LET ME KNOW IF YOU FIND TYPOS (send to

Simple one-dimensional potentials

Lecture #1. Review. Postulates of quantum mechanics (1-3) Postulate 1

One-dimensional harmonic oscillator. -motivation. -equation, energy levels. -eigenfunctions, Hermite polynomials. -classical analogy

Physics 137A Quantum Mechanics Fall 2012 Midterm II - Solutions

CHAPTER 6 Quantum Mechanics II

Chemistry 532 Problem Set 7 Spring 2012 Solutions

Brief review of Quantum Mechanics (QM)

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

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

Quantum Mechanics for Scientists and Engineers. David Miller

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

Ae ikx Be ikx. Quantum theory: techniques and applications

Quantum mechanics. Chapter The quantum mechanical formalism

PHYS-454 The position and momentum representations

Exercises : Questions

Quantum Mechanics: Postulates

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

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

Introduction to Quantum Mechanics PVK - Solutions. Nicolas Lanzetti

Ch 125a Problem Set 1

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

CHAPTER 8 The Quantum Theory of Motion

d 1 µ 2 Θ = 0. (4.1) consider first the case of m = 0 where there is no azimuthal dependence on the angle φ.

Problems and Multiple Choice Questions

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

Quantum Physics (PHY-4215)

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

Mass on a Horizontal Spring

The quantum state as a vector

By following steps analogous to those that led to (20.177), one may show (exercise 20.30) that in Feynman s gauge, = 1, the photon propagator is

van Quantum tot Molecuul

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

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

Linear Algebra in Hilbert Space

Quantum mechanics in one hour

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

Q U A N T U M M E C H A N I C S : L E C T U R E 5

Physics-I. Dr. Anurag Srivastava. Web address: Visit me: Room-110, Block-E, IIITM Campus

General Physical Chemistry II

Physics 217 Problem Set 1 Due: Friday, Aug 29th, 2008

E = hν light = hc λ = ( J s)( m/s) m = ev J = ev

POSTULATES OF QUANTUM MECHANICS

Basic Quantum Mechanics Prof. Ajoy Ghatak Department of Physics Indian Institute of Technology, Delhi

Physics 115/242 Comparison of methods for integrating the simple 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

which implies that we can take solutions which are simultaneous eigen functions of

Lecture 10. Central potential

The 3 dimensional Schrödinger Equation

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

01 Harmonic Oscillations

Physics 443, Solutions to PS 2

Chemistry 532 Practice Final Exam Fall 2012 Solutions

The Quantum Harmonic Oscillator

Harmonic Oscillator I

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

Quantum Mechanics for Scientists and Engineers

Why quantum field theory?

arxiv: v1 [quant-ph] 15 Dec 2011

MP463 QUANTUM MECHANICS

Transcription:

Lecture 3 Dynamics 29

30 LECTURE 3. DYNAMICS 3.1 Introduction Having described the states and the observables of a quantum system, we shall now introduce the rules that determine their time evolution. In classical mechanics, the state of a single particle system is described by the classical trajectory, i.e. by the knowing the position of the particle as a function of time: x(t). 1 The time-dependence of the position is described by Newton s law: f = ma, (3.1) where a(t) =d 2 x(t)/dt 2 is the acceleration of the system, and f is the external force acting on it. Newton s law is a dynamic equation, relating the time evolution of the state of the system to the force acting on it. It is a second order di erential equation, which can be solved (at least in principle) once we know the initial condition for the position and the velocity of the particle at some given time t 0. As discussed in the previous chapter, in quantum mechanics the concept of classical trajectory cannot be defined, and the state of the system is described by the wave function. In this lecture we describe the quantum analogue of Newton s law, i.e. the equation that describes the time evolution of the wave function. 3.2 Schrödinger equation 3.2.1 Equation of motion The time evolution of a quantum state is dictated by Schrödinger s equation: i~ @ @t (x, t) =Ĥ (x, t), (3.2) where Ĥ = ˆT + ˆV is the Hamiltonian operator. Eq. (3.2) replaces Newton s equation of classical mechanics. 2 The operator Ĥ is obtained from the classical Hamiltonian, by replacing the position and momentum with the corresponding operators ˆX and ˆP in Cartesian coordinates. Schrödinger s equation states that the Hamiltonian operator determines the time shifts of the wave function, i.e. Ĥ = i~ @ @t. Similarly, the momentum operator determines the space 1 If the particle evolves in more than one space dimension, then the position is identified by a vector x(t). 2 It was introduced by the Austrian physicist Erwin Schrödinger in 1926, following the wave-particle duality suggested by de Broglie.

