phy 416, quantum mechanics - siue

277
PHY 416, Quantum Mechanics Notes by: Dave Kaplan and Transcribed to L A T E X by: Matthew S. Norton December 13, 2008

Upload: others

Post on 12-Feb-2022

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: PHY 416, Quantum Mechanics - SIUE

PHY 416, Quantum Mechanics

Notes by: Dave Kaplan and Transcribed to LATEX by: Matthew S. Norton

December 13, 2008

Page 2: PHY 416, Quantum Mechanics - SIUE

Contents

1 First Class: Quantum Mechanics 61.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61.2 Interference: Review of Evidence for the Wave Theory of Light − Thomas Young’s

Interference Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81.3 Two-slit interference, Quantitative Calculation of Intensity distribution at the Screen 10

2 Second Class: More on the Behavior of Light 122.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122.2 Single Slit Diffraction, continued . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132.3 Method of Phasors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142.4 Position Resolution − Why ∆x ∼ λ . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

3 Third Class: Quantum Bahavior 21

4 Fourth Class: Double Slit and Quantum Mechanics 224.1 Double Slit Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224.2 Orthodox (“Copenhagen”) Interpretation . . . . . . . . . . . . . . . . . . . . . . . . 244.3 Probability Rule ( Born, 1927) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264.4 Normalization Condition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264.5 Analogies with light . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

5 Fifth Class: What Would Be a Valid State Function For a Free Particle? 305.1 Some Experimentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 305.2 Partially Localized States . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 325.3 A Single Moving Group − “Wave Packet” . . . . . . . . . . . . . . . . . . . . . . . . 345.4 Some Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 365.5 Toward a Governing Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

6 Sixth Class: A Governing Equation 416.1 Complex Plane Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 416.2 What is the Meaning of a Complex Wave Function? . . . . . . . . . . . . . . . . . . 426.3 Probabilities and Normalization Revisited . . . . . . . . . . . . . . . . . . . . . . . . 436.4 Plane Wave and Uniformity of Space . . . . . . . . . . . . . . . . . . . . . . . . . . . 456.5 A First Example of Using the Schrodinger Equation . . . . . . . . . . . . . . . . . . 45

1

Page 3: PHY 416, Quantum Mechanics - SIUE

7 Seventh Class: Time Dependence Through the Governing Equation 497.1 A Review of the Governing Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . 497.2 Persistence of Normalization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 527.3 The Reason for the Persistence of the Normalization . . . . . . . . . . . . . . . . . . 537.4 Determinism in Quantum Mechanics − A Comment . . . . . . . . . . . . . . . . . . 55

8 Eight Class: Probability 578.1 Probability Current . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 578.2 Choice of Free Particle Plane Wave States . . . . . . . . . . . . . . . . . . . . . . . . 588.3 Circular and Spherical Definite Energy State Functions . . . . . . . . . . . . . . . . 598.4 Superposition States and Probabilities . . . . . . . . . . . . . . . . . . . . . . . . . . 608.5 Packet States − Physical Meaning . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

9 Ninth Class: Uncertainty in Quantum Mechanics 659.1 Certainty and Uncertainty . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

9.1.1 Meaning of ∆x . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 659.1.2 Meaning of ∆p . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

9.2 Example: Plane-Wave states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 679.3 Example: Position Measurement by Aperture . . . . . . . . . . . . . . . . . . . . . . 689.4 Expectation Values . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

10 Tenth Class: Heisenberg Uncertainty Principle and Fourier Analysis 7410.1 Heisenberg Uncertainty Principle − Another Example . . . . . . . . . . . . . . . . . 7410.2 Fourier Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

10.2.1 Basic Fourier Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7610.2.2 Complex Orthogonal Functions . . . . . . . . . . . . . . . . . . . . . . . . . . 7710.2.3 Complex Fourier Series on [0, 2π] . . . . . . . . . . . . . . . . . . . . . . . . . 7810.2.4 Complex Fourier Series on [−`, `], [0, L] . . . . . . . . . . . . . . . . . . . . . 7910.2.5 Special Case of Real f(x) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80

10.3 Fourier Expansion of Complex Non-periodic Functions − Fourier Transforms . . . . 81

11 Eleventh Class: Fourier Analysis and the Heisenberg Uncertainty Principle 8311.1 Fourier Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8311.2 A More General Line of Reasoning Leading to the Fourier Bandwidth Theorem . . . 8611.3 Heisenberg Uncertainty Principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8811.4 A Square-Normalized f(x) Implies a Square-Normalized g(k) . . . . . . . . . . . . . 8911.5 Fourier Analysis of “Arbitrary” f(x, t) . . . . . . . . . . . . . . . . . . . . . . . . . . 9011.6 Dirac Delta Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9111.7 The General Solution to Schrodinger’s Equation for a Free Particle . . . . . . . . . . 93

12 Twelfth Class: Momentum Space 9612.1 Quantum Initial-Value Problem, Case of Free Particle . . . . . . . . . . . . . . . . . 96

12.1.1 State Function of Definite Momentum . . . . . . . . . . . . . . . . . . . . . . 9912.1.2 State Function of Definite Position . . . . . . . . . . . . . . . . . . . . . . . . 99

12.2 Momentum-Space State Function − General Potential Energy Profiles . . . . . . . . 10012.3 Expectation Value Revisited . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102

2

Page 4: PHY 416, Quantum Mechanics - SIUE

12.4 Reality of 〈p〉 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10312.5 Hermitian Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10412.6 Normalization of Free-Particle Definite Momentum States . . . . . . . . . . . . . . . 105

12.6.1 “Box Normalization” . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10612.6.2 “Dirac Normalization” of Free-Particle Definite Momentum State . . . . . . . 106

12.7 Normalization of a State of Definite Position . . . . . . . . . . . . . . . . . . . . . . 10712.8 Completeness of Free-Particle Definite Momentum States . . . . . . . . . . . . . . . 107

13 Thirteenth Class 109

14 Fourteenth Class: 110

15 Fifteenth Class: States of Definite Energy 11115.1 Eigenstate Property . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11115.2 Time Dependence of Energy Eigenstates . . . . . . . . . . . . . . . . . . . . . . . . . 11515.3 Time Independent Schrodinger Equation . . . . . . . . . . . . . . . . . . . . . . . . . 11615.4 Energy Eigenstates are “Stationary States” . . . . . . . . . . . . . . . . . . . . . . . 116

15.4.1 Normalization of ψ(x) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11715.5 A Possible Means to Further Solution? (“Separation of Variables”) . . . . . . . . . . 11715.6 Spectrum of Eigenstates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11915.7 General Solution to the Schrodinger Equation . . . . . . . . . . . . . . . . . . . . . . 120

15.7.1 Completeness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12115.8 A Concrete Example: The Infinite Square Well Revisited − Eigenfunctions . . . . . 121

16 Sixteenth Class: Examples and Orthogonality 12516.1 A Concrete Example of the Formalism: Infinite Square Well Revisited . . . . . . . . 12516.2 The “Reason” for Energy Quantization . . . . . . . . . . . . . . . . . . . . . . . . . . 12816.3 Zero-Point Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13416.4 Orthogonality and Completeness of Eigenfunctions . . . . . . . . . . . . . . . . . . . 13616.5 The Quantum Initial Value Problem − Particle Not Free . . . . . . . . . . . . . . . . 13816.6 Energy Measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140

17 Seventeenth Class: Connections With Classical Physics and Optics 14317.1 The Connection with Everyday Physics . . . . . . . . . . . . . . . . . . . . . . . . . 14317.2 State Functions for Incidence on Abrupt but Finite Changes in V . . . . . . . . . . . 14717.3 Reflection and Transmission Coefficients . . . . . . . . . . . . . . . . . . . . . . . . . 15117.4 Classically Forbidden Region . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152

18 Eighteenth Class: More on the Potential Step 15318.1 More on Free-Particle Packet Motion and Group Velocity . . . . . . . . . . . . . . . 15318.2 Wave Packet Incident on a Potential Step, case E < V0 . . . . . . . . . . . . . . . . . 15518.3 Particle “Incident” on Potential Step, Case E > V0 . . . . . . . . . . . . . . . . . . . 158

19 Nineteenth Class: The Barrier Potential 16019.1 Barrier Penetration − Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160

19.1.1 A Feel for the Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16419.2 Useful Approximations for Rectangular Barrier Penetration . . . . . . . . . . . . . . 164

3

Page 5: PHY 416, Quantum Mechanics - SIUE

19.2.1 “High and Wide Barrier” . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16519.2.2 “Thin Barrier” . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165

19.3 “Arbitrary” Shaped Barrier (E < V0) . . . . . . . . . . . . . . . . . . . . . . . . . . . 166

20 Twentieth Class: Applications of Barrier Potentials 16820.1 Applications of Gamow’s Approximation . . . . . . . . . . . . . . . . . . . . . . . . . 16820.2 Application to Radioactive Alpha Decay . . . . . . . . . . . . . . . . . . . . . . . . . 169

20.2.1 The Basic Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17020.2.2 Summary of Goals for the Calculation . . . . . . . . . . . . . . . . . . . . . . 17220.2.3 Calculation of Collision Factor f . . . . . . . . . . . . . . . . . . . . . . . . . 17220.2.4 Calculation of T (E) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17320.2.5 Comparison with Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . 175

20.3 Another Application − Fusion in the Sun’s Core . . . . . . . . . . . . . . . . . . . . 17620.4 Another Example of Barrier Penetration − Ammonia Molecule . . . . . . . . . . . . 17620.5 Another Application − Covalent Molecular Bonding . . . . . . . . . . . . . . . . . . 177

21 Twenty-first Class: Potential Barrier and Finite Square Well 17921.1 Potential Barrier, Case E > V0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17921.2 Resonance Transmission . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18121.3 The Finite Square Well . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182

21.3.1 Finite Square Well, Incident Plane Wave State With E > 0 . . . . . . . . . . 18321.3.2 Bound States of the Finite Square Well . . . . . . . . . . . . . . . . . . . . . 18421.3.3 Bound States for Finite Square Well − Qualitative Observations . . . . . . . 18521.3.4 Parity of Eigenfunctions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18621.3.5 Finite Square Well, Quantitative Treatment of Bound States . . . . . . . . . 187

22 Twenty-Second Class: The Finite Square Well 19022.1 Brief Review of Quantitative Solutions for Finite Square Well . . . . . . . . . . . . . 19022.2 Some Intuition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19122.3 Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192

22.3.1 Application − Nuclear Potential and the Deuteron . . . . . . . . . . . . . . . 19322.4 Further Intuition into Bound State Eigenfunctions . . . . . . . . . . . . . . . . . . . 19622.5 Qualitative Eigenfunctions for Wells . . . . . . . . . . . . . . . . . . . . . . . . . . . 198

23 Twenty-third Class 201

24 Twenty-Forth Class: The Simple Harmonic Oscillator 20224.1 Eigenstate in a Harmonic Oscillator Potential Energy Profile . . . . . . . . . . . . . 202

24.1.1 Intuition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20324.1.2 Mathematical Treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204

24.2 Harmonic Oscillator Potential, Continued . . . . . . . . . . . . . . . . . . . . . . . . 208

25 Twenty-Fifth Class: Further Properties of Harmonic Oscillator Eigenfunctions21125.1 Expansion Postulate for Harmonic Oscillator . . . . . . . . . . . . . . . . . . . . . . 212

25.1.1 Mathematical Aside . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21325.2 Further Important Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214

4

Page 6: PHY 416, Quantum Mechanics - SIUE

25.3 The Classical Limit in the Harmonic Potential −A First Look . . . . . . . . . . . . . 21525.4 Behavior of Harmonic Oscillator Superposition States . . . . . . . . . . . . . . . . . 216

25.4.1 Motion of Gaussian Wave Packet in Harmonic Potential . . . . . . . . . . . . 21725.5 The Time Dependence of Expectation Values . . . . . . . . . . . . . . . . . . . . . . 22025.6 When Does Classical Mechanics Apply? . . . . . . . . . . . . . . . . . . . . . . . . . 225

25.6.1 Exceptions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22525.7 Another Example of an Operator − Angular Momentum . . . . . . . . . . . . . . . . 226

26 Twenty-Sixth Class: Quantum Mechanics and the Hydrogen Atom 22826.1 An Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22826.2 Spherically Symmetric Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23126.3 Other Spherically Symmetric Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . 23326.4 The Road to More General Hydrogen Atom Eigenfunctions . . . . . . . . . . . . . . 23426.5 Hydrogen Atom Eigenstate Functions . . . . . . . . . . . . . . . . . . . . . . . . . . 236

26.5.1 Boundary Conditions on Φ(φ) . . . . . . . . . . . . . . . . . . . . . . . . . . . 23726.6 The Θ Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23826.7 Spherical Harmonics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239

27 Twenty-Seventh Class: More on the Hydrogen Atom 24227.1 The Radial Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24227.2 More on the Hydrogen atom Radial Equation . . . . . . . . . . . . . . . . . . . . . . 24427.3 A check . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24627.4 Hydrogen Atom − Eigenstates, Completeness, Probabilities . . . . . . . . . . . . . . 248

28 Twenty-Eigth Class: Resolutions of Paradoxes 25128.1 The Location of the Electrons in the Hydrogen Atom . . . . . . . . . . . . . . . . . . 25128.2 Probability Distributions for Excited States . . . . . . . . . . . . . . . . . . . . . . . 25228.3 Angular Momentum and Hydrogen Atom States . . . . . . . . . . . . . . . . . . . . 25528.4 Compatible and Incompatible Quantities . . . . . . . . . . . . . . . . . . . . . . . . . 25828.5 Angular Dependence of States and Probabilities . . . . . . . . . . . . . . . . . . . . . 26028.6 Schwarz Inequality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26128.7 The Uncertainty Principle − Proved . . . . . . . . . . . . . . . . . . . . . . . . . . . 263

29 Twenty-Ninth Class: Four Last Things 26629.1 What’s Special About the z-Axis? . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26629.2 A Brief Basis for Chemical Bonding − Atomic “Orbital” in Different Directions . . . 26729.3 The Necessary “Bohr-Limit” . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26929.4 If an Eigenstate is Forever, Why Do Atoms Radiate? . . . . . . . . . . . . . . . . . . 27029.5 Transition Rates and Spin Angular Momentum of Photons . . . . . . . . . . . . . . . 272

5

Page 7: PHY 416, Quantum Mechanics - SIUE

Chapter 1

First Class: Quantum Mechanics

Tuesday, August 26, 2008

1.1 Introduction

Welcome to Physics 416! Quantum Mechanics is, almost needless to say, anamazing subject! We will, of corse, see many amazing things in studying it.

Let me begin by briefly telling you about an example about the utility ofquantum mechanics for us.

As we learned from Albert Einstein, energy and mass are interchangeable.Also, we now know that it is nuclear fusion that powers the sun. The followingsequence of reactions is believed to contribute to this.

p+ p→ 1H2 + e+ + ν + 1.44MeV (1.1)

1H2 + p→ 2He

3 + γ + 5.49MeV (1.2)

2He3 + 2He

3 → 2He4 + p+ p+ 12.49MeV (1.3)

Notice that this sequence is a chain reaction. It is called the proton-protoncycle, or the “p-p cycle”. But, consider the first step: the protons in the coreplasma of the sun repel each other according to Coulomb’s law. Unless theyare touching (r ∼ 10−15m).

Fig 1We would like you to show that, at touching, the potential energy ∼1MeV

(for a pair of protons, Coulomb P.E.).Do the protons have this much energy? The sun’s core is hot. Ac-

cording to quantum mechanics, we know how hot it is (≈ 15.7×106K. Is this

6

Page 8: PHY 416, Quantum Mechanics - SIUE

hot enough? Recall that, for independent particles in thermal equilibrium,⟨1

2mv2

⟩=

3

2kBT (1.4)

where kB ≡ “Boltzmann’s Constant” = 1.381×10−23J/k = 8.617×10−5eV/k(recal that 1eV = 1.6× 10−19J).

Now, the plasma at the center of the sum is roughly, we shall say, “like”an ideal gas of protons, neutrons, and electrons. Consider a proton in thisplasma. For it⟨

1

2mv2

⟩≈(

8.7× 10−5eV

k

)(1.6× 107K

)∼ 1.4× 103eV 1MeV (1.5)

Conclusion: According to classical physics, even in the core of the sun, itis too cold for nuclear fusion to occur; therefore, the sun should not shine!The protons should not penetrate the Coulomb barrier! Yet, according toquantum mechanics, particles can “appear” in forbidden regions − in fact,we use the theory to predict exactly the rate of this “forbidden appearance”,and from this, the rate of the p-p cycle. From this, we can predict the meanpower output of the sun (∼ 3.8× 1026W ), and from this, the mean expectedtemperature of the earth − and it works well!

Here is another example. Suppose we throw an eraser; it goes from “here”to “there” on a parabolic path.

fig2You say that this follows Newton’s law, but where is this from? Suppose

a particle is observed to be at x1 at time t1, and later at x2 at time t2. Howdoes it get from (x1, t1) to (x2, t2)?

According to orthodox quantum mechanics: Between the observa-tions, the particle does not possess all the attributes of reality. In particular,it does not possess the attribute of position. Thus, there is no definite tra-jectory. Rather, “in potentia”, all possible trajectories, even very wild ones,contribute to the result of the second observation.

fig3As we will see, these trajectories interfere with each other in potentia. As

we will also see, in the limit of macroscopic mass, the interference is destruc-tive for all paths except those in a narrow band around the classical parabolicpath. In this band, the “superposition of possibilities” is constructive.

7

Page 9: PHY 416, Quantum Mechanics - SIUE

fig4Other examples

1. Light incident on glass

2. Photon counting in the lab

Clearly, an understanding of interference is important for embarking tounderstand quantum mechanics. We turn to a quick review of this next. Forinterest, we place this in a historical context.

1.2 Interference: Review of Evidence for the Wave Theory ofLight − Thomas Young’s Interference Experiment

Maxwell’s equations and the theory of electrodynamics provide, as you know,very strong theoretical evidence for the electromagnetic wave nature of light(1865). However, the earlier results of Thomas Young are, from an experi-mental point of view, extremely important.

Background:

1. Newton: Corpuscular Theory of Light (1670’s)

2. Huygen: Early Wave Theory of Light (1670’s)

Situation in the 1700’s: Great weight of Newton’s opinion hung like a shroudover all of Europe.

Along comes 22 year old Thomas Young in 1802. He says that he can provethat light is a wave, and to boot that he can measure its wavelength! Howdid he do this? To understand, we must ask: What properties do waveshave that is (are) different than those of particles? They can interfere e.g.You can have wave1 + wave2 = 0 if the crests of wave1 overlap the troughsof wave2. But, “particle + particle” cannot equal zero. Then Young had toface the issue of how to use this interference idea effectively to distinguishthe two possibilities. for this, he resurrected Huygen’s principle: all points ona wavefront can be considered as point sources for the production of a new(“secondary”) spherical waves. After the elapsing of time, the new wavefrontis the common surface of tangency to these secondary spherical waves.

fig5

8

Page 10: PHY 416, Quantum Mechanics - SIUE

The next ingenious idea was to attempt to effectively isolate only “two”spherical waves that would then be allowed to interfere with each other. It ispresumably, mush easier to keep track of the interference of two waves thanof an infinite number of waves. The idea then is to force the two waves tocombine out of phase with each other by different amounts at different pointson a viewing screen, thus causing the brightness of illumination along thescreen to vary. To accomplish this, he used a setup that forced the two wavesto start out in phase, but to meet out of phase due to an imposed differencein the distance traveled.

fig6Let’s now fix these ideas semi-quantitativelyfig7For P to be a max: s1b = mλ

d sin(θ) = mλ or sin(θ) =mλ

d, m = 1, 2, . . . (1.6)

For P to be a min

d sin(θ) = (m+ 1)λ, m = 1, 2, . . . (1.7)

som yound, knowing d and θ could find λ, thus proving that light is a wave!But, no one believed him.

About what is the wavelength of light? Example: green light

• Slit spacing: d ≈ 0.1mm

• Distance to screen: D ≈ 20cm

• First minimum found at 0.16

sin(θ) =12λ

d=

λ

2d(1.8)

Therefore, λ = 5460A. World’s first (of course) determination of wavelengthof light − now you see how we know it! (micro-world property − amazingfor 1802!)

9

Page 11: PHY 416, Quantum Mechanics - SIUE

1.3 Two-slit interference, Quantitative Calculation of Intensitydistribution at the Screen

At a given point on the screen we have the superposition of say, E1 =E0 sin(ωt) and E2 = E0 sin(ωt+ φ), φ = φ(θ). We wish to add these:

Ep = E1 + E2 = E0 [sin(ωt) + sin(ωt+ φ)] (1.9)

By using the trigonometry identity

sinA+ sinB = 2 sin

(A+B

2

)cos

(A−B

2

)(1.10)

by using this identity, one can obtain

Ep = 2E0 cos

2

)sin

(ωt+

φ

2

)(1.11)

The 2E0 cos(φ2

)term depends only on φ, which depends on θ, so this is an

amplitude that varies a P is moved along the screen. The sin(ωt+ φ

2

)term

is a simple harmonic variation at the same frequency ω. Of course, the eyeresponds to intensity, not field. Recall that, for the electromagnetic wave,the intensity depends on the magnitude of the Poynting vector, ~S, where

~S =1

µ0

~E × ~B (1.12)

For an electromagnetic wave in free space, ~B ⊥ ~E and B = Ec , so

~S =1

µ0cE2 (1.13)

Now S2 varies in time according to its factor sin2(ωt + α), which oscillatesat the frequency, ω − some 1015Hz − far to fast for the eye to follow. So,what we need is the time average over the resolution time of the eye. Sincethe variation is simple-harmonic, all cycles are the same, so it is suffices toaverage over one cycle. Thus, intensity at point P on the screen is

IP =1

µ0c4E2

0 cos2(φ

2

)⟨sin2(ωt+ φ)

⟩t

(1.14)

10

Page 12: PHY 416, Quantum Mechanics - SIUE

We drop the 1µ0c

since we will be taking ratios for which it will cancel. Now,⟨sin2(ωt+ φ)

⟩t

= 12 , so

IP = 4E20 ·

1

2cos

2

)(1.15)

Now again, φ = φ(θ). How? To find out, set up the ratio

φ

2π=p.l.d

λ⇒ φ

2π=d sin)θ)

λ⇒ φ(θ) =

2πd sin(θ)

λ(1.16)

thus, in its full glory (except for the supposed factor 1µ0c

),

IP =1

2· 4E2

0 cos2(πd sin(θ)

λ

). (1.17)

It is often of important interest to compare the result for IP when both slitsare open to the intensity at P when only one slit is open and one is blocked.For this purpose, defineI0(θ) to be the intensity at P (θ) with only one slitopen and the other blocked. Then

I0(θ) =⟨E0 sin2(ωt)

⟩t

=1

2E2

0 , (1.18)

thus, the intensity with both slits open is

IP = 4I0 cos2(φ

2

)(1.19)

This is interesting and very important. Let’s plot it:fig8As we say, this is very important: If I0 is the intensity with just one slit

open, note that we do not get 2I0 = I0 + I0 (as you might expect)1 byopening both slits. Rather, I(θ) varies between 0 and 4I0 as θ changes. It’sas if 1 + 1 = 4 or 1 + 1 = 0!

1You do get this if you use a flashlight instead of a laser as a source. Why is this?

11

Page 13: PHY 416, Quantum Mechanics - SIUE

Chapter 2

Second Class: More on the Behaviorof Light

Thursday, August 28, 2008

2.1 Introduction

The wave picture of light, coupled with the interference idea, already leadsus to conclusions like the possibilities of 1 + 1 = 4 and 1 + 1 = 0, whichmay seem natural in the wave picture, but which become very strange in thephoton picture: it is the job of quantum mechanics to explain this.

There are other really amazing things. These occur with nearly monochro-matic light (which is freed from the mixings and jumblings of everyday light).The laser is a good source of such light.

Suppose, for example, you illuminate a ball bearing with a indexlaser laser.With “true light”, right at the center of the shadow, where you expect thingsto be the darkest (the “penumbra” of the shadow with ordinary light), thereis a bright spot! This is evidence that true light does not “travel in straightlines”, as we are led to believe! Some must have curved to get to the center!

fig1-5These results are rather amazing! We do not expect these from geomet-

rical optics. Maybe you don’t believe them. That’s o.k., you’ll do it in labsometime. Return to the shadow of the ball bearing, note the bright spota the center − certainly weird. In 1818, Fresnel, a young Frenchman, (29years old) presented an essay to the Paris academy. The judging committeehad many of the prominent physicists of the day, including, Poisson, Arago,

12

Page 14: PHY 416, Quantum Mechanics - SIUE

Laplace, and Biot. Poisson, an advocate of the corpuscular theory, workedout Fresnel’s results and found the bright spot. This resulted in the cor-puscular theory dying in one day, and was not revived again until the 20th

century.Now back to the “expanding” behavior of light. We can get a first inkling

from the following example. As a definite case, consider monochromaticplane waves incident on a single aperture of width a in one dimension (andwidth a in the perpendicular direction, which is in and out of the page infigure 2.6). This figure shows what we call diffraction, which is really justthe interference of many circular waves according to Huygen’s principle.

fig6

2.2 Single Slit Diffraction, continued

Semi-quantitative ConsiderationWe could calculate the diffraction pattern quantitatively by summing (or

integrating) over the infinite number of infinite similar-spaced points of emis-sion along the slit. However, before we get involved, in detailed mathematics,let us see if we can figure out what we expect the answer to look like (roughlyat least) in advance. Where possible, this procedure is always a good idea.

fig7The point Q at the center will clearly be bright (Why?). Now consider a

point P such that the following happensfig8Each point on top half is canceled by a point in the bottom half, therefore,

we have found a minimum. So, a sin(θ) = mλ are minima. Somethinghappens at mλ. Now consider an angle θ such that the following is true

fig9I cancels II, and the light reaching the screen at P is effectively 1

3 of thetotal light from the slit. Therefore, we expect a maximum there1, but not abright as the central maximum. by this kind of preliminary logic, we expectan intensity pattern that looks something like this:

fig101At least roughly midway between the first and second minima. It turns out that, in fact, the maxima are not

exactly given by a sin(θ) =`m+ 1

2

´λ. This will be addressed in the homework.

13

Page 15: PHY 416, Quantum Mechanics - SIUE

We now attempt to consider this more quantitively. The paradigm prob-lem: A plane wave is incident on a small2 aperature.

fig11We want the calculate the result at P . To model the problem, try breaking

the aperture into an infinite number of infinitesimal “subapertures”; eachcontributes a phasor of infinitesimal amplitude to P . So we must “add up”(or integrate) the contributions from all of the little “subslits”. This meansthat we have to “add up” a very large number of harmonically varying electricfields (each with a different phase) at P . Although technically this is anintegral, E(P ) ∝

∫dy sin(ωt + φ(y)), we will develop a neat picture method

insteadfig12

2.3 Method of Phasors

In order to understand our results, in a quantitative way, we will need analternate, powerful method for adding sinusoidal oscillations of the samefrequency. You will recall that we did this via “trigonometry identity” forsin(ωt)+sin(ωt+φ) for double slit interference. Let us consider that problemagain. Consider an electric field, E1 = E0 sin(ωt). This is the “vertical com-ponent” of the rotating vector in figure 13.a. Likewise, E2 = E0 sin(ωt + φ)is the “vertical component” of the rotating vector in figure 13.b. Considerthen, the construction shown in figure 13.c. A moment’s thought convincesyou that the vertical component of ~E, the vector sum, is the magnitude ofER = E0 sin(ωt) + E0 sin(ωt + φ), which is what we want. Further, to findthe phase of ER, we can use simple geometry!

fig13a,b,cfig14We see that α = φ

2 . Also, using the law of cosines, the magnitude,

E2R = E2

0 + E20 − 2E2

0 cos(180 − φ)

= E20(1 + cos(φ)

= 4E20 cos

2

)2As we saw earlier, “small” means small compared to the wavelength of light

14

Page 16: PHY 416, Quantum Mechanics - SIUE

so,

ER = 2E0 cos

2

)(2.1)

So

EP = sin(ωt) + sin(ωt+ φ)

= 2E0 cos

2

)sin

(ωt+

φ

2

)just as we had figured earlier by trigonometry identity. The great advantageof this technique is that it is easily extended to the addition of many har-monic oscillations of the same frequency − this would be very painful by trigidentity, but if all else fails, with this method, you could do it graphically. Itwas invented by M Co?????, circa 1880.

fig 15For our “single slit” problem, we will have essentially, to add together an

infinite number of phasors, which is a tantamount to doing an integral “bypictures”. In the following, we’ll sketch how this works.

Now let us look at our “single slit” problem at the next level. We use thephasor method. An outline of this fllows:

• Divide slit up into many “elements”

• Adjacent elements have phase difference ∆φ

where,∆φ

2π=p.l.d

λ⇒ ∆φ =

λ∆x sin(θ) (2.2)

fig16Now let’s look at the situation in the θ = 0 direction. there, all path length

are the same, so all the phasors point the same way and add up to somethingbig − a very bright max of intensity.

fig17Now we move P slightly up (or down) the screen. Then each successive ray

has a slight path length difference from the previous one, so each successivephasor at P is inclined at a slightly greater angle. Thus, the resultant (Eθ infigure 18) is not as big as when all the phasors point in the same direction.

15

Page 17: PHY 416, Quantum Mechanics - SIUE

Figure 2.1: Figure 21: The intensity as a function of θ at point P .

fig18And if θ increases further, so does δφ − until we reach a point when the

phasors close in on themselves!fig19Then we have a minimum. And if we increase θ further,fig20,we reach a secondary maximum, with clearly less resultant than the straight

ahead max. And so on, a pattern of maxima and minima, with the maximagetting dimmer and dimmer. our expectations from this again predict a curvethat looks something like what is shown in figure 2.22.

But we still need to be more quantitative. Now we make a fully quantita-tive calculation − we “integrate by picture”. To see how this goes, considerthe phasor situation at arbitrary θ

fig22

Eθ = 2R sin

2

)(2.3)

16

Page 18: PHY 416, Quantum Mechanics - SIUE

but S = rθ, so Em = RΦ and R = EmΦ so

Eθ =2Em

Φsin

2

)⇒ Eθ = Em

sin(Φ

2

)(Φ2

) (2.4)

But, what is Φ?Φ

2π=p.l.d

λ⇒ Φ =

λa sin(θ) (2.5)

So our final result for single-slit diffraction is

I(θ) = I0

[sin(Φ

2

)(Φ2

) ]2

, whereΦ

2=πa

λsin(θ) (2.6)

The function(

sin(x)x

)is called sinc(x), so this is

I(θ) = I0sinc2[πaλ

sin(θ)]

(2.7)

fig23but, what is meant by “I(θ)”? of course, the energy sent off at exactly

angle θ is zero. What we mean is I(α)dα = energy diffracted into the in-finitesimal angular range dα centered on value α, where

α ≡ πa

λsin(θ). (2.8)

We see that this function has the expected maxima and minima. What areare minima? We wish sin

(Φ2

)= 0 so Φ

2 = π, 2π, 3π, · · · 3

πa

λsin(θ) = mλ m = 1, 2, 3, · · · (2.9)

Therefore, the condition for minima in single slit diffraction is given by

a sin(θ) = mλ (2.10)

which is as we anticipated.Most Important 1 As it is obvious from the plot, most (almost all) of

the light falls within θ < θfirst min. So, the angle of the first minimum isespecially important. It is given by

θfirstmin = sin−1(λ

a

)(2.11)

3not Φ2

= 0, since Φ2

is also in the denominator

17

Page 19: PHY 416, Quantum Mechanics - SIUE

Most Important 2 Note how this depends on the ration of λ to a. Threecases are plotted in figure 2.24.

fig24Most Important 3 Remember the figure from earlier where it showed the

waves diverging from apertures of widths a = λ, a = 3λ, and a = 5λ? Makesure that you understand the correlation between the situations shown in thatfigure and those shown in figure 24. This shows the effects of increasing anddecreasing λ and a relative to each other. As you will see, that understandingis extremely important. A good rule of thumb to remember is the half widthof the central maximum is ∼ λ

a .

2.4 Position Resolution − Why ∆x ∼ λ

If we look at a very small object with light (or other electromagnetic radia-tion), the spatial resolution, ∆x is approximately λ. This limiting resolutionis due to diffraction! Let’s understand this. We saw that, for monochromaticplane wave radiation incident on a long aperture of width a (single slit), theenergy spreads in angle (diffracts), but is mostly contained within an angularrange ∆θ = 2 sin

(λa

)(the range between the first zeros). If the aperture is

circular, the details of the results are different, but the main result is verysimilar − you get a diffraction pattern consisting of bright and dark bands(in this case, concentric rings). If the screen is far enough from aperture, thesituation is called “Fraunhofer diffraction” results − in this case, for a circu-lar aperture, you get a very strong and large central maximum surroundedby alternating light and dark rings. This is shown in figure 25.

fig25For a circular aperture, under Fraunhofer conditions, the first minimum

occurs, not exactly at θ = sin−1 (λa

), but rather at

sin(θ) = 1.22λ

d(2.12)

where d is the diameter of the aperture. Now suppose you look at somethingwith a lens − say you look at a point-like source (i.e. star at great distance)with a telescope. Say we concentrate on one wavelength component. Then,because of its very great distance, the light from the star is essentially a

18

Page 20: PHY 416, Quantum Mechanics - SIUE

plane wave at the telescope’s objective lens. This lens effectively “cuts out”a circular part of the plane wave fronts, passes them on, and rejects the rest.thus, in addition to its properties related to geometric optics, the lens actslike an aperture. Thus because its diameter is not infinite, the image formedon the “screen” is not a point, but rather is a somewhat spread out diffractionpattern.

fig26Limit of Resolution This raises an interesting issue in resolution: sup-

pose we look at two distance point sources. Then, if they are too closelyspaced, if we look at both, they will not be individually resolved.

We will take as a resolution criterion “Rayleigh’s criterion” (actually, onecan do a bit better) − two objects are barely resolved if the central maxi-mum of the diffraction pattern of one source falls on the first minimum ofthe diffraction pattern of the other. Thus, by this criterion, the minimumresolvable angular separation, or the angular limit of resolution is

∆θmin = 1.22λ

d(2.13)

for aperture size d.fig27Let us look at this in a bit more detail. Suppose we have two objects (O1

and O2) to be imaged by a lens. Let the separation between them be δ; letthe central maxima of their image blurs be centered at I1 and I2, respectively,as in figure 2.28. So, how far apart do O1 and O2 have to be for their distinctimages I1 and I2 to be resolvable? We use Rayleigh’s criterion − that themaximum of I1 coincide with the first minimum of I2. This means that O2AI1

differs in path length from O2BI2 by distance λ, To see where this conditioncomes from, recall that at the first minimum of a diffraction pattern, say froma rectangular slit, the rays coming from the extreme ends of the slit differ inbath length by λ (with this, dividing the slit into two zones ensured that foreach ray in zone I, there is a corresponding ray in zone II that is π radiansout of phase with it).

fig28Now, from figure 227, O2A is shorter than O1A by length δ sin

(α2

), and

O1A = O2B, and O1B is shorter than O2B by δ sin(α2

). Thus, O2B−O2A =

19

Page 21: PHY 416, Quantum Mechanics - SIUE

δ sin(α2

)≈ O2BI1 −O2AI1.Thus for O1 and O2 to be resolvable,

2δ sin(α

2

)≥ λ⇒ δ ≥ λ

2 sin(α2

) (2.14)

Thus, the limit of resolution is

∆x =λ

2sin(α

2

)(2.15)

of course, we can make this as small as we want by either using λ as small aspossible (gamma rays) or by making α large (big diameter lens). However, ifwe insist on using long wavelength light, we have a price to pay in resolution.Since the largest α

2 can be is π2 , we see that

∆x ∼ λ (2.16)

as claimed.

20

Page 22: PHY 416, Quantum Mechanics - SIUE

Chapter 3

Third Class: Quantum Bahavior

Tuesday, September 2, 2008The notes for this section are taken directly from The Feynman Lec-

tures on Physics, Volume III, Chapter 1 (Addison-Wesley, c©1965.

21

Page 23: PHY 416, Quantum Mechanics - SIUE

Chapter 4

Fourth Class: Double Slit andQuantum Mechanics

Thursday, September 4, 2008

4.1 Double Slit Experiment

Let us recall some features of the double slit experiment with electrons:fig1Recall salient features

1. Hear clicks from a detector: no half clicks

2. Move detector around: rate of clicking changes, but still no half clicks

3. Lower gun temperature: clicking slows down, still no half clicks

4. Use two detectors at once: never hear both go off simultaneously

Conclusion: Upon detection, the electron shows up like a particle − at onepoint in space, not spread out in space.

fig2As we will see, in the case with photons, the “final picture” is not dimly

present during the intermediate stages (as it would for a wave-caused buildup).Rather, each electron is detected at one place according to a probability rule.Yet, P12 results! P12 is a statistical pattern − it results from the accumu-lation of data over time − even if (and especially for the purposes of thisexperiment) the gun only sends one electron at a time.

22

Page 24: PHY 416, Quantum Mechanics - SIUE

Therefore, if we send in a single electron, we can’t predict definitely whereon the screen it will wind up. From P12(x) we can only predict the relativeprobability of winding up at different points. this illustrates the central fea-ture of quantum mechanics − we cannot, even in principle, predict exactlywhere a given electron will wind up on the screen. We can only talk aboutrelative probabilities. This is in marked contrast to classical physics wherein principle, we could have followed the trajectory of a given bullet and thenused Newton’s laws to predict the subsequent trajectory.

Between creation at the electron gun and detection at a point at the screen,the electron seems to not exist as a particle, but propagates in accordancewith the mathematics of wave motion. That is, it propagates as if it isstatistically guided by a wave. (We will see later that the “wave” does notreally exist as a spread out physical entity in space − it is a mathematicalconstruct).