3.2. SCHRÖDINGER EQUATION 31 translations - remember that ˆP = i~ @ @x. In a four-dimensional Minkowski space-time this correspondence can be written as a four-vector relation: ˆP µ = i~ @ @x µ. (3.3) It is interesting to note that special relativity can be easily embedded in the rules of quantum mechanics! The Standard Model of particle physics, which describes the interactions between the fundamental constituents of matter very successfully, is based on this fact. This will become more clear when you study quantum field theory. A digression on Schrödinger s equation Let us discuss briefly a heuristic explanation of the identification Ĥ = i~ @ @t by considering a one-dimensional plane wave (x, t) = const e i(!t kx). (3.4) The plane wave describes e.g. the propagation of monochromatic light. Applying the operator i~ @ @t we obtain: i~ @ @t (x, t) =~! (x, t) =h (x, t) (! =2 ), (3.5) which is precisely the energy of the photon. Moreover we know that probabilities must be normalized to one. The mathematical translation of this property is that: Z dx (x, t) (x, t) =1, for all t. (3.6) Hence the derivative with respect to time of the integral above must vanish: Z apple @ @ dx (x, t) (x, t)+ (x, t) (x, t) = 0 ; (3.7) @t @t the latter equation shows that @ @t must be represented by an anti-hermitean operator. The Hamiltonian Ĥ is the operator associated to the energy of the system and therefore is hermitean. This explains the factor of i that appears in Schrödinger s equation, it is necessary to connect @ @t to an Hermitean operator. Finally the factor ~ is needed on dimensional grounds. Other numerical factors e.g. factors of 2,, etcetera cannot be derived by these simple arguments, but are explained by the heuristic derivation above. Time evolution using Dirac notation If we use Dirac s notation to represent the quantum states by kets, Schrödinger s equation becomes: i~ @ @t (t)i = Ĥ (t)i. (3.8) The equation above simply states that the time-evolution of the vector (t)i is generated by acting on the vector with the operator Ĥ.

32 LECTURE 3. DYNAMICS 3.2.2 Eigenstates of the Hamiltonian Let us consider a time-independent Hamiltonian: Ĥ = Ĥ( ˆX, ˆP ), (3.9) where there is no explicit dependence on time. The time evolution for an energy eigenstate can be worked out easily. equation becomes simply: Schrödinger s i~ @ @t (x, t) =E (x, t). (3.10) Integrating Eq. (3.10) with respect to time: apple (x, t) = E (x)exp i Et ~, (3.11) where E (x) is the spatial wavefunction corresponding to the eigenvalue E. Such states are known as stationary states because the associated probability density is independent of time: E(x, t) 2 E(x, t) E(x, t) = E(x)exp{iEt/~} E (x)exp{ iet/~} = E (x) 2. Likewise, the expectation value of any operator is also independent of time: Z h E(t) Ô E(t)i = dx E(x, t) Ô E(x, t) (3.12) Z = dx E (x) e iet/~ Ôe iet/~ E(x) (3.13) Z = dx E (x) e iet/~ e iet/~ Ô E (x) (3.14) Z = dx E (x) Ô E (x) (3.15) = h E(0) Ô E(0)i. (3.16) Note that formally we can write the solution of the Schrödinger equation as: apple i (t)i =exp (0)i, (3.17) ~Ĥt h i i even for a state that is not an eigenstate of the energy. However in this case exp ~Ĥt is a complicated operator. It can be written explicitly using a Taylor expansion: apple i exp ~Ĥt = X 1 i n n! ~ Ĥn t n. (3.18) n