It is important to realize that over pattern P12, though collected over manytrials, reflects the interference of the wave function of a single electron withitself. Let us denote this “fictitious” or “proxy” wave function “guiding”the electron as Ψ(x, y, x, t). That is, we will hypothesize that, before theposition measurement at the screen, the electron is “represented” by the”state function” (or “wave function”) Ψ(x, y, z, t). After doing the experimentwith a very, very large number of electrons and collecting the data, we willthen hypothesize the following interpretation:

Let dx be a very small distance along the screen centered at positionx. Then we will say that the probability that any electron in a singletrial of the experiment will materialize in dx, P (x)dx, is proportionalto the intensity in that small region. In particular, we will take as aworking postulate that:

P (x)dx = |Ψ(x)|2 dx (4.1)

where P (x) is the probability for a single electron to materialize in dx centeredon x.

This postulate (“Born probability rule”) is clearly suggest by the resultsof the double-slit experiment for electrons.

23

Page 25: PHY 416, Quantum Mechanics - SIUE

4.2 Orthodox (“Copenhagen”) Interpretation

As we say, before (in between) position measurement(s), the electron behavesas if “represented” by a wave function Ψ(x, y, z, t). But there is, as we say, acrucial difference between this and the “water waves” case − as we say, Ψ isjust a mathematical construct − for electrons, there is no real physical wavefunction in three dimensional space.

Further, the electron double-slit experiment results are consistent with theCopenhagen interpretation that we cannot say an object (“electron”) even

posses any definite position or partaken of any definite trajectory unless ameasurement occurs (e.g. plug up one of the holes or use a light source to forcea through only the one aperture). Likewise, according to the Copenhageninterpretation, the act of detection at the screen is a “position-materializingmeasurement” that forces the electron to “pick” a position along the screento materialize at. In the words of another author1 “If we cannot know thelocation of a particle until we actually look for it, it is hard to justify the claimthat it even has a location until we look for it.” this brings us to a major“philosophical” point in the modern consensus interpretation of quantummechanics.

According to the standard modern (“Copenhagen”) interpretationof quantum mechanics, generally the particle does not possess aposition until the act of measurement by detector forces to “mate-rialize” at a fixed position (at a point − not spread out).

This is because− according to current standard quantum mechanics,the electron does not exist in full reality between measurements withdetection devices. It’s not that the electron between detections issomewhere, but we just don’t know where it is − rather, generally,between detections the electron is “nowhere” because itdoes not exist in full reality.

This is a definite break with all classical Western thought.As Harris explains: “Early in the quantum age, many physicists, most no-

tably, Albert Einstein himself, asserted that [the theory must then simply] beincomplete, that some modification is needed to allow ‘real’ quantities, such

1R. Harris, p. 122 of Modern Physics, 2nd Ed. (Pearson Addison-Wesley)

24

Page 26: PHY 416, Quantum Mechanics - SIUE

as [the] position [of a particle] to have definite values at all times. However,the modern consensus, known as the Copenhagen interpretation, is that untilexperiment actually localizes it, a particle simply does not have a location.”

Again, it is not that the electron is “somewhere” or even really phys-ically spread out in space. Rather, we say that before the detectionthe electron generally does not possess even the concept ofposition (“partial reality”). In the orthodox quantum mechanicstheory it is believed that the act of measurement forces the previ-ously only “non-objectively real” electron to materialize in positionalreality, selecting a position in the real world according to a certain“preassigned” probability function, |Ψ(x, y, z, t)|2.

Its as if, before the detection measurement the electron is like a ghost holdinga deck of cards, each card labeled with position in the real world. Consideran interval of positions dx starting at x = xa, say then the number of cardsholding this position in this interval is weighted compared to the numberof cards holding positions in other intervals of equal width dx accordingto the electron’s preassigned function ψ∗(x)ψ(x) ≡ |ψ|2. The measurementapparatus then says to the ghostly electron − “Hurry and pick one of thecards at random − then appear in the world of humans at that point!”

Example: Suppose that before measurements, the state function of anelectron is, in some region, say 0 ≤ x ≤ L, at time t is

Ψ(x, y, z, t) = A sin(kx)e−iωt (4.2)

for some constants k and ω2. Where can the electron represented by this statefunction be forced into materialization, and with what relative probabilities,say at t = 0? Answer: Probability of materialization in a small interval, dxcentered on x at time t is:

P (x, t)dx = |Ψ(x, t)|2 dx (4.3)

This is an example of a general rule:2At the moment, we do not worry about where this state function “comes from” or under what conditions it is

valid. Of course, these are very important questions. We will deal with them at length later.

25

Page 27: PHY 416, Quantum Mechanics - SIUE

4.3 Probability Rule ( Born, 1927)

Suppose that, between measurements of positions, the particle is representedby the state function, Ψ(x, y, z, t). Then, upon a position measurement attime t, the probability that the particle will materialize in the small volume(dx dy dz) centered on the point (x, y, z) is

P (x, y, z, t) dx dy dz = |Ψ(x, y, z, t)|2 dx dy dz. (4.4)

In our example,

P (x, t = 0) = |Ψ(x, t = 0)|2 = |A|2 sin2(kx) (4.5)

since e−iω0 = 1. We plot the relative probability to materialize at differentpoints.

fig1If governed by this state function, the particle will definitely not ma-

terialize at x = 0, x = πk , x = 2π

k , · · · , In addition, the most likely places tomaterialize are x = π

2k,3π2 k,

5π2 k, · · · .

4.4 Normalization Condition

Now, the total probability of the particle’s materializing anywhere in spaceon position measurement must be one (i.e. 100%). This, with Born’s rule,leads to the normalization condition, leads to the equation∫ +∞

+∞

∫ +∞

+∞

∫ +∞

+∞|Ψ(x, y, z, t)|2 dx dy dz = 1. (4.6)

Adjusting unspecified constants in a candidate state function in called “nor-malizing” the state function; you will do an example of this in the homework.

4.5 Analogies with light

Let us now look at some experimental results with light3. As you know, insome experiments (e.g. double-slit interference, diffraction, etc.) light showswavelike properties.

3By “light”, I mean monochromatic, coherent light.

26

Page 28: PHY 416, Quantum Mechanics - SIUE

From a reading of Einstein’s 1905 paper, it appears (though it is notexplicitly so stated) that by then he viewed a quantum of definite energy asbeing localized in space and used this to explain the photoelectric effect. By1909, he seems to have arrived at a view that the localized energy packets(“photons”) are guided statistically by the “associated” classical EM wave.

In this picture, which preceded the Copenhagen view, the intensity ofradiation striking a surface is proportional to the number of photons incidentper unit area per second on the surface. (Recall that I =

⟨ energyarea·sec

⟩and the

energy of each photon of frequency f is fixed at value hf).In this view, at any instant of time, the number of photons present in

an (infinitesimal) volume, dV of space is proportional to the square of theelectric field value there. Part of the reasoning leading to this: recall that,classically, the energy in dV is nhf (we assume a single frequency f), wheren is the number of photons. Thus, n(x, y, z, t) ∝ E2(x, y, z, t).

Thus, if a region of large E2 moves through space (as around say a crestof the E field in a traveling sinusoidal EM wave), then the photons tendto follow along. Thus, this pre-Copenhagen view of Einstein’s is a sort of“pilot-wave”/particle viewpoint − in it, the photon really exists all alongas a localized bundle (“particle”) and is statistically guided by the “reallyexisting” associated wave.

Superficial support for this view comes from noting that detections in thedouble-slit experiment with the photons essentially come at “random” placesone at a time, as is shown in figure 2, rather than from a dim version ofthe entire final pattern that gets progressively brighter, as it would if theclassical EM wave theory were correct. Further support at this level seems tocome from the development of images built up progressively again, the initialstages show seemingly randomly placed dots, rather than a dim replica of thefinal image.

fig2Further evidence for the particle-like nature of light seems to come from

further consideration of the double-slit experiment. If we line the screen withseparate photon detectors, sending a very low level of light in, we find thattwo detectors never fire at once.

fig3

27

Page 29: PHY 416, Quantum Mechanics - SIUE

This is further evidence that, in interaction, light is not a wave. However,if we let the experiment accumulate data, we find a predictable statisticalpattern of strikes. In particular, the relative number of strikes in any detector(compared to the others) when accumulated, follows the distribution expectedfrom interference waves − e.g., there are detectors that never fire (located atpositions of “destructive interference”).

fig4However, most4 physicists now believe in accordance with the Copenhagen

view: Although the pictures in figure 4 show that very dim light is detected“like a particle” (“photons”), the simple ‘existing all along particle guidedby an existing “pilot wave” ’ is incorrect. Nowadays, this pilot-wave picturecould be called a “semiclassical picture”.

Although the orthodox Copenhagen interpretation has been mainstreamsince 1927-1928, it is only quite recently that we are approaching direct ex-perimental evidence for it! Consider the experiments shown in figures 4.5through 4.8.

First a “wide angle version of the double-slit experiment”:

1. “Single-Photon, Wide-Angle Experiment5”

fig5

2. Anti-coincidence Experiment: Consider the following experiment per-formed in 1986.

fig6

This is a “space-separated” analog of our “multiple detector detectors onthe screen” thought experiment with the double-slit apparatus. Again,one photon is sent in at a time. Result: for any given event, either D1

fires or D2 fires, never both. This experiment shows the “particle natureof light”.

3. Interferometer Experiment6

a Showing wavelike behavior4But not all5Lie and Diels, Journal of Optical Society of America B9, page 2290 (1992). The brief discussion here is from

Fundamentals of Physics, 8th Ed. by Halliday, Resnick, & Walker, pp. 1067-8.6Discussion taken from Physics, 5th Ed., by Halliday, Resnick, and Krane, p.p. 1024-1025

28

Page 30: PHY 416, Quantum Mechanics - SIUE

fig7

b Showing Particle-Like Behavior

fig8

c Delayed Choice Experiment: Experiments bearing on this very ques-tion have recently been done!

fig9

What can we conclude from this? Quoting7 another author.fig10Thus, the probability wave has no physical existence. It is, so to speak, an

artificial calculation device. Nor can the “photons” have a physical existence,like that of a little ball, while they are propagating (“in transit”). We willreturn to these issues later in the course.

7Amit Goshamu in Quantum Mechanics, 2nd Ed., p. 114 (Custom Publishing)

29

Page 31: PHY 416, Quantum Mechanics - SIUE

Chapter 5

Fifth Class: What Would Be a ValidState Function For a Free Particle?

5.1 Some Experimentation

We have seen interference effects with single electrons incident on a Youngdouble-slit experiment (or its equivalent). So far as we now know, the electronis a fundamental particle, not composed of other “smaller” constituent parti-cles. The questions then arises − would a composite system also show clearlypredictable simple interference in a Young double-slit apparatus? Considerfor example, the neutron, which is a composite bound state of quarks. Yet,they show predictable (according to their deBroglie wavelength) interferenceand diffraction:

fig1fig2Even entire atoms run through a double-slit apparatus show interference.

This works for Neon and other relatively light atomsfig3However, for larger particles, interference is difficult to observe for at least

two regions

1. The associated deBroglie wavelength becomes so short that the dimen-sions of the apparatus become impractically small

2. Decoherence effects (which quickly wash out “quantum behavior” andcause a system to behave in a Newtonian way) become very strong.

30

Page 32: PHY 416, Quantum Mechanics - SIUE

Let us recall deBroglie’s proposal − it is that, associated with a free1

particle of (definite) momentum, ~p, and (definite) energy, E is a wavelengthand frequency given by

λ =h

p(5.1)

and

f =E

h. (5.2)

Since ω ≡ 2πf and k = 2πλ , these are p = hk and E = hω, where h ≡ h

2π .These relations argue that a plane wave is to be associated with a free

particle of definite energy E and definite momentum ~p. Of course, fromthe results of our interference experiments, we now know (at least in theorthodox Copenhagen interpretation) that the associated wave does not havea physical existence in real physical space, but rather is a probability wave( “state function”). Thus, we tentatively assume that the state functionassociated with a free particle of definite (E, ~p = px) is

Ψ(x, t) = A cos

(p

hx± E

ht

), (5.3)

the choice of the sign depending on the direction of the momentum. Noticesomething interesting about this “wave function”, it fills all space. (Its wave-fronts are planes perpendicular to x). This means that the position associatedwith such a state is “completely unknown” (or better, “completely does notexist”). On the other hand, for such a state (of nonobjective reality), themomentum, and hence the speed, is precisely known (partial reality state).

Now, if this plane wave (even though its not a real wave that actuallyexists in physical space − perhaps, we should call it the “proxy-wave”) rep-resenting the state of a free electron of definite energy, E is to be a reasonablerepresentation of the state, then its velocity should match the “velocity ofthe particle” calculated from vp = p

m . Consider, though, as you know, the(proxy) wave velocity is

vφ = fλ =E

h· hp

=E

p=

√c2p2 +m2c4

p= c

√1 +

(mc

p

)2

> c. (5.4)

1By “free” particle, we mean one that is not subject to any forces of course, this is an idealization.

31

Page 33: PHY 416, Quantum Mechanics - SIUE

We seem to have a problem − the particle somehow “can’t keep up with”the state function that is supposed to represent it. We must deal with thisproblem.

5.2 Partially Localized States

Another issue (which will turn out, as we’ll see) to be related, also presentsitself: While our deBroglie plane wave is useful as an idealization, it is “veryrare” that the position of the particle it is to represent is completely nonex-istent. Even in the most extreme case − if the universe is finite, the regionof unknown position is also no infinite. thus, the plane state function is anidealization, never really occurring in practice. More typically, the position,while not defined within a certain range, is known to be zero (or very closeto zero) outside this range. That is to say, upon the measurement, in mostsituation, the particle will not materialize “just anywhere”, but only at apoint within a certain range2.

what would be a valid proxy wave representing a “moving, partially local-ized state?” Consider, for example, the possibility of a proxy wave that is amoving pulse, say

Ψ(x, t) = Ae−(x−vt)2

2σ2 (5.5)

Now, the function

f(x) = Ae−x2

2σ2 (5.6)

is a “Gaussian”, or “bell-shaped” function of standard deviation, σ.fig4Then, by the rule for making rigidly moving pulses,

Ψ(x, t) = Ae−(x−vt)2

2σ2 (5.7)

would represent this Gaussian shape moving rigidly down the x-axis at speedv.

Perhaps this is the sort of probability state function we should use inthe theory to represent a free particle? It is the nice feature that adjusting σ

2e.g. the electron is somewhere within the resolution of a measuring instrument, etc.

32

Page 34: PHY 416, Quantum Mechanics - SIUE

adjusts the range of “unknownness” in the position. Interestingly, the answerto this question is no!

As we will see, not all functions are valid state functions in quantummechanics − and this one is not. (As we will see, valid state functions mustbe solutions of a partial differential equation analogous to, but not the sameas, the classical wave equation.) Since we do not yet (in this course) knowthis differential equation, we must look for another route to localized statefunctions for moving particles.

For this, we invoke the principle of superposition, which says that,under the right conditions, the sum of two valid state functions should alsobe a valid state function − evidence for this is the interference pattern we getat the screen in a double-slit electron experiment. (ψt = ψ1 +ψ2 is a solution;|ψt|2 = |ψ1 + ψ2|2, which shows the interference “cross term”.)

So, to build more complicated solutions, which will hopefully lead to adegree of localization, we begin by superposing two of our basic deBroglieplane waves solutions, each corresponding to a definite, but different value ofmomentum.

Ψ(x, t) = A cos

(p1

hx− E1

ht

)+ A cos

(p2

hx− E2

ht

)(5.8)

Ψ(x, t) = A cos (k1x− ω1t) + A cos (k2x− ω2t) (5.9)

We note the trigonometry identity,

cos(α) + cos(β) = 2 cos

(β − α

2

)cos

(β + α

2

)(5.10)

which tells us that

Ψ(x, t) = 2 cos

(k2x− ω2t− k1x− ω1t

2

)cos

(k2x− ω2t+ k1x− ω1t

2

)(5.11)

Ψ(x, t) = 2 cos

(∆k

2x− ∆ω

2t

)cos(kx− ωt

)(5.12)

where ∆k ≡ k2 − k1, ∆ω ≡ ω2 − ω1, k = 12 (k1 + k2), and ω = 1

2 (ω1 + ω2).Let us consider the case where ω2is only slightly greater than ω1 and k2 is

only slightly greater than k1. Then:

∆k k2 or k1, ∆k k (5.13)

33

Page 35: PHY 416, Quantum Mechanics - SIUE

∆ω ω2 or ω1, ∆ω ω (5.14)

Then:

Ψ(x, t) = 2 cos

(∆k

2x− ∆ω

2t

)cos(kx− ωt

)(5.15)

The 2 cos(∆k

2 x−∆ω2 t)

term is sinusoidal relatively long wavelength traveling“envelope” and the cos

(kx− ωt

)term is a higher frequency “inside envelope”

oscillations.fig5,6These are “groups” of wave. What is the speed of the wave? It is

∆ω2

∆k2

=∆ω

∆k= venv (5.16)

This is not the same as the speed of the motion of the “inside oscillations”,which is

vinside =ω

k. (5.17)

This is still not very localizing, though − the groups repeat in space. To dealwith this problem requires a bit more work, which we will do next.

5.3 A Single Moving Group − “Wave Packet”

It turns out3 that if we suppose more (an infinite number) of sinusoidal trav-eling waves, we can arrange that there is basic constructive interference over asmall ∆x and essentially total destructive interference everywhere else. Thatis, we get one moving group only. Such a moving group is called a “wavepacket”.

fig7This happens if you superpose an infinite number of different sine waves

that span only a finite interval of wave numbers. That is, if δk is very small,3see or hear a course on a wave physics − e.g., late in the semester in PHY 301. We will also probably have a

derivation of this a bit later in this course.

34

Page 36: PHY 416, Quantum Mechanics - SIUE

(likewise δω is small),

Ψ(x, t) = A cos(k1x− ω1t) + A cos([(k1 + δk)x− (ω1 + δω)t]

+ A cos [(k1 + 2δk)x− (ω1 + 2δω)t]

+ A cos [(k1 + 3δk)x− (ω1 + 3δω)t]

+ · · ·+ A cos[k2x− ω2t],

where ω = ω(k). We really want the limit of this as δk → 0, but k1 and k2

remain of the finite separation (in k). Then,

Ψ(x, t) = A

∫ k2

k1

cos(kx− ωt) dk (5.18)

This gives us our wave packet.fig8The integral above is called a “Fourier integral4” In it, ω is a function of

k. How so? We have

E =1

2mv2

particle =p2

2m(5.19)

but, according to deBroglie,

E = hf = hω

p =h

λ= hk

so,h2k2

2m= hω (5.20)

This leads to the “dispersion relation” for deBroglie waves for a free particle,

ω =hk2

2m(5.21)

4Parts of the more general fourier integral will be developed later in the course; for now, we are drawing onbackground from your previous exposure to the general concept from a “modern physics” course.

35

Page 37: PHY 416, Quantum Mechanics - SIUE

As is shown in a waves course (e.g. PHY 301), the group (region of construc-tive interference) moves a velocity

vgroup = limδω→+0

δω(k)

δk=dω(k)

dk(5.22)

This is called the group velocity (Vg).

vg =dω(k)

dk(5.23)

Let us, following deBroglie, evaluate this “group velocity”

vg =dω

dk=

2hk

2m=

2p

2m(5.24)

since p = hλ = hk

vg =p

m=mvparticle

m= vparticle!! (5.25)

I believe it was Einstein who called this, “the most amazing result of alltime”!

5.4 Some Problems

Let us now return to considering the idealized basic plane-wave state functioncos(phx−

Eh t+ φ

). Several problems occur with it; among them are:

1. A free particle represented by this state function ought to have equalprobability to materialize anywhere along the x-axis. But, as we’ve seen,this plane wave state function, at any time, represents bumps and valleysin spatial distribution of materialization probability.

2. Suppose we attempt to describe a free particle with momentum in the+x direction by, say

Ψ→(x, t) = A sin(kx− ωt), k =p

h. (5.26)

Now, invoking the superposition principle,

Ψ(x, t) = Ψ→(x, t) + Ψ←(x, t) = A sin(kx− ωt) +A sin(kx+ ωt) (5.27)

36

Page 38: PHY 416, Quantum Mechanics - SIUE

then must be a possible state. By expanding out the right hand side,you can easily show that this is

Ψ(x, t) = 2A sin(kx) cos(ωt) (5.28)

(which is a standing wave). But, this is clearly not satisfactory, for att = π

2ω , this Ψ(x, t) vanishes for all x, thus, violating the normalizationcondition! (A similar problem exists for Ψ← = A cos(kx+ ωt).)

Thus, we can only conclude that

f(x, t) = A cos(kx± ωt+ φ) (5.29)

is not a valid free particle state function!

5.5 Toward a Governing Equation

So, we now have a catalog with zero valid state functions. Perhaps findingan equation that governs all state functions will help us find some that arevalid.

Here is another reason why we need to have a governing equation for astate function:

As we will discuss in a bit more detain later on, a measurementchanges a state function. For example, a “perfectly accurate” (i.e.very small resolution) position-materializing measurement suddenlychanges the state function to a tall, very narrow spike centered onthe central measured value; this is shown in figure 5.9

fig9This is called the “collapse of the state function”; it is usually

taken as a postulate in the orthodox Copenhagen interpretation.The figure also shows that, after the measurement, while “on itsown”, that state function changes in time. It is very importantto understand how state functions change in time betweenmeasurements (“time evolution of state function”).

If the governing equation is a partial differential equation in the variables xand t, then presumably (if we can solve it), it will tell us exactly how state

37

Page 39: PHY 416, Quantum Mechanics - SIUE

functions evolve in time. So we seek a governing equation. How can we findout if one exists, and if so, what it is? Let us think, we need:

1. A linear equation. This allows for the possibility that some solutionsmay be superpositions of other solutions. We saw that the double-slitexperiment for electrons requires this.

2. We know that the energy of a particle is (nonrelativistically),

E =1

2mv2 + V (x, t) (5.30)

or

E =p2

2m+ V (x, t) (5.31)

Now, in deBroglie language, this is

hω =h2k2

2m+ V (x, t) (5.32)

This uses the dispersion relationship we found earlier. Our governingequation must be consistent with this.

3. Something that allows plane and spherical wave solutions

• Now the appearance of V (x, t) in the dispersion relationship suggeststhat it must appear in the equation. It cannot appear alone because,for the sum of two solutions to be a solution of the equation must behomogeneous.So it must appear as

V (x, t)Ψn(x, t). (5.33)

But, for the same reason, the equation must be linear; hence, n = 1.Thus, we expect a term

V (x, t)Ψ(x, t). (5.34)

• The hω in the dispersion relation (with a plane wave) suggests one deriva-tive in time

• The h2k2

2m suggests (again, with a plane wave) a second derivative in x.

38

Page 40: PHY 416, Quantum Mechanics - SIUE

Putting these ideas together, we try

a∂2Ψ(x, t)

∂x2 + V (x, t)Ψ(x, t) = b∂Ψ(x, t)

∂t(5.35)

For a free particle, we would have V = V0constant, which we could take aszero. Then, trying cos(kx−ωt) or sin(kx−ωt) in our trial governing equation,it doesn’t work (as you can easily show). This is a good sign, since we alreadyknow from other arguments that these can’t be valid state functions. For afree particle, let’s try the superposition state

Ψ→(x, t) = cos(kx− ωt) + γ sin(kx− ωt) (5.36)

where γ is unknown as yet (but hopefully, a constant). Then by using ourtrial governing equation:

−ak2 cos(kx− ωt)− aγk2 sin(kx − ωt) + V0 cos(kx− ωt) + γV0 sin(kx− ωt) =

ωb sin(kx − ωt)− bγω cos(kx− ωt).

which, simplifies to

[−ak2 + V0 + bγω] cos(kx− ωt) = [ak2γ − V0γ + bω] sin(kx− ωt). (5.37)

For this to hold for all x and t, we must have

1. −ak2 + V0 = −bγω

2. −ak2 + V0 = bωγ

By subtracting both of these equations from each other,

0 = −βγω − βω

γ(5.38)

γ = −1

γ(5.39)

γ2 = −1 (5.40)

This is very interesting − we see that the imaginary number

i ≡√−1

39

Page 41: PHY 416, Quantum Mechanics - SIUE

is required. However, for the moment, let us continue: we plug in γ = ±iback into our trial governing equation, getting

−ak2 + V0 = ∓ibω (5.41)

Now this looks like a relation between ω and k, and so it must agree (if thisform of the wave equation is to be consistent) with our original dispersionrelation.

h2k2

2m+ V0 = hω (5.42)

Comparing these, we find

a = − h2

2m(5.43)

b = ±ih (5.44)

We will choose the positive sign for β; then our governing equation is

− h2

2m

∂2Ψ(x, t)

∂x2 + V0Ψ(x, t) = ih∂Ψ(x, t)

∂t(5.45)

We now assume that this promising-looking equation is valid wether V isconstant or not. Then we have

− h2

2m

∂2Ψ(x, t)

∂x2 + V (x, t)Ψ(x, t) = ih∂Ψ(x, t)

∂t(5.46)

This is the famous Schrodinger equation. We will be dealing with it for therest of the semester.

40

Page 42: PHY 416, Quantum Mechanics - SIUE

Chapter 6

Sixth Class: A Governing Equation

Thursday, September 11, 2008Let’s begin today by recalling where we have arrived. We saw that the

basic candidate plane wave function cos(phx−

Eh t+ φ

)for several reasons

cannot be valid as a state function in quantum theory. Therefore, neithercan superpositions of these be valid. And, as we remarked, nor is a rigidly

moving shape, e.g., f(x, t) = Ae−(x−vt)2

2σ2 . Therefore, we have a catalogue ofzero valid state functions. Our hope was that finding a “governing equation”might help us find valid state functions; presumably they are solutions of it.

6.1 Complex Plane Waves

A very interesting point has come up. In the process of our “derivation” wehave seen that the original deBroglie form of the state function for a freeparticle of definite energy E and definite momentum p is not valid

Ψ(x, t) 6= A cos

(pihx− Ei

ht

)(6.1)

for this case! Rather, the appropriate proxy wave function for this case isapparently

Ψ(x, t) = A

[cos

(pihx− Ei

ht

)+ i sin

(pihx− Ei

ht

)](6.2)

41

Page 43: PHY 416, Quantum Mechanics - SIUE

6.2 What is the Meaning of a Complex Wave Function?

It makes no sense to say “the length of this table is 27i inches” or “thevoltage difference was 3i volts.;; Likewise, it makes not sense to say “ thewave amplitude at this point of this stretched string is now 5i meters.” So,how can we deal with a complex wave function in quantum mechanics? Toquote another author who puts it sell:

fig1Thus, the state function (“wave function”) is a (generally) complex (in

mathematical sense) function that represents (mathematically) the particlewhile it is in partial reality. We would like to write the equation in a morecompact form; to do this, we recall Euler’s formula:

eiθ = cos θ + i sin θ (6.3)

A convenient way to demonstrate this is by series expansions. We know

ex = 1 + x+x2

2!+x3

3!+ · · · (6.4)

so,

eiθ = 1 + iθ +i2θ2

2!+i3θ3

3!+i4θ4

4!+ · · · , (6.5)

which is

eiθ = 1 + iθ − θ2

2!− iθ3

3!+θ4

4!+iθ5

5!

=

(1− θ2

2!+θ4

4!− · · ·

)+ i

(θ − θ3

3!+θ5

5!− · · ·

)= cos θ + i sin θ

Thus, the valid deBroglie plane wave for definite momentum and energy is

Ψ(x, t) = A [cos(kx− ωt) + i sin(kx− ωt)]

Aei(kx−ωt) (6.6)

42

Page 44: PHY 416, Quantum Mechanics - SIUE

6.3 Probabilities and Normalization Revisited

The probability of an occurrence must, of course, be real and ≥ 0. How-ever, as we see from the Schrodinger equation, with its explicit

√−1, state

functions are, in general, complex. Now, earlier, to form the probability ofmaterialization at position x1 at time t we took Ψ2(x, t)dx. But, if Ψ iscomplex, Ψ2 will generally be complex as well; hence it cannot be used aprobability density. For example, if

Ψ(x, t) = Ae−(kx−ωt)

then,

Ψ2(x, t) = A2e2i(kx−ωt)

which is

Ψ2(x, t) = A2 [cos(2kx− 2ωt) + i sin(2kx− 2ωt)]

which is not real except at certain instants.On the other hand, the interference pattern obtained in the double-slit

electron experiment shows that that we must use a form bilinear in Ψ for theprobability density function. What option then do we have?

Consider the function Ψ∗(x, t)Ψ(x, t), where Ψ∗ is the “complex conjugate”of Ψ.

If a complex number, z ≡ a + ib, where a and b are real, thenz∗ ≡ a− ib. thus, if f(x, t) is a complex-valued function of the realvariables x and t, f(x, t) ≡ u(x, t) + iv(x, t) where u and v are realvalued functions, then f ∗(x, t) ≡ u(x, t)− iv(x, t). Then,

f ∗(x, t)f(x, t) ≡ |f(x, t)|2 = [u(x, t)− iv(x, t)] [u(x, t) + iv(x, t)]

= u2(x, t)− iv(x, t)u(x, t) + u(x, t) · iv(x, t) + v2(x, t)

Since, (−i)(i) = i2 = −(−1) = +1. Thus,

f ∗(x, t)f(x, t) = u2(x, t) + v2(x, t) (6.7)1we mean “in the dx centered on x.”

43

Page 45: PHY 416, Quantum Mechanics - SIUE

and note that this is always ≥ 0 since it is the sum of squares of realfunctions!

Consider also, the polar form of the representations:

f(x, t) = Reiθ(x,t)

eiθ(x,t) = cos θ(x, t) + i sin θ(x, t)(eiθ(x,t)

)∗= cos θ(x, t)− i sin θ(x, t)

= e−iθ(x,t)

where θ(x, t) is a real valued function of (x, t). Thus,(eiθ)∗

= e−θ =1

eiθ(6.8)

The complex conjugate of a product is, as you can easily show, the productof the conjugates. Thus, if f(x, t) = Reiθ(x,t) where R and θ are real, then,f ∗(x, t) = R∗e−iθ(x,t). Thus, in this case,

|f(x, t)|2 ≡ R∗e−iθReiθ = R∗Re−iθeiθ = R∗R ≡ |R|2 (6.9)

(If all of this does not make full, familiar sense to you, get ahold of a stan-dard undergraduate book with a section or chapter on complex numbers andreview/learn it. )

With the above in mind, we consider the (bilinear) function |Ψ(x, t)|2 ≡Ψ∗(x, t)Ψ(x, t) as a candidate for the positional probability density function.Since it is real and greater then or equal to zero, it is a good candidate,and, in fact, its use in this role shows agreement or consistency with allof the experiment data we have. In fact, formalizing this guess, in 1926,Max Born postulated the following, which has since been accepted as partof the standard orthodox interpretation of quantum mechanics: PositionalProbability Postulate (Born, 1926)

If, at time t, a position measurement is made for a particle associatedwith the state function Ψ(x, t), then the probability that the particlewill materialize in a small range, dx,centered on x is given by

P (x, t)dx = Ψ∗(x, t)Ψ(x, t) dx (6.10)

44

Page 46: PHY 416, Quantum Mechanics - SIUE

This means that out previously stated normalization condition must nowbe accordingly modified to∫ ∞

−∞Ψ∗(x, t)Ψ(x, t) dx = 1 (6.11)

at all times t.

6.4 Plane Wave and Uniformity of Space

Now we note another bonus of our complex plane wave states. With the realplane waves, e.g. cos(kx− ωt), the probability density was bumpy”

fig2This is a clear violation of the homogeneity of empty space2, as we re-

marked some time ago. However, with the complex plane wave and the newprobability rule,

P (x, t)dx = Ψ∗(x, t)Ψ(x, t) = e−i(kx+ωt)ei(kx−ωt) = 1. (6.12)

This is constant in space, which accords with this basic principle of homo-geneity of space.

6.5 A First Example of Using the Schrodinger Equation

We want to show a first example illustrating use of the Schrodinger equationto find what state functions are allowed for a specific situation. For this,we will use the familiar “particle-ini-a-box” (or “infinite square well”). Thesituation is familiar from a previous “modern physics” course, but the methodwe use here today is a little different:

Say we have a certain force field that the particle is “in.” Then, wehave a potential energy profile, V (x, t). Say, for example, the profileis that of the familiar infinite square well:

V (x) =

0 if 0 ≤ x ≤ L,

∞ otherwise.(6.13)

2This is a cherished, basic principle of physics. Violation of it would be a very serious issue.

45

Page 47: PHY 416, Quantum Mechanics - SIUE

Then, inside the well, V = 0, so the Schrodinger equation is

−h2

2

∂2Ψ(x, t)

∂x2 = ih∂Ψ(x, t)

∂t(6.14)

inside the well.

How do we solve this? For now, instead of using formal mathematicalmethods, we note that inside the well the particle is free, so, if we assumea fixed, definite energy, E, the general solution should be a superposition ofright and left moving quantum plane waves

ΨEe(x,t) = Aei√

2mEh x−iEth +Be−i

√2mEh x−iEth (6.15)

We emphasize that this is the solution inside the well for only one energy.Solutions of this form for different energies can, by principle of superposition,be superposed; we will discuss that later on.

We are not done, however − recall that we have two boundary conditions:

1. Ψ(x = 0, t) = 0 for all t

2. Ψ(x = L, t) = 0 for all t

As you can show, application of the Schrodinger equation on our quantumplane waves leads to

ΨT = A sin

(√2mE

hx

)e−i

Eth (6.16)

Them application of the second boundary conditions on this equation leadsto √

emEL

h= nπ, n = 1, 2, 3 · · · (6.17)

Which is the familiar quantization of energy

En =n2π2h2

2mL2 , n = 1, 2, 3, · · · (6.18)

Thus, the possible state functions for definite energy are

Ψn(x, t) =

An sin

(nπL x)

e−in2π2h2mL2 t, 0 ≤ x ≤ L

0, x < 0, x > L(6.19)

46

Page 48: PHY 416, Quantum Mechanics - SIUE

since the spatial parts of these state functions happen (in this case) to bepurely real and we can plot them − they are plotted in figure 3, along withthe associated probability density function Ψ∗(x, t)Ψ(x, t) for each.

fig3Note that, for the state function Ψn(x, t) given by out possible state func-

tions, the probability densities, while dependent on position (x) are all inde-pendent of time, since ∣∣e−ict∣∣2 = e+icte−ict = 1 (6.20)

for c real, as in our possible state function.The coefficients An can be found through the application of the normal-

ization condition requiring 1 =∫

Ψ∗(x, t)Ψ(x, t) dx here is

1 = |An|2∫ L

0sin2

(nπxL

)dx (6.21)

SinceΨ is zero outside the well and since the time dependence from the expo-nential factor goes away in the probability density. Thus, the normalizationcondition is independent of time − it can be applied at any time to determinethe normalization constant An, and once An is determined, it remains con-stant in time. (We will see later that there is a very beautiful theorem thatensures this time independence of the normalization is true for all state func-tions valid for single particles states in non-relativistic quantum mechanics.)Now, you can show, ∫ L

0sin2

(nπxL

)dx =

L

2(6.22)

for all integer n − note that it is independent of n. Thus, equation (6.21)

tells us that for the infinite square well, An =√

2L for all n. Thus, the state

functions for definite energy are

Ψn(x, t) =

2L sin

(nπL x)

e−in2π2h2mL2 t 0 ≤ x ≤ L

0 x < 0, x > L(6.23)

Since the Schrodinger equation is linear and homogeneous, the sum of solu-tions − in fact, an arbitrary linear combination of solutions

Ψ(x, t) =∞∑i=1

cnΨn(x, t) =∞∑i=1

cn sin(nπxL

)e−

iEnth (6.24)

47

Page 49: PHY 416, Quantum Mechanics - SIUE

is a solution inside the well. (of course, for “superposition solutions” likethis, the cn’s must be scaled so that the overall Ψ(x, t) is correctly normal-ized.) Of course, for this sort for superposition state function there is nosingle value of definite energy, since each Ψn in the sum corresponds to adistinct energy. This raises an important interpretational question about themeaning of superposition states like our superposition state. We have muchto say about this later in the course. In fact, we’ll also have much furtherto say about the definite-energy infinite square-well states and their physicalmeaning later. Here our purpose was to give you a first illustration of theuse of the Schrodinger equation to find valid state function.

48

Page 50: PHY 416, Quantum Mechanics - SIUE

Chapter 7

Seventh Class: Time DependenceThrough the Governing Equation

Tuesday, September 16, 2008

7.1 A Review of the Governing Equation

Our governing equation for state functions is the Schrodinger equation

Ψ(x, t) = ih∂Ψ(x, t)

∂t= − h2

2m

∂2Ψ(x, t)

∂x2 + V (x, t)Ψ(x, t) (7.1)

Recall that the governing equation plays two (overlapping) roles:

1. It guides us in determining what state functions are permissible for whatsystems under what conditions and what what functions aren’t permis-sible − i.e., any permissible state function must be a solution of theSchrodinger equation, and through this,

2. It tells us how the state function Ψ(x, t) evolves between measurements(recall that measurements causes a sudden change, or “collapse”, of thestate function). This is so because the Schrodinger equation is providingthe time-rate-of-change of Ψ(x, t) (as long as the right-hand-side of theSchrodinger equation can be calculated).

We illustrated these roles by using the Schrodinger equation to find statefunctions of definite energy for a particle in an infinite square-well poten-tial energy profile. In this case, the time evolution was through the fac-tor e−

iEnth in each state function. Further, in this case, the probability den-

sity function, P (x) = Ψ∗(x, t)Ψ(x, t), has no explicit time dependence since

49

Page 51: PHY 416, Quantum Mechanics - SIUE

(e

+iEnth

)(e−iEnth

)= 1. Thus, in this case, at all x, the probability density

is constant in time. As another example in the homework, you show thatthe functions e

±iEnth are valid free particle state functions provided a certain

dispersion relationship is obeyed. As yet another example, some time ago weclaimed that a rigidly moving “bell-shaped” curve,

f(x, t) = Ae−(x−vt)2

2σ2 (7.2)

with v and σ real and constant is not a valid state function − and it is clearwhy not. Suppose that f(x, t) is purely real for all x and all t. Then theapplication of the Schrodinger equation for f(x, t) produces an imaginaryfunction on one side and a real function on the other − and it is impossiblefor these to be equal to each (unless they are both zero). Since f(x, t) =

AE− (x−vt)2

2σ2 with v and σ constant and real is purely real for all x and t, it istherefore not a valid quantum mechanical state function.

A situation that sometimes arises is one in which you know the statefunction at a specific time only (either by knowledge or by an assumption)and you want to find out how it changes in time. This is called the quantuminitial value problem. Here is a quick example (without details) that willgive you a first idea of how this sort of thing works. Suppose we have a freeparticle (V (x, t) = 0 all x and t) and we somehow know that at the specifictime t = 0, the state function Ψ(x, t), is Gaussian (bell-shaped).

Ψ(x, t = 0) = Ae−x2

2σ2 (7.3)

At a given instant, this can be a valid state function for a free particle.(In fact, as we will see in a few classes hence, any normalizable1 continuousfunction f(x) is a possible state function for a free particle at single instantof time2).

In fact, this is an example of the sort of state function that can result atthe instant of position measurement, when the resolution of the measurementis 2σ. If σ is small, this looks like a tall, narrow spike of “width” 2σ, as shownin figure 1. Suppose Ψ(x, t = 0) is this. The question is − what happens to

1“Normalizable” meansR +∞−∞ Ψ∗(x, t = 0)Ψ(x, t = 0) dx <∞

2Fourier theory will show this

50

Page 52: PHY 416, Quantum Mechanics - SIUE

Ψ as time goes by in the absence of or before another measurement? Mustit change? (yes)

Fig1Could it change in time by moving rigidly? As we’ve already seen, no.

To find out how Ψ evolves in time, we need to find solution, Ψ(x, t) to theSchrodinger equation that reduces down to equation (7.3) at t = 0.

Since the Schrodinger equation is first order in time, there is at most onelinearly independent function that will do this. (see later if you are notfamiliar with this fact about first-order differential equations), In the nearfuture, we’ll see3 (also from the theory of Fourier analysis) how to find thisfunction, but that is not our goal here. Therefore, for now, I’ll just tell youthe answer, so we can work with it.

Ψ(x, t) = Ce− ax2

[1+( 2ihatm )]√

1 +(2ihat

m

) (7.4)

where a ≡ 12σ2 and C is a constant. If you have the patience, you can verify

that this equation works in free particle Schrodinger equation, and by settingt equal to zero, we see that it does reduce to our previous function at timeequal to zero. So, this equation is correct. This question now is − how do weinterpret it? A plot would help, but since Ψ(x, t) is complex, we can’t plot it.Anyway, what is more important than Ψ is the probability density function.By letting θ = 2hat

m , we have

Ψ = Ce−

ax2

(1+θ)

√1 + iθ

⇒ Ψ∗ = C∗e−

ax2

(1−θ)

√1− iθ

(7.5)

so,

P (x, t) = Ψ∗Ψ = |C|2 e−ax2[ 1

1+iθ+ 11−iθ ]

√1 + θ2

=|C|2√1 + θ2

e− 2ax2

(1+θ2) (7.6)

To interpret this expression, we approximate by ignoring the denominator(i.e. approximate the denominator by 1. Show that this is valid to about

3if you are curious now, see Griffths problem 2.22, which assumes you can handle Fourier transforms.

51

Page 53: PHY 416, Quantum Mechanics - SIUE

10% accuracy as long as t <√

m2

20h2a2 .) Then

P (x, t) ≈ |C|2e− ax2

(1+ 4h2a2t2

m2 ) (7.7)

which is again a Gaussian (“bell-shaped” curve) whose width increases intime. Thus, as we claimed in the last class, between position measurementsthe probability distribution spreads or “diffuses”.

fig4

7.2 Persistence of Normalization

Suppose we have a state function Ψ(x, t0) and we normalize it at a partic-ular time, say t = t0. The state function changes in time according to theSchrodinger equation (since the Schrodinger equation has a ∂

∂t). Then itwould seem unlikely that Ψ remains normalized. But, if it doesn’t, we havea problem! We investigate this: Let

P (t = t0) =

∫ +∞

−∞Ψ∗(x, t0)Ψ(x, t0) dx = 1. (7.8)

Then

dP (t)

dt=

d

dt

∫ +∞

−∞Ψ∗(x, t)Ψ(x, t) dx

=

∫ +∞

−∞

∂t[Ψ∗(x, t)Ψ] dx

=

∫ +∞

−∞

[∂Ψ∗(x, t)

∂tΨ + Ψ∗(x, t)

∂Ψ

∂t

]dx

But, according to the Schrodinger equation,

∂Ψ

∂t=

ih

2m

∂2Ψ

∂x2 −i

hVΨ

∂Ψ∗

∂t= − ih

2m

∂2Ψ∗

∂x2 +i

hVΨ∗

52

Page 54: PHY 416, Quantum Mechanics - SIUE

By plugging these in, we get

dP (t)

dt=

ih

2m

∫ +∞

−∞

(Ψ∗∂2Ψ

∂x2 −∂2Ψ∗

∂x2 Ψ

)dx

=ih

2m

∫ +∞

−∞

∂x

[Ψ∗∂Ψ

∂x− ∂Ψ∗

∂xΨ

]dx

=ih

2m

[Ψ∗(x, t)

∂Ψ

∂x− ∂Ψ∗

∂xΨ

]+∞

−∞

Now, in any sensible theory, Ψ and Ψ∗ −→ 0 as x −→ ±∞. Thus, dPdt = 0,

and “once normalized, always normalized” (unless measurement intervenes).

7.3 The Reason for the Persistence of the Normalization

We saw that, in the absence of measurement, if

P (t0 =

∫ +∞

−∞Ψ∗(x, t0)Ψ(x, t) dx = 1, (7.9)

thendP (t)

dt=

d

dt

∫ +∞

−∞Ψ∗(x, t)Ψ(x, t) dx = 0. (7.10)

Key to this proof was the use of the SchrodingerSchrodinger equation

ih∂Ψ(x, t)

∂t= − h2

2m

∂2Ψ(x, t)

∂x2 + V (x, t)Ψ(x, t). (7.11)

As you can convince yourself by going through the proof, again, what “madeit work” is the SchrodingerSchrodinger equation is first order in time. Asyou can convince yourself (try it), if the SchrodingerSchrodinger equationwere second order in time, dP

dt would not by zero and then we would haveproblems. Only a first order (in time) equation is consistent withconservation of probability. Actually, this is pretty obvious (from thetheory of differential equations) without going through the proof. A firstorder differential equation has only one “arbitrary constant” in its general

53

Page 55: PHY 416, Quantum Mechanics - SIUE

solution and this is usually “f(t = 0)”. For example, consider say,

df(t)

t= 2t

f(t) =

∫2t dt = t2 + const.

f(t) = t2 + f(t = 0)

or say,

df

dt= −3f(t) ⇒ f(t) = Ce−3t ⇒ C = f(t = 0).

So, for a first order equation, only f(t = 0) can be arbitrarily specified as aninitial condition. However, a second order equation, say

d2f(t)

dt2= 2t

f(t) =1

3t3 + at+ b

where a and b are constants.clearly, b = f(0) and a = f ′(0). So for a uniquetime development f(t), we need to specify both f(t = 0) and f ′(t = 0). Or,say,

d2f(t)

dt2= −ω2f(t)

f(t) = A cos(ωt+ φ)

= C sin(ωt) +D cos(ωt)

Then, f(0) = D and f ′(0) = ωC. We see that, with second order differentialequations, we can independently and arbitrarily specify both f(0) and f ′(0)− and both are needed to specify a unique time development of f(t). Thus, ifthe Schrodinger equation were second order in time, we could independentlyand arbitrarily specify Ψ(x, t = 0) and ∂Ψ(x,t=0)

∂t . This would mean that we

could set ∂∂t [Ψ∗(x, t = 0)Ψ(x, t = 0)] and d

dt

∫ +∞−∞ Ψ∗(x, t = 0)Ψ(x, t = 0) dx

to be nonzero. This would destroy conservation of probability; hence theSchrodinger equation must be first order in time!

54

Page 56: PHY 416, Quantum Mechanics - SIUE

7.4 Determinism in Quantum Mechanics − A Comment

It is often remarked that quantum mechanics has scarified the determinism(i.e. “predictability-in-principle”) of Newtonian mechanics. While this istrue, in another sense non-relativistic quantum mechanics, with its first orderin time governing equation, is very deterministic for the state function itself.That is, it follows from our discussion of the last class that, in the absence ofintervening measurement, knowledge of Ψ(x, t0) for all x at a single instantof time (t0) is sufficient to determine Ψ(x, t) anywhere for all t > t0

4. Thus,the time evolution of the state function can be said to be deterministic.

Such a statement is not true for solutions of governing equations that aresecond order (or higher order) in time. Consider, for example, the familiarclassical wave equation.

∂2Ψ(x, t)

∂x2 =1

v2

∂2Ψ(x, t)

∂t2(7.12)

Which governs many classical wave phenomena (e.g. wave on dispersionlessstretched string). For this case, (classical wave obeying the CWE) supposethat Ψ(x, t = 0) = cos(kx). Then, there are many possibilities for Ψ(x, t) −among them, Ψ→ = cos(kx−ωt) and Ψ← = cos(kx+ωt). For a second order(in time) equation, to be assured on a unique Ψ(x, t), one must specify att = 0 both Ψ and ∂Ψ

∂t . Thus for a second order (in time) equation, knowledgeof only Ψ at t = 0 is not enough to predict Ψ(x, t > 0). All this being said, thequestion then arises as to wether one can write an explicit “general solution”of the Schrodinger equation valid for all potential energy profile functionsV (x, t) that directly express Ψ(x, t) in terms of the values of Ψ at all pointsx′ at an earlier time, say t = 0. Such an equation would be in the form of ageneral formula Ψ(x, t) =??, rather than in the form of a riddle (differentialequation relating derivatives of Ψ to each other that must be solved). Basedon what we have said above, in principle this is possible. Such an equationtakes the form:

Ψ(x, t) =

∫ +∞

−∞K(x′, x, t)Ψ(x′, t) dx′ (7.13)

4TO do this involves a sort of conversion of the Schrodinger equation to an integral equation. The results arefascinating and to be returned to if time allows.

55

Page 57: PHY 416, Quantum Mechanics - SIUE

where the function K is called the propagator, represents the influence of“old values” if Ψ at x′ (i.e.Ψ(x′, 0)) at the point x at the present time (t).[In this sense, K(x′, x, t) is like the contribution at point x at time t madeby the state function at x′ at the earlier time t = 0.] It turns out that theexplicit form of the propagator is known for the case of a free particle

Kfree(x′, x, t) =

√m

2πihtei

(x−x′)2m2ht (7.14)

This technique5, while conceptually important, is uncommon in elementarycourses, largely because the propagator is known in closed analytical form a safunction for only a few potential energy profiles (of course, including the freeparticle case V (x, t) = 0) and even in those cases, the mathematics involvedcan be quite formidable. For this reason, we will mostly or exclusively staywith the standard differential equation approach.

5The propagator approach to quantum mechanics, largely pioneered by R. P. Feynman, is very important inmodern quantum field theory. Nonrelativistic quantum mechanics is developed in his book Quantum Mechanicsand Path Integrals.

56

Page 58: PHY 416, Quantum Mechanics - SIUE

Chapter 8

Eight Class: Probability

Thursday, September 18, 2008

8.1 Probability Current

In the last class, we looked into the rate of change of the total probability formaterialization (or position measurement) anywhere on the x-axis and foundthat it was zero;

dPtotdt

=ih

2m

[Ψ∗∂Ψ

∂x− ∂Ψ∗

∂xΨ

]+∞

−∞= 0 (8.1)

by using the normalization boundary conditions,

Ψ(x, t)→ 0∂Ψ(x, t)

∂x→ 0 as x→ ±∞ (8.2)

Suppose, however that we consider the rate of change of the total probabilityto materialize in the finite region, say from x1 to x2. This would then be

∂P (x1, x2, t)

∂t=

ih

2m

[Ψ∗∂Ψ

∂x− ∂Ψ∗

∂xΨ

]x2

x1

(8.3)

which in general, is not zero. In this case, we can define a function

J(x, t) =h

2mi

(Ψ∗(x, t)

∂Ψ(x, t)

∂x−Ψ(x, t)

∂Ψ∗(x, t)

∂x

)(8.4)

and then∂P (x1, x2, t)

∂t= J(x1, t)− J(x2, t) (8.5)

57

Page 59: PHY 416, Quantum Mechanics - SIUE

Interpretation: The rate of change of probability in [x1, x2] is the rate at whichprobability enters [x1, x2] at x1, take away the rate at which probability leavesat x2

1.Thus, the quantity J(x, t) represents the “flow of probability” (in the di-

rection of increasing x) at position x at time t. It is called the probabilityflux. It is also often symbolized as S(x, t) ≡ J(x, t)

8.2 Choice of Free Particle Plane Wave States

We now return to look more closely at an issue we glossed over in the past.For a free particle, we’ve been using, as our basic “right-moving” plane-wavestate,

Ψ(1)→ (x, t) = Aei(kx−ωt). (8.6)

However, perhaps we should use for this purpose

Ψ(2)→ (x, t) = Aei(kx−ωt) = Ae−ikxe+iωt (8.7)

or, can be use these “interchangeably” or even mix them in calculation? Wealso have different choices for “left-moving” plane-waves states:

Ψ(1)← (x, t) = Aei(kx−ωt) or Ψ(2)

← (x, t) = Ae−i(kx+ωt). (8.8)

Let’s look into this: say we have the choice Ψ(1)← = Aei(kx−ωt). With this, we

could, for left-moving state, a prıori still use either Ψ(1)← or Ψ(2)

← above or both.Say we pick Ψ(1)

← and Ψ(1)→ . Then, according to the principle of superposition,

Ψ(x, t) = ei(kx−ωt) + ei(kx+ ωt) (8.9)

should be a possible state. But this is

Ψ(x, t) = eikx(eiωt + e−iωt

)= 2eikx cos(ωt), (8.10)

which has the same problem we were trying to avoid, (namely, there is atime when Ψ = 0 everywhere). So, suppose we try Ψ(1)

→ (x, t) = Aei(kx−ωt) andΨ(2)← (x, t) = Ae−i(kx+ωt). Then the sum of these two is

Ψ(x, t) = eiωt(eikx + e−ikx

)= 2 cos(kx)e−ωt, (8.11)

1Note that because J can be a positive or negative at any point, the statement is equivalent to [J(x1, t)−J(x2, t)],is the net rate at which probability enters [x1, x2] at both ends minus the net rate at which probability leaves [x1, x2]

at both ends. Since this is equal to ∂P (x1,x2,t)∂t

, probability is neither spontaneously created or destroyed in [x1, x2];any change in it must follow in or out at the boundaries.

58

Page 60: PHY 416, Quantum Mechanics - SIUE

which avoids the problem! (Why?) Likewise, Ψ(2)→ (x, t) = Aei(kx−ωt) combined

with Ψ(1)← = Aei(kx−ωt) avoids this problem (show this).

What if sometimes use Ψ(1)→ (x, t) = Aei(kx−ωt) and sometimes we use Ψ(2)

→ (x, t) =Aei(kx−ωt)? This leads to trouble, as you can show by considering their su-perposition. So we need to make a choice, In this, we follow the QuantumMechanics Convention for free particle plane wave states:

Ψ→(x, t) = Aei(kx−ωt) (8.12)

Ψ←(x, t) = Ae−i(kx+ωt) (8.13)

Here is a quick way to remember this: the time part is always e−iEth . (This is

true in quantum mechanics for states of definite energy).

8.3 Circular and Spherical Definite Energy State Functions

Consider the case of a free (no force field acting on it) electron of definiteenergy E incident on a “pin-hole” aperture. On the downstream side of theaperture, the state function should be (in three dimension) a spherical wave

Ψ(~r, t) = A(r)ei(pr−Et)

h (8.14)

where r is the distance from the pin-hole to the point ~r (origin taken as thepinhole). Note that A, the amplitude, depends on r − it must decrease withr, in fact. Why? Because the probability spreads out in space. For example,for the three dimensional case (spherical wave)

A(r) ∝ 1

r(8.15)

while, for the two dimensional case, (circular wave)

A(r) ∝ 1√r. (8.16)

Thus, if we consider the electron double slit experiment with the slits aspin-holes, we have, downstream of the pin-hole plane

Ψ(~r, t) = Ψ1(~r, t) + Ψ2(~r, t) (8.17)

59

Page 61: PHY 416, Quantum Mechanics - SIUE

where Ψ1 is a spherical wave centered on hole 1, and Ψ2 is a spherical wavecentered on hole 2. If we put a measurement device (“proximity device”)near hole 1 to determine if the electron “went through” hole 1, then if thedevice registers a signal, after the measurement, Ψ(~r, t) is “collapsed” toΨ(~r, t) = Ψ1(~r, t).

For the case of a plane-wave state function incident on a long, infinitesi-mally narrow slit, the result downstream is a cylindrical wave state:

Ψ(~r, t) =A√r

ei(kr−ωt) (8.18)

where r is measured in cylindrical coordinates. An example of cylindricalwaves is shown in figures 8.1 and 8.2.

fig1fig2As we see, for this case, r is just the direct perpendicular distance from

the slit line to the nearest point on the wavefront.

8.4 Superposition States and Probabilities

Now let’s consider a superposition state of plane waves:

Ψ(x, t) = A1ei(k1x−ω1t) + A2e

i(k2x−ω2t) (8.19)

Here, we have an interesting situation: each of the two components corre-sponds to a different, definite value of momentum (p1hk1, p2 = hk2). Thus,the superposition state no longer corresponds to a definite value of momen-tum! What then happens if we measure the momentum for a particle repre-sented by this superposition state function? Answer: If we do so, we definitelyobtain

either p1 = hk1

or p2 = hk2

How likely is each probability? In this case, the probability of getting resultp1 is:

|A1|2

|A1|2 + |A2|2(8.20)

60

Page 62: PHY 416, Quantum Mechanics - SIUE

the probability of getting result p2 is”

|A2|2

|A1|2 + |A2|2(8.21)

and the probability of getting either p1 or p2 is:

|A1|2

|A1|2 + |A2|2+

|A2|2

|A1|2 + |A2|2=|A1|2 + |A2|2

|A1|2 + |A2|2= 1 = 100% (8.22)

A similar sort of rule applies to superposition states having components cor-responding to definite values of energy. As an example, consider the case ofthe infinite square-well. We found that the states of definite energy are

Ψn(x, t) =

√2

Lsin(nπxL

)e−

iEnth (8.23)

However, since the Schrodinger equation is linear and homogeneous, superpo-sitions of these are possible valid quantum states. Such superposition stateswill then not correspond to definite energy. Consider, for example, the su-perposition state

Ψ(x, t) = c1Ψ1(x, t) + c2Ψ2(x, t). (8.24)

(Here, c1 and c2 must be chosen such that Ψ(x, t) is normalized; we will worryabout the details of how to choose c1 and c2 to ensure this, later.) Thus

Ψ(x, t) = c1

√2

Lsin(πxL

)e−

iE1th + c2

√2

Lsin(πxL

)e−

iE2th (8.25)

If a measurement of energy is made on a particle represented by this statefunction, then we definitely obtain energy equal to either E1 or E2. Recall

En =n2π2h2

2mL2 (8.26)

The probability of getting definite energy value E1 is

E1 =|c1|2

|c1|2 + |c2|2(8.27)

and the probability of getting definite energy value E2 is

E2 =|c2|2

|c1|2 + |c2|2(8.28)

61

Page 63: PHY 416, Quantum Mechanics - SIUE

and the probability of getting either E1 or E2 is

|c1|2

|c1|2 + |c2|2+

|c2|2

|c1|2 + |c2|2= 1 = 100%. (8.29)

8.5 Packet States − Physical Meaning

In the past, we tried to construct localized moving “wave packets” by addingtogether many, very closely spaced (in wave number, or momentum) planewave states. It is now clear that this must be done with complex exponentials,rather that with real cosines and/or sines. Thus, a general packet state wouldlook (mathematically) like

Ψ(x, t) = limN→∞,∆k→0

N∑n=1

Anei[(k1+n(δk))x−(ω1+n(δω))t] (8.30)

=

∫ k2

k1

A(k)ei(kx−ωt) dk (8.31)

(8.32)

Such a packet is made up of a whole band of different plane waves andtherefore, a whole band of wave numbers. Thus, the possible results of mea-surement of momentum are spread over a band of width we call “∆p”.

“Orthodox ” Interpretation:If we do measure the momentum, however, the particle will materialize

one and only one value of momentum. The rule for this is the following:

Ψ(x, t) =

∫ k2

k1

A(k)ei(kx−ωt) dk (8.33)

where k = ph . Then, the probability of obtaining a value of momentum in a

small range dp centered on p is

P (p) dp ∝ |A(k)|2 dk (8.34)

We’ll work out the constant of proportionality later. Even without this, wecan, of course, will work the relative probabilities (i.e., ratios of probabilitiesfor which the constant of proportionality cancels).

62

Page 64: PHY 416, Quantum Mechanics - SIUE

Of course, for this to be really useful, we need to know how to calculateA(k) for a given Ψ(x, t). for that, we need Fourier analysis; we will see howthat works in the next class.

Comment: “Materialization of momentum p” does not imply materializa-tion at a definite position. In fact, there exists no state which simultaneouslyallows materializing atoms with precise p and precise x. (We will prove thislater).

In the meantime, let us look at an example of something conceptuallysimpler− the evaluation of an integral like that in equation (8.33) to constructΨ(x, t).

Example:

Suppose we want to evaluate the functional form (in x, say at t = 0)of an equal-amplitude mix of plane waves with equal amplitudesin some range [k1, k2] and zero amplitudes for k outside this range.Thus, the “bandwidth in wave number” for this mix is ∆k = k2−k1.We suppose that the band of wave numbers is centered on k = 0,thus k1 = k2. Then k2 = ∆k

2 and k1 = −∆k2 . A plot of A(k) vs. k

would then be flat from = ∆k2 → +∆k

2 . Figure 8.3 shows this plot.

fig3

We chose the amplitude of A(k) to be 1√∆k

so that the area underthe plot is 1 for “neatness”. Thus, this mix has half the planewaves propagating to the left (negative values of k) and, with equal

63

Page 65: PHY 416, Quantum Mechanics - SIUE

amplitudes, half to the right. We evaluate it at t = 0:

Ψ(x, t = 0) =1√∆k

∫ +∆k2

−∆k2

eikx dk (8.35)

=1√∆k

[1

ixeikx]+∆k

2

−∆k2

(8.36)

=2

x√

∆k

[ei(

∆k2 )x − e−i(

∆k2 )x

2i

](8.37)

=2√

∆kxsin

(∆k

2x

)(8.38)

=

√∆k(∆k

2

)x

sin

(∆k

2x

)(8.39)

=√

∆k sinc

(∆k

2x

)(8.40)

Schematically,fig 4Question for now. (Qualitatively), what do you expect will happen to the

shape of Ψ(x, t) as t increases from zero? Why?

64

Page 66: PHY 416, Quantum Mechanics - SIUE

Chapter 9

Ninth Class: Uncertainty in QuantumMechanics

Tuesday, September 23, 2008

9.1 Certainty and Uncertainty

As we have explained, for a particle in “partial reality” and represented bystate function Ψ(x, t), generally it is unpredictable or “uncertain” where a sin-gle position-producing measurement will cause materialization1. The rangeof this uncertainty for position materialization, called “2∆x” (for the one-dimensional case) is defined or, in principle, is determined experimentally2,as follows

9.1.1 Meaning of ∆x

:Imagine preparing an ensemble of very many3 identical particles, all with

the same state function, Ψ(x, t). A position materialization measurement isperformed for each system, Each particle materializes at a different position;the distribution of positions follows that of Ψ∗(x, t)Ψ(x, t). ∆x representsthe standard deviation of this distribution of positions. Applied to the statefunction for a single particle before materialization on position measurement,

1Of course, the procedure described is as “in principle” or in the nature of a “thought experiment”. Still, it definesthe meaning of ∆x.

2The only exception would be if the state function is an infinitely narrow spike function at the time of measurement,which never occurs in practice, but which represents an idealization we will further discuss later on.

3Technically, an infinite number

65

Page 67: PHY 416, Quantum Mechanics - SIUE

2∆x represents the “uncertainty” in our knowledge of where we expect theparticle to materialize. A schematic “cartoon” of this is shown in figure 1.

fig1Note then, that in theory, ∆x is equal to standard deviation of Ψ∗(x, t)Ψ,

not that of Ψ.It is also true, as we’ve seen, that for a particle in partial reality and

represented by state function Ψ(x, t), generally, if a momentum-producingmeasurement is made, the value of momentum materialized is unpredictableor “uncertain” over a ranger called 2∆p. The “in-principle” experimentalprocedure for determining ∆p is a follows.

9.1.2 Meaning of ∆p

:Imagine preparing an ensemble of very many identical particles, all with

the same state function, Ψ(x, t). A momentum materializing measurement isperformed for each system. Generally each particle will materialize a differ-ent value of momentum. The symbol ∆p represents the standard deviation ofthis distribution of momenta. Applied to a single particle with state functionΨ before momentum materialization, and 2∆p represents the “uncertainty”in our knowledge of what value of momentum we expect the particle to ma-terialize with. Special cases are those of the plane waves states

Ψ(x, t) = Aei(phx−

Eh t) and Ae−i(

phx+E

h t) (9.1)

For each of these states the momentum is definite(+ phx and − p

hx), respec-

tively); thus for each of these states ∆p = 0. We now come to somethingvery important and fundamental. Heisenberg Uncertainty Principle −Statement

There exists no valid state function in quantum mechanics forwhich the product ∆x∆px is less then h

2 . That is, for any valid statefunction

∆x∆px ≥h

2(9.2)

In the next class we’ll demonstrate the Heisenberg Uncertainty Principle(HUP) using Fourier analysis; in the meantime we will illustrate the resultwith a few preliminary examples.

66

Page 68: PHY 416, Quantum Mechanics - SIUE

9.2 Example: Plane-Wave states

Consider the plane wave states Ψ(x, t) = Aei(phx−

Eh t) and Ae−i(

phx+E

h t). Asremarked earlier, for each, ∆px = 0. However, for each Ψ∗Ψ = |A|2 =constant in x, ∆x = ∞. Thus, for these states, ∆x∆px =“∞ · 0”, which isnot inconsistent with the Heisenberg Uncertainty Principle .

Example:Consider the equal amplitude superposition of plane wave states we dis-

cussed in the last class − namely

Ψ(x, t = 0) =1√δk

∫ + δk2

− δk2eikx dk (9.3)

Strictly speaking, we should use the standard deviation for the wave numberdistribution; as you can show4, it is ∆k = 2δk√

12for a flat distribution. Then,

since for each plane wave component p = hk → ∆p = h∆k = 2hδk√12

. As wesaw last time, the integral evaluates to

Ψ(x, t = 0) = constant · sinc

(δk

2x

). (9.4)

Thus,

|Ψ(x, t = 0)|2 ∝ sinc2(δk

2x

)(9.5)

Strictly speaking, for ∆x we should use the standard deviation of |Ψ(x, t =0)|2. However, this would lead us into a (now) distracting side calculation −we seek an understanding, not exactness. Therefore, we will estimate ∆x asthe half-width to the first zero of |Ψ|2 for this case. The first zero occurs at

δk

2x = π ⇒ x =

δk. (9.6)

Calling this ∆x, we have

∆x ≈ 2π

δk=

4π√12∆k

≈ 1.15π

∆k≈ π

∆k(9.7)

4see later. Also, since the wave number distribution A(k) is flat, the standard distribution for |A(k)|2 (∝ toprobability of momentum p = hk) is the same as that of A(k).

67

Page 69: PHY 416, Quantum Mechanics - SIUE

i.e. ∆x∆k ≈ π. Thus,

∆px∆x = (h∆k)(∆x) ≈ πh ≥ h

2, (9.8)

consistent with the uncertainty principle.Notice an interesting point: since ∆x ∝ 1

∆p , the narrower we make wave

number bandwidth ∆k, the broader we make the positional uncertainty ∆x5.Conversely, the broader we make ∆k, the narrower we make ∆x. The Heisen-berg Uncertainty Principle says that this is general − for any class of statefunctions

∆x ≥ h

2∆p∝ 1

∆px(9.9)

∆px ≥h

2∆x∝ 1

∆x(9.10)

More examples will be shown in the near future. Actually, the HeisenbergUncertainty Principle holds in three dimensions − so there are really threerelations

∆x∆px ≥h

2(9.11)

∆y∆py ≥h

2(9.12)

∆z∆pz ≥h

2(9.13)

We illustrate today next with a physical example.

9.3 Example: Position Measurement by Aperture

Suppose we prepare an electron so that its state function is a plane wave ofwave number k. This is accomplished by making a very 6 precise measurementof momentum − if the result p = hk is achieved, then the immediatelyafterward, the state function is the plane wave. We take this as

Ψ(x, t) = Aei(kx−ωt) (9.14)5with a flat distribution, but a narrower ∆k, you still get a sinc2 function for |Ψ|2, its just broader in x.6actually, infinitely precise − this is an idealization

68

Page 70: PHY 416, Quantum Mechanics - SIUE

where k = ph , ω = E

h . In this state, px = hk (definitely) and py is definitelyzero. However, ∆x = ∞ and ∆y = ∞ (plane waves extends over x andy). So ∆y∆py = “∞ · 0”, which is not inconsistent with the HeisenbergUncertainty Principle ∆y∆pt ≥ h

2 . Now suppose we attempt to measure ywith an aperture:

fig2Immediately after “passing through the aperture”, ∆y = a, the aperture

size. Thus, we have reduced ∆y. This means that the particle now has a newstate function − this one with ∆y ≥ a. (This is general − a measurementor a disturbance on the particle causes a change in the state function. Moreon this later). In fact, we can reduce ∆y as much a we like (by making a assmall as we lke).

However, in the new state function, ∆py is no longer zero! This isbecause the state function has diffracted thru the aperture!

fig3The arrows show “possible trajectories” to points on the screen. (Remem-

ber, though, a single trajectory is not actually taken unless we force a one bymeasurement − recall the Feynman double-slit experiment − the situation issimilar here.) But, p = h

λ , so, since ∆py = p∆θ = pλa , ∆py = hλ ·

λa = h

a . Now,at any time after passage thru the aperture, ∆y > a, so on the far side of theaperture, the new state function obeys

∆y∆py ≥ h ≥ h

2(9.15)

Thus, our attempt to reduce ∆y in the new state function has been compen-sated with an increase in ∆py, in accordance with Heisenberg UncertaintyPrinciple .

9.4 Expectation Values

We’ve established, that if you prepare an ensemble of very many identical par-ticles all with the same state function Ψ(x, t), position-producing measure-ments for each will generally produce different materialization places spreadof a “range” 2∆x. It is often important to know what average would be

69

Page 71: PHY 416, Quantum Mechanics - SIUE

obtained if all these position measurements were (hypothetically) to be per-formed. This average is called the expectation value of x, denoted by 〈x〉or x.

It is easy to find a general formula for 〈x〉. As your text shows (p.p. 5-10),if P (x) is the probability density function, then

〈x〉 =

∫ +∞

−∞xP (x) dx (9.16)

since P (x) = Ψ∗(x, t)Ψ(x, t), 〈x〉 is a function of t and is

〈x〉 =

∫ +∞

−∞x |Ψ(x, t)|2 dx =

∫ +∞

−∞Ψ∗(x, t)xΨ(x, t) dx (9.17)

similarly,

〈x〉2 =

∫ +∞

−∞Ψ∗(x, t)x2Ψ(x, t) dx (9.18)

and for an analytical function, f(x),

〈f(x)〉 =

∫ +∞

−∞Ψ∗(x, t)f(x)Ψ(x, t) dx (9.19)

What about the expectation value for momentum from momentum-producingmeasurements? For the time being, we will assume7 that

〈p〉 = md 〈x〉dt

(9.20)

Now

d 〈x〉dt

=d

dt

∫ +∞

−∞|Ψ(x, t)|2 dx (9.21)

=

∫ +∞

−∞x

(∂

∂t|Ψ(x, t)|2

)dx (9.22)

We’ve already seen that

∂t|Ψ|2 =

∂x

[ih

2m

(Ψ∗∂Ψ

∂x− ∂Ψ∗

∂xΨ

)](9.23)

7We will prove this assertion later, again, through Fourier analysis.

70

Page 72: PHY 416, Quantum Mechanics - SIUE

sod 〈x〉dt

=ih

2m

∫ +∞

−∞x∂

∂x

(Ψ∗∂Ψ

∂x− ∂Ψ∗

∂xΨ

)dx (9.24)

Integrating this by parts and using Ψ → 0 as x → ±∞ (and also ∂x∂x = 1)

yieldsd 〈x〉dt

= − ih

2m

∫ +∞

−∞

(Ψ∗∂Ψ

∂x− ∂Ψ∗

∂xΨ

)dx (9.25)

The second term can be integrated by parts again, and again throwing awaythe boundary term, it is equal to the first term, so

d 〈x〉dt

= −ihm

∫ +∞

−∞Ψ∗∂Ψ

∂xdx (9.26)

so

〈p〉 =h

i

∫ +∞

−∞Ψ∗(x, t)

∂Ψ(x, t)

∂xdx (9.27)

We can write this as

〈p〉 =

∫ +∞

−∞Ψ∗(x, t) · h

i

∂xΨ(x, t) dx (9.28)

along with

〈x〉 =

∫ +∞

−∞x |Ψ(x, t)|2 dx =

∫ +∞

−∞Ψ∗(x, t)xΨ(x, t) dx (9.29)

From these we say that the “position basis” (what we are so far working in)the (multiplicative) “operator” x = x “represents” position, and the operatorp = h

i∂∂x “represents” momentum. (We will see in the future that the utility of

this operator concept, seemingly superfluous here, has significance far beyondthat for these simple expectation values.)

Similarly, as you can show, if f(p) is any polynomial in p (e.g. KE = p2

2m),then

〈f(p)〉 =

∫ +∞

−∞Ψ∗(x, t)f

(h

i

∂x

)Ψ(x, t) dx (9.30)

Note also the action of p (or pop or p) on a plane wave state:

pei(phx−

Eh t) =

h

i

∂x

[ei(

phx−

Eh t)]

=h/

i/

i/p

h/

[ei(

phx−

Eh t)]

(9.31)

= pei(phx−

Eh t) (9.32)

71

Page 73: PHY 416, Quantum Mechanics - SIUE

We see that, for a plane-wave state, p on it is equivalent to multiplication bythe number p.

Now, any function of p is also an operator. For example, consider thekinetic energy K = p2

2m . Then

K =p

2m=h

i

∂x

(h

i

∂x

)= − h2

2m

∂2

∂x2 (9.33)

Thus, (in one dimension),

〈K〉 =

∫ +∞

−∞Ψ∗(x, t)KΨ(x, t) dx (9.34)

=

∫ +∞

−∞Ψ∗(x, t)

(− h2

2m

∂2Ψ(x, t)

∂x2

)dx (9.35)

Which is

〈K〉 = − h2

2m

∫ +∞

−∞Ψ∗(x, t)

∂2Ψ(x, t)

∂x2 dx (9.36)

Now consider the total energy E = K + V . Then,

E = − h2

2m

d2

dx2 + V (x, t) (9.37)

And, consider the Schrodinger equation

ih∂Ψ(x, t)

∂t= − h2

2m

∂2Ψ(x, t)

∂x2 + V (x, t)Ψ(x, t) (9.38)

which is then

EΨ(x, t) = ih∂

∂tΨ(x, t) (9.39)

Thus, the energy operator is also represented by

E = ih∂

∂t(9.40)

so

〈E〉 =

∫ +∞

−∞Ψ∗(x, t)

[− h2

2m

∂2

∂2x+ V (x, t)

]Ψ(x, t) dx (9.41)

= − h2

2m

∫ +∞

−∞Ψ∗(x, t)

∂2Ψ(x, t)

∂x2 dx (9.42)

+

∫ +∞

−∞Ψ∗(x, t)V (x, t)Ψ(x, t) dx (9.43)

72

Page 74: PHY 416, Quantum Mechanics - SIUE

and also

〈E〉 = ih

∫ +∞

−∞Ψ∗(x, t)

∂Ψ(x, t)

∂tdx (9.44)

73

Page 75: PHY 416, Quantum Mechanics - SIUE

Chapter 10

Tenth Class: Heisenberg UncertaintyPrinciple and Fourier Analysis

Thursday, September 25, 2008

10.1 Heisenberg Uncertainty Principle − Another Example

Here is another example. Suppose that we prepare a particle in a plane wavestate (or any other state with substantial ∆x) and decide to see if we can forcea violation of the Heisenberg Uncertainty Principle by actually “looking” atthe particle with a “super-microscope” and “seeing” were it materializes. Ifwe can do this to accuracy ∆x < h

2∆p , where ∆p is the resulting uncertaintyin momentum, then we would be able to violate the Heisenberg UncertaintyPrinciple . Let us see what we can do.

For this, we use the simplest “microscope” − one lens.fig1Now, because of the finite diameter of the lens, the probability density at

the screen of materialized particle is, as you know, not a point, but rather alittle diffraction pattern

fig2As we saw in discussing physical optics, based on this diffraction distribu-

tion, two objects a distance δ apart can be resolved only if

2δ sin

2

)≥ λ

74

Page 76: PHY 416, Quantum Mechanics - SIUE

δ >λ

2 sin(φ2

) (10.1)

Thus, if the limit of resolution on where the electron materializes is

∆x =λ

2 sin(φ2

) (10.2)

Of course, we can make this as small as we want by either using λ as smallas possible (gamma rays) or by making φ large (big diameter lens). In anycase, the positional size of the electron wave packet after this observation isgiven by this equation (or is larger).

Now, let us inquire into the momentum of the electron after the viewing.In the initial state, we had ∆p = 0. But, in the measurement, a certainamount of momentum is transferred to the electron. How much? To attemptto keep track of this, we assume a photon-electron collision (i.e. Comptonscattering) of a single photon of known momentum (pγ = E

c = hλ)with the

electron.But, after the collision, the photon could have entered the lens anywhere,

thus, the x-component of the photon momentum after the collision is uncer-tain by amount ∆px = h

λ sin θ.fig3Thus, after the collision, the electron is in a packet state such that for it

∆px ≥h

λsin θ (10.3)

This is equal only if the initial e− momentum is zero. Considering, then, thefinal packet state for the electron, we have

∆x∆px ≥λ

2 sin θ· hλ

sin θ (10.4)

or

∆x∆px ≥h

2>h

2(10.5)

in full accord with the uncertainty principle.In the context of this example, a key “reason” for the result is Einstein’s

photon hypothesis, according to which electromagnetic radiation is delivered

75

Page 77: PHY 416, Quantum Mechanics - SIUE

(e.g., to the electron being forced to materialize at a point) is indivisiblepackets, each of momentum px = h

λ . If it were possible that pγ < hλ for

wavelength λ, then we could scatter photons from electrons with sufficientminimal unknowable x-momentum transfer to violate the Heisenberg Uncer-tainty Principle .

Many examples of this sort have been investigated; a real possibility ofviolating the Heisenberg Uncertainty Principle has never been found.

10.2 Fourier Analysis

Considering our general “wave packet”

Ψ(x, t) =

∫ k2

k1

A(k)ei(kx−ωt) dk, (10.6)

where k = ph and ω = E

h . We mentioned that the probability of obtaining avalue of momentum (on momentum measurement) in the infinitesimal rangedp centered on p is

P (p)dp ∝ |A(k)|2 dk, dk =dp

h(10.7)

As we remarked, in order for this to be really useful, we would like to findout how to deduce, for a given Ψ(x, t), what function A(k) is1. For this, weneed Fourier analysis:

10.2.1 Basic Fourier Analysis

We will begin by recalling that any function f(x) with period 2π can beexpanded as

f(x) =a0

2+ a1 cosx+ a2 cos(2x) + a3 cos(3x) + a4 cos(4x) + · · ·

+ b1 sinx+ b2 sin(2x) + b3 sin(3x) + b4 sin(4x) + · · ·

or

f(x) =a0

2+∞∑n=1

an cos(nx) +∞∑n=1

bn sin(nx). (10.8)

1Of course, we need also to be assured that, if given an arbitrary reasonable state function Ψ(x, t), that A(k)exists for it. Fourier analysis answers this question, as we will see.

76

Page 78: PHY 416, Quantum Mechanics - SIUE

If the function is periodic with a different period (call it 2L), the the expansionbecomes

f(x) =a0

2+∞∑n=1

an cos(nπxL

)+∞∑n=1

bn sin(nπxL

)(10.9)

We recall that the coefficients are found from

an =1

L

∫ L

−Lf(x) cos

(nπxL

)dx (10.10)

bn =1

L

∫ L

−Lf(x) sin

(nπxL

)dx (10.11)

where the integrals could be over any full period of f(x); I’ve arbitrarily chosen−L→ L. Recall also that, key to establishing the formulas for the coefficientsare orthogonality relations∫ L

−Lcos(mπx

L

)cos(nπxL

)dx = 0 if m 6= n (10.12)∫ L

−Lsin(mπx

L

)cos(nπxL

)dx = 0 if m 6= n (10.13)∫ L

−Lsin(mπx

L

)sin(nπxL

)dx = 0 if m 6= n (10.14)

∫ L

−Lsin2

(nπxL

)dx =

∫ L

−Lcos2

(nπxL

)dx = L (10.15)

Thus, the set of functions1√2L,

1√L

sin(nπxL

),

1√L

cos(nπxL

)(10.16)

form an orthonormal set on any interval of length 2L.

10.2.2 Complex Orthogonal Functions

Consider two complex valued functions of the real variable x:

f(x) = u(x) + iv(x) (10.17)

g(x) = w(x) + is(x) (10.18)

77

Page 79: PHY 416, Quantum Mechanics - SIUE

where u(x), v(x), w(x), s(x) are real. f(x) and g(x) are said to be orthogonalon the interval [a, b] if∫ b

a

f ∗(x)g(x) dx =

∫ b

a

g∗(x)f(x) dx = 0. (10.19)

This is a direct generailization of the definition of orthogonality for real func-tions. It is interesting and easy to show that complex exponentials, also froman orthogonal set. To see this, consider (say, on interval [0, 2π]).

I =

∫ 2π

0yn(x)y∗m(x) dx (10.20)

where yn(x) = einx. We have

I =

∫ 2π

0einxe−imx dx (10.21)

=1

i(n−m)ei(n−m)x|2π0 = 0 if m 6= n. (10.22)

If n = m, ∫ 2π

0einxe−inx dx =

∫ 2π

01 dx = 2π (10.23)

Thus, another orthonormal set on [0, 2π] is1√2π

einx, n = −∞→ +∞ in integer steps. (10.24)

10.2.3 Complex Fourier Series on [0, 2π]

This leads us to an interesting conjecture − if f(x) is an complex “arbitrary”periodic function of period 2π, perhaps we can expand this as

f(x) = c0 + c1eix + c−1e

−ix + c2e2ix + c−2e

−2ix + · · · ? (10.25)

then,

f(x) =+∞∑

n=−∞cne

inx ? (10.26)

78

Page 80: PHY 416, Quantum Mechanics - SIUE

Let us see if we can find coefficients cn consistent with this. Assuming thatequation (10.25) is true, by integrating both sides we get,

1

∫ 2π

0f(x) dx =

1

∫ 2π

0c0 dx+

1

∫ 2π

0c1e

ix dx+1

∫ 2π

0c1e−inx dx+ · · ·

However, all integrals on the right except the first vanish.∫ 2π

0eix dx =

1

i

[eix]2π

0 =1

i

[e2πi − e0] =

1

i[1− 1] = 0 (10.27)

Thus,

c0 =1

∫ 2π

0f(x) dx (10.28)

To find cn, multiply equation (10.25) by e−inx and integrate:

1

∫ 2π

0f(x)e−inx dx =

c0

∫ 2π

0e−inx dx+

c1

∫ 2π

0e−inxeix dx

+c−1

∫ 2π

0e−inxe−ix dx+ · · ·

=c0

∫ 2π

0e−inx dx+

∞∑m=1,m 6=n

∫ 2π

0ei(m−n)x dx

+

∫ 2π

0e(n−n)x dx

These integrals are all zero except the last one, which is 2π; thus

cn =1

∫ 2π

0f(x)e−inx dx (10.29)

10.2.4 Complex Fourier Series on [−`, `], [0, L]

In the interval is [−`, `], the series is π → nπ`

f(x) =+∞∑

n=−∞cne

inπx` (10.30)

79

Page 81: PHY 416, Quantum Mechanics - SIUE

with

cn =1

2L

∫ `

−`f(x)e

−inπx` dx (10.31)

From this, if the period is L (i.e. we are considering an “reasonable” complex-valued function of real variable x such that f(x+L) = f(x)), then the seriescan be gotten by the replacement in our above work of 2` by L (2`→ L)

f(x) =+∞∑

n=−∞cne

2inπxL (10.32)

with

cn =1

L

∫ L

0f(x)e−

2inπxL dx (10.33)

It can be shown that these series converge for any “reasonable” f(x). Thus,the set

einx, n = −∞, · · · ,∞ is a complete set.

10.2.5 Special Case of Real f(x)

If f(x) is real, then

c∗n =

(1

L

∫ L

0f(x)e−

2inπxL dx

)∗=

1

L

∫ L

0f(x)e

2inπxL dx = cn (10.34)

In this case, our series

f(x) =∞∑n=1

cne2inπxL +

0∑n=−∞

c−ne− 2inπx

L (10.35)

becomes

f(x) =∞∑n=0

(cne

2inπxL + n∗ne

− 2inπxL

)(10.36)

Noting that in this case,

cn =1

L

∫f(x)e−

2inπxL dx

=1

L

∫f(x) cos

(2nπx

L

)− i

L

∫f(x) sin

(2nπx

L

)80

Page 82: PHY 416, Quantum Mechanics - SIUE

where 1L

∫f(x) cos

(2nπxL

)is an and − i

L

∫f(x) sin

(2nπxL

)is ibn. Then

cn = an − ibn (10.37)

andc∗n = an + ibn (10.38)

Then, equation (10.36) becomes

f(x) =∞∑n=0

(an − ibn)

[cos

(2nπx

L

)+ i sin

(2nπx

L

)]+ (an + ibn)

[cos

(2nπx

L

)− i sin

(2nπx

L

)]which is

f(x) =∞∑n=0

an cos

(2nπx

L

)+ bn sin

(2nπx

L

)(10.39)

which itself is the familiar Fourier sine and cosine series for a real functionf(x).

10.3 Fourier Expansion of Complex Non-periodic Functions − FourierTransforms

For a complex function with a period L, we saw that the Fourier expansion

f(x) =+∞∑

n=−∞cne

2inπxL (10.40)

and

cn =1

L

∫ L

0f(x)e−

2inπxL (10.41)

is valid as long as f(x) is “reasonable”.What if our expansion function f(x) is not periodic? “Not periodic” means

the same as “period → ∞”. Therefore, we let L → ∞ in our equationsabove, taking the interval (length L) as

[−L

2 ,L2

]. Then, defining k ≡ 2nπ

L ,and ∆k ≡ 2π

L ,

f(x) =+∞∑

k=−∞

cneikx =

+∞∑k−∞

L

2πcne

ikx∆k (10.42)

81

Page 83: PHY 416, Quantum Mechanics - SIUE

We now take the limit L→∞; the sum then becomes

f(x) =

∫ +∞

−∞

1√2π

Lcn√2π

eikx dk (10.43)

Since,

cn =1

L

∫ L2

−L2f(x′)e−

2inπx′L dx′, (10.44)

f(x) =

∫ +∞

−∞

1√2π·(

1√2π

∫ +∞

−∞f(x′)e−ikx

′dx′)

eikx dk (10.45)

Let us define the quantity

g(k) ≡ 1√2π

∫ +∞

−∞f(x)e−ikx dx (10.46)

Then we see that we have established a very interesting and very importanttheorem: the Fourier Transform Theorem:

Theorem:Suppose that f(x) is a square-integrable function i.e.,∫ +∞

−∞|f(x)|2 dx <∞ (10.47)

Then, there exists a function g(k) such that

f(x) =1√2π

∫ +∞

−∞g(k)eikx dk (10.48)

where

g(k) =1√2π

∫ +∞

−∞f(x)e−ikx dx (10.49)

82

Page 84: PHY 416, Quantum Mechanics - SIUE

Chapter 11

Eleventh Class: Fourier Analysis andthe Heisenberg Uncertainty Principle

Tuesday, September 30, 2008

11.1 Fourier Analysis

In the last class, we established the very important Fourier Transform theo-rem:

Theorem: Suppose that f(x) is a complex-valued square integrable func-tion of real variable x − i.e.

∫ +∞−∞ |f(x)|2 dx < ∞. Then, there exists a

function g(k) such that

f(x) =1√2π

∫ +∞

−∞g(k)eikx dk (11.1)

where

g(k) =1√2π

∫ +∞

−∞f(x)e−ikx dx (11.2)

Since quantum mechanical state functions are always square-integrable, wesee that this theorem addresses our concerns about the existence of “A(k)”for any state function Ψ(x, t0) at any particular time t0 − e.g. “A(k)” at timet = 0 is just 1√

2πg(k) above. (As we will see, for a free particle, A(k) does

not change in time. Of course, both f(x) and g(k) can be complex valuedfunctions. We’ll see later that if f(x) is square-integrable, so is g(k).

83

Page 85: PHY 416, Quantum Mechanics - SIUE

Example: Suppose that, at t = 0,

Ψ(x, t = 0) = f(x) =

1√a|x| ≤ a

2

0 |x| > a2

(11.3)

fig 1Then

g(k) =1√2π

∫ +∞

−∞f(x)e−ikx dx

=1√2πa

∫ a2

−a2e−ikx dx

=1√2πa

[− 1

ike−ikx

]a2

−a2

=1√2πa

sin(ka2

)k2

g(k) =

√a

2πsinc

(ka

2

)fig 2Now you might think − O.K. − let me check this − if I were to take this

g(k) and Fourier transform it, I should get f(x) back − that is,

1√2π

∫ +∞

−∞

√a

sin(ka2

)ka2

eikx dk =

1√a

if |x| ≤ a2

0 if |x| > a2

(11.4)

This is true, but unfortunately, actually showing this by evaluating the inte-gral on the left is not easy and requires an advanced technique called “contourintegration”. We won’t get involved in that. Consider, however, an example(that we really did some time ago) that will serve as a kind of “complemen-tary example” to the previous − say we now start with a flat distribution ofwave numbers and ask what f(x) this makes (“Fourier synthesis”)

fig 3

g(k) =

1√∆k

if |k| < ∆k2

0 if |k| > ∆k2

(11.5)

84

Page 86: PHY 416, Quantum Mechanics - SIUE

Then

f(x) =1√∆k

∫ ∆k2

−∆k2

1√∆k

eikx dk (11.6)

=1√

2π∆ksinc

(x∆k

2

)(11.7)

fig 4Let us look at an aspect of these examples: In the first, we have for the

“full width” of g(k) (spread between first zeros on either side of the center),

∆k =4π

a(11.8)

If we take the full-width in x (call it ∆x) as a, we then have, for this example,

∆k∆x = 4π (11.9)

For the second example, as you easily show, we also have, for the full widths,∆k∆x = 4π. Thus in both examples,

∆k ∝ 1

∆x(11.10)

and∆k∆x ≥ 1 (11.11)

In order to see how general these are, let’s look at another example.Example: Suppose we take a Gaussian

f(x) = Ne− x2

2σ2x (11.12)

fig 5Here N is a normalization constant that we won’t bother with yet, σ is

the standard deviation. Then

g(k) =N√2π

∫ +∞

−∞dxe

− x2

2σ2x eikx (11.13)

This integral is doable (we won’t bother); the result is

g(k) =N√2π

√π

2σ2x

e−σ2xk

2

2 (11.14)

85

Page 87: PHY 416, Quantum Mechanics - SIUE

or

g(k) =N

2σxe−

σ2xk

2

2 (11.15)

This result is very interesting − we see that g(k) is also a Gaussian! Let’sfind out what its standard deviation width (σk). To do this, we put the result(g(k)) into standard Gaussian form:

g(k) = N ′e− k2

2σ2k =

N

2σxe−

σ2xk

2

k (11.16)

Thus

k2

2σ2k

=k2σ2

x

2(11.17)

σ2k =

1

σ2x

(11.18)

σk =1

σx(11.19)

Again the widths in x and k space are reciprocally related:fig 6We will have you work out a third example for homework: again you will

find that ∆x ∝ 1∆k . This should make it plausible that this is general prop-

erty. The statement σxσk ∼ 1 is called the : “Fourier Bandwidth Theorem”.Following, we will indicate a more general line of reasoning leading to thisimportant mathematical result.

11.2 A More General Line of Reasoning Leading to the FourierBandwidth Theorem

Consider the function f(x) synthesized as

f(x) =1√2π

∫ +∞

−∞g(k)eikx dk (11.20)

where g(k) is real and symmetrically peaked around k = k0 with standarddeviation δk. Then, as we know, schematically f(x) looks like

fig 7

86

Page 88: PHY 416, Quantum Mechanics - SIUE

peaked at x = 0. Suppose we seek to understand the “qualitatively”.We ask: Why is f(x) biggest at x = 0? It’s biggest at x = 0 since allcontributions eikx for all the different k’s in the wave number spectrum arein phase at x = 0 (all eik·0 = 1), so big resultant there. However, at anyother x 6= 0, f(x) must be smaller than f(0). Why? Consider the synthesisat x = x0 6= 0. The integral

f(x0) =

∫ +∞

−∞g(k)eikx0 dk (11.21)

contains contributions from an infinite number of “phasors” eikx0 dkone (ofinfinitesimal magnitude) for each value of k the integral sweeps over. But,if x0 6= 0, each such phasor carries a different phase, so there is a degree ofdestructive interference in the sum implied by the integral, and thus f(x0) <f(0). As we initially increase (or decrease) x0 from 0, the greater the phasedifference between “successive” phasors eikx0 dk (for neighboring k’s) in theintegral. Now suppose we go out to the “end of the packet in position space”,i.e. x = x1. There, the phasors

eokx1

for the different k’s uniformly cover a

phase interval of 2π − “for every phasor pointing in a given ‘direction’ (say 20’clock), there’s another pointing opposite (8 o’clock); so we get cancellation”.This is shown in figure 11.8.

fig 8So, at x1, ∆kx1 = 2π, where ∆k is the full-bandwidth. Then if the full-

width in x is 2x, we have

∆xfw ≈2π

∆kfw(11.22)

or∆xfw∆kfw ≈ 2π (11.23)

Let’s say that we want the product σxσk instead of the product of full-widths.At least approximately, for symmetric g(k) and f(x), σx or k ≈ 1

3∆fw(or ∆kfw).Of course, the exact ratio of the standard deviations to full-widths dependson the exact shapes of g(k) and f(x); here we are only working toward a“rule-of-thumb” understanding. If we take σ ∼ 1

3∆fw, then we have

σxσk ∼2π

9≥ 1

2(11.24)

We will take σxσk ≥ 12 as our rule-of-thumb statement.

87

Page 89: PHY 416, Quantum Mechanics - SIUE

11.3 Heisenberg Uncertainty Principle

Let us now apply the forgoing to Quantum Mechanics. Let φ(k) be theFourier transform of ψ(x) ≡ Ψ(x, t0) for some particular time t0. Them wesee that, in general, for state functions,

∆φ ∝1

∆φ(11.25)

where ∆xψ is the x-width of φ(x) and ∆kψ is the k-width of φ(k). Now,for the Gaussian case, we saw that, if we use ∆xψ = σx, ∆kψ = σk, then∆xψ∆ψ = 1 (Gaussian, for full-widths). It turns out that the Gaussian caseachieves the minimum possible value for σxσk. (We will prove this later inthe course). However, in quantum mechanics, the Heisenberg UncertaintyPrinciple refers to probabilities, not to state functions themselves. Now, theprobability of materialization per unit x as a function of x is

dP (x)

dx= |ψ(x)|2 (11.26)

Likewise, the probability on momentum measurement of materializing of mo-mentum p per unit interval of k is

dP (k)

dk= |φ(k)|2 φ(k) ≡ g(k) (11.27)

Now, for the Gaussian

ψ(x) = Ne− x2

2σ2x ⇒ |ψ(x)|2 = |N |2 e

− x2

σ2x (11.28)

Thus,

σx|ψ|2 =1√2ψxψ (11.29)

Likewise

σx|φ|2 =1√2σkφ (11.30)

Thus letting,

∆x ≡ σx|ψ|2 (11.31)

∆k ≡ σx|φ|2 (11.32)

88

Page 90: PHY 416, Quantum Mechanics - SIUE

We have

∆x∆k =1

2(11.33)

since p = hk, we have, for the Gaussian case,

∆x∆px =h

2(11.34)

As I remarked above, this product is minimum of all cases for the Gaussian.Thus, in general

∆x∆px ≥h

2(11.35)

This, again, is the Heisenberg Uncertainty Principle . We have already lookedat physical examples (single-aperture, Bohr microscope, etc.) that explicateor illustrate its physical meaning.

Note: Note that ∆x and ∆p in the Heisenberg Uncertainty Principle refer

to standard deviations (σ’s), strictly, these are defined by ∆x =⟨

(x− x)2⟩ 1

2

,

where

〈x〉 = x =

∫ψ∗(x) x ψ(x) dx (11.36)

∆x =

[∫ψ∗(x) (x− x) ψ(x) dx

] 12

(11.37)

similarly for ∆[. For Gaussian distributions these definitions correspond to σ(“half-width”). Quoting another author, “The specific value of the numberon the right hand side is not [so] important.” What is important is theHeisenberg Uncertainty Principle viewed as an order-of-magnitude relation,inequality. (reciprocally proportional quantities)

11.4 A Square-Normalized f(x) Implies a Square-Normalized g(k)

There is a handy little theorem, called Parseval’s Theorem by mathemati-cians, that shows this. It follows easily from the Fourier Transform Theoremas follows:∫ +∞

−∞g(∗(k)g(g) dk =

∫ +∞

−∞g∗(k)

[1√2π

∫ +∞

−∞f(x)e−ikx dx

]dk (11.38)

89

Page 91: PHY 416, Quantum Mechanics - SIUE

As physicists we cavalierly interchange orders of integrations, thus∫ +∞

−∞g∗(k)g(k) dk =

∫ +∞

−∞

[1√2π

∫ +∞

−∞g∗(k)e−ikx dk

]f(x) dx(11.39)

=

∫ +∞

−∞f ∗(x)f(x) dx (11.40)

Thus, if f(x) is square-normalized, so is g(k). And vice-versa. Nice thisParseval’s theorem!

11.5 Fourier Analysis of “Arbitrary” f(x, t)

The Fourier theorem says that “any” f(x) can be synthesized as

f(x) =1√2π

∫ +∞

−∞g(k)eikx dk (11.41)

with

g(k) =1√2π

∫ +∞

−∞f(x)e−ikx dk (11.42)

Suppose you have an arbitrary continuous square -normalizable function oftwo variables f(x, t). Think of f(x, t) as defining a different function f(x)for each t. Then, for each t, the pair of equations above are

f(x, t) =1√2π

∫ +∞

−∞g(k, t)eikx dk (11.43)

where

g(k, t) =1√2π

∫ +∞

−∞f(x, t)e−ikx dx (11.44)

The “catch” is that g is now different at each time t; there is no general rulerelating g(k, t) to g(k, t = 0). Thus, it is, for example, not true that “any”function of x and t can be written as superposition of plane waves − i.e.,there exists no g(k) such that generally

f(x, t) =1√2π

∫ +∞

−∞g(k)ei(kx−ωt) dk

for ω some function of k.

90

Page 92: PHY 416, Quantum Mechanics - SIUE

In quantum mechanics there is, however, one class of very important ex-ceptions to this warning: As we will see shortly (in the coming section) “Gen-eral Solution to Schrodinger equation for a free particle”’, for a free particle(only) in quantum mechanics, g(k, t) is easily predictable from g(k, t = 0),which (as we’ll see) allows a plane wave expansion for any free particle statefunction Ψfree(x, t). That simplifies things considerably for the case of a freeparticle.

11.6 Dirac Delta Function

We found that, if ψ(x) is a square-integrable function i.e.,∫ +∞−∞ |ψ(x)|2 dx <

∞, then there exists a function g(k) such that

f(x) =1√2π

∫ +∞

−∞g(k)eikx dk (11.45)

where

g(k) =1√2π

∫ +∞

−∞ψ(x)e−ikx dx (11.46)

f and g are Fourier transforms of each other. By combining the two aboveequations, we can write

f(x) =1

∫ +∞

−∞eikx

[∫ +∞

−∞f(x′)e−ikx

′dx′]dk (11.47)

by changing the order of integration1 this is

f(x) =

∫ +∞

−∞f(x′)

[1

∫ +∞

−∞eik(x−x′) dk

]dx′ (11.48)

for the quantity in brackets we use the notation

1

∫ +∞

−∞e−k(x−x′) dk = δ(x′ − x) (11.49)

Thus equation (11.48) becomes

f(x) =

∫ +∞

−∞δ(x′ − x)f(x′) dx′ (11.50)

1again, as physicists we cavalierly do this, more on this later

91

Page 93: PHY 416, Quantum Mechanics - SIUE

We see that the object δ(x′ − x) is the Dirac delta function, since equation(11.50) (“sifting property”) defines the delta function.

For graduate students2: There is “another” way to see that equation(11.49) is a representation of the Dirac delta function. Consider the function

gb(x) =

1b for |x| < b

2

0 for |x| > b2

(11.51)

fig 9For this function

∫ +∞−∞ gb(x) dx = 1. If we take the limit of this as b → 0,

we get δ(x):δ(x) = lim

b→0gb(x) (11.52)

To check that this reproduces equation (11.50), note that the for “any” ar-bitrary function f(x),

limb→0

∫ +∞

−∞gb(x)f(x) dx = lim

b→0

∫ b2

− b2

1

bf(x) dx = f(0) (11.53)

where we take the limit after closing the integral, instead of (as we “should”)before doing the integral. (That is we can be cavalier like this about inter-changing the order of integrals and limits was “creatively intuited” by PaulDirac in his famous 1930 book on quantum mechanics; it was finally provedlogically rigorous by Laurent Schwartz in the late 1930’s.)

Now consider the Fourier transform of gb(x):

gb(k) =1√2π

∫ +∞

−∞gb(x)e−ikx dx

=1√2π

∫ b2

− b2

1

be−ikx dx

=1√2π

sin(kb2

)kb2

thus

limb→0gb(k) =1√2π

(11.54)

2and also for interested undergrads

92

Page 94: PHY 416, Quantum Mechanics - SIUE

Now, if we take the Fourier transform of limb→0 gb(k), we must get backlimb→0 gb(x) = δ(x). Doing this:

δ(x) = F[limb→0

gb(k)]

= F[

1√2π

]=

1√2π

∫ +∞

−∞

1√2π

eikx dk (11.55)

Which corroborates our previous result that

δ(x) =1

∫ +∞

−∞eikx dk (11.56)

similarly

δ(x− x0) =1

∫ +∞

−∞eik(x−x0) dk (11.57)

δ(k − k0) =1

∫ +∞

−∞eiu(k−k0) du (11.58)

(end of section “for graduate students”.)

11.7 The General Solution to Schrodinger’s Equation for a FreeParticle

As we know, all free-particle state functions must solve and obey the freeparticle Schrodinger equation

− h2

2m

∂2Ψ(x, t)

∂x2 = ih∂Ψ(x, t)

∂t(11.59)

In the past, we put together a fairly general solution of this as a superpositionof plane waves,

Ψ(x, t) =

∫ k2

k1

A(k)ei(kx− hk

2

2m t)dk (11.60)

We now ask if all free-particle solutions are of this form, of if there are addi-tional solutions. The following very nice treatment (from Ohanian3

3Ohanian, Hans. Principles of Quantum Mechanics. Benjamin Cummings, 1989.

93

Page 95: PHY 416, Quantum Mechanics - SIUE

We will now find the general solution of the Schrodinger waveequation for a free particle. By Fourier’s theorem, at one given timet, any normalizable wave function can be written as

Ψ(x, t) =1√2π

∫ +∞

−∞g(k, t)eikx dk (11.61)

We have indicated a time dependence in the Fourier transform g(k, t),because although Ψ(x, t) always can be expressed in the form givenin the previous equation, the Fourier transforms at different timeswill be different. We call g(k, t) the amplitude in momentum space.Next, we have to find g(k, t). For this we substitute the above equa-tion into the Schrodinger equation

− h2

2m

[1√2π

∫ +∞

−∞−k2g(k, t)eikx dk

]= ih

[1√2π

∫ +∞

−∞

∂tg(k, t)eikx dk

]If we compare the Fourier transforms, or the coefficients of eikx, onboth sides of this equation, we obtain

h2k2

2mg(k, t) = ih

∂tg(k, t) (11.62)

this has the obvious solution

g(k, t) = g(k, 0)e−ihk2t2m (11.63)

org(k, t) = g(k, 0)e−

iEth (11.64)

where

E =h2k2

2m(11.65)

is the energy that corresponds to the momentum p = hk. Equation(11-61) then becomes

Ψ(x, t) =1√2π

∫ +∞

−∞g(k, 0)e

i(kx− hk

2

2m t)dk (11.66)

This is the general solution of the Schrodinger’s wave equation fora free particle. The amplitude g(k, 0) is arbitrary, but must satisfy

94

Page 96: PHY 416, Quantum Mechanics - SIUE

the normalization condition∫ +∞

−∞|g(k, 0)|2 dk = 1 (11.67)

By Parseval’s theorem, this will imply the correct normalization forΨ(x, t). Thus, according to equation (11.66), the general solution ofthe free particle Schrodinger equation is a (continuous) superposi-tion harmonic waves. This means that the free-particle Schrodingerdoes not produce any solutions except those that can be directlyconstructed by superposition − the equation gives us nothing new.

If we chose the coefficients g(k, 0) in equation (11.66) correctly,we can construct a wave function Ψ that is nonzero only in somesmall region. Such a wave function is called a wave packet.

95

Page 97: PHY 416, Quantum Mechanics - SIUE

Chapter 12

Twelfth Class: Momentum Space

Thursday, October 2, 2008

12.1 Quantum Initial-Value Problem, Case of Free Particle

We showed that for a free particle (only) the most general state-function is(the most general solution to the free-particle Schrodinger equation)

Ψ(x, t) =1√2π

∫ +∞

−∞g(k, 0)e

i(kx− hk

2

2m t)dk (12.1)

where g(k, 0) is the Fourier transform of Ψ(x, t = 0). As we will illustrateby example shortly, this is clearly the formal solution to the free-particleinitial value particle problem (find Ψ(x, t) given Ψ(x, t = 0)) in quantummechanics. Prior to doing an example of this, we first note an interestingfact of it: Evaluating equation (12.1) at t = 0 yields

Ψ(x, t = 0) =1√2π

∫ +∞

−∞g(k, 0)eikx dk (12.2)

But, by Fourier’s theorem, any continuous square-normalizable function of xcan be represented in this form with the proper g(k, 0). Therefore, we havenow shown something we assumed in the past − that, for a free particle,Ψ(x, t = 0) can be (depending on the physical/experimential situation) anycontinuous square-normalizable function of x. (Of course, given Ψ(x, t = 0),Ψ(x, t) is not arbitrary − it is controlled by the Schrodinger equation. Nowwe discuss examples of the initial value problem:

96

Page 98: PHY 416, Quantum Mechanics - SIUE

Example: Suppose Ψ(x, t = 0) = cos(k0x) and the represented particle isfree. What is subsequent Ψ(x, t)?

Solution: Following our prescription, we Fourier analyze Ψ(x, t = 0) overthe set

eikx

. This Fourier analysis is

cos(k0x) =1

2eik0x +

1

2e−ik0x (12.3)

Note: This also comes out of formal Fourier series: setting up

f(x) = cos(k0x) =+∞∑

n=−∞cne

iknx

as you can show by evaluating, the integral expression for cn, wherekn = nk0. all cn = 0 except c1 = c−1 = 1

2 . You can also use theformal transform prescription − if you evaluate

g(k) =1√2π

∫ +∞

−∞cos(k0x)e−ikx dx

you will find

g(k) =1√2π

[δ(k − k0) + δ(k + k0)] (12.4)

which implies plane wave contributions for k = k0 and k = −k0 only

Then

Ψ(x, t = 0) =

√2π√2π

∫ +∞

−∞

1

2δ(k − k0)e

ikx dx+1√2π

∫ +∞

−∞

1

2δ(k + k0)e

ikx dx

=1

2

[eik0x + e−ik0x

]= cos(k0x)

so

Ψ(x, t) =1

2

[ei(k0x−ω0t) + e−i(k0+ω0t)

](12.5)

where ω0 = hk20

2m , which of course, shows that Ψ(x, t) 6= cos(k0x− ω0t).Example: Free particle1

1this is from Griffiths, worked example 2.6 pp 62-63

97

Page 99: PHY 416, Quantum Mechanics - SIUE

SupposeΨ(x, t = 0) is a “flat-top function” of full-width 2a (perhaps resultof measurement)

fig 1Find Ψ(x, t).Solution: First, normalize: A = 1√

2a. Then find g(k, 0):

g(k, 0) =1√2π

1√2a

∫ a

−ae−ikx dx =

1√πa

sin(ka)

k=

√a

πsinc(ka) (12.6)

then

Ψ(x, t) =

√a

π

∫ +∞

−∞sin(ka)e

i(kx− hk

2

2m t)dk (12.7)

The integral can be evaluated numerically, the result looks like we wouldexpect

fig 2Looking at limiting cases is instructive:

a) lima→0 ⇒ Ψ(x, t = 0) is a spike (∆x→ 0), Ψ(x, t = 0) = δ(x− x0) Then

sin(ka) ≈ ka

so

g(k) ≈√a

π

and this is flat (∆k →∞). Note that

F [δ(x− x0)] =1√2π

∫ +∞

−∞δ(x− x0)e

−ikx dx = e−ikx0 (12.8)

and this is flat.

b) lima→∞, then

g(k, 0) =

√a

πsinc(ka)

has max at k = 0, the first zero is at ka = π ⇒ k1 = πa → 0⇒ g(k, 0) ∝

δ(k − k0) (∆x→∞, ∆k → 0).

Example: In the past we had Ψ(x, t = 0) = Ae−ax2

, then I told you thatΨ(x, t) was also a Gaussian, but with standard deviation, σ, that gets widerin time. (this is how I get it, (in fact, its Griffiths 2.22)).

98

Page 100: PHY 416, Quantum Mechanics - SIUE

12.1.1 State Function of Definite Momentum

A state of definite momentum hk0 must have g(k, 0) = Aδ(k − k0) where Ais a normalization we won’t worry about. Let us see what this implies for afree particle,

Ψ(x, t) =1√2π

∫ +∞

−∞g(k, 0)e

i(kx− hk

2

2m t)dk (12.9)

Ψ(x, t) =A√2π

∫ +∞

−∞δ(k − k0)e

i(kx− hk

2

2m t)dk (12.10)

=A√2π

ei

(k0x−

hk20

2m t

)(12.11)

as we would expect. This shows the consistency of part of the theory.

12.1.2 State Function of Definite Position

This should beΨ(x, t = 0) = δ(x− x0) (12.12)

(spike in position at x0), again we don’t yet worry about the normalization.Then

g(k) =1√2π

∫ +∞

−∞δ(x− x0)e

−ikx dx (12.13)

=A√2π

∫ +∞

−∞δ(x− x0)e

ikx dx (12.14)

=A√2π

eikx0 (12.15)

Therefore,

|g(k)|2 =|A|2

2π(12.16)

99

Page 101: PHY 416, Quantum Mechanics - SIUE

this is flat in k, we we would expect by the bandwidth theorem. Futher, weshould have

Aδ(x− x0) = Ψ(x, t = 0)

=1√2π

∫ +∞

−∞g(k, 0)eikx dk

=1√2π

∫ +∞

−∞

A√2π

eik0xeikx dk

=A

∫ +∞

−∞ei(−k0)x dk

= Aδ(x− x0)

since, as we showed in last class

1

∫ +∞

−∞ei(k−k0)x dk = δ(x− x0) (12.17)

12.2 Momentum-Space State Function − General Potential En-ergy Profiles

As mentioned in the last class, by Fourier’s theorem, we can always Fourieranalyze any state function (free-particle or not) at any time t as

Ψ(x, t) =1√2π

∫ +∞

−∞g(k, t)eikx dk (12.18)

where

g(k, t) =1√2π

∫ +∞

−∞Ψ(x, t)e−ikx dx (12.19)

Here, Parseval’s theorem tells us that g(k, t) is normalized at any time t ifΨ(x, t) is normalized. g∗(k, t)g(k, t) dk is not just proportional to, but isequal to the probability that, upon momentum producing measurement, themomentum will materialize in range hdk centered on hk. We thus see thatin our early work where we dealt with packets for a free particle of the form

Ψ(x, t) =

∫A(k)ei(kx−ωt) dk (12.20)

100

Page 102: PHY 416, Quantum Mechanics - SIUE

that |A(k)|2 is indeed proportional but not equal to it − we were merelymissing the factor 1√

2π(g(k) = 1√

2πA(k) normalizes it).

now, in quantum mechanics, it is customary to recast equations (12.18)and (12.19) as integrals over a measurable quantity (momentum) rather thenover k. If we do this, equations (12.18) and (12.19) are

Ψ(x, t) =1√2π

1

h

∫ +∞

−∞(g [p(k), t] e

iph x dp (12.21)

g [p(k), t] =1√2π

∫ +∞

−∞Ψ(x, t)e−

iph x dx (12.22)

While correct, part of the symmetry is lost between equations (12.21) and(12.22) in this form. Therefore, it is conventional to rescale g(k, t) by defininga function φ(p, t) as

φ(p, t) ≡ 1√hg(k, t) (12.23)

In terms of φ, we have then

Ψ(x, t) =1√2πh

∫ +∞

−∞φ(p, t)e

iph x dp (12.24)

φ(p, t) =1√2πh

∫ +∞

−∞Ψ(x, t)e−

iph x dx (12.25)

which restores the symmetry. Since∫ +∞

−∞|g(k, t = 0)|2 dk = 1 (12.26)

it follows that ∫ +∞

−∞|φ(p, t = 0)|2 dp = 1 (12.27)

(show this) and also ∫ +∞

−∞|φ(p, t)|2 dp = 1 (12.28)

Thus, φ∗(p, t)φ(p, t) dp is the probability that the momentum will materializein the range dp centered on p

101

Page 103: PHY 416, Quantum Mechanics - SIUE

The function φ(p, t) is called the “momentum-space state function” or “thestate function in momentum-space”. Since, by equation (12.24), we can re-construct Ψ(x, t) from knowledge of φ(p, t), φ(p, t) provides for any system,a description completely equivalent to that of Ψ. Thus, it is possible to com-pletely do quantum mechanics in momentum space without ever mentioningΨ. We do not pursue this in this course.

If a system is in state Ψ(x, t) [φ(p, t)] and if a measurement is performed toproduce momentum at time t, the probability that the momentum producedis in the infinitesimal interval dp centered on p is

P (p, t) dp = φ∗(p, t)φ(p, t) dp (12.29)

We will take this as an axiom of quantum mechanics for the special caseof free particle. We note that, since

φfree(p, t) = φfree(p, 0)e−iEh t (12.30)

φ∗free(p, t)φfree(p, t) = φ∗free(p, 0)φfree(p, 0) (12.31)

12.3 Expectation Value Revisited

Now, if the above interpretation of φ∗(p, t)φ(p, t) is correct then we must beable calculate the expectation value of p from

〈p〉 =

∫ +∞

−∞pφ∗(p, t)Ψ(x, t) dp (12.32)

Let us check this, say at t = 0. Then if φ(p) ≡ φ(p, t = 0), we must have

〈p〉t=0 =

∫ +∞

−∞φ∗(p)pφ(p) dp (12.33)

and likewise, for any analytic function of p, s(p),

〈s(p)〉 =

∫ +∞

−∞φ∗(p)s(p)φ(p) dp (12.34)

where I’ve suppressed the t = 0 subscript on 〈 〉. Let us express equation(12.33) in terms of ψ(x) ≡ Ψ(x, t = 0). Using

φ(p) =1√2πh

∫ +∞

−∞ψ(x)e−

iph x dx (12.35)

102

Page 104: PHY 416, Quantum Mechanics - SIUE

we have

〈p〉 =

∫ [1√2πh

∫ψ(x)e−

iph x dx

]∗p

[1√2πh

∫ψ(x′)e−

iph x′dx′]dp

=1

2πh

∫ ∫ ∫dx dx′ dp ψ∗(x)ψ(x′)pe

ip(x−x′)h

=1

2πh

∫ ∫ ∫dx dpψ∗(x)ψ(x′)

(−hi

d

dx′eip(x−x′)

h

)dx′

integrating by parts on x′ and using ψ, ψ∗ → 0 as x→ ±∞.

〈p〉 =1

2πh

∫ ∫ ∫dx dx′ dpψ∗(x)e

ip(x−x′)h

(h

i

d

dx′

)ψ(x′)

=

∫ ∫dx dx′ψ∗(x)δ(x− x′)

(h

i

d

dx′

)ψ(x′)

〈p〉 =

∫ +∞

−∞dxψ∗(x)

h

i

d

dxψ(x) (12.36)

Likewise, as you can show, if f(p) is any polynomial in p,

〈f(p)〉 =

∫ +∞

−∞dxψ∗(x)f

(h

i

d

dx

)ψ(x) (12.37)

That these are the same results that we obtained some time ago shows theconsistency of momentum probability density interpretation of |φ(p)|2.

12.4 Reality of 〈p〉

Now, in the coordinate representation, we have

〈p〉t =

∫ +∞

−∞Ψ∗(x, t)

h

i

∂xΨ(x, t) dx (12.38)

which explicitly contains i ≡√−1. On the othr hand, since 〈p〉 is a measure-

ment result, it had better be real! Let us check this explicitly. From equation(12.38),

〈p〉∗ =

∫ +∞

−∞Ψ(x, t)

(h

−i∂Ψ∗

∂x

)dx (12.39)

103

Page 105: PHY 416, Quantum Mechanics - SIUE

By integrating by parts,

〈p〉∗ = −∫ +∞

−∞

(h

−i

)∂Ψ

∂xΨ∗ dx− h

i[Ψ∗Ψ]+∞−∞ (12.40)

=

∫ +∞

−∞

h

iΨ∗∂Ψ

∂xdx (12.41)

〈p〉∗ = 〈p〉 (12.42)

So, in spite of the i in the operator, 〈p〉 is real as it must be.

12.5 Hermitian Operators

In fact, the expectation value of any operator that represents a physical quan-tity (position, momentum, energy, angular momentum, etc.) must, of course,be real. Operators that have real expectation values for any permissible statefunction are called Hermitian operators.

Theorem: Let Ω be a Hermitian operator. Then, for any admissible statefunction This is also true for H. Let Ω = H or p or any other operator suchthat ∫ +∞

−∞Ψ∗ΩΨ dx =

∫ +∞

−∞

(ΩΨ)∗

Ψ dx (12.43)

Proof: If Ω is Hermitian, ⟨Ω⟩

=⟨

Ω⟩∗

∫Ψ∗ΩΨ =

(∫Ψ∗ΩΨ

)∗=

∫Ψ(

ΩΨ)∗

dx

=

∫ +∞

−∞

(ΩΨ)∗

Ψ dx

Theorem: If Ω is a Hermitian operator, and if Ψ1(x, t) and Ψ2(x, t) are anytwo admissible state functions, then∫ +∞

−∞Ψ∗1ΩΨ2 dx =

∫ +∞

−∞

(ΩΨ1

)∗Ψ2 dx (12.44)

104

Page 106: PHY 416, Quantum Mechanics - SIUE

Proof : We sketch the proof; you can carry out the details: Consider Ψ(x, t) =Ψ1 + cΨ2 for arbitrary complex constant c. Then∫

(Ψ∗1 + c∗Ψ∗2) Ω (Ψ1 + cΨ2) =

∫Ψ∗1ΩΨ1 + c∗c

∫Ψ∗2ΩΨ2

+ c

∫Ψ∗1(ΩΨ2) + c∗

∫Ψ∗2(ΩΨ1)

must be real. The The first two terms on the right we know are real by thelast theorem; consequently, the sum of third and forth terms must also bereal and hence equal to its complex conjugate:

c

∫Ψ∗1(ΩΨ2) + c∗

∫Ψ∗2(ΩΨ1) = c∗

∫Ψ1(Ω

∗Ψ∗2) + c

∫Ψ2(Ω

∗Ψ∗1)

Applying this equation twice, once for real c = a real constant b and once forc = ib and adding yields:∫

Ψ∗1ΩΨ2 dx =

∫(ΩΨ1)

∗Ψ2 dx (12.45)

which is the same as before.

12.6 Normalization of Free-Particle Definite Momentum States

Consider the free-particle definite momentum state at time t = 0:

ψp(x) = Aeiph x (12.46)

several different normalizations are used, depending on circumstances. Twoof them are discussed briefly in the following sections

105

Page 107: PHY 416, Quantum Mechanics - SIUE

12.6.1 “Box Normalization”

In this normalization scheme, we simply normalize to (large) length L. (“Theuniverse is finite”, “My lab is finite”, etc.). Then we simply require∫ L

2

−L2ψ∗ψ dx = 1 (12.47)

|A|2∫ L

2

−L21 dx = 1 (12.48)

A =1√L

(12.49)

Ψp(x, t = 0) =1√L

eiph x (12.50)

12.6.2 “Dirac Normalization” of Free-Particle Definite Momentum State

Although technically ψp(x) = Aeiph x is not square integrable, consider the

inner product

(ψp′(x), ψp(x)) ≡ 〈ψp′ (x)|ψp(x)〉 =

∫ +∞

−∞ψp′(x)ψp(x) dx (12.51)

= |A|2∫ +∞

−∞ei(p−p′)

h x dx (12.52)

= |A|2 · 2πhδ(p− p′) (12.53)

We pick A = 1√2πh

, then

ψp(x) =1√2πh

eiph x (12.54)

Then, as you can easily show, the normalization of ψp(x) is

〈ψp′ |ψp〉 = δ(p− p′) (12.55)

this is the continuum analogy of the Kronecker delta-orthonormality and iscalled “Dirac Normalization”

106

Page 108: PHY 416, Quantum Mechanics - SIUE

12.7 Normalization of a State of Definite Position

Let ψx0(x) be a state function of definite position x = x1. Then

ψx1(x) = Aδ(x− x1) (12.56)

Then consider the inner product for states of spike at x1 and x2:

〈ψx1|ψx2〉 =

∫ +∞

−∞ψ∗x1

(x)ψx2(x) dx

= |A|2∫ +∞

−∞δ(x− x1)δ(x− x2) dx

= |A|2 δ(x1 − x2)

we pick A = 1 Then〈ψx1

|ψx2〉 = δ(x1 − x2) (12.57)

This is Dirac Normalization of a state of definite position. Note that the setof states of definite position is a complete set: For any ψ(x)

ψ(x) =

∫ +∞

−∞c(x0)δ(x− x0) dx0 (12.58)

where c(x0) is the expansion coefficient and δ(x − x0) is the basis vector.Therefore, ψx0

(x) is a complete set. That the last expansion is valid canbe seen by working with the right hand side:

ψ(x) =

∫ +∞

−∞c(x0)δ(x− x0) dx (12.59)

comparing this to the known truth

ψ(x) =

∫ +∞

−∞ψ(x0)δ(x− x0) dx0 (12.60)

shows that c(x0) exists and is equal to ψ(x0).

12.8 Completeness of Free-Particle Definite Momentum States

The set ψp(x) =

1√2πh

eiph x

(12.61)

107

Page 109: PHY 416, Quantum Mechanics - SIUE

is also a complete set− as we know from Fourier’s theorem, any state functionψ(x) can be expanded as

ψ(x) =

∫ +∞

−∞c(p)Ψp(x, t) dp (12.62)

with coefficient

c(p) =1√2πh

φ(p) (12.63)

the the completeness statement is just

ψ(x) =1√2πh

∫ +∞

−∞φ(p)e

iph x dp (12.64)

with φ(p) is the expansion coefficient and eiph x is the basis function. And we

know this to be true!These expansions are analogous to those of an ordinary vector in 3-space.

~A = Axı+ Ay + Azk (12.65)

The expansion coefficients are

Ax = ~A · ıAy = ~A · Az = ~A · k

and ı, , and k are the basis vectors.

108

Page 110: PHY 416, Quantum Mechanics - SIUE

Chapter 13

Thirteenth Class

Tuesday, October 7, 2008 Exam 1

109

Page 111: PHY 416, Quantum Mechanics - SIUE

Chapter 14

Fourteenth Class:

Thursday, October 9, 2008 No Class

110

Page 112: PHY 416, Quantum Mechanics - SIUE

Chapter 15

Fifteenth Class: States of DefiniteEnergy

Tuesday, October 14, 2008

15.1 Eigenstate Property

As we have seen, the Schrodinger equation is[p2

2m+ V (x, t)

]Ψ(x, t) = ih

∂Ψ(x, t)

∂t(15.1)

The operator in the left, p2

2m + V , we call the “Hamiltonian”, (H) as itrepresents the total energy. Now, for any particular function V (x, t), theSchrodinger equation has, in general, a large variety of solutions. How canwe hope to be guided in identifying these solution? In the case of the freeparticle, we know that

1. Momentum is classically conserved (~F = d~pdt , but ~F = 0) and total energy

is conserved

2. States of definite momentum are also states of definite energy − theplane waves states

3. The general solution of the Schrodinger equation is the general superpo-sition of the definite energy and definite momentum states

However, if the particle is not free, the situation is quite different. For one,if the particle is not free(V (x, t) 6= constant in x and t), then momentum is

111

Page 113: PHY 416, Quantum Mechanics - SIUE

not conserved (classically, dpdt = F = −∂V

∂x , so V = V (x)⇒ dpdt 6= 0.) However,

total energy still is conserved. This motivates us to see if we can identify orconstruct, from the maze of all solutions to the non-free Schrodinger equation,some general properties of solutions that represent definite energy.

As we will no show, the definite energy states obey the equation

HΨE(x, t) = EΨE(x, t) (15.2)

where E is the (“sharp” or definite) value of the energy associated with thestate function ΨE(x, t) that satisfies this equation. An equation in the generalform

QΨQ(x, t) = CΨQ(x, t) (15.3)

where Q is any operator and where C is any constant (real or complex) iscalled an “eigenvalue equation” in the mathematical parlance. The constantC is called the “eigenvalue”, and the states ΨQ(x, t) are called the “eigen-states” (of operator Q). So, solutions to equation (15.2) are the eigenstatesof the energy operator (Hamiltonian). In fact, we will show that a statefunction is sharp in energy if and only if it obeys equation (15.2).

To show that definite energy state function satisfy equation (15.2), we notethat, for such a state, σ2

E, the variance of the results of energy producingmeasurements, must be zero. Thus, if ΨE(x, t) is sharp in energy,

σ2E =

⟨(E − 〈E〉)2

⟩=

∫ +∞

−∞ΨE(x, t)

[H − 〈E〉

]2ΨE(x, t) dx (15.4)

Now, since H is a Hermitian operator and 〈E〉 is a constant,(H − 〈E〉

)is

also a Hermitian operator. Recall that if Q is a Hermitian operator, then

〈Q〉∗ = 〈Q〉 (15.5)

(reality of expectation value − required if the operator corresponds to aphysical observable), which implies also that∫ +∞

−∞Ψ(x, t)

[QΨ(x, t)

]∗dx =

∫ +∞

−∞Ψ∗(x, t)QΨ(x, t) dx (15.6)

for any valid state function Ψ(x, t) (not just for eigenstates of Q) and also∫ +∞

−∞Ψ∗1(x, t)QΨ2(x, t) dx =

∫ +∞

−∞

[QΨ1(x, t)

]∗Ψ2(x, t) dx (15.7)

112

Page 114: PHY 416, Quantum Mechanics - SIUE

for any pair of valid quantum mechanical state functions Ψ1(x, t) and Ψ2(x, t).Thus, for the condition in equation (15.4) is,

0 =

∫ +∞

−∞ΨE

[(H − 〈E〉

)(H − 〈E〉

)ΨE

]dx (15.8)

where ΨE is like Ψ1 is equation (15.7) and(H − 〈E〉

)ΨE is like Ψ2 is equation

(15.7).

0 =

∫ +∞

−∞

[(H − 〈E〉

)ΨE

]∗ [(H − 〈E〉

)ΨE

]dx (15.9)

0 =

∫ +∞

−∞

∣∣∣(H − 〈E〉)ΨE

∣∣∣2 dx (15.10)

Now, the integrand cannot be negative for any x, so the only way that thewhole integral can vanish is if the integrand is zero for all x, which impliesthat (

H − 〈E〉)

ΨE = 0 ⇒ HΨE = 〈E〉ΨE (15.11)

Since, for this state, 〈E〉 = E (sharp), we have established that if ΨE is astate of sharp E, then it satisfies the eigenvalue equation

HΨE(x, t) = EΨE(x, t) (15.12)

Now let us establish the converse, suppose that

HΨ(x, t) = CΨ(x, t) (15.13)

where C is any constant. Consider then, 〈E〉 in such a state

〈E〉 =

∫Ψ∗HΨ dx =

∫Ψ∗CΨ dx = C

∫Ψ∗Ψ dx = C (15.14)

thus,C = 〈E〉 (15.15)

Consider, also the variance of measurements of the energy in such a state:Recall from chapter one of Griffith,

σ2E = σ2

H=⟨H2⟩− 〈H〉2 (15.16)

113

Page 115: PHY 416, Quantum Mechanics - SIUE

Now,H2Ψ = HHΨ = HCΨ = CHΨ = CCΨ = C2Ψ (15.17)

Therefore ⟨E2⟩ =

∫Ψ∗H2Ψ dx = C2

∫Ψ∗Ψ dx = C2 (15.18)

soσ2E =

⟨E2⟩− 〈E〉2 = C2 − C2 = 0 (15.19)

Thus, in such a state, the energy is sharp (variance zero) and C = E.So

energy is sharp ⇐⇒ HΨE(x, t) = EΨE(x, t) (15.20)

We see that Ψ obeysHΨE(x, t) = EΨE(x, t) (15.21)

if and only if the energy is sharp. We note that the same sort of thing is truefor momentum: p = h

iddx is the momentum operator. Then Ψ(x, t) obeys the

eigenvalue equationpΨ(x, t) = pΨ(x, t) (15.22)

if and only if Ψ(x, t) is a state of sharp momentum.Proof :

In our previous proofs for the operator H, replace H by p and ev-erything works analogously − in fact, the proof works for any Her-mitian operator Q. We will have more to say about this generalproperty of Hermitian operators and their eigenstates later in thecourse. for now, except for the one example immediately following,we will worry about eigenstates of energy.

ExampleFor a free particle, the plane wave state

Ψ(x, t) = Aei(phx−

Eh t) (15.23)

is an eigenstate of both the Hamiltonian (H) and the momentum operator(p): To see this, note

H[Aei(

phx−

Eh t)]− ih ∂

∂t

[Aei(

phx−

Eh t)]

= ih · −ihEΨ (15.24)

114

Page 116: PHY 416, Quantum Mechanics - SIUE

orHΨ = EΨ (15.25)

and

p[Aei(

phx−

Eh t)]

=h

i

∂x

[Aei(

phx−

Eh t)]

=h

i· ihpΨ (15.26)

pΨ = pΨ (15.27)

It turns out that only for a free particle is an eigenstate of H also and eigen-state of p.1 We now return to our discussion of energy eigenstates.

15.2 Time Dependence of Energy Eigenstates

For eigenstates of energy

HΨE(x, t) = EΨE(x, t) (15.28)

which is

ih∂ΨE(x, t)

∂t= EΨE(x, t) (15.29)

which is∂ΨE(x, t)

∂t= − i

hEΨE(x, t) (15.30)

orΨE(x, t) = f(x)e−

iEth (15.31)

where f(x) is a function of x only. Notation: We call f(x), “ψ(x)”. Com-ment: We see that ψ(x) = Ψ(x, t = 0).

Terminology: ψ(x) is called the eigenfunction (or more correctly, theenergy eigenfunction) associated with Ψ(x, t). We will see the reason forthis terminology very soon. Thus,

ΨE(x, t) = ψE(x)e−iEth (15.32)

For Eigenstates only1This makes sense in light of the fact that energy is conserved for a non-free particle, but momentum is not

conserved. Thus, a state of definite energy for a non-free particle will have its momentum amplitude change in time.

115

Page 117: PHY 416, Quantum Mechanics - SIUE

15.3 Time Independent Schrodinger Equation

We now established an important property of the energy eigenfunctions,ψE(x): Of course, Ψ(x, t) must obey the Schrodinger equation

ih∂Ψ(x, t)

∂t= − h2

2m

∂2Ψ(x, t)

∂x2 + V (x, t)Ψ(x, t) (15.33)

thus, we can put equation (15.32) into the Schrodinger equation. We do thisfor the (most natural) case of V (x) not depending on time. Then

− h2

2m

∂2

∂x2

[ψ(x)e−

iEth

]+ V (x)ψ(x)e−

iEth = ih

(−iEh

)ψ(x)e−

iEth (15.34)

We can cancel the common exponential, then this is

− h2

2m

d2ψ(x)

dx2 + V (x)ψ(x) = Eψ(x) (15.35)

This equation must be obeyed by the spatial factor ψ(x) of an eigenstate of theHamiltonian. It is called the Time Independent Schrodinger Equation

Here is another way of phrasing the same argument: Suppose that ΨE(x, t)is an energy eigenstate and suppose that V depends on x but not on t. Thenagain

HΨE(x, t) = EΨE(x, t) (15.36)

Putting Ψ(x, t) = ψee− iEth into this equation and canceling the e−

iEth yields

Hψ(x) = Eψ(x) (15.37)

which is the time independent Schrodinger equation.

15.4 Energy Eigenstates are “Stationary States”

We now discuss an important general property of eigenstates, without ref-erence to any particular example. First note that that for a general stateΨ(x, t) that is not an eigenstate, the probability density at any point x gen-erally depends on time and thus fluctuates:

P (x) = Ψ∗(x, t)Ψ(x, t) (15.38)

116

Page 118: PHY 416, Quantum Mechanics - SIUE

depends on time, even though the integral∫ +∞

−∞Ψ∗(x, t)Ψ(x, t) dx = 1 (15.39)

does not depend on time.However, for the eigenstates, P (x) does not depend on time. (We say it is

“stationary”, hence eigenstates are also called “stationary states”.)Proof : If Ψ(x, t) is an eigenstate, Ψ(x, t) = ψ(x)e−

iEth Thus,

P (x) = Ψ∗(x, t)Ψ(x, t) (15.40)

=(ψ(x)e−

iEth

)∗ (ψ(x)e−

iEth

)(15.41)

= ψ∗(x)e+ iEth ψ(x)e−

iEth (15.42)

= ψ∗(x)ψ(x)e+ iEth e−

iEth (15.43)

= ψ∗(x)ψ(x) (15.44)

which clearly has not time dependence.Thus, in a eigenstate, there can be no “left-right” motion of probability.

15.4.1 Normalization of ψ(x)

We already know that ∫ +∞

−∞Ψ∗(x, t)Ψ(x, t) dx = 1

If Ψ(x, t) is an eigenstate of energy, the same logic we used above implies (asyou can easily show) that ∫ +∞

−∞ψ∗(x)ψ(x) dx = 1 (15.45)

is necessary; thus. the energy eigenfunctions obey the same normalizationcondition as does the state function.

15.5 A Possible Means to Further Solution? (“Separation of Vari-ables”)

We still only have certain solution (i.e. energy eigenstates) of the Schrodingerequation. We found that these solutions are factored in their space and time

117

Page 119: PHY 416, Quantum Mechanics - SIUE

dependence − ie.ΨE(x, t) = ψE(x)e−

iEth (15.46)

Perhaps, there are other factorizable solutions,

Ψ(x, t) = f(x)g(t) (15.47)

To find out, we put this equation into the Schrodinger equation an see whathappens. It makes sense to try this separation of variables method, as itis a common mathematical technique for generating solutions of differentialequation. Putting this equation into the Schrodinger equation yields

ihdg(t)

dt= − h

2mg(t)

d2f(x)

dx2 + V (x, t)f(x)g(t) (15.48)

which is (divide through by f(x)g(t)

ih1

g(t)

dg(t)

dt= − h2

2m

1

f(x)

d2f(x)

dx2 + V (x, t) (15.49)

To make progress with this technique, we must again assume that V dependsonly on x and not on t. In that case, a very interesting thing happens − theleft side of the equation depends only on t while the right side depends onlyon x. Thus, each side can only equal something that depends on neither xnor t − c constant − the same constant, so the equation can be split intotwo ordinary differential equations:

ih1

g(t)

dg(t)

dt= C (15.50)

and

− h2

2m

d2f(x)

dx2 + V (x)f(x) = Cf(x) (15.51)

Now equation (15.51) isHf(x) = Cf(x) (15.52)

Which we recognize as an energy eigenvalue equation; therefore, C = E andapparantly f(x) is nothing but our energy eigenfunction ψ(x) (or at leastproportional to it). furthermore, the solution to equation (15.50) is

g(t) = g(t = 0)e−iCth = g(t = 0)e−

iEth (15.53)

118

Page 120: PHY 416, Quantum Mechanics - SIUE

Now, g(t = 0) is just a constant, so we may as well absorb it into ψ(x)(which has to be normalized anyway). So, interestingly, we have found noth-ing new − all time/space factorizable (“seperable”) solutions to theSchrodinger equation are energy eigenstates!

15.6 Spectrum of Eigenstates

In general, for a given Hamiltonian H (i.e. a given potential energy functionV (x)), there are a number of eigenfunctions (solution to the time independentSchrodinger equation), say,

ψ1(x), ψ2(x), ψ3(x), · · ·

In general, each of these will be a solution of the time independent Schrodingerequation for a different value of the energy. Thus, we have a spectrum ofallowed energies

E1, E2, E3, · · ·(A concrete example of this is for the infinite square well. We will discussthis and other examples much further soon.

This means that the eigenstate solutions of the full (original) Time De-pendent Schrodinger Equation

ih∂Ψ(x, t)

∂t= − h2

2m

∂2Ψ(x, t)

∂x2 + V (x, t)Ψ(x, t)

are many − in general (depending on the specific form of V (x))

Ψ1(x, t) = ψ1(x)e−iE1th

Ψ2(x, t) = ψ1(x)e−iE2th

Ψ3(x, t) = ψ1(x)e−iE3th

...

Again, how many linearly independent eigenfunctions (and hence eigenstates)there are in a given situation onthe potential energy profile V (x).

119

Page 121: PHY 416, Quantum Mechanics - SIUE

15.7 General Solution to the Schrodinger Equation

Now look again at the Schrodinger equation

ih∂Ψ(x, t)

∂t= − h2

2m

∂2Ψ(x, t)

∂x2 + V (x, t)Ψ(x, t)

It is linear (in Ψ)− remember− we designed it that way. This means that thesum of two solutions is also a solution: If Ψ1(x, t) and Ψ2(x, t) are solutions,then Ψsum(x, t) = Ψ1(x, t) + Ψ2(x, t) is also a solution. This can be shown bydirect substitution.

In fact, since the equation is linear, any linear combination

Ψ(x, t) = C1Ψ1(x, t) + C2Ψ2(x, t) + C3Ψ3(x, t) + · · ·

is also a solution.Thus, for a given V (x) there are N eigenstates, a quite general solution of

the time dependent Schrodinger equation for this potential is

Ψ(x, t) =N∑n=1

cnΨn(x, t) =N∑n=1

cnψne− iEnth (15.54)

This is really “quite general” because for every time we change at least oneof the cb’s. we get a different solution. Generally, N = ∞, and in fact, weassert (without proof for now) that the most general solution to thetime dependent Schrodinger equation is

Ψ(x, t) =∞∑n=1

cnψne−−iEnth (15.55)

Thus, if you find all the eigenfunctions of the time independent Schrodingerequation, you have completely solved the time dependent Schrodinger equa-tion!

Unfortunately, we only know how to do that analytically for a relativelyvery few cases!

The reader who has familiarity with wave physics will note the stronganalogy to classical normal mode theory − the energy eigenstates are reallythe normal modes of the Schrodinger equation and the time independent

120

Page 122: PHY 416, Quantum Mechanics - SIUE

Schrodinger equation is nothing but its Helmholtz equation. one difference isthat the mode time dependence in quantum mechanics is e−iωt

(ω ≡ E

h

)and

not cos(ωt) as in classical wave physics. Another difference, of course, is theprobability interpretation of the quantum mechanical state functions.

15.7.1 Completeness

Let us look at the assertion of equation (15.55) in “another” way. If any validstate function Ψ(x, t)2 can be written in the “eigenstate expansion form”,(equation (15.55)), then at t = 0, any valid state function can be expressedas

Ψ(x, t = 0) =∞∑n=1

cnψn(x) (15.56)

Thus, in principle, one way of solving the initial value problem for a non-free particle is the following: Given Ψ(x, t = 0), solve the time independentSchrodinger equation for the relevant V (x) − i.e, find all the eigenfunctionsψi(x), then expand Ψ(x, t = 0) over these eigenfunctions as in equation

(15.56); then tack a factor e−iEit

h to each ψi(x) and the result is the expansionin equation (15.55). Both equation (15.55) and (15.56) then, are statementsof completeness. It may seem miraculous that one can always find the setcn for any f(x) = Ψ(x, t = 0), but this completeness property can beproven3 from the theory of “Storm-Liouville” equations.

15.8 A Concrete Example: The Infinite Square Well Revisited −Eigenfunctions

Fig 1Here V (x) =∞ unless 0 ≤ x ≤ L, in which case V (x) = constant (which

we will take to be zero). You already found the eigenstates for this problemby combining traveling wave solutions of the time dependent Schrodingerequation; here we look at this through the time independent Schrodingerequation formalism. Outside the well (and at its boundaries x = 0 andx = L) clearly ψ(x) = 0, otherwise there would be considerable probability

2For a particular potential energy profile V (x), of course.3We won’t do this, we’ll just use the result. More on this next class.

121

Page 123: PHY 416, Quantum Mechanics - SIUE

for the particle to materialize at “±∞”, which is not valid. inside the well,the time independent Schrodinger equation is

− h2

2m

d2ψ(x)

dx2 = Eψ(x) (15.57)

(since V = 0) ord2ψ(x)

dx2 = −2mE

h2 ψ(x) (15.58)

This is exactly the equation for harmonic oscillation in space, so

ψ(x) = C cos

(√2mE

h2 x+ φ

)(15.59)

where φ is a phase angle and where C is a constant. We rewrite this in themore convenient form

ψ(x) = A sin

(√2mE

h2

)+B cos

(√2mE

h2 x

)(15.60)

Applying the boundary condition ψ(x = 0) = 0⇒ B = 0 (why?) so

ψ(x) = A sin

(√2mE

h2 x

)(15.61)

Applying the boundary condition ψ(x = L) = 0, yields√2mE

h2 = nπ, n = 1, 2, 3, · · · (15.62)

(Why not n = 0?) Then

En =n2π2h2

2mL2 (15.63)

The eigenstate of the infinite square well have the same shapes as those ofthe string, but their amplitude is fixed.

Fig 2We see that the application of the boundary conditions has led to quanti-

zation of energy − only certain values of energy are allowed. (Of course, wereached the same conclusion in the past when we “solved” this problem using

122

Page 124: PHY 416, Quantum Mechanics - SIUE

the original Schrodinger equation). We note, in particular, that the energycannot be zero. Let us try to get some insight into this. The HeisenbergUncertainty Principle tells us tat, for any state function,

∆x∆px ≥h

2(15.64)

Consider now the lowest energy state or “ground state”. For it

∆x ≈ L (15.65)

Thus

∆px/geh

2L(15.66)

But, inside the well E = p2

2m then, if E = 0, then p is definitely zero. But, ifp is sharp at zero, then ∆p = 0, in violation of the Heisenberg UncertaintyPrinciple . Hence, E cannot be sharp at value zero. The minimum energy,

E1 =π2h2

2mL2 (15.67)

is called the “zero-point” energy. In general, the ground state for any boundsystem has a zero-point energy that is not zero. As you can show, the nor-malization condition requires that, unlike the classical case of standing waveson a stretched string, the amplitude is not arbitrary; rather

A

√2

L(15.68)

Thus

ψn(x) =

√2

Lsin(nπxL

)(15.69)

Hence

Ψn(x, t) =

√2

Lsin(nπxL

)e−

iEnt

h2 (15.70)

or

Ψn(x, t) =

√2

Lsin(nπxL

)e−i

n2π2h2mL2 t (15.71)

inside the well. Now, the time dependent (i.e. “original”) Schrodinger equa-tion is linear, hence a more general solution of it for the infinite square wellis, as we already mentioned some time ago,

123

Page 125: PHY 416, Quantum Mechanics - SIUE

Ψ(x, t) =∞∑n=1

cn

√2

Lsin(nπxL

)e−i

n2π2h2mL2 t (15.72)

if 0 ≤ x ≤ L and zero otherwise and cn is an arbitrary complex constant.

124

Page 126: PHY 416, Quantum Mechanics - SIUE

Chapter 16

Sixteenth Class: Examples andOrthogonality

Thursday, October 16, 20081

16.1 A Concrete Example of the Formalism: Infinite Square WellRevisited

Recall that definite energy state function can be written in factorized formas

ΨE(x, t) = ψ(x)e−iEth (16.1)

where ψ(x), the (energy) eigenfunction, obeys the time independent Schrodingerequation

− h2

2m

d2ψ(x)

dx2 + V (x)ψ(x) = Eψ(x) (16.2)

if V is a function only of x (and not of t). The simplest example to try theformalism out on is the infinite square well − recall that, for it V (x) = ∞unless 0 ≤ x ≤ L, in which case V (x) = a constant (which for convenience wetake to be zero).Working from the time dependent Schrodinger equation, westarted with the assumption that the general definite energy state functioninside the well is

Ψ(x, t) = Aei(phx−

Eh t) +Be−i(

phx−

Eh t) (16.3)

1This includes the “make-up” of the missed 14th class

125

Page 127: PHY 416, Quantum Mechanics - SIUE

In the eigenfunction approach, we start with the time independent Schrodingerequation. Inside the well, the time independent Schrodinger equation is

− h2

2m

d2ψ(x)

dx2 = Eψ(x) (16.4)

(since V=0) ord2ψ(x)

dx2 = −2mE

h2 ψ(x) (16.5)

This is exactly the equation for the harmonic oscillation in space, so

ψ(x) = C cos

(√2mE

h2 x+ φ

)(16.6)

where φ is a phase angle and where C is a constant. We rewrite this in themore convenient form

ψ(x) = A sin

(√2mE

h2 x

)+ B cos

(√2mE

h2 x

)(16.7)

Note that, a priorı2, ψ(x) could be complex, and a priorı, A, B, and C canall be complex. Note also that an alternative, and (as you can easily show)completely equivalent way to write the general solution to this free timeindependent Schrodinger equation is

ψ(x) = Dei(√

2mEh x

)+ Fe

−i(√

2mEh x

)(16.8)

for convenience, we will work with equation (16.7). Applying the boundarycondition, ψ(x = 0) = 0, B = 0 (why?), so

ψ(x) = A sin

(√2mE

h2 x

)(16.9)

Applying the boundary condition ψ(x = L) = 0, yields√2mE

h2 L = nπ, n = 1, 2, 3, · · · (16.10)

2without proof for now

126

Page 128: PHY 416, Quantum Mechanics - SIUE

(why not n = 0?) Solving for E, we obtain

En =n2π2h2

2mL2 (16.11)

Thus, the energy eigenstates are

ψn(x, t) = An sin

(√2mE

hx

)e−

in2π2h2mL2 t (16.12)

where the normalization condition determine An to be√

2L for all n. Figure 1

remind us of the shapes of the eigenfunctions and the associated probabilitydensity profiles; since the states are stationary, the probability density doesn’tchange in time. So,

Ψn(x, t) =

√2

Lsin(nπLx)

e−in2π2h2mL2 t (16.13)

fig 1The analogous classical system is that of the stretched string which is

bound down at both its ends (x = 0, L). The energy eigenstates here (nor-mal modes of the Schrodinger equation) are almost exactly mathematicallyanalogous to the normal modes of the classical string − the “only” mathemat-

ical differences are the mode amplitudes (fixed at√

2L in quantum mechanics

by probability normalization, arbitrary in the classical stretched string nor-mal modes) and the time dependences (cos(ωnt + φn)) in the classical case,which means that Ψn = 0 for all x at certain times in the classical case (atthose times the energy is all kinetic) and eiωnt i nthe quantum case, whichprevents this from ever happening since it is never zero (as it must − oth-erwise conservation of total probability would be violated). That the modeamplitudes An are arbitrary in the classical case is a reflection that a givenmode can possess arbitrary energy; in the quantum case, the amplitude isfixed and the energy in the mode is independent of it and is quantized sinceit is proportional to the quantized frequency. In the quantum case, the fixedamplitude of each node fixes the associated probability density profile.

127

Page 129: PHY 416, Quantum Mechanics - SIUE

16.2 The “Reason” for Energy Quantization

We see that the application of the boundary conditions has led to quanti-zation of energy − only certain values of energy are allowed. (Of course,we reached the same conclusion in the past when we “solved” this problemusing the original Schrodinger equation). It is easy to see the essence of themathematical reason for the energy quantization − the Schrodinger time in-dependent equation, being a second order ordinary differential equation, hadtwo arbitrary constants in its general solution: “A” and “B” in

ψ(x) = A sin

(√2mE

hx

)+ B cos

(√2mE

hx

)(16.14)

But, there are three conditions to impose:

1. Boundary Condition I

ψ(x = 0) = 0 ⇒ B = 0

2. Overall normalization

A =

√2

L

3. Boundary Condition IIψ(x = L) = 0

To satisfy condition 3), E must be considered as “adjustable” to the condition− and hence, no longer arbitrary − thus, only certain values of E work. Itshould be plausible to you (we’ll investigate the reasons more deeply later)that this same sort of thing occurs for an arbitrary potential − of course, theSchrodinger equation is different for each V (x), but generally there are twolinearly independent solutions (hence two adjustable constants) and threeconditions,

1. ψ(x)→ 0 as x→ +∞ (or some other point)

2. ψ(x)→ 0 as x→ −∞ (or some other point)

3.∫ +∞−∞ ψ∗(x)ψ(x) dx = 1

128

Page 130: PHY 416, Quantum Mechanics - SIUE

Thus, it should be plausible to you at this stage that, in general, for a “boundstate” scenario (V (x) < 0 in a region where, by convention, V (x) → 0 asx → ±∞) there are solutions to the time independent Schrodinger equa-tion(i.e. eigenstates exist) only for certain values of the energy − the figure2 schematically shows an example of this

fig 2In order to try to get some preliminary insight into this, let’s discuss the

general situation just briefly now. Consider a general potential energy profilethat would lead to bound states. (The phrase “bound state” means thatthe particle is permanently essentially3 confined to a localized region of finiteextent.)

fig 3In figure 3, V (x) is plotted and a possible energy value (E) on the same

plot. First, let us review what would happen classically to a particle withthe shown energy E moving in this potential energy profile. Whenever thepotential energy changes with position there is a force given by the negativeof its slope

F (x) = −dV (x)

dx(16.15)

Thus, whenever the potential energy on the curve is becoming shallower, theparticle is pushed back to the region of lower V (x). Thus, the acceleration isgreatest where the slope of V (x) is greatest in magnitude. From the energypoint of view, since KE = E − V (x), the greater the difference E − V , thegreater the kinetic energy. Thus, the points xa and xb, where the particlehas zero kinetic energy, are classical turning points − approaching them,the particle slows, stops and turns around. Thus, classically, the particleoscillates between xa and xb and remains absolutely confined between thesepositions. Thus, in regions I and III in figure 3 are classically forbidden (inthem KE < 0, which classically is nonsense) while region II is classicallyallowed. In quantum mechanics, the situation is somewhat different. Thetime independent Schrodinger equation

− h2

2m

d2ψ(x)

dx2 + V (x)ψ(x) = Eψ(x) (16.16)

3the reason for the qualifying “essentially” will become apparent shortly and clear later

129

Page 131: PHY 416, Quantum Mechanics - SIUE

must be obeyed in all regions, and hence, acceptable state functions subjectto certain conditions4.

1. ψ(x) must everywhere be single-valued

2. ψ(x) must everywhere be continuous

3. dψ(x)dx must be single-valued everywhere

4. dψ(x)dx must be continuous everywhere

The reason for condition (1) is that there must be a well defined probabilitydensity ψ∗(x)ψ(x) at all x. The reason for condition (4) can be easily seenby rewriting the time independent Schrodinger equationas

d2ψ(x)

dx2 =2m

h2 [V (x)− E]ψ(x) (16.17)

Thus, for any finite and continuous V (x), ψ′′(x) must be defined and finite aslong as ψ(x) is. For ψ′′(x) to be defined and finite at all places, ψ′(x) must besingle-valued, finite, and continuous; this thus establishes conditions (3) and(4). For ψ′(x) to be finite and defined everywhere, ψ(x) must be continuous,thus establishing condition (2). Generally, nonzero solutions to the timeindependent Schrodinger equationexist in all three regions, and these mustjoin smoothly to each other to make a solution valid for all x that accordswith conditions (1) through (4).

Example: Consider the finite square well. We consider the possibility ofa bound state with E < V0

fig 4In region II, the time independent Schrodinger equationis

d2ψ(x)

dx2 = −√

2mE

hψ(x) (16.18)

and the solution is sinusoidal

ψII)(x) = A sin

(√2mE

hx

)+ B cos

(√2mE

hx

)(16.19)

4in addition to the normalization conditions mentioned earlier

130

Page 132: PHY 416, Quantum Mechanics - SIUE

while in regions I and III, the time independent Schrodinger equationis

d2ψ(x)

dx2 =

√2m (V0 − E)

hψ(x) (16.20)

Since V0 > E, the coefficient of ψ(x) on the right hand side is a positivenumber, and thus the solution to this equation are exponential:

ψI,III(x) = De+κx + Fe−κx (16.21)

where κ =

√2m(v0−E)

h > 0. In region I, F must be zero; in region III, D mustbe zero; otherwise ψ(x) blows up as x→ ±∞. Thus

ψI(x) = De+κx (16.22)

ψII(x) = A sin

(√2mE

hx

)+ B cos

(√2mE

hx

)(16.23)

ψIII(x) = Fe−κx (16.24)

We will discuss how to match pieces up to make one continuous ψ(x) extend-ing from x = −∞ to x = +∞, but for now, our main points are ψ(x) is notzero in the classically forbidden regions I and III and that the matching-upcan be done. Thus, for example, the ground state eigenfunction for the finitesquare well looks like

fig 5We will discuss the physical implications of the penetration of ψ(x) into

the nonclassical region later. Now return to consideration of the “general”potential V (x). From elementary calculus, d2ψ(x)

dx2 represents the curvature of

ψ(x). Where d2ψ(x)dx2 > 0, ψ(x) is concave up, where d2ψ(x)

dx2 < 0, ψ(x) in concavedown. Thus, according to the time independent Schrodinger equation

d2ψ(x)

dx2 =2m

h[V (x)− E]ψ(x) (16.25)

in a classically allowed region (E > V ), ψ′′(x) has the opposite sign as doesψ(x) and hence always curves ψ(x) toward the ψ(x) = 0 axis, when in aclassically forbidden region (E < V ) ψ′′(x) has the same size as ψ(x) andhence always curves away from the axis; these possibilities are shown in figure6.

131

Page 133: PHY 416, Quantum Mechanics - SIUE

fig 6thus, for example, for the finite square well, the concavities of the ground-

state eigenfunction segments are as marked in figure 7fig 7How do we know that this represents the ground state? In the classically

allowed region the curvature is

d2ψ(x)

dx2 = −Eψ(x) (16.26)

and its magnitude is least when E is least (taking V = 0 to be the bottomof the well). Other ways to do the matching of segments are shown in figure8. Clearly the curve marked ψ1(x) has the lest magnitude of curvature andis hence the ground state. (The other curves are ψ2(x) with E2 > E1 andψ3(x) with E3 > E2(> E1).)

fig 8Also, it should be clear from the above considerations that, for the V (x)

profile shown in figure 9. We know this must be at lest qualitatively correctin depicting the shape of the ground state

fig 9Armed with these insights, we can now understand the reason for the

quantization of energy for cases other than the infinite square well. Con-sider again then, the symmetric potential energy profile shown below and itsground state eigenfunction.

fig 10Suppose we attempt to solve the time independent Schrodinger equationfor

an energy E just slightly lower then Eground. Then, according to our discussionabove, the curvature of the solution in the classically allowed region−x1 → x1

is shallower than it was for Etextrmground. Consequently, the new solution(labelled “ψ(x)” for Et < Egroundin the figure, and which we call ψt(x) (tfor “trial”)), which we assume has the same amplitude at x = 0 as didthe ground state eigenfunction ψ1(x), arrives at the turning point xm, with ahigher amplitude than does ψ1(x). Now in region III, ψ1 must also be concaveup, but, since according the time independent Schrodinger equation

d2ψ(x)

dx2 =2m

h[V (x)− E]ψ(x) (16.27)

132

Page 134: PHY 416, Quantum Mechanics - SIUE

the curvature magnitude, d2ψ(x)dx2 ∝ ψ(x) and ∝ [V (x) − E], the region III

curvature magnitude is greater than that of ψ1(x) on two counts. Thus,as shown in figure 10, the new trial solution diverges as x → +∞, and,by symmetry, also as x → −∞. This is unacceptable behavior. Now, sinced2ψ(x)dx2 ∝ ψ(x), we could curve the divergent behavior by simply rescaling ψ(x)− i.e., reducing the amplitude everywhere by the same factor f . It shouldbe apparent that there does exist an f such that, for the new E < Eground,ψrescaled = fψt does go smoothly to zero as x → ±∞. This is perfectlyacceptable behavior, bur now there is a new problem − ψrescaled does not obeythe probability normalization condition! By this logic it is apparent that noenergy near Eground other than Etextrmground itself is possible. Question tothing about: What would have gone wrong if we had tried raising the energya bit instead of lowering it? If you look back now at our original statement(some time ago)

generally, there are two linearly independent solutions (hence twoadjustable constants) and three condition

1. ψ(x)→ 0 as x→ +∞ (or some other point)

2. ψ(x)→ 0 as x→ −∞ (or some other point)

3.∫ +∞−∞ ψ∗(x)ψ(x) dx = 1

Thus, it should be plausible to you· · ·

as an intuitive explanation for energy quantization, you will see that the moredetailed “numerical integration” argument given above, while increasing ofour insight, is essentially the same argument. The situation “from here on”is well described by another author5:

quoteExample: The potential energy profile shown below below has the spec-

trum of energies shown on the same plot − discretely spaced bound stateenergies and a continuum of possible energies for E > 0.

fig 115Robert Scherrer in Quantum Mechanics: An Accessible Introduction, pp 57-58

133

Page 135: PHY 416, Quantum Mechanics - SIUE

16.3 Zero-Point Energy

Now back to the infinite square well: We note, in particular, that the energycannot be zero. Let us try to get some insight into this. the HeisenbergUncertainty Principle tells us that, for any state function

∆x∆p ≥ h

2(16.28)

Consider now the lowest energy of “ground” state. For it,

∆x ≤ L√12

=L

2√

36 (16.29)

thus

∆p ≥√

3h

L(16.30)

But, inside the well E = p2

2m if E = 0, then p is definitely zero. but if p issharp at zero, then ∆p = 0, in violation of our limitation on ∆p in equation(16.16). Hence E cannot be sharp at value zero. The minimum energy

E1 =π2h2

2mL2 (16.31)

is called the “zero-point” energy. In general, the ground state for any boundsystem has a zero-point energy that is not zero. We can, in fact, use theHeisenberg Uncertainty Principle to estimate the ground state energy for aparticle in a box. We have, for the ground state

E = EK = EK =p2

2m(16.32)

where E ≡ 〈E〉. Now, from its definition,

∆p =

√p2 − p2 (16.33)

p is zero,since it would be equally likely to be positive as negative (homo-geneity of space inside the well − V = 0 everywhere inside). Thus

∆p =

√p2 (16.34)

6∆xground = L

2√

3if ψ∗ψground is flat inside the well, since it is peaked, we expect ∆xground <

L

2√

3; in actuality it

works out to be ∆x ≈ 0.18L. Here we are only “estimating.”

134

Page 136: PHY 416, Quantum Mechanics - SIUE

orp2

2m= E =

(∆p)2

2m(16.35)

using ∆p ≥ h√

3L ,

Eground ≥3h2

2mL2 (16.36)

which compare favorably with the actual value

Eground =π2h2

2mL2 (16.37)

If one want to turn the logic around and use the ground state as an illustrationof the Heisenberg Uncertainty Principle , one finds, after calculation

∆x =

√x2 − x2 ≈ 0.181L

∆p =

√p2 − p2 =

πh

L

so

∆x∆p = (0.181L)

(πh

L

)≈ 0.569h ≥ h

2. (16.38)

As we have already mentioned, the (time dependent) Schrodinger equation islinear and homogeneous, thus, an arbitrary linear combination of eigenstatesis also a solution. In fact, as we already mentioned on more than one occa-sion, the general solution to the time dependent Schrodinger equation is anarbitrary linear combination of the eigenstates; hence the general solution forthe infinite square well potential is

Ψ(x, t) =∞∑n=1

cn

√2

Lsin(nπLx)

e−in2π62h2mL2 t (16.39)

for 0 ≤ x ≤ L and 0 otherwise and the cn’s are determined (at least in ration)by the overall normalization requirement.

135

Page 137: PHY 416, Quantum Mechanics - SIUE

16.4 Orthogonality and Completeness of Eigenfunctions

From our general solution to the infinite square well potential. we can seethat

Ψ(x, t = 0) =∞∑n=1

bn sin(nπLx)

(16.40)

From the theory of Fourier series, we know that the Fourier series on theright-hand side can, with appropriate bn’s represent any square-normalizablefunction that vanishes at x = 0 and x = L. Hence, as in the free parti-cle case, Ψ(x, t = 0) can be any square-normaliable function obeying theboundary conditions. As well we will see, a similar statement is true for anyreasonable V (x). We can get a first idea of why the last sentence is truefor the orthogonality principle of eigenfunctions corresponding to differenteigenfunctions that we now demonstrate for “general” V (x).

Remember − the completeness of ı, , k in ordinary 3-space partlydepends on their orthonormality − using e1 ≡ ı, e2 ≡ , and e3 ≡ k,

em · en = δm,n ≡

0 m 6= n

1 m = n(16.41)

As we will see in the Fourier case, a key to the completeness is the orthogo-nality property:

Theorem: For any continuous V (x), eigenfunctions of H belongingto different eigenvalues are orthogonal − i.e.∫ +∞

−∞ψ∗m(x)ψn(x) dx = 0 if m 6= n. (16.42)

Proof : ψn(x) and ψm(x) are respectively solutions to the time independentSchrodinger equation.

− h2

2m

d2ψn(x)

dx2 + V (x)ψn(x) = Enψn(x)

− h2

2m

d2ψm(x)

dx2 + V (x)ψm(x) = Emψm(x)

136

Page 138: PHY 416, Quantum Mechanics - SIUE

Taking the complex conjugate of the lower equation

− h2

2m

d2ψ∗m(x)

dx2 + V (x)ψ∗m(x) = Emψ∗m(x)

We then multiply the the top equation by ψm(x) and the above equation byψn, then subtract them from each other and integrate

2m

h(Em − En)

∫ +∞

−∞ψ∗m(x)ψn(x) dx =

∫ +∞

−∞

[ψ∗m

d2ψn(x)

dx2 − ψnd2ψ∗m(x)

dx2

]dx

This is

2m

h(Em − En)ψ

∗m(x)ψn(x) dx =

∫ +∞

−∞

d

dx

(ψ∗m(x)

dψn(x)

dx− ψn(x)

dψm(x)

dx

)dx

The right hand side is[ψ∗m(x)

dψn(x)

dx− ψn(x)

dψ∗m(x)

dx

]+∞

−∞= 0 (16.43)

thus since Em 6= En,∫ +∞

−∞ψ∗m(x)ψn(x) dx = 0 when m 6= n (16.44)

If the ψn’s are normalized, then they form an orthonormal set∫ +∞

−∞ψ∗m(x)ψn(x) dx = δm,n (16.45)

where δm,n = 0 for m 6= n and δm,n = 1 for m = n. Based on this, we nowstate a postulate: Postulate: For any continuous potential energy functionV (x), the set of eigenfunctions is a complete set − i.e. any function f(x)obeying the boundary conditions associated with V (x) can be expanded as

f(x) =N∑n=1

cnψn(x) (16.46)

This is known as the completeness postulate.Example: As we already remarked, the infinite square well eigenfunctions

are complete.

137

Page 139: PHY 416, Quantum Mechanics - SIUE

Example: Consider the case V (x) = 0 everywhere (free particle). Thenthe eigenfunctions of H are the set

eikx

.Proof :

Heikx =p2

2meikx =

1

2m

(h

i

)2d2

dx2

(eikx)

= −k2h2

2meikx (16.47)

orHeikx = Eeikx. (16.48)

The completeness of these eigenfunctions is just the statement

f(x) =1√2πh

∫ +∞

−∞φ(k)eikx dk (16.49)

which we know already to be true.Comments

1. The completeness postulate stated here (for any V ) can actually beproved (“Sturm-Liouville Theory” in differential equation).

2. The completeness postulate stated here is actually a special case of amore general completeness postulate concerning eigenfunctions of anyHermitian operator.

16.5 The Quantum Initial Value Problem − Particle Not Free

The above considerations relate to the quantum initial value problem: de-termining Ψ(x, t) from Ψ(x, t = 0) for general V (x). Now, as you recall,we already discussed and solved7 this problem for the case of a free parti-cle. There, the key was to expand Ψ(x, t = 0) over a complete set of basisfunctions (i.e. the set

eikx

) and then evolve each basis function forwardin time − this works if the basis functions do evolve independently in timeand if we know how they evolve. For the case of “general” V (x), it is nowclear that we can attempt to proceed analogously: we begin, with licensefrom the completeness postulate, by expanding the given Ψ(x, t = 0) over

7at least in principle

138

Page 140: PHY 416, Quantum Mechanics - SIUE

the complete set of eigenfunctions of H(V ).

Ψ(x, t = 0) =∞∑n=1

cnψn(x) (16.50)

But: How can we determine the coefficients cn?Here we use the orthonormality condition. We first multiply both sides of

this equation by ψm(x) and integrate:∫ +∞

−∞ψ∗m(x)Ψ(x, t = 0) dx =

∞∑n=1

cn

∫ +∞

−∞ψ∗mψn(x) dx (16.51)

=∞∑n=1

nnδm,n = cm (16.52)

i.e.

cm =

∫ +∞

−∞ψ∗m(x)Ψ(x, t = 0) dx (16.53)

Note that or expansion for Ψ(x, t = 0) and its derivation are exactly analogousto how we find Fourier coefficients − that is nothing but a special case of this!Thus, in general (i.e. for any continuous V (x)), we have

Ψ(x, t) =∞∑n=1

[∫ +∞

−∞ψ∗n(x

′)Ψ(x′, 0) dx′]ψn(x)e−

iEnth (16.54)

This is the formal solution to our problem. Note how it explicitly show thatknowledge of Ψ(x, t = 0) at all x determines Ψ(x, t).

Example (Griffiths example 2.2, infinite square well):Say Ψ(x, t = 0) = Ax(a − x) for 0 ≤ x ≤ a (where a is the extent of the

well)Fig 2What is Ψ(x, t) for subsequent time t?Normalizing, you find

A =

√30

a5 (16.55)

139

Page 141: PHY 416, Quantum Mechanics - SIUE

From Fourier,

cn =

∫ a

0Ψ(x, t = 0)

√2

asin(nπax)dx (16.56)

=

√2

a

√30

a5

∫ a

0x(a− x) sin

(nπax)dx (16.57)

which works out to

cn =4√

15

(nπ)3 [cos(0)− cos(nπ)] (16.58)

which is

cn =

0 n is even8√

15n3π3 n is odd

(16.59)

So

Ψ(x, t) =

√30

a

(2

π

)3 ∑n=1,3,5,···

1

n3 sin(nπax)

e−n2π2h2ma2 t (16.60)

16.6 Energy Measurement

Let us turn now to another important issue. Suppose that, for some potentialenergy function V (x), the system is in some state Ψ(x, t). Suppose the energyis measured. What values are possible and what are their probabilities?

Expectation ValueWe start by calculating 〈E〉 for the general superposition state

Ψ(x, t) =∞∑n=1

cnψn(x)e−iEnth (16.61)

We have

〈E〉 =⟨H⟩

=

∫ +∞

−∞Ψ∗(x, t)HΨ(x, t) dx

=

∫ +∞

−∞

[∑m

c∗mψ∗m(x)e+ iEmt

h

]H

[∑n

cnψn(x)e−iEnth

]dx

140

Page 142: PHY 416, Quantum Mechanics - SIUE

which is

〈E〉 =

∫ +∞

−∞

∑m

∑n

c∗mcnψ∗me

iEmth e−

iEnth Hψn dx (16.62)

Now Hψn = Enψn, so

〈E〉 =∑m

∑n

c∗mcnEnei(Em−En)t

h

∫ +∞

−∞ψ∗m(x)ψn(x) dx (16.63)

=∑m

∑n

c∗mcnEnei(Em−En)t

h δm,n (16.64)

〈E〉 =∑n

|cn|2En (16.65)

This suggests that, if you measure the energy, |cn|2 is at least proportional tothe probability of materializing result En

8. In fact, we can find the probabilityconstant by finding the sum

∞∑n=1

|cn|2 (16.66)

We know that ∫ +∞

−∞Ψ∗Ψ(x, t) dx = 1 (16.67)

which is

1 =

∫ +∞

−∞

∞∑m=1

∞∑n=1

c∗mψ∗m(x)e

iEmth cnψn(x)e−

iEnth (16.68)

=∞∑m=1

∞∑n=1

c∗mcnei(Em−En)t

h

∫ +∞

−∞ψ∗m(x)ψn(x) dx (16.69)

=∞∑m=1

∞∑n=1

c∗mcnei(Em−En)t

h δm,n (16.70)

1 =∞∑n=1

|cn|2 (16.71)

Thus, the probabilities are normalized; hence: If a particle is representedby the state function Ψ(x, t) then, if a measurement of energy of

8Because we should have , in analogy to finding the average of any discrete distribution, 〈E〉 =P

n|cn|2EnP|cn|2

.

141

Page 143: PHY 416, Quantum Mechanics - SIUE

energy is made, one and only one of the eigenvalues En (for thegiven potential) materializes.

The probability of obtaining the result En is |cn|2 = c∗ncn, where

Ψ(x, t) =∑n

cnψn(x)e−iEnth (16.72)

This is taken as a postulate; actually it is a special case of a more generalpostulate concerning probabilities of eigenvalues of Hermitian operator. Onthe other hand, a similar argument can be made for the eigenvalues of anyHermitian operator (time permitting, we’ll do this later in the course). Forthe case of momentum, for example, it strongly suggests the conclusion thatwe already assumed, namely, that the probability of obtaining momentum inthe range dp centered on p at time t is

φ(p, t)dp = φ∗(p, t)phi(p, t) (16.73)

where

φ(p, t) =1√2πh

∫ +∞

−∞Ψ(x, t)e−

ipxh dx (16.74)

142

Page 144: PHY 416, Quantum Mechanics - SIUE

Chapter 17

Seventeenth Class: Connections WithClassical Physics and Optics

Tuesday, October 21, 2008

17.1 The Connection with Everyday Physics

This is a good question that one should continually ask − what is the con-nection between the quantum mechanics we’ve learned and the everyday realworld of large objects? After all, classical physics must result as some limitof quantum mechanics; otherwise quantum mechanics is wrong! This is anissue we will return to in stages. Let us begin to think about this in thecontext of our infinite square well example. The stationary states and theirresulting stationary probability densities are shown in figure 1

fig 1We know ask: What is the connection of this with the behavior we expect?

Classically, we expect the particle to bounce back and forth between the walls,moving at constant speed (since V = 0 inside the well).

fig 2But, the behavior we’ve found doesn’t look anything like this at all! In our

normal modes, nothing moves right or left − the probability to materializein any small region is constant in time (“stationary states”). We can makesome progress on this problem as follows: We have

Ψn(x, t) =

√2

asin(knx)e−iωnt (17.1)

143

Page 145: PHY 416, Quantum Mechanics - SIUE

where ωn = Enh . Now,

sin θ =eiθ − e−iθ

2i(17.2)

So,

Ψn(x, t) =1

2i

√2

a

[ei(knx−ωnt) − e−i(knx−ωnt)

](17.3)

This is a combination to oppositely directed traveling waves. (In fact, startingwith this form was our first approach.) While this is somewhat comforting,it is still true that the probability density for this state is stationary. Whatwe would really like to see is a moving “lump” of probability.

fig 3Let us look at this issue from another point of view, however, our state

functions give us information on the probability of materializing the particlein any small region dx in the box [ψ∗(x)ψ(x)]. Classically, this probability isproportional to the time spent (by the bouncing back-and-forth) particle indx, call this dt. Then

dt =dx

vparticle= constant (17.4)

call it C ′. SoPclassical(x) ∝ C ′ (17.5)

call thisPclassical = C (17.6)

Now ∫ a

0Pclassical(x) dx = 1 ⇒ C

∫ a

0dx = 1 ⇒ C =

1

a(17.7)

so

Pclassical =1

a(17.8)

Now, quantum and classical physics are supposed to agree in the large nlimit. This is known as the “correspondence limit”. Let us think on this:

Example: An electron confined to a thin metal wire segmentfig4mimics a quantum “particle in a box”. If the segment is of macroscopic

length (say 1cm), and the system is at equilibrium at room temperature, we

144

Page 146: PHY 416, Quantum Mechanics - SIUE

expect to have a “classical limit” (if one is possible for trapped electrons).Here is an interesting example on this:

fig5Conclusion: In any classical-like situation, n is huge! Figure 6 shows the

plot of ψ∗(x)ψ(x) for n = 10fig 6For n larger, the peaks get closer together, and if n is large enough, we

can’t resolve them. So, as n→∞, we see an average “flat” ψ∗(x)ψ(x):

〈P (x)〉ave = (ψ∗(x)ψ(x)) =

⟨2

asin2 (knx)

⟩=

2

a

⟨sin2 (knx)

⟩=

2

a· 1

2

=1

a

which agrees with the classical prediction. Still, this is not yet really satisfying− we would still like to see a “moving lump” of probability.

Again. “so far, so good”, but we would like to see a “moving lump” ofprobability. Can we somehow obtain this? We know that we are allowed tohave superposition (of normal mode) states. We now look into this. We knowwe really should do this in the classical limit (n large), but to keep thingssimple, we consider only an equal amplitude superposition of the two loweststates:

Ψ(x, t) = A sin(πLx)

e−iω1t + A sin

(2π

Lx

)e−iω2t (17.9)

This superposition is shown for t = 0 along with its P (x) in figure 6.fig 7Now, E2 = 4E1 (En ∝ n2), so ω2 = 4ω1. Also

T1 =2π

ω1=

2πE1

h

=2πh

E1=

h

E1

Thus at t = 12T1 = h

2E1, the relative signs of the two modes reverse, as shown

in figure 8.fig 8

145

Page 147: PHY 416, Quantum Mechanics - SIUE

We thus see that, even for this low-n superposition, we do have a “lumpof probability density” sloshing back and forth.

fig 9However, we expect to be able to construct something that moves back

and forth at the classical particle velocity. For this, we need to put a narrowwave packet state in the well, and it has to be a narrow wave packet that iscarefully chosen: How can we do this? The method (in outline) should be asfollows:

1. We choose an appropriate Ψ(x, t = 0)

2. We calculate Ψ(x, t) from Ψ(x, t = 0) via our “propagation formula”

Ψ(x, t) =∑n

[∫ +∞

−∞ψ∗n(x

′)Ψ(x′, 0) dx′]ψn(x)e−

iEnth (17.10)

Here

Ψn(x) =

2a sin

(nπa x)

0 ≤ x ≤ a

0 otherwise(17.11)

3. We plot |Ψ(x, t)|2 at various times t.

While this is easy to say, there are a number of subtleties involved in actuallycarrying this out: What do we choose for Ψ(x, t = 0)? We have two mainrequirements for it:

1. We want Ψ(x, t = 0) to be a peaked function that is narrow comparedto the well width (a).

fig 10

2. We want Ψ(x, t = 0) to have only right-moving (i.e. positive) momentumcomponents

Of course, Ψ(x, t = 0) must obey the boundary conditions (vanish at x = 0and x = L). Therefore, it cannot be just a simple Gaussian function, sincethat has infinite tails. Rather, it is usual to start with

Ψ(x, t = 0) =∑n

cnψn(x) (17.12)

146

Page 148: PHY 416, Quantum Mechanics - SIUE

and choose the cn’s first so as to satisfy our second condition. This in-volves a fairly complicated procedure that we do not describe here. The re-sults of such a procedure for a moderately tight “packet” initial state areshown on the web at site http://www.optics.rochester.edu/~stroud/

animations/, click on Decay of High Momentum Wave Packet undersection “infinite square well wave packets”. It is fascinating to contemplatethat such classical limit behavior results from a superposition of eigenstates(correctly chosen, of course) none of which individually involves any left orrightward motion.

It is further profound that what causes “bouncing off the wall” in theclassical limit is not anything “put in by hand”, but rather just the timeevolution of superpositions of eigenstates, each involving its own phase fac-tor e−

iEnh t, and obeying the boundary conditions. From this, much of our

“classical experience” results!

17.2 State Functions for Incidence on Abrupt but Finite Changesin V

As you know, any really successful theory of quantum mechanics must dealwith the effects of forces on state functions. So far, we have only investigatedthis in one case − that where the change in potential energy is infinitelyabrupt and also infinite in magnitude − infintie square well. Of course, bothof these conditions represent idealizations. Consider, for example, a neutronapproaching an atomic nucleus in the atmosphere of the sun. The potentialenergy profile it experiences may look like

fig 11where the “change” occurs over a distance of order 1fm = 10−15m − very

abrupt on a macroscopic scale, not so abrupt compared to the size of theneutron, but still abrupt compared to the neutron’s DeBroglie wavelength ifits energy is low enough. To get a first handle on these sort of importantpractical (for physics), we consider possible state functions for the case ofabrupt, but finite, changes in V . The simplest case is that of a sudden step-like change in the potential energy profile in x:

fig 12

147

Page 149: PHY 416, Quantum Mechanics - SIUE

V (x) =

0 x < 0 (“region I”)

V0 x ≥ 0 (“region II”)(17.13)

Now, we would like to deal with the realistic situation of a narrow packetstate incident on the step, say from negative-x. (You will deal with incidencefrom the right (“potential-cliff” problem) in the homework). Since the packetstate, at any given time, is a superposition of energy eigenstates (recall thecompleteness postulates), we first consider only the energy eigenstates. Wethen have two cases, E < V0 and E > V0. We treat the first case first: E < V0

In region I, the time independent Schrodinger equationis

− h2

2m

d2ψI(x)

dx2 = EψI(x) (17.14)

since V = 0 in region I. This is

d2ψI(x)

dx2 = −2mE

h2 ψI(x) (17.15)

which has the general solution

ψI(x) = Aeik1x + Be−ik1x (17.16)

where k1 =√

2mEh (any E is allowed − we see here explicitly also, this is note

a bound state). Note: this is the same as

ψI(x) = C cos (k1x+ φ) (17.17)

Thus, the traveling wave form of Ψ(x, t) in region I is

ΨI(x, t) = Aei(k1x−ωt) + Be−i(k1x+ωt) (17.18)

where ω = Eh . We see what “physical situation” this corresponds to: we

would like to say that this is the situation of a plane wave state “incident’from negative-x” and undergoing a degree of “reflection” at x = 01. In regionII, the time independent Schrodinger equationis

− h2

2m=d2ψ(x)

dx2 + V0ψ(x) = Eψ(x) (17.19)

1More strictly speaking, however, the temporal ordering of “reflection following incidence” occurs with asuperposition-state traveling wave packet incident from the left, but as I say, first we’ll just deal with the eigen-states − for them, the math is simpler.

148

Page 150: PHY 416, Quantum Mechanics - SIUE

which isd2ψ(x)

dx2 =2m

h2 (V0 − E)ψ(x) (17.20)

since V0 > E, the general solution to this is

ψII = Ceκ2x +De−κ2x (17.21)

where κ2 =

√2m(V0−E)

h Now consider the behavior of this as x → ∞. Thefirst term blows up, which is not possible for the resulting probability density.Hence, C = 0. Thus

ψII = De−κ2x (17.22)

andΨII(x, t) = De−κ2xe−

iEth (17.23)

The issue now is how to join ψ1 and ψ2 at the boundary x = 0. For this, wecall on the following general theorem:

Theorem: To be acceptable, any eigenfunction ψ(x) must obey

1. ψ(x) must be finite for all x

2. dψ(x)dx must be finite for all x

3. ψ(x) must be continuous

4. dψ(x)dx must be continuous

Proof : for the time being, we offer the following simple “proof”. thisproof actually has a loophole, and there is a case in which that loophole isrealized, but we’ll worry about that later. If ψ(x) and dψ(x)

dx were not finiteand continuous everywhere, then

S(x, t) = − ih

2m

[Ψ∗(x, t)

∂Ψ(x, t)

∂x−Ψ(x, t)

∂Ψ∗(x, t)

∂x

](17.24)

would not be continuous everywhere. but, a discontinuity in S(x, t) wouldrequire a source of “sink” of probability, which, in this theory makes no sense.We now return to our step potential problem. We have

ψ(x) =

Aeik1x + Be−ik1x x < 0

De−κ2x x > 0(17.25)

149

Page 151: PHY 416, Quantum Mechanics - SIUE

continuity of ψ(x) across x = 0 requires

A+ B = D (17.26)

continuity of dψ(x)dx requires

−κ2D e−κ2x∣∣x=0 = ik1A eik1x

∣∣x=0 − ik1B e−ik1x

∣∣x=0 (17.27)

oriκ2

k1D = A− B (17.28)

Subtracting equations (17.26) and (17.28) yields

A =D2

(1 +

iκ2

k1

)B =

D2

(1− iκ2

k1

)B =

k1 − iκ2

k1 + iκ2A

D =2k1

k1 + iκ2A

we choose to use the last two equations and let A be set by normalization.Thus, the eigenfunction for energy is

ψ(x) =

Aeik1x +

(k1−iκ2

k1+iκ2

)Ae−ik1x x ≤ 0

2k1

k1+iκ2Ae−κ2x x ≥ 0

(17.29)

Thus

ΨE(x, t) =

Aei(√

2mEh x−Eh t

)+(k1−iκ2

k1+iκ2

)Ae−i(√

2mEh x+E

h r)

x ≤ 0

2k1

k1+iκ2Ae−

√2m(V0−E)

h xe−iEth x ≥ 0

(17.30)

We see that if Ψ(x, t) has a component which is incident from the left, there isnecessarily a reflected wave. The leads us to expect that Ψ(x, t) is a standingwave in region I. Indeed it is; as you can show, the eigenfunction ψI can berewritten as

ψI(x) = D cos(k1x)−Dκ2

k1sin(k1x) (17.31)

150

Page 152: PHY 416, Quantum Mechanics - SIUE

which isψI(x) = E cos(k1x+ φ) (17.32)

(find E and φ in terms of D, k1, and κ2). Finally, we show a plot of theeigenfunction for the case D real

fig 13

17.3 Reflection and Transmission Coefficients

Let us calculate the ratio of “reflected” probability flux to incident probabilityflux. we will cal this the “reflection coefficient” (R). Now,

S(x, t) = − ih

2m

(Ψ∗∂Ψ

∂x−Ψ

∂Ψ∗

∂x

)(17.33)

For a general plane wave

ei(kx−Eh t) (17.34)

as you can easily show,

S(x, t) =hk

mA∗A =

p

m|A|2 = v |A|2 (17.35)

Thus, for our eigenfunction in the region x < 0,

SI(x, t) = vA∗A− vB∗B (17.36)

where v ≡ hk1

m . Thus,

R =vB∗BvA∗A

=B∗BA∗A

=(k1 − iκ2)

∗ (k1 − iκ2)

(k1 + iκ2)∗ (k1 + iκ2)

(17.37)

which isR = 1 (17.38)

This agrees with the classical prediction, since classically the incident particledefinitely reflects.

151

Page 153: PHY 416, Quantum Mechanics - SIUE

17.4 Classically Forbidden Region

We notice something strange, there is a region of nonzero probability of theparticle to materialize in the classically forbidden region x > 0 where E < V !How can this be?

Firstly, the probability of materializing in the forbidden region is “con-siderable” only in the range of x, ∆x of order 1

2κ2(since ψII(x) ∝ e−κ2x ⇒

ψ∗IIψII ≈ e−2κ2x). Thus,

∆x ≈ 1

2κ2=

h

2√

2m(V0 − E)(17.39)

In the classical limit, (often but not always), m(V0 − E) h2, so at leastin that limit, this effect goes away. Suppose, however, that the situation isnot “in the classical limit”. Suppose measurement does force the particle tomaterialize in the forbidden region. Then, after the measurement, thereis new state function associated with the “particle” and for this newstate function

∆x ≤ h

2√

2m(V0 − E)(17.40)

This means that for the new state function

∆p/geh

∆x≥√

2m(V0 − Eoriginal) (17.41)

Thus, for the new state function

∆Enew ∼(∆p)2

2m≈ V0 − Eoriginal (17.42)

Thus, it is no longer possible to say with certainty that Enew < V0. In thisway, the paradox “disappears”. Upon contemplation, examples like this showhow thin the edge is on which the consistency of theory hangs together. But,the universe depends on it.

152

Page 154: PHY 416, Quantum Mechanics - SIUE

Chapter 18

Eighteenth Class: More on thePotential Step

Thursday, October 23, 2008

18.1 More on Free-Particle Packet Motion and Group Velocity

In order to prepare for a discussion of a wave packet incident from x = −∞on a step potential, we need first to fill in some detail previously glossed overabout the motion of free particle wave packets. As a start, in particular,back toward the beginning of the course I claimed that a reasonably tightpacket envelope moves at the group velocity vg = dw

dk

∣∣k=k+0 (where k0 is the

“central” value of k in the mix), but really only demonstrated this for the caseof two plane waves (rather than an infinite number of plane waves). Thus,we consider the case of the packet state

Ψ(x, t) =1√2π

∫ +∞

−∞g(k)ei(kx−ωt) (18.1)

where g(k) is at least fairly sharply and symmetrically peaked around a cen-tral value k = k0. Our remarks will hold true for any continuous dispersionrelation ω = ω(k), hence, they apply e.g., to water wave packets, electro-magnetic wave packets, etc., as well as deBroglie wave packets. Since ω is acontinuous function of k, we can expand it around k = k0:

ω(k) = ω(k0) +dω

dk

∣∣∣∣k=k0

(k − k0) +1

2

d2ω

dk2

∣∣∣∣k=k0

(k − k0)2 + · · · (18.2)

153

Page 155: PHY 416, Quantum Mechanics - SIUE

For our first approximation, we ignore all but the first two terms of thisexpansion. This this is not unreasonable approximation is justified by theassumed sharpness of g(k) around k0. Putting this into equation (18.1) wehave

Ψ(x, t) =1√2π

∫ +∞

−∞g(k)eikxe−iω0te−iω

′0(k−k0)t dk (18.3)

It is convenient to change variables from k to s ≡ k − k0 in this integral

Ψ(x, t) =e−iω0t

√2π

∫ +∞

−∞g(k0 + s)ei(k0+s)xe−iω

′0st ds (18.4)

Multiplying by 1 = eik0ω′0te−ik0ω

′0t, and this is

Ψ(x, t) =e−i(ω0−k0ω

′0)t

√2π

∫ +∞

−∞g(k0 + s)ei(k0+s)xe−iω

′0(k0+s)t ds (18.5)

or

Ψ(x, t) =1√2π

e−i(ω0t+k0ω′0t)∫ +∞

−∞g(k0 + s)ei(k0+s)(x−ω′0t) ds (18.6)

which tells us that

Ψ(x, t = 0) =1√2π

∫ +∞

−∞g(k0 + s)e−(k0+s)x ds (18.7)

comparing equations (18.6) and (18.7), we see that

Ψ(x, t) = e−i(ω0−k0ω′0)tΨ(x− ω′0t, 0) (18.8)

except for an uninteresting phase factor that accumulates (and which cancelsout in the probability density Ψ∗Ψ anyway), Ψ(x, t) is the same as Ψ(x −ω′0t, 0)). Thus, in the approximation in which we ignore the quadratic termin the expansion of ω(k) around k = k0, the packet moves rigidly withoutchanging shape at velocity

vg = ω′0 =dω

dk

∣∣∣∣k=k0

(18.9)

The effect of the quadratic term is then the spreading of the packet while itmoves. As shown in the homework (set 6G), the effects of the quadratic termare ignorable as long as

t m (∆x)20

h(18.10)

154

Page 156: PHY 416, Quantum Mechanics - SIUE

where (∆x)0 is the spatial extent of the packet at t = 0. This is the “coherencetime” for the packet.

Now we are ready to consider the case of a fairly tight packet incident ona potential step with mean energy less than the step height

18.2 Wave Packet Incident on a Potential Step, case E < V0

[This section of mathematics is for graduate students]Let us consider a tightly peaked packet incident on the step from negative

x1

fig 1We arrange things so that the center of the packet reaches the step at

t = 0. We consider the case where all energy components are less than V0.We write the eigenfunction in region I as

ψI = Aeik1x + Be−ik1x (18.11)

we have seen that|A| = |B| (18.12)

therefore,B = eiφ(k1)A (18.13)

for some φ(k1). It will convenient to write this as

B(k1) = e−2iθ(k1)A(k1) (18.14)

where θ ≡ −φ2 . We will see that the phase difference between B and A can

(surprisingly to some people) have an observable effect. Using

BA

=k1 − iκ2

k1 + iκ2, k1 =

√2mE

h2 , κ2 =

√2m(V0 − E)

h2 (18.15)

and defining K0 =√

2mV0

h2 , a little algebra leads to

tan (θ(k)) =

√K2

0 − k2

k(18.16)

1I am partly following here the treatment of Quantum Mechanics, Volume I, by Cohen-Tannovdji, DavidLaloe (see list of books in syllabus), pp. 79-85.

155

Page 157: PHY 416, Quantum Mechanics - SIUE

Now let’s put together a superposition (of energy eigenstates) wave packet.At time t = 0, this can be expressed as

ΨI(x, t = 0) =1√2π

∫ k=K0

0g(k)

[eikx + e−2iθ(k)e−ikx

]dk (18.17)

ΨI(x, t = 0) =1√2π

∫ K0

0g(k)eikx dk +

1√2π

∫ K0

0g(k)e−i[kx−2θ(k)] dk (18.18)

we choose g(k) to have a narrow peak centered on k = k0 < K0. We alsochoose g(k) to be real (e.g. narrow Gaussian in k). Thus, by completing each“eigenfunction seed”, using

ω(k) =hk2

2m(18.19)

Ψ(x, t) =1√2π

∫ K0

0g(k)ei[kx−ω(k)t] dk

+1√2π

∫ K0

0g(k)e−[kx+ω(k)t+2θ(k)] dk

Of these, the first is the incident packet; the second is the reflected packet.The incident packet moves in from x = −∞ with speed dω

dk

∣∣k=k0

= hk0

m thus its

center xinc is xinc = hk0

m t. This packet is thus “on course” to arrive at x = 0 att = 0. Now consider the phase function θ(k). Since g(k) is narrowly peakedaround k0, over the important range of k.

θ(k) ≈ θ(k0) +dθ

dk

∣∣∣∣k=k0

(k − k0) (18.20)

Thus, the reflected packet is

Ψrefl(x, t) =1√2π

∫ K0

0g(k)ei[kx+ω(k)t+2θ0+2θ′0(k−k0)] (18.21)

where θ0 ≡ θ(k0), θ′0 ≡ dθ

dk

∣∣k=k0

Defining ω0 ≡ ω(k0),

Ψrefl(x, t) ≡ 1√2π

e−2iθ0eiω0t

∫ K0

0g(k)e−i[kx+(ω−ω0)t+2θ′0(k−ko)] dk (18.22)

156

Page 158: PHY 416, Quantum Mechanics - SIUE

Since the important range of k is narrow, for all important k

k − k0 ≡ δk ≈ dk

∣∣∣∣k=k0

δω =dk

∣∣∣∣k=k0

(ω − ω0) (18.23)

So

Ψrefl(x, t) ≈ 1√2π

ei(ω0t−2θ0)∫ K0

0g(k)e

−i[kx+(ω−ω0)

(t− 2θ′0

dω/dk|k=k0

)]dk (18.24)

or

Ψrefl(x, t) ≈ 1√2π

e2i(ω0t−θ0)ei

(2θ′0vg

) ∫ K0

0g(k)e

−i[kx+ω

(t− 2θ′0

vg

)]dk (18.25)

where

vg ≡dω

dk

∣∣∣∣k=k0

(18.26)

The phase factors in front do not affect the probability density |Ψ|2, so

Ψrefl(x, t) ∼∫ K0

0g(k)e

−i[kx+ω

(t− 2θ′0

vg

)](18.27)

Now consider a hypothetical wave packet which leaves x = 0 at t = 0 andtravels left; it would be

Ψ(x, t) ∝∫g(k)e−i(kx−ωt) dk (18.28)

Comparing equations (18.27) and (18.28), we see that the reflected packetmoves in the x direction at the classical velocity vg = dω

dk

∣∣k=k0

, but which isdelayed in leaving x = 0 by an amount of time

τ =2θ′0vg

(18.29)

(to see this, note that t in equation (18.28) is replaced in equation (18.27) byt− τ). A little algebra shows that

τ =2m

hk0√K2

0 − k20

(18.30)

157

Page 159: PHY 416, Quantum Mechanics - SIUE

Thus, in disagreement with classical physics, the particle is not reflectedimmediately − it is said that “the particle spends time τ in the classicallyforbidden region (x > 0) before heading back.” The following sequence offigures illustrates this.

fig 2[end of graduate student section]Note the relatively long interval of time during which the packet interacts

with the step − this is a reflection (no pun intended) of the delay time τ .The sharp interference “fringes” occur as a result of the extreme sharpness

of the step − they are the result of interference between “incident” and“reflected” components of the packet. With a smoother change from V = 0to V = V0, the interference fringes are much less pronounced.

18.3 Particle “Incident” on Potential Step, Case E > V0

fig 3The eigenfunction for the energy E is of the form (as you should show)

ψI = Aeik1x + Be−ik1x k1 =

√2mE

h(18.31)

ψII = Ceik2x +De−ik2x k2 =

√2m(E − V0)

h(18.32)

For definiteness, we specify that a particle is “incident from the left”, thus weset D (“there is nothing out at x = +∞ to cause back reflections”). Thus wehave three unknowns (A, B, C). Continuity of ψ and ψ′ give two equations,thus one unknown is unspecified. We take this one to be A (arbitrary incidentamplitude). The algebra then yields

B =k1 − k2

k1 + k2A (18.33)

C =2k1

k1 + k2A (18.34)

158

Page 160: PHY 416, Quantum Mechanics - SIUE

We define a “reflection coefficient” and a “transmission coefficient” as

R =Sreflected

Sincident=|B|2

|A|2(18.35)

T =Stransmitted

Sincident=k2 |C|2

k1 |A|2=v2 |C|2

v1 |A|2since k ∝ v) (18.36)

R =

(k1 − k2

k1 + k2

)2

(18.37)

T =4k1k2

(k1 + k2)2 (18.38)

As it must be, the sum R+T = 1 (a given particle is either transmitted or itis reflected). Note the definite break with classical physics here − as we say,a given incident particle is either reflected or transmitted − it never splits.The probability of reflection is given by R, and R decreases with increasingincident energy. The probability of transmission T increases with incidentenergy (as you can see in figure 18.5). Note that, in the case E < V0, R = 1and T = 0. this makes good sense − R = 1since the reflected amplitudediffers from the incident amplitude only by a phase factor, as we saw lastclass. T = 0 for E < V0 since then ψII is not a traveling wave, so SII = 0.forming an incident wave packet for the case Einc > V0 leads to splitting ofthe packet − part is transmitted and part is reflected. Remember, however,that the reflected packet (or its modulus squared) represents the probabilityof reflection in a given case. If I send in a beam of identical particles all inthe same packet state,

fig 4 2

R is the fraction of particles reflected and T is the fraction transmitted. Agiven particle either reflects or transmits. How does a given particle “know”if it must reflect or transmit? Good question − this is quantum mechanics!

fig 5

2Figure from Cohen-tannoudji et.al

159

Page 161: PHY 416, Quantum Mechanics - SIUE

Chapter 19

Nineteenth Class: The BarrierPotential

Thursday, October 30, 2008

19.1 Barrier Penetration − Theory

We last dealt with the “step potential”; now we terminate the step beforex→∞ and thus consider the rectangular “barrier potential”

Fig 1We consider today the case E < V0 and treat the eigenfunction. Of course,

a classical particle with energy E < V0 and incident from the left would simplybounce back from the first wall with 100% probability. In quantum mechan-ics, as with the step potential, there is probability (but not 100%, as withthe step) of reflection at the left barrier, but here there is, as we will see,finite probability that a traveling probability current will be excited in theregion to the right of the barrier (region III in figure 19.1) in spite of theintervening classically forbidden region. As I mentioned to you, without theexistence of this usually very, very small effect, called “quantum tunneling”,the sun would not shine, and therefore, we would not live. Other applicationsof the effect abound, conduction of electrons in solids (tunneling of electronsthrough lattice ions, ammonia masers, radioactive decay, field emission pro-cess, new semiconductor devices, the “scanning tunneling microscope”, etc.)Let us find the eigenfunction: the time independent Schrodinger equationis

− h2

2m

d2ψ(x)

dx2 + V (x)ψ(x) = Eψ(x) (19.1)

160

Page 162: PHY 416, Quantum Mechanics - SIUE

so, in region I, the Schrodinger equation becomes

ψ′′I (x) = −2mE

h2 ψI(x) (19.2)

with solutionψI(x) = Aeikx + Be−ikx (19.3)

where k =√

2mEh . In region II, the Schrodinger equation is

ψ′′(x) = +2m(v0 − E)

h2 ψ(x) (19.4)

with solutionψII(x) = Ce−κx +De+κx (19.5)

with κ =

√2m(V0−E)

h here we must keep C since x does not go to infinity inregion II. In region III,

ψIII(x) = Feikx + Ge−ikx (19.6)

where k =√

2mEh . The second part of this a reflection; since there is nothing

out at infinity to cause this, we must have G = 0. Now we must use theboundary conditions to stitch together the three parts of the eigenfunction.The boundary conditions are, of course

• continuity of ψ at x = 0

• continuity of ψ′ at x = 0

• continuity of ψ at x = a

• continuity of ψ′ at x = a

As you can easily show, applications of these four conditions leads to fourequations

1. A+ B = C +D

2. ikA− ikB = −κC + κD

3. Ce−κa +Deκa = Feika

4. −κCe−κa + κDeκa = ikFeika

161

Page 163: PHY 416, Quantum Mechanics - SIUE

We have five unknowns and only four equations, so we could express B, C,D, and F in terms of A, which can later be set by the overall normalizationcondition1. However, since the algebraic situation is a little complicatedhere, it is best to focus on a more specific goal − of the greatest interest isdetermining the fraction of the probability current that leaks into region III,since that is a measure of the “tunneling”. That fraction is the “transmissioncoefficient”.

T =current in region III

current in region I=vIII |F|2

vI |A|2=|F|2

|A|2(19.7)

since vIII = vI. Here is an outline2 of a relatively quick route to this goal:

1. Eliminate B: multiply our first boundary condition by ik and add theresult to the second boundary condition, obtaining

2ikA = (ik − κ)C + (ik + κ)D (19.8)

2. Now multiply the third boundary condition by −κ and add the result tothe fourth boundary condition, we obtain

C =κ− ik

2κeκaeikaF (19.9)

F =κ+ ik

2κe−κaeikaF (19.10)

Now, putting both of these into equation (19.8), the result is

FA

=4iκk

[(κ+ ik)2e−κa − (κ− ik)2eκa]e−ika (19.11)

thus

T (E) =

∣∣∣∣FA∣∣∣∣2 =

16κ2k2

|(κ− ik)2 − (κ+ ik)2e−2κa|2e−2κa (19.12)

where |denominator|2 ≡ (denominator)∗(denominator), and where we’veemphasized that T depends on E (through k and κ).

1Thus, any value of E < V0 is possible − the energy isn’t quantized.2from Understanding Quantum Physics by Michael Morrison, Prentice Hall, p.314

162

Page 164: PHY 416, Quantum Mechanics - SIUE

although equation (19.12) is fairly unwieldily form due to the | |2 in thedenominator, it will be useful to us. Further algebra leads to the form mostcommonly seen in textbooks:

T (E) =

1 +1

4

(1 + k2

κ2

)2

k2

κ2

sinh (κa)

−1

(19.13)

As you can easily show, an alternate form of equation (19.13) that shows thedependence on the energy explicitly is

T (E) =

1 +sinh2 (κa)

4(EV0

)(1− E

V0

)−1

(19.14)

=

1 +

sinh2(√

2mV0a2

h2

(1− E

V0

) )4 EV0

(1− E

V0

)−1

(19.15)

We’ll look at a plot of this later, but for now, the main point is that T > 0for E < V0! The math reason “why” this occurs is that the combination ofrising and decaying exponentials in region II is not dead at the door to regionIII.

The reflection coefficient, R = |B|2

|A|2 is not unity, as it is with the potential

step, since the incoming current splits at the first wall − some goes on toregion II, and some heads back. Thus, a given incident particle will definitelyeither reflect or tunnel; quantum mechanics can only tell us the probabilities.

Of course, this is highly nonclassical behavior for a particle, but it isfamiliar behavior with classical waves − in fact, the same phenomenon occursin optics, as predicted by the wave theory (of optics) − Newton discoveredthat light can “tunnel” from one prism to another even with an air-gap inbetween − the wave is not traveling, but only “evanescent” on the air-gap −this interesting phenomenon is called “frustrated total internal reflection”.

fig 2In quantum mechanics, the reflection probability is

R(E) = 1− T (E) (19.16)

163

Page 165: PHY 416, Quantum Mechanics - SIUE

Let us now look at the probability density for this eigenfunction.fig 3P (x, t) is constant in region III, since the eigenstate is a pure traveling

wave there. Note how the amplitude is diminished there compared to regionI − because of the exponential drop (essentially) in region II. In any case, wesee clear possibility of “tunneling through the classically forbidden region”.The “surviving” state function in region III is a pure traveling wave. Notethat, in region I, P (x, t) has peaks and valleys. This is because it is a mixtureof right and left traveling waves. If the reflection coefficient were 1, as it isfor the simple step potential, then this would be a pure standing wave andthe valleys of P (x, t) would go down to zero. But, here, R 6= 1 (R = 1− T ),so we don’t get a pure standing wave − we get a mixed standing-travelingwave (in region I).

19.1.1 A Feel for the Numbers

Typical cases occur in atomic physics; there, energies and barrier heights areeV order, barrier widths are, say, molecular size (∼ 0.5mm).

Example3: Say electrons of kinetic energy 1eV encounter a barrier height

2eV and width 0.5nm. Then, the parameter K ≡√

2mV0

h2 a which appears in

T has value

K =

√2(0.511× 106 eV

c2

)(2eV )

197.3eV · mmc· (0.5mm) ≈ 3.6 (19.17)

Then for this case

T =1

1 +sinh2(3.623

√1−0.5)

4(0.5)(0.5)

≈ 0.024 (19.18)

so, about 1 electron in 40 will penetrate the barrier.

19.2 Useful Approximations for Rectangular Barrier Penetration

Of course, one can always calculate T (E) from equations (19.12) or (19.13)or (19.15) and get a numerical answer, but for insight in applications4 it’s

3From Professor R. Hill’s book manuscript Basic Quantum Mechanics4one or two of which we’ll discuss in the next class

164

Page 166: PHY 416, Quantum Mechanics - SIUE

useful to have simple − form approximations in each of two limiting cases.These are

19.2.1 “High and Wide Barrier”

(criterion κa 1) The criterion makes sense because

κa =

√2mV0

h2

(1− E

V0

)· a (19.19)

In this case, by expanding the exponentials on the sinh in equation (19.15),and keeping only through the term factor linear in κa, we have (as you shouldverify)

T ≈ 16E

V0

(1− E

V0

)e−2κa (19.20)

(high/wide approximation) When this approximation is valid, T 1. (showthis). In most applications involving “high and wide” barriers, the expo-

nential factor is so tiny that the “prefactor” 16 EV0

(1− E

V0

)is relatively quite

unimportant. For example, in applications reflecting the classical limit, evenif the prefactor has value, say, 10−6, the exponential is so much smaller thanthis that even an “error” of 6 orders of magnitude is unimportant in all butexact-seeking calculations. Thus, very often, if not usually,

T ∼ e−eκa (19.21)

where ∼ no longer means “order of magnitude”, but means “is dominatedby.”

19.2.2 “Thin Barrier”

(criterion κa 1) In this case

e−2κa ≈ 1 (19.22)

In the limit of an extremely thin barrier of finite height (κa→ 0), we expectT → 1. From equation (19.12)

T → 16k2κ2

|(κ− ik)2 − (κ+ ik)2|2=

16k2κ2

16k2κ2 = 1 (19.23)

165

Page 167: PHY 416, Quantum Mechanics - SIUE

However, we want a better approximation than this − T should be a bit lessthan 1, and we want an easy to use approximation that will tell us how muchless than 1. For this, we use the following5 method:

T =16k2κ2

|(κ− ik)2 − (κ+ ik)2e−2κa|2e−eκa (19.24)

we ignore the e−2κa in the denominator, but not the overall factor of e−2κa.Then

T ≈ 16k2κ2

16k2κ2 e−2κa ≈ e−2κa (19.25)

Therefore, the approximation of the transmission coefficient for the thin,finite square barrier is

T ≈ e−2κa (19.26)

Thus, in the case of the thin, narrow barrier, the exponential e−2κa not onlydominates T , but is actually approximately equal to it.

19.3 “Arbitrary” Shaped Barrier (E < V0)

Now consider the case of an arbitrary shaped barrier6. This can be approxi-mated as a sequence of thin barriers:

fig 3The points x1 and x2 delineate the boundaries of the classically forbidden

region. (Thus, the “barrier” extends from x1 to x2). In approximation7 theoverall transmission factor is, then

T ≈ T1T2T3 · · ·TN (19.27)

Here for Ti we use the thin barrier formula

Ti ≈ e−2κi∆xi (19.28)5Hans C. Ohanian, Principles of Quantum Mechanics, pages 87-886We use the method of Ohanian, which is related to the original method of Ganow (1928). Ohanian’s treatment

is distinguished by it clarity; many unclear treatments exist in the literature, Amore rigorous treatment requires theWKB approximation (see Griffiths chapter 8 for more information)

7The simple combining of transmission factors we are doing ignores back and forth reflections within the barrier− we are assuming that once a reflection occurs, the particle has turned back.

166

Page 168: PHY 416, Quantum Mechanics - SIUE

thus

T ≈N∏i=1

e−2κi∆xi = e−2[κ1∆x1+κ2∆x2+···+κN∆xN ] (19.29)

Taking the limit as ∆xi → 0, N →∞, we have, for our approximation

T ≈ e−2∫ x2x1

κ(x) dx (19.30)

or

T ≈ e−2h

∫ x2x1

√2m[V (x)−E] dx (19.31)

The formula above, Garnow’s approximation, has many applications. Ourderivation of it has been nonrigorous, but the same result is obtained us-ing a somewhat more sophisticated approximation scheme called the “WKBApproximation”, which is treated in an elementary way in chapter 8 of theGriffiths text.

Note: Many book derive Garnow’s result for a high and wide but arbi-trarily shaped barrier in a less rigorous way − by approximating the barrieras a sequence of “moderately wide” rectangular barriers, dropping the pref-actor for the transmission for each, and then multiplying the Ti’s and thenpretending that the moderately wide sub-barriers are thin enough to takeover to the integral in a “limit”.

167

Page 169: PHY 416, Quantum Mechanics - SIUE

Chapter 20

Twentieth Class: Applications ofBarrier Potentials

Tuesday, November 4, 2008

20.1 Applications of Gamow’s Approximation

We developed Gamow’s approximation for a non-rectangular high and widebarrier.

fig 1

T ≈ e−2∫ x2x1

κ(x) dx (20.1)

or

T ≈ e−2h

∫ x2x1

√2m[V (x)−E] dx (20.2)

This approximation is especially useful in applications, for example, to typesof radioactive decay (see following), to nuclear fusion, to field emission ofelectrons from a metal sample (see homework) and more. In many of theseapplications, one deals with the classically forbidden (but quantum-allowed)escape of a particle from a finite confined region in which it is free across abarrier of height greater than the energy of the particle and of substantialwidth.

fig 2Figure 20.2 represents this schematically. In it, r1 and r2 are the classical

turning points − the positions at which a classical particle approaching thebarrier from either classically allowed region would be turned back − these

168

Page 170: PHY 416, Quantum Mechanics - SIUE

are the places where V (r) = E, so the kinetic energy is zero. Frequently inapplications one “starts” with the particle bound inside a finite classicallyallowed region (as mentioned above) in which the particle is “free”, and oneis interested in the probability per second of tunneling across the barrier. tothe “outside world of freedom” (“escape rate”). In these applications, oneoften adopts a simple semiclassical picture in which the particle, in the initialconfined region (r = 0 to r = r1, as in figure 20.2) travels repeatedly backand forth due to “bounces” (reflection probability) off the “wall” of the well.In this semiclassical picture, the escape rate is given by

escape rate = (# of collisions with the “wall” per second) × (probabilityof tunneling per collision)

The “probability of tunneling per collision” is just the transmission coef-ficient T (E) supplied, in approximation, by gamow’s formula.

Of course, the semiclassical picture of a “bouncing back-and-forth particlein the well” is not rigorous − it is merely a semiclassical picture. The reallycorrect way to approach the problem, however, is quite mathematical1. Weuse the semiclassical picture because it is simpler and because, to the estima-tion accuracy we are interested in, it works (e.g. the following discussion).

20.2 Application to Radioactive Alpha Decay

Many heavy atomic nuclei disintegrate (“decay”) spontaneously, emittingparticles (this is the phenomenon of radioactivity)2. One such type of decay,called alpha decay, occurs when a heavy nucleus emits an alpha particle (anintact Helium nucleus) and, in the process, becomes the (lighter) nucleus ofanother element. An example of this is

88Ra226 → 86R224 + 2He4

the half life (τ) for this is 1622 years! (this is incredibly long on what we nowknow as the typical time scale for nuclear interactions − the time it wouldtake a light-speed impulse to cross a nucleus is ∼ 10−23 seconds!) but, somealpha emitting have much shorter half-lives e.g. 84Po212 has τ ≈ 3 × 10−7

1see, e.g. Quantum Theory by David Bohm (Dover), chapter 12.2Our discussion here is based on that of E. Wichmann (class handout); the original work was done by George

Gamow around 1928.

169

Page 171: PHY 416, Quantum Mechanics - SIUE

seconds! at the other extreme, 92U238 has τ ≈ 4.2 × 109 years! Any good

theory will have to explain this enormous range of lifetimes!Another fact: Each decaying isotope emits (usually) its alpha particle

with a unique energy, usually in the range 4-10 MeV (K.E.) It was foundthat there is a rough correlation between the lifetime and the energy of theemitted alpha particle.

log(τ) ∼ 1√E

(roughly) (20.3)

A good theory should explain this also.So, we have some incredibly daunting demands for any theory that seeks

to “explain” alpha decay., This was the challenge that Gamow and Codontook up with the then fledgling theory of quantum mechanics.

20.2.1 The Basic Model

1. Pretend that the alpha particle exists intact inside the nucleus. It is heldin by the very strong nuclear force, which presents a formidable barrierpotential energy profile against the alpha particle leaving the nucleus.

2. The short-range strong nuclear force (range ≈ 10−15m), while essentiallyzero outside the parent nucleus, must be very strong inside and must varystrongly with position inside the nucleus. Nevertheless, for simplicity, themodel assumes that the V (r) that the “alpha particle inside the nucleus”is subject to is flat (constant) − in other words, replace the actual V (r)inside the nucleus with its average value for 0 < r ≤ R (R is the radiusof the nucleus).

3. Since the strong nuclear force on an alpha particle is essentially zero ifit is outside the nucleus, the force on the alpha-particle outside is purelyelectromagnetic. Since Zα = 2, if we let Z ′ be the charge of the daughternucleus (Z ′ = Z − 2), then outside the nucleus,

V (r) =2e2Z ′

4πε0rr > R (20.4)

thus, the picture in this simple model is the following: The alpha-particle iscaught in a deep square well, inside of which it feels no force. Thus, in a

170

Page 172: PHY 416, Quantum Mechanics - SIUE

semiclassical picture, it bounces back and forth between the boundaries ofthe parent nucleus. V (r) looks like the (in the model)

fig 3fig 4

Each time the alpha particle collides with the “nuclear wall”, since itis “merely” very, but not infinitely high (in V ) there is some small (verysmall) probability to leak out (transmission coefficient for penetration of theCoulomb barrier from inside the square well). With enough collisions withthe wall, the accumulated chance to leak out should be “easily” relatable tothe observed half-life for emission of the alpha-particle. Calculating this wasGamow’s (and Codon’s) program − it was a means, essentially, of testingquantum mechanics.

Let’s take a moment to check the scale of figure 20.4 for V (r). Let Rc

be the classical turning point for a hypothetical particle of energy E (sameas the emitted alpha particle energy) trying to get into the well from theoutside. Then

E =1

4πε0

2Z ′e2

Rc(20.5)

This can be solved for Rc:

Rx =1

4πε0

2Z ′e2

E(20.6)

Let’s see what this is for the case of the alpha decay of 88Ra226 :

Z = 88, Z ′ = 86, E = 4.8 MeV

Rc ≈ 50× 10−13cm = 50f

Now, recall from Rutherford scattering

R ≈ r0A13 , r0 = 1.2× 10−13cm

soA = 226 ⇒ R ≈ 7f < 50f (20.7)

So, qualitatively, the picture is correct: however, the barrier should havedrawn much thicker and much higher!

171

Page 173: PHY 416, Quantum Mechanics - SIUE

20.2.2 Summary of Goals for the Calculation

Before getting involved in the details of the calculation, let us summarize themain goals: We want to calculate, via the model, half-lives for alpha decayfor radioactive nuclei and compare the predictions to the experimental data.To do this we need to

1. Calculate the approximate transmission coefficient T (E) using Gamow’sapproximation

2. Multiply T (E) by the factor f , the frequency of collisions with the barrierwall. (Of course, to do this, we first have to calculate, or estimate, f).The result of this is the decay rate R = T · f

escape probability per second = probability per “knock”×# of knocks per second

3. We must the convert R to the half life. For this, we note that themean lifetime τ = 1

R . The half life, τ1/2 is then related to τ by anarithmetical factor: since radioactive decay follows the exponential decaylaw, N(t) = N(t = 0)e−t/τ , as you can easily show, τ1/2 and τ are relatedby a factor of log(2).

20.2.3 Calculation of Collision Factor f

Following Gamow, we approximate as follows:We pretend that the nucleus is a hallow and that the alpha particle before

emission, bounces around inside it like a free particle3 bouncing off the nuclearwall (i.e., we completely ignore the strong nuclear force of the alpha particleinside the nucleus!)

fig 5At first sight, it may seem outrageous to “ignore” the strong force inside

the nucleus. We can rationalize this somewhat, however, by noting thatnucleons (protons and neutrons) in a nucleus are tightly packed, thus, analpha particle in the interior of the nucleus might be imagined to feel equalforces in all directions, for a net force of zero. With this model, then, if R is

3of energy E = the observed energy of the emitted alpha particle after the decay − i.e., about 6 MeV for the Radecay we’re taking as typical

172

Page 174: PHY 416, Quantum Mechanics - SIUE

the nuclear radius, the “back and forth time” (τ0) is

τ0 =2R

v(20.8)

where v =√

2Emα

, so

τ0 = 2R

√mα

E(20.9)

The collision frequency f is 1τ0

= 12R

√Emα

in this model.

f =1

τ0=

1

2R

√E

mα(20.10)

It is helpful to get an idea of the magnitude of f for a typical case. We takethe case of 88Ra226 :

f ≈ C

2R

√E

mαC2 (20.11)

with

R ≈ 7× 10−15 m ⇒ 2R ≈ 1.4× 10−14 m

E ≈ 4.8 MeV

mαc2 ≈ 4 GeV = 4000 MeV

(since each alpha particle has 4 nucleons). Then,

f ≈ 3× 108 m/s

1.4× 10−14 m

√4.8

4000∼ 1021 Hz (20.12)

We see that f , the “number of knocks per second” is typically very large.

20.2.4 Calculation of T (E)

(Gamow, 1928)Recall that

ln (T ) ≈ −2

h

∫ x2

x1

√2m [V (x)− E] dx (20.13)

173

Page 175: PHY 416, Quantum Mechanics - SIUE

Here, V (r) = 2e2Z ′

4πε0rfor the barrier region, so

ln(T ) ≈ −2

h

∫ r2

r1

√2mα

[2e2Z ′

4πε0r− E

]dr (20.14)

Here, r1 and r2 are the two turning points. Now, E = 14πε0

2Z ′e2

r2(conservation

of energy), so

ln(T ) ≈√

2mE

h

∫ r2

r1

√r2

r− 1 dr (20.15)

This integral is doable, the result is

ln(T ) ≈√

2mE

h

[r2

2− sin−1

(√r1

r2

))−√

(r2 − r1) r1

](20.16)

In our case, r1 r2, so

ln(T ) ≈√

2mE

h

[π2r2 − 2

√r1r2

](20.17)

again using E = 14πε0

2Z ′e2

r2⇒ r2

14πε0

2Z ′e2

E and evaluating the constants gives

ln(T ) ≈

(e2

4πε0

π√

2m

h

)· Z

′√E−√

e2

4πε· 4√m

h·√Z ′r1 (20.18)

or

ln(T ) ≈ 1.980Z ′√

E (in MeV)− 1.485 f−1/2 ·

√Z ′r1 (20.19)

where f=Fermi=10−15 meters. As a rough approximation to the above, wetake Z ′ ≈ 86, R ≈ 7.3f for all alpha-decaying nuclei. (Recall that R ∼1.1A1/3f for all nuclei, for and archetypal nucleus 88Ra226 , A = 226 ⇒ R ≈7.3f). converting, then, to base 10 logarithms (customary for plotting later),one winds up with, as a rough approximation,

log10(T ) ≈ − 148√E (in MeV)

+ 32.5 (20.20)

174

Page 176: PHY 416, Quantum Mechanics - SIUE

20.2.5 Comparison with Experiment

We not put the pieces together to compare with experimental data. We haveτ=mean lifetime= 1

R = τ0T (since f = 1

τ0. As we remarked, the values of τ

for diferent nuclear species subject to alpha decay vary over many orders ofmagnitude; therefore, for a global comparison it is good to deal with log(τ)rather than τ . In fact, it is traditional to deal with log10(τ). thus, from theabove,

log10(τ) = log10(τ0)− log10(T ) (20.21)

From equation (20.20), this is

log10(τ) = log10(τ0) +148√

E (in MeV)− 32.5 (20.22)

Now, as we saw, for the case of 88Ra226 , f = 1τ0∼ 1021, so, for this case,

log10(τ0) = −21. As you can convince yourself by looking at values of E fordifferent parent nuclei, log10(τ0) varies at most ±1 or so from case to case;this variation is very minor compared to the other two terms of equation(20.22). Therefore, since we are only interested in an approximate answer,we may as well approximate log10(τ0) as −21 in all cases. thus, the theorypredicts,

log10(τ) ≈ 148√E (in MeV)

− 53.5 (20.23)

In spite of some rough approximations on the way to equation (20.23), we4

compare it to the data for known alpha emitters in figure 20.6.fig 6

The results are a spectacular confirmation of quantum mechanics noticethat the trend5 of the theory is correct over 23 orders of magnitude!

The dotted line (theory prediction) is actually for the mean lifetime; whereasthe experimental points are half-lives. As we remarked,

τ1/2 = (ln(2)) τmean ≈ (0.69)τmean (20.24)

however, given the approximations of our discuaaion and the latitude in theplot from the dotted line, the factor (ln(2)) is not worth bothering with here.

4The plot in figure (20.6) is from Quantum Mechanics (Berkeley Physics Series, Volume IV) by Evyinf Wich-mann (McGraw-Hill)

5of course, for individual cases (see plot) the erros can be order factor of 100 or even 1000, but the importantthing is the trend

175

Page 177: PHY 416, Quantum Mechanics - SIUE

20.3 Another Application − Fusion in the Sun’s Core

For fusion, protons have to get together to fuse, but, as we remarked in thebeginning of the course, there is a ∼ 1MeV Coulomb barrier in the way, and,in the core of the Sun, the mean proton energy is only ∼ 1

1000 of this.fig 7Here we have the “opposite” of the alpha decay problem, we just treated

a proton tries to get into the deep square well of attraction presented by theother proton from the outside, but in the way is a very high coulomb barrier!Now, even with quantum mechanics a simultaneous fusion of the 4 protonsnecessary is very unlikely − we must look for a chain. We believe that thatcorrect chain for the sun is (Bethe)

11H + 1

1H → 21H + e+ + ν

21H + 1

1H → 32He + γ

32He + 3

2He → 42He + 1

1H + 11H

This is called p-p cycle.the ν’s produced in the core of the sun should escape (they interact so

weakly that their mean free path in most matter is ∼ light years!). For thelast 25 years, R. Davis has tried to detect these (actually these ν’s are toolow in energy − he looks for higher energy ν’s from the p-p III cycle) in ashielded tank in Homestake Nums, S.D. filled with 100,000 gallons of C2Cl4(cleaning fluid) via 37Cl + ν → 37A , which is radioactive. Only one problem− he only sees about 35% of what he should! This is one of great unsolvedproblem of astrophysics (until recently).

20.4 Another Example of Barrier Penetration −Ammonia Molecule

Another (of many) example of barrier penetration is offered by the Ammoniamolecule (NH3. The location of the N ion is determined by the conditionthat there the energy is a minimum; there are two equivalent positions forthis (as in figure 20.8). thus, in the region between the two minima there isa higher potential energy.

fig 8

176

Page 178: PHY 416, Quantum Mechanics - SIUE

The potential energy profile is shown in figure 20.9fig 9As we will see in the near future, a state with the N-ion “on one side” is

not an eigenstate of the Hamiltonian for this potential energy profile. Conse-quently, if a measurement materializes the N on one side, the state functionwill change over the course of time. As we will see, the initial state functionhas a certain probability per unit time to “diffuse” through the forbiddenregion. and. given enough time, can end up almost completely on the otherside − in fact, analysis shows that the N-ion will tunnel back and forth acrossthe barrier at a fixed frequency! A calculation of this frequency (which wedo not do here) shows it to be

f = 2.387× 1010 Hz (20.25)

This forms the basis of the Ammonia MASER. A maser of this sort wasused in the discovery of the cosmic microwave background radiation thatpermeates the universe. Now, the N nucleus is charged, and as you know,classically, a charge moving back and forth in space emits radiation. Thisradiation, at this specific frequency, has been detected by radio telescopesas originating in interstellar space, thus signaling the presence of Ammoniamolecules in the interstellar medium.

20.5 Another Application − Covalent Molecular Bonding

Consider, e.g., the H+2 molecule − two protons and one electron. An electron

sees a Coulomb barrier between the protonsfig 10Thus, classically, an electron “near” one proton must stay there. (by

“near”, we mean somehow localized by measurement). However, accordingto quantum mechanics, tunneling is possible and the probability density canflow back and forth, leading to the mechanism of the covalent bond.

Other applications of tunneling include:

• superconductors

• possible vaporization of cosmic back holes

177

Page 179: PHY 416, Quantum Mechanics - SIUE

• “scanning-tunneling microscope”, etc.

fig 11Finally, the following sequence of figures shows the situation when a wave

packet with E = 12V0 is incident on a barrier. The strong interference fringes

are an artifact of the sharpness of the barrier.fig 12

178

Page 180: PHY 416, Quantum Mechanics - SIUE

Chapter 21

Twenty-first Class: Potential Barrierand Finite Square Well

Thursday, November 6, 2008

21.1 Potential Barrier, Case E > V0

fig 1For this case we have traveling wave eigenfunctions in all three regions:

ψI(x) = Aeik1x + Be−ik1x k1 =

√2mE

h(21.1)

ψII(x) = Ceik2x +De−ik2x k2 =

√2m(E − V0)

h(21.2)

ψIII(x) = Feik3x + Ge−ik3x k3 = k1 (21.3)

where G = 0 because there is nothing at infinity to cause a reflection back.Above, we’ve assumed a plane wave incident from region I. Say, we wantto know the transmission factor (T) to region III. We could go through thebusiness of applying the boundary conditions to this, but we don’t have to!We already done it! To see why, consider that the only difference in the formof our set of eigenfunctions to the eigenfunctions for the case V < V0 seemsto be in region II: For E < V0 in region II we had

ψII(x) = Ceκ2x +De−κ2x κ2 =

√2m(V0 − E)

h(21.4)

179

Page 181: PHY 416, Quantum Mechanics - SIUE

This is still valid if E > V0! (“The math doesn’t know the difference”).However, if E > V0, κ2 is now imaginary:

κ2 =

√2m(V0 − E)

h=

√(−1)2m(E − V0)

h= i

√2m(E − V0

h

and now (E − V0) > 0 so κ2 is imaginary. Therefore, if we define ik2 ≡ κ2

with k2 real, then indeed

ψII(x) = Ceik2x +De−ik2x (21.5)

Thus, we can take our old expression for TE<V0, nameley,

TE<V0=

1 +sinh2(κ2a)

4 EV0

(1− E

V0

)−1

(21.6)

and wherever we see κ2, replace it by ik2, and we have TE>V0! Now

sinh(κ2a) =eκ2a − e−κ2a

2→ eik2a − e−ik2a

2= i sin(k2a) (21.7)

sosinh2(κ2a)→ (i)2 sin2(k2a) = − sin2(k2a) (21.8)

so

TE>V0=

1 +sin2(k2a)

4 EV0

(EV0− 1)−1

(21.9)

(note that I changed the sign in the denominator) If you like, you can plug

in k2 =

√2m(E−V0)

h . Again, R works out to (As it must) R = 1T .Example: Photons incident on a slab of glass − 4% are reflected back.

This is barrier penetration with some probability of reflection. With wavepackets, recall that the packet spits, but the particle is definitely either re-flected or transmitted.

180

Page 182: PHY 416, Quantum Mechanics - SIUE

21.2 Resonance Transmission

Look at the expression for TE>V0:

TE>V0=

1

1 +sin2(k2a)

4 EV0

(EV0− 1) (21.10)

Notice that it has maxima (“resonances”) when sin2(k2a) = nπ− then T =100% (perfect transmission − no reflection!)! Let us investigate this, thesetransmission resonances occur when

k2a = nπ n = 1, 2, 3, · · · (21.11)2π

λ2a = nπ (21.12)

λ2 =2a

n(21.13)

How can we understand this physically? Consider a DeBroglie plane waveincident from the left − part of it reflects at the “front wall” (call that partr1) and part reflects at the back wall (call that part r2).

fig 2r1 and r2 combine and interfere with each other in region I − and their

resultant is the net reflection off the barrier 1/ our condition for resonanttransmission, λ2 = 2a

n , means that the two reflections r1 and r2 are in phase,and hence add, in region I. And so are higher order reflections (“back andforth”) r3, r4, · · · But wait − that means that a resonance the reflections aremax, which means that the transmission should be min, not max! So what iswrong? What is wrong is that we have neglected the fact that there is a 180

phase shift for r2 when it is formed at the bounce off wall 2. Why is therea 180 phase shift there? To see why, let’s look at a situation in classicalwave physics. Say a wave pulse is incident from a less dense string to a moredense string. Then, at the junction, there is partial transmission and partialreflection − and, as you know, the reflection is inverted.

fig 31ignoring, for the moment, back and forth multiple reflections resulting in an r1r2r1

181

Page 183: PHY 416, Quantum Mechanics - SIUE

The signal for inversion is an increase in wave number in growing frommedium 1 to medium 2. You can check that k increases by noting that the

phase velocity vφ =√

Fµ = λf = ω

k decreases in going to medium 2, and since

the frequency is the same in both media, k increases (λ decreases). (The samesort of thing happens when light waves travel from a medium with index ofreflection n1 to a medium with index n2.) Now look at deBroglie waves. Wehave, for any region with constant V ,

K =1

h

√2m(E − V ) (21.14)

Thus, the smaller V , the bigger k. So, for our barrier, kIII > kII, and therefore,the reflection r2 suffers “inversion” (180 phase shift). However, r1 does not.(Why?). So at resonance, rnet = 0 ⇒ T = 1. Some plots showing thebehavior of T are shown for typical cases in figure 21.4. Of course, at allvalues of E

V0, R = 1− T .

fig 4

21.3 The Finite Square Well

We now consider the effects of an attractive, rather then a repulsive, energyprofile. We consider only rectangular profiles (called finite square wells); thepotential energy profile is shown in figure 21.5.

fig 5Some books2 take the bottom of the well to be V = 0 and the top to be

V = V0 > 0; however, in many applications, it is more convenient to define theenergy zero so that the top of the well (and hence, the “outside world”) is atV = 0 − this makes sense in, e.g., scattering problems in nuclear and particlephysics and is in accord with the convention that V → 0 as x → ±∞; wewill thus follow this second convention3. With this, the bound-state energiesare negative.

fig 6As in classical physics, all observable results are the for either convention;

2e.g., R. Hill, to be referred to later. This convention is common in elementary books.3so does your Griffith text (section 2.6) in accord with his particle physics background

182

Page 184: PHY 416, Quantum Mechanics - SIUE

to pass from one to the other, simply note that

E → E + V0 (21.15)

On the left, E is positive and less than V0, and on the right, E is positiveand less than V0 (since E < 0).

21.3.1 Finite Square Well, Incident Plane Wave State With E > 0

We considerfig 7We note that we choose to center the well at x = 0 (so it runs from

x = −a to x = a). This is done to simplify the search for solutions by takingadvantage of “parity” − see later. And, consider the case of an incident planewave with E > 0. Then, for the eigenfunction,

ψI(x) = Aeik1x + Be−ik1x k1

√2mE

h(21.16)

ψII(x) = Ceik2x +De−ik2x k2

√2m(e+ V0)

h(21.17)

ψIII(x) = Fe−k1x (21.18)

The easiest way to derive the transmission coefficient T ≡ |F|2

|A|2 is to start with

the expression for T for the potential barrier and replace.fig 8thus

TE>0 =

1 +sin2(2k2a)

4 EV0

(1 + E

V0

)−1

(21.19)

this expression for T is in mixed notation

[k2 =

√2m(E+V0)

h

]. It also provides

the ratio of transmitted to incident currents, and hence, the transmissionprobability, since k3 = k1. The reflected probability is, of course

R = 1− T (21.20)

183

Page 185: PHY 416, Quantum Mechanics - SIUE

We see that, as wit the repulsive barrier, there are situations in which thewell is “transparent” − i.e., T → 1 (and R→ 0), namely, when

sin(2k2a)→ 0 ⇒ 2k2a = nπ, n = 1, 2, 3, · · · (21.21)

Naturally, this is the same condition as for the square barrier, namely nλ2 =4a = back and forth distance in region II. As you can show, this says thatfor incident energies,

E = −V0 +n2π2h2

8ma2 (21.22)

there is perfect transmission. This perfect transmission was first observedexperimentally. In 1921 (before quantum mechanics was discovered!) byRamsauer & Townsend when they scattered low energy electrons off of nobleatoms (like xenon) nuclei − at the right values of energy (given by the lastequation), the electrons essentially pass right though the nuclei as if it weren’tthere! Experiments electron energy of ≈ 1 eV, and then working backwardthough the formula showed that the effective well depth presented by thexenon ion is ≈ 10 eV. (We use a ∼ 2A − as the diameter of the ion.) foran attractive square well, depending of the ratio of well width to depth, theresonances on T can be very sharp, as in figure 21.9 and figure 21.10.

fig 9fig 10We conclude this section with a look at three cases (different mean en-

ergies) of Gaussian wave packets approaching and “scattering” from finitesquare wells. (in each case, E > 0

fig 11fig 12fig 13

21.3.2 Bound States of the Finite Square Well

We take the well to “run from −a to a” for bound states E < 0 (where thewell depth is to −V0.

fig 14As we remarked some time ago, since E < 0, regions I and III are for-

bidden; we expect most of the state function amplitude to be in region II

184

Page 186: PHY 416, Quantum Mechanics - SIUE

with (for eigenfunctions) only a bit of leakage into regions I and III. Thus,for example, we might expect an eigenfunction to look, say, something likethis

fig 15These states are bound states − they hang around in the well region.

21.3.3 Bound States for Finite Square Well − Qualitative Observations

Firstly, we note that the the solution to the time independent Schrodingerequation(i.e., the eigenfunction) in region II can be written equivalently inany of the following ways:

ψII(x) = Feikx + Ge−ikx (21.23)

ψII(x) = A sin(kx) + B cos(kx) (21.24)

ψII(x) = Yeikx + Z cos(kx) (21.25)

These are all equivalent. The first form is convenient for considering proba-bility currents (since its composed of traveling wave pieces). However, for thebound states we are interested in, the net current must be zero (no left-rightprobability flux) and so the sin-cosine form is more convenient. Another rea-son the sine-cosine form is more convenient is that these (sin and cos) arefunctions of definite parity (odd and even). As in the case of the infinite well,if the ordinate axis is centered, the eigenfunctions must have definite parity.Mathematical proof of this assertion is in Professor Hill’s book (section 4.11)however, a “physics proof” of it is simply the following: As you know, in aneigenstate, probability density at any x is independent of time. Since the po-tential energy is left-right symmetric around x = 0, then so must ψ∗(x)ψ(x)− i.e., ψ∗(x)ψ(x) must be an even function. There are only two ways forthis to happen − either ψ(x) or ψ∗(x) is odd. Which possibility corresponds

to the ground state? As noted above, in region II, k =√

2m(E+V0)h2 . How E

means lok k, which means long wavelength. We note, for example, thatfig 16In either case, though, these have to map onto the region I and region III

pieces of the eigenfunction.

185

Page 187: PHY 416, Quantum Mechanics - SIUE

21.3.4 Parity of Eigenfunctions

Recall that inside the well (region II), we choose to write the general form ofthe eigenfunctions as

ψII(x) = A sin(kx) + B cos(kx) (21.26)

This is especially convenient as the form in light of our recent result that, ifV (x) is symmetric around some point (call it x = 0, so well runs −a to a),then the eigenfunctions must have definite (plus or minus) parity around thispoint. Sine and cosine both have definite parity. So in region II, either

ψ(x) = A sin(kx) (21.27)

orψ(x) = B cos(kx) (21.28)

but not any linear combination of both.So, with the well centered on x = 0, the ground state must be of even

parity, and hence, inside the well must be of the form cos(kx) with k =√2m(E+V0)

h2 . In regions I and III, the time independent Schrodinger equationis

− h2

2m

d2ψ

dx2 = Eψ ⇒ d2ψ

dx2 = −2m

h2 Eψ (21.29)

has solutions that are exponential (since E < 0), namely

ψI(x) = Ae+κx (21.30)

ψIII(x) = De−κx (21.31)

where κ =√−2mEh > 0. Since e−κx blows up at x → −∞ and e+κx blows up

as x→ +∞. So the tails in the “forbidden” regions are simply exponentialssine the ground state is even, D = A. We choose the even cosine in region II(since it has the longest wavelength) and match it to the external exponentialfall offs. So, the ground state looks like

fig 17The first excited state has next shorter wavelength in the well, hence, by

similar logic, it must look likefig 18

186

Page 188: PHY 416, Quantum Mechanics - SIUE

Why does this state have higher (but still negative) energy than the groundstate? Answer: The wavelength is shorter in region II, and λ ∝ 1√

E

λ =2π

k2, k2 =

√−2m(E + V0)

h(21.32)

Thus, E2 is more than infinitesimally greater than E1. Similar logic showsthat successively higher bound states will have successively higher energies(E3, E4, · · · ) all of which are still negative, and eigenfunctions that alternatein parity. Of course, only certain values of energy are “allowed”.

21.3.5 Finite Square Well, Quantitative Treatment of Bound States

fig 19Based on our previous discussion4

ψII(x) =

A cos(kx) even

A sin(kx) odd(21.33)

ψIII(x) = Be−κx (21.34)

ψI(x) =

ψIII(−x) even

−ψIII(−x) odd(21.35)

where k =√

2m(E+V0)h2 and κ =

√−2mEh2 . Since we have definite parity for

each eigenfunction, we need only apply the boundary conditions at x = a

and they are automatically taken care of at x = −a (even or odd). We dothe even case.

The continuity of ψ(x):

A cos(ka) = Be−κa (21.36)

The continuity of ψ′(x):

−kA sin(ka) = −κBe−κa (21.37)

Dividing these,k tan(ka) = κ (even case) (21.38)

4Reference: Basic Quantum Mechanics by Professor R. Hill, Sections 7.8, 7.9, howver, we do not use hispotential energy convention (as noted earlier)

187

Page 189: PHY 416, Quantum Mechanics - SIUE

which is really an (transcendental) equation for the energy:√2m(E + V0)

h2 tan

(√2m(E + V0)

h2 a

)=

√−2mE

h2 even case (21.39)

For the odd case (as you can easily verify),

−k cot(ka) = κ (odd case) (21.40)

We now simplify that form of equation (21.38) and (21.40) by defining di-mensionless energy and potential energy variables (following Professor Hill,but in our convention)

Let E ′ be the difference of E and the bottom of the well, i.e., E ′ ≡ E+V0

fig 20Then, define a new dimensionless variable ν by

ν2 ≡ E ′

E infgnd

(21.41)

where E infgnd is the amount by which the ground-state energy in the equivalent

width (2a) infinite square well is above the floor of that well.

E infgnd =

π2h2

2m(2a)2 (21.42)

thus

ν2 =E ′

π2h2

8ma2

(21.43)

We also define another dimensionless less variable ν by

ν2 ≡ V0

E infgnd

(21.44)

Thus

µ2 =8m

π2h2V0a2 (21.45)

so µ is a measure of the “strength” of the well − i.e., µ ∝√V0a. µ is then also

dimensionless. Equations (21.38) and (21.40) can then be written in terms

188

Page 190: PHY 416, Quantum Mechanics - SIUE

of µ and ν as

ν tan(π

2ν)

=√µ2 − ν2 even solutions (21.46)

−ν cot(π

2ν)

=√µ2 − ν2 odd solutions (21.47)

These can be solved graphically (see figure 21.21) by plotting the left-handsides of equations (21.46) and (21.47) and also their right hand side versusν − solutions for each type (odd or even) are defined by where these curvesintersect. One nice aspect of this is that , since µ is a constant (µ ∝

√V0a,√

µ2 − ν2 is a quarter circle of radius µ on such a plot.fig 21

Figure 21.21 shows the situation for µ = 2.5 so V0 = (2.5)2π2h2

8ma2 . So

V0a2 ≈ 0.78π2h2 (21.48)

(in general. µ is, as we remarked earlier, a measure of a√V0). We see that,

for this value of µ (i.e., this product of a√V0) thre are 3 bound (E < V0)

states. Based on our parity considerations, these bound states must look (atleast semi-quantitatively) as shown in figure 21.22.

fig 22To find the intersection points you can use trial and error with a pocket cal-

culator, or, if you prefer, Newton-Raphson iteration in a spreadsheet such asMicrosoft Excel. (This is explained in loving detail in section 7.9 of ProfessorHill’s book manuscript). From the values of ν at the intersection points youcan easily obtain the energy for each bound state, and from that, the valuesof k and κ for the state. Thus, without bothering with the overall normaliza-tion, you can, if called to, write out the entire state functions. We note that,as we expected, the energies are always a bit lower than the correspondingenergies for an infinite square well of the same width.

fig 23We note an interesting point − by looking at figure 21.21 we see that, no

matter how shallow the well (i.e., how small µ is) there is always at least onebound state. This, incidentally, is only true in one-dimensional wells − inthe real world, three dimensional wells, there can be no bound states.

189

Page 191: PHY 416, Quantum Mechanics - SIUE

Chapter 22

Twenty-Second Class: The FiniteSquare Well

Tuesday, November 11, 2008

22.1 Brief Review of Quantitative Solutions for Finite Square Well

Let’s recall some features of the path to our solution for the case E < V01.

fig 1The eigenfunctions must have definite parity (even or odd), thus

ψII(x) =

A cos(kx) (even)

A sin(kx) (odd)(22.1)

ψIII(x) = Be−κx (22.2)

ψI(x) =

ψIII(−x) (even)

−ψIII(−x) (odd)(22.3)

Since higher energies implies more curvature of ψ, and since the odd solutionshave ψ = 0 at x = 0,the ground state must be of even parity (cosine inthe well) and hence the first excited state is of odd parity, etc. Since wehave definite parity for each eigenfunction, we need only apply the boundaryconditions at x = a and they are automatically taken care of at x = −a (evenor odd). We do the even case.

Continuity of ψ: A cos(ka) = Be−κa

Continuity of ψ′: −kA sin(ka) = −κBe−κa

1reference: Basic Quantum Mechanics by R. Hill, sections 7.8 and 7.9

190

Page 192: PHY 416, Quantum Mechanics - SIUE

or dividing

k tan(ka) = κ (even) (22.4)

−k cot(ka) = κ (odd) (22.5)

We solved these transcendental equations graphically − we defined dimen-sionless variables

ν2 ≡ E ′

E infgnd

(22.6)

µ2 ≡ V0

E infgnd

(22.7)

where E ′ = E − (−V0) = E + V0 and E infgnd is the ground state energy of an

infinite square well of same width (2a). (µ ∝√V0a). In terms of which the

transcendental equations become

ν tan(π

2ν)

=√µ2 − ν2 (even) (22.8)

−ν cot(π

2ν)

=√µ2 − ν2 (odd) (22.9)

and the plotted left-hand sides and the right-hand side on one plot versus ν;from the intersection points you can easily deduce the energies.

fig 2From knowing the energies you can then deduce the full form of each of

the eigenstates ψn(x, t); probability you will do an example of this in thehomework.

22.2 Some Intuition

Let’s now seek some intuition about why only certain energies are allowed.Of course, we already understand the basic reason for this2, but here, we canprovide a bit of additional insight for the finite square well. We consider theground state. Inside the well the solution is a cosine; outside it is exponential. We must have both ψ(x) and ψ′(x) continuous across the two boundaries(x = ±a).

2section B of class notes for class of Thursday, October 16, especially the discussion of curvature of ψ and itsimplications therein.

191

Page 193: PHY 416, Quantum Mechanics - SIUE

Now, there is no problem in making ψ continuous at x = ±a for anywavelength (i.e., for any value of energy) in region II − both ψI and ψIII

have their own (same) constant multiplier that can be arbitrarily adjustedto match any amplitude at x = a (the amplitude at x = −a is the same asthat at x = a since ψ is even. However, in general, continuity of ψ′ is notsimultaneously achieved. To see this: If E is chosen too low3, the slope of theexponential is too steep at the edge of the well (or, if you prefer, the slope ofthe cosine is not steep enough):

k =

√2m(E + V0)

h2 E to neg. ⇒ low k ⇒ low cosine slope (22.10)

κ =

√−2mE

h2 E too neg. ⇒ high κ ⇒ high exponential slope (22.11)

So, if E is too negative, the situation is shown in figure 22.3.fig 3Which has discontinuous slope at x = ±a. If E is too high, the situation

is shown in figure 22.4/fig 4which also has discontinuous slope. Only for a particular value of E do we

get a ground state that is continuous in both ψ and ψ′

fig 5Now let us think about the energies of the states. How do they compare

to the energies of an infinite well of the same width? We expect they be abit lower than those of the infinite square well. Why? Because in the groundstate we don’t fit half a wavelength in the well, whereas in the infinite wellwe do. So, for the finite well, the wavelength in the interior to the well regionis a little longer than that for the infinite well, so the energy is a little lower.This is true for each eigenstate.

22.3 Applications

An understanding of the finite square well is extremely important for under-standing many applications, both in nature and in technology. For example

3i.e., too negative, or too little bigger than −V0.

192

Page 194: PHY 416, Quantum Mechanics - SIUE

many of the properties of the conduction of electricity in metals (quantumtheory of conduction) can be understood on the basis of the “quasi-free elec-tron model”’ in which “free” electrons are moving in a (three-dimensional)finite square well.

fig 6Most of these applications involve three-dimensional wells and considera-

tion of states of many particles in the well at once. We’re not quite ready forthese, so we defer detailed consideration of them to the next course. There is,however, one important application that we are about ready to briefly treat− that of a “free” neutron inside an atomic nucleus.

22.3.1 Application − Nuclear Potential and the Deuteron

A good application pertains to a simplified picture of a neutron inside thenucleus. We consider it relatively free to move about the interior of the (weassume spherical) nucleus, but it is pulled back strongly if it tries to reachthe edge, which is at a few ×10−13 cm. We assume such a potential energycurve looks more or less like figure 22.7.

fig 7The “smoothed out” version shown in figure 22.7, at least in rough ap-

proximation. For “simplicity”, we assume that the neutron is in a sphericalwell (sharp edged)

V (r) =

−V0 0 ≤ r ≤ b

0 r > b(22.12)

For this we need the Schrodinger equation in three dimensions. We willdevelop this a pit later in the course, but so as to not get distracted by thatdevelopment now. I will tell you that, for state functions with no dependenceon direction (only depends on r, distance from an origin), if we define

u(r) ≡ rψ(r) (22.13)

then u(r) obeys a very very familiar Schrodinger equation

− h2

d2u(r)

dr2 + V (r)u(r) = Eu(r) (22.14)

193

Page 195: PHY 416, Quantum Mechanics - SIUE

which is identical to the one-dimensional time independent Schrodinger equa-tionwe are used to. Here µ is the “reduced mass”, which should be familiarfrom classical mechanics. (If it is not, just think of µ as the mass). So, forour problem of a “free” neutron inside a nucleus,

− h2

d2u(r)

dr2 − V0u = −Bu r ≤ b (22.15)

− h2

d2u(r)

dr2 = −Bu r > b (22.16)

where B is the binding energy (−B < 0 is the actual energy of the neutronin the nucleus). Note that our finite well is now, of course, “one-sided”:

fig 8Now, for r ≤ b, the solution is

u(r) = A sin(kr) + B cos(kr) (22.17)

where k =

√2µ[−B−(−V0)]

h =

√2µ(V0−B)

h . Note that the first boundary conditionis

u(r) = 0atr = 0 (22.18)

so the ground state cannot be cosine − nor can cosine be present in any ofthe bound states4 − B = 0, so

u(r) = A sin(kr) (22.19)

since u(r) = rψ(r),

ψin(r) =Ar

sin(kr) r ≤ b (22.20)

Now, for r > b, the solution is

u(r) = Ce−κr + Fe+κr (22.21)

where κ =√

2µBh2 . Since ψ(r) must goto zero as r goes to infinity, F = 0.

Now, we have to match the solutions at r = b

1. Continuity of ψ(r = b):

Ab

sin(kb) =Cb

e−κb ⇒ A sin(kB) = Ce−κb (22.22)

4There is not problem with our “parity theorem” here the well is not symmetric around r = 0

194

Page 196: PHY 416, Quantum Mechanics - SIUE

2. Continuity of ψ′(r = b):

kA cos(kB) = −κCe−κb (22.23)

dividing,k cot(kb) = −κ (22.24)

This is exactly the same as the condition for the odd solutions of the one-dimensional finite square well. (as you will see by reading section 7.10 ofHill5, this is not accident). It follows from this that a three dimensional wellcan be too shallow to have any bound states! If there is a bound state, theinterior sine wave function u(r) must pass π/2 in phase at r = b so that it isfalling and can connect on to the decreasing exponential outside as in figure22.9.

fig 9As a concrete example, we consider the deuteron (bound state of neutron

and proton, nucleus of deuterium). We take b ≈ 1.4 × 10−13 cm (recall thatR ∼ 1.2A1/3 from Rutherford) and B ≈ 2.225 MeV (measured as energy ofthe photon (γ) in n + p → d + γ). (note: Here, the reduced mass, µ =

mp

2since mp ≈ mn). Then our condition

k cot(kb = −κ (22.25)

becomes √2µ(V0 −B)

hcot

(√2µ(V0 −B)

hb

)= −√

2µB

h(22.26)

As we will see in a moment, V0 B (binding energy small compared to welldepth − deuteron is “barely bound”). Therefore, to ger an idea of what isgoing on, we begin by setting V0 −B ≈ V0. Then, our condition is

√2µV0

hcot

(√2µV0

hb

)≈ −√

2µB

h(22.27)

or

cot

(√2µV0

hb

)≈ −

√B

V0(22.28)

5Please consider this a reading assignment

195

Page 197: PHY 416, Quantum Mechanics - SIUE

Since, B V0, as a first approximation

cot

(√2µV0

hb

)≈ 0 (22.29)

The bound state then corresponds to√

2µV0

hb ≈ π

2(22.30)

or

V0 ≈π2h2

8µb2 ≈π2h2

4mpb2 (22.31)

Let’s put some numbers into this: We have

h2 ≈(6.6× 10−16 eV·s

)2

b ≈ 1.4× 10−15 m

µ ≈ 0.5 GeV

c2 =5× 108 eV

c2

(µ ≈ mp/2). So,

V0 ≈(10)

(44× 10−32

)eV2 · s2

(9× 1016

)m2/s2

4 (109eV) (2× 10−30m2)(22.32)

V0 ≈ 50MeV (22.33)

(believed to be about right) This confirms our approximation V0 B. Actu-

ally, cot(√

2µV0

h b)

is a little less than zero. Therefore,√

2µV0

h b is a little greater

than zero, so

√2µ(V0−B)

h b = kB is a little greater than π/2. As it must be tomap onto the decaying exponential outside the well.

22.4 Further Intuition into Bound State Eigenfunctions

Frequently − usually, in real problems we have to deal with potential wellswith non-square shapes. Very typically, the “bottom” of the well is notconstant. Such problems, while ubiquitous in research, can be quite hard tosolve analytically, and for this reason they are usually ignored in elementarytexts6 − this is unfortunate!

6a notable exception is Quantum Physics by Edwin Taylor.

196

Page 198: PHY 416, Quantum Mechanics - SIUE

As a concrete example, consider the “simple” problem of a particle in anuniform force field − e.g., a particle in the gravitational field new earth’ssurface. Then V (y) = mgy. If the particle is confined to a region (y1, y2

boundaries), then the potential energy profile is shown in figure 22.10.fig 10Another example: consider the “free” electron in a metal wire segment

running from x1 to x2. As we’ve already remarked, in the “quasi-free electronmodel”, the electron in the metal are considered to be able to more freely ina finite potential well.

fig 11Now suppose we apply a constant electric field − e.g., from a battery

hooked up to the ends of the wire. Them the potential energy varies linearlyalong the length of the wire (V = −kx ⇒ E = −dV

dx = k = constant). Theelectrons are now in a linearly ramped well.

fig 12Of course, this leads to the conduction of electric current. An under-

standing of the quantum mechanics of this problem leads to the first stepsin understanding the quantum theory of conduction! While obviously veryimportant, gaining a full understanding of this not simple, and today, welimit ourselves to just a few semi-quantitative first remarks about the E < 0eigenfunctions in such a well. As a paradigm for this, take V = gx, where gis a constant. Then, the Schrodinger equation is

− h2

2m

d2ψ(x)

dx2 + gxψ(x) = Eψ(x) (22.34)

We can simplify the appearance of this by defining

ξ ≡(x− E

g

)3

√2mg

h2 (22.35)

the the Schrodinger equation is

d2ψ(ξ)

dξ2 = −ξψ(ξ) (22.36)

This equation looks very harmless, but, in fact, the solution to it cannot bewritten in terms of a finite number of elementary functions! The solution to

197

Page 199: PHY 416, Quantum Mechanics - SIUE

equation (22.34) turns out to be the so-called “Airy Integral”

Ai(ξ) ≡ 1

πlimε→0

∫ ∞0

e−εu cos

(u3

3+ ξu

)du (22.37)

In this course, we won’t get involved with this function. however, it wouldbe nice to know if we could intuit anything about the result without goingthrough the mathematics. As I hope to show you, indeed we can, and ourthought son this will apply much more generally than to just this problem.

22.5 Qualitative Eigenfunctions for Wells

Suppose first that we have a “split-level” well and we want the bound-stateeigenfunction.

fig 13Now, we could solve this exactly, but let’s see what we can get by just

“thinking”.

1. In the internal regions (I and II), ψ is sinusoidal and

k2 =2m(E − V )

h2 =2m

h2 (K.E.) =2m

h2p2

2m=p2

h2 =2π

λ2 (22.38)

Thus, large E − V ⇒ small λ. So, λI < λII.

2. In the internal regions we can write, for some phases ϕi,

ψi(x) = Ai sin (kix+ ϕi) i = I,II (22.39)

Now consider the slope

Si =dψi(x)

dx= kiAi cos (k1x+ ϕi) (22.40)

soSik1

= Ai cos (kix+ ϕi) (22.41)

so

Siki

+ ψ2i = A2

i cos2 (kix+ ϕi) +Ai62 sin2 (kix+ ϕi) = A2i (22.42)

now since kII < kI, and since both S and ψ are continuous across thestep in the well bottom, it follows that AII > AI!

198

Page 200: PHY 416, Quantum Mechanics - SIUE

3. Number of nodes: We know that from our previous experience that ψnshould have n−1 nodes inside the well. So suppose someone asks for thefifth lowest energy level in our split level well. We see in advance that itmust look pretty muck like what is shown in figure 22.14.

fig 14

Now, any shape well can be approximated by a succession of narrow stepwells. Consider for example, the “ramp-bottom” well

fig 15(the dotted line shows an approximation to it). Then, from our consider-

ations, the say, fifth and seventh levels in this well have eigenfunctions thatmust look like figure 22.16.

fig 16Here’s a way of remembering that the amplitude is bigger in the shallower

part of the well. Consider the classical limit in the sense of large n. Theeigenfunction would then have a large number of zig zags. On which side ofthe well, shallow or deep, must ψ have greater amplitude? A classical particlespends more time where it is moving slowly. In the well in figure 22.16, theclassical acceleration would be to the left (F = −dV

dx ). Therefore, the classicalparticle spends more time (and is thus more likely to be found, in a randomlytimed snapshot) on the right side. Hence P (x) = ψ∗(x)ψ(x) must be greateron the right side, therefore, the amplitude of ψ is greater on the right side.Hence, shallower side of well ↔ greater amplitude for ψ. Something of areversal of this logic gives us insight into where classical physics comes from:a narrow wave packet centered on very large n (“classical’) accelerates onmoving from the shallow to the deep end of the well − this is because that“local phase velocity” of any component of the packet is vϕ(k, x) = ω

k(x) , and

k(x) increases for each component as the packet moves to the deep end ofthe well. (If the phase velocity of every component increases as the packetmoves, then the center of the packet is accelerating.) Now, if V = V (x),

classically there is a force F = −dV (x)dx toward the deeper part of the well, so

this looks like the center of the packet is obeying a ∝ F − i.e., this looks likeNewton’s second law a = F

m . Of course, we haven’t shown that our “a” (we

really mean d2〈x〉dt2 ) for that packet is proportional to 1

m and othr details, but,a bit later in our course we hope to show that, indeed, for this situation, as

199

Page 201: PHY 416, Quantum Mechanics - SIUE

well as for many others

md2 〈x〉dt2

= −d 〈V 〉dx

∼ “F” (22.43)

(The general result is called Ehrenfest’s theorem). Now let us return to ourconstant force problem. If this is true in a limited region (x = −a→ x = a,say) as it always would be, then V (x) looks like

fig 17This is exactly what we just looked at! Now, let’s look at the exact Airy

function. It looks like figure 22.18.fig 18

Note that we are plotting against ξ = constant×(x− E

g

). ψ(x) = ψ (ξ(x))

must be unravelled from this. Since we’ve using this problem only as athought spring board, we won’t worry about that.

200

Page 202: PHY 416, Quantum Mechanics - SIUE

Chapter 23

Twenty-third Class

Thursday, November 13, 2008 Exam 2

201

Page 203: PHY 416, Quantum Mechanics - SIUE

Chapter 24

Twenty-Forth Class: The SimpleHarmonic Oscillator

Tuesday, November 18, 2008

24.1 Eigenstate in a Harmonic Oscillator Potential Energy Profile

We consider states for a potential energy profile that is quadratic, i.e.

V (x) =1

2kx2 (24.1)

where k is a constant.fig 1Classically, then, there is a return force

F (x) = −dV (x)

dx= −kx (24.2)

which leads to

x(t) = A cos

(√k

mt+ φ

)(24.3)

We define ω ≡√

km , which is the classical frequency. In quantum mechanics,

the Schrodinger equation for this case is

− h2

2m

d2ψ(x)

dx2 +1

2kx2ψ(x) = Eψ(x) (24.4)

To construct the general state function in this potential energy profile, wewould start by finding the eigenstate through the solution of the Schrodingerequation.

202

Page 204: PHY 416, Quantum Mechanics - SIUE

24.1.1 Intuition

However, before we get involved in any detailed mathematically it is well toask what we expect the eigenfunctions to look like. In our previous consid-erations of “split-level” wells with flat segments, we found rigorously thatthe amplitude is greater and the curvature less in the shallower part. Ourquadratic potential is no composed of rectangular segments, so we cannotrigorously expect similar conclusions to apply to lowest-lying eigenfunctionshere, but we can appeal to expectations for the classical limit − as we dis-cussed, on these grounds, we expect the amplitude to be greater and thecurvature less in the shallower parts of V (x) (compared to E), certainly forhigh n. Also we expect ψn to have (n− 1) nodes, and for there to be leakageof probability onto the classically forbidden region. However, in the classi-cal (large n) limit, the “percentage leakage” should go to zero (since in thatlimit, we must recover classical physics). Let us now think through thatstates “one-by-one”.

1. Parity: Since V (x) is even, all eigenfunctions must have definite parity(even or odd).

2. Ground-State: To minimize curvature (and hence, to minimize the en-ergy), must be a smooth, continuous even function of x with no nodes.thus, for the ground state, the amplitude cannot be greater in the shal-lower parts of V (x). However, that is not a contradiction since this isnot large n and not a flat bottomed split-level well. So, the ground statemust look something like

fig 2

where ±x1 are the classical turning points.

3. First Excited State: To minimize curvature (but be more curved thanthe ground state) must be a smooth continuous odd function of x withone node which must be at x = 0. Thus, again, it is not possible for theamplitude to be always greater in more shallow regions

fig 3

4. Second Excited State: ψ must be even with two nodes, and hence

203

Page 205: PHY 416, Quantum Mechanics - SIUE

it is now geometrically possible for the amplitude to be greater in theshallow parts of the well (see figure 23.4)

5. Further Excited States: In summary, we expect that the eigenfunc-tions should qualitatively look like those shown in figure 23.4

fig 4

The vertical lines indicate that limits of the classically allowed region ineach case. Note that the spacing increases as n does, since E increasesas n does, so the intersection points of E with V move further apart (seefigure 23.5).

fig 5Figure 23.5 shows, of course, the actual eigenfunctions. It is clear that

these look qualitatively as we expect (amplitudes for n = 2, n = 3 bigger inshallower parts of well, etc.). The figure 23.6 shows |ψ100(x)|2, and in it youcan see both the expected amplitude increase and the expected “wavelength”decrease as you move to the shallower parts of the well for the fixed value ofenergy.

fig 6

24.1.2 Mathematical Treatment

Now that we have some intuition, we are ready for the math. We return tothe Schrodinger equation

− h2

2m

d2ψ(x)

dx2 +1

2kx2ψ(x) = Eψ(x) (24.5)

Simplify the algebra, we define two dimensionless variables

ξ ≡√mω

hx (24.6)

K ≡ 2E

hω(24.7)

this leads to energy in units of hω2 . In terms of these, the Schrodinger equationis

d2ψ(ξ)

dξ2 =(ξ2 −K

)ψ(ξ) (24.8)

204

Page 206: PHY 416, Quantum Mechanics - SIUE

which is another of these simple-looking differential equations that is quitehard to solve directly. To approach this, we look at some limits to get someinsight. Consider the limit of very large ξ (very large x). Then equation(23.8) becomes

d2ψ(ξ)

dξ2 ≈ ξ2ψ(ξ) (24.9)

Even this equation cannot be solved exactly, but an approximate solution is

ψ(ξ) = Aeξ2

2 + Be−ξ2

2 (24.10)

To check this, note

d2ψ(ξ)

dξ2 = A(1 + ξ2) e

ξ2

2 − B(1− ξ2) e−

ξ2

2 (24.11)

so, for ξ 1d2ψ(ξ)

dξ2 ≈ Aξ2eξ2

2 + Bξ2e−ξ2

2 = ξ2ψ (24.12)

Now, we want ψ(ξ) to be well-behaved as ξ → ∞, therefore, A = 0, so forlarge ξ

ψ(ξ)→ Be−ξ2

2 (24.13)

Now we really want a solution to equation (23.8). Spurred by our insight in

equation (23.13), let’s factor e−ξ2

2 out of ψ, so we write

ψ(ξ) ≡ h(ξ)e−ξ2

2 (24.14)

Of course we sill don’t know h(ξ). Therefore, let us find the differential

equation it obeys. Now ψ(ξ) must obey equation (23.8). To get d2ψdξ2 we work

through h(ξ)

dξ=

dh

dξe−

ξ2

2 − ξh(ξ)e−ξ2

2 =

(dh(ξ)

dξ− ξh(ξ)

)e−

ξ2

2 (24.15)

d2ψ

dξ2 =

(d2h

dξ2 − 2ξdh

dξ+(ξ2 − 1

)h

)e−

ξ2

2 (24.16)

so equation (23.8) is

d2h(ξ)

dξ2 − 2ξdh(ξ)

dξ+ (K − 1)h(ξ) = 0 (24.17)

205

Page 207: PHY 416, Quantum Mechanics - SIUE

Now this is an equation that is known to mathematicians and physicists.It was first studied by Charles Hermite in the 19thcentury, and is therefore,called “Hermite’s equation”. To solve it, we use the method of power series:We try a power series solution.

h(ξ) = a0 + a1ξ + a2ξ2 + a3ξ

3 + · · · =∞∑j=0

ajξj (24.18)

Our goal is to determine the aj’s. Now equation (23.18) leads to

dh

dξ=

∞∑j=0

jajξj−1 (24.19)

so

2ξdh

dξ= 2

∞∑j=0

jajξj (24.20)

andd2h

dξ2 =∞∑j=0

(j − 1)(j)ajξj−2 =

∞∑j=0

(j + 1)(j + 2)aj+2ξj (24.21)

since the first two terms of the left-hand sum are zero. We put these intoequation (23.17), finding∑

j=0

∞ [(j + 1)(j + 2)aj+2 − 2jaj + (K − 1)aj] ξj = 0 (24.22)

since this must be true for all ξ, the coefficient of each power of ξ is separatelyzero, so

(j + 1)(j + 2)aj+2 − 2jaj + (K − 1)aj = 0 (24.23)

for all j. This leads to

aj+2 =2j + 1−K

(j + 1)(j + 2)aj (24.24)

A relation like this is called a recursion relation. From it, we can express alleven-numbered a’s in terms of a0 and all odd-numbered a’s in terms of a1:

a2 =1−K

2a0, a4 =

5−K12

a2 =(5−K)(1−K)

24a0, · · · (24.25)

a3 =3−K

6a1, a5 =

7−K20

a3 =(7−K)(3−K)

120a1, · · · (24.26)

206

Page 208: PHY 416, Quantum Mechanics - SIUE

Thus, we have two independent series solution of equation (23.17); therefore,the general solution of (23.17) is a linear combination of them

h(ξ) = heven