3.2. SCHRÖDINGER EQUATION 33 When applied to an eigenstate of the energy it turns out to be a simple complex phase!! You will see some examples in problem sheet 2. The eigenvalue equation Ĥ n (x) =E n n (x) (3.19) yields the eigenvalues E n, i.e. all the possible values of the energy of the system, and the stationary states n (x). Eq. (3.19) is often referred to as time-independent Schrödinger equation. Using the explicit definition of Ĥ in Eq. (1.35), we can write the time-independent Schrödinger equation as a second order di erential equation: apple ~ 2 2m @ 2 @x 2 + V (x) u n(x) =E n u n (x). (3.20) Eigenvalues and eigenstates are found by looking for normalizable solutions of the di erential equation, taking into account the appropriate boundary conditions. Note that some authors use the notation {u n (x)} for the eigenfuncitons of the energy. We are going to use this convention below. 3.2.3 General restrictions on (x, t) We see from the form of Eq. (3.20) that its solutions have to satisfy a number of properties: 1. (x, t) must be a single-valued function of x and t. 2. (x, t) must be a continuous function of x and t. @ 3. @x (x, t) must be a continuous function of x. There is an exception to this set of rules. 1. Restriction 3 does not apply when the potential energy function V (x) has infinite discontinuities e.g. the particle-in-a-box problem met in Physics 2, for which 1 if x>aor x<0 V (x) = 0 otherwise We shall see see later a detailed description of this system.

34 LECTURE 3. DYNAMICS 3.2.4 Example: the quantum harmonic oscillator In order to give a concrete example of solutions of the Schrödinger equation, let us review the quantum harmonic oscillator. Remember that the harmonic oscillator is of paramount importance in physics, because every system close to the minimum of the potential is described by a quadratic potential, i.e. it behaves like a harmonic oscillator. Classically, the equation of motion for a particle of mass m, subject to a restoring force proportional to the displacement, x, is mẍ = k x (3.21) where the constant of proportionality, k, is usually called the spring constant. This has oscillatory solutions of angular frequency! = p k/m. We can obtain the energy equation by writing ẍ = v dv dx, (3.22) and integrating to give 1 2 mv2 + 1 2 kx2 = E, (3.23) so that the potential energy function, V, is given by V (x) = 1 2 kx2 = 1 2 m!2 x 2. (3.24) The evolution of the position of the oscillator as a function of time x(t) is shown in Fig. 3.1 for two di erent values of!. The total energy of the oscillator is E = 1 2 m!2 L 2,whereL is the maximum amplitude of the oscillations. The larger the amplitude of the oscillations, the larger the energy of the system. In the case depicted in Fig. 3.1 the amplitude L = 1 for both curves. For the quantum harmonic oscillator, the time-independent Schrödinger equation is thus apple ~ 2 2m d 2 dx 2 + 1 2 m!2 x 2 u n (x) =E n u n (x). (3.25) We will not discuss the details of the solution of this di erential equation, although you may have already seen it in a mathematical methods course and/or book. In any case we will develop a much more elegant method to find its solutions in a later lecture. It su ces to say that acceptable solutions only arise for certain values of the total energy i.e. the energy is quantised. The energy eigenvalues are as follows: E n =(n + 1 2 )~! n =0, 1, 2, 3,...1 (3.26)

3.2. SCHRÖDINGER EQUATION 35 1.0 x t 0.5 2 4 6 8 t 0.5 1.0 Figure 3.1: Classical trajectory of the harmonic oscillator for! = 1 (blue curve), and! =2 (red curve). The state of lowest energy, or ground state, is labelled by n = 0 and has an energy 1 2 ~! rather than zero. The corresponding eigenfunctions may be written as u n (x) =C n exp{ 2 x 2 /2} H n ( x) (3.27) where C n is a normalisation constant, 2 m!/~ and H n ( x) are polynomials of degree n, known as Hermite polynomials. They satisfy the orthogonality relation Z 1 1 exp{ s 2 }H m (s)h n (s)ds =2 np n! mn, (3.28) implying that the energy eigenfunctions are orthogonal (orthonormal with an appropriate choice of the constant, C n ). The first few Hermite polynomials are H 0 (s) = 1 H 1 (s) = 2s H 2 (s) = 4s 2 2 H 3 (s) = 8s 3 12s (3.29) Note that the polynomials, and hence the energy eigenfunctions, are alternately even and odd functions, i.e. are of either even or odd parity. If we introduce a parity operator, ˆP, whose action on a function, f(x), is defined by ˆP f(x) =f( x), (3.30)

36 LECTURE 3. DYNAMICS so that the eigenvalues of ˆP are ±1 only, corresponding to the even and odd functions f + (x) 1 2 [f(x)+f( x)] and f (x) 1 2 [f(x) f( x)]. (3.31) You can readily check that: ˆP u n (x) =( 1) n u n (x), (3.32) and we say that the parity is ( 1) n. The probability density to find the oscillator at a given position x is u n (x) 2. The probability densities for the first five levels are shown in Fig. 3.2. Note that for higher values of the energy, we get a larger probability to find the oscillator at larger values of x. 5 4 3 2 1 4 2 2 4 Figure 3.2: Probability to find the oscillator at x for the first five levels of the harmonic oscillator for = 1. Each curve is shifted upwards by n + 1 2 to improve the readability of the plot. In order to quantify the last sentence we can compute the average position, and the fluctuations of the position. The expectation value of the position of the quantum harmonic oscillator is given by: hxi n = Z +1 1 dx x C n exp{ 2 x 2 /2} H n ( x) 2. (3.33) You can easily convince yourself that hxi = 0, because the integrand is always an odd function of x. This is what we would have naively expected, since the system is symmetric under parity. The quantum oscillator is equally likely to be to the left or to the right of the origin at x = 0.

3.2. SCHRÖDINGER EQUATION 37 It is more interesting to compute the mean square deviation hx 2 i, which measures the amplitude of the oscillations. You can compute the integrals explicitly for the first few Hermite polynomials, or simply look up the answer in a table of integrals: hx 2 i n =(n + 1 2 ) 1 2. (3.34) Once again this is the quantitative expression of our intuitive understanding from classical mechanics: the larger the energy, the larger the oscillations!! 3.2.5 Expansion in stationary states Given the expansion of the wave function at t =0inthe{u n (x)} basis: (x, 0) = X n a n u n (x), a n = hu n (0)i, (3.35) we can express the evolved wave function at time t as: (x, t) = X n a n e ient/~ u n (x). (3.36) If the Hamiltonian does not depend explicitly on time, then the energy levels E n are constant in time: d dt E n = d dt h n(x, t) Ĥ n(x, t)i = d Z dx n(x, t) Ĥ n(x, t) dt = 1 Z h i dx n(x, t) Ĥ,Ĥ n(x, t) = 0 (3.37) i~ An observable such that: for any state (x, t) is called a conserved observable. d h (t) Ô (t)i =0, (3.38) dt A time-independent operator is conserved if and only if hĥ, Ôi = 0. This is a consequence of the equation that determines the time-dependence of a matrix element: i~ d i hô, dt h (t) Ô (t)i = h (t) Ĥ (t)i. (3.39) The proof of this result is left as an exercise in problem sheet 2.

38 LECTURE 3. DYNAMICS 3.3 Summary Let us conclude this chapter by summarizing the main concepts introduced in this lecture. Time-dependent Schrödinger equation. Eigenstates of the Hamiltonian. Time-independent Schrödinger equation. Physical characterization of the stationary states. First properties of the solutions of the Schrödinger equation. The harmonic oscillator. Analytic solution, comparison between the quantum, and the classical behaviours. Completeness of the energy eigenstates. Expansion in stationary states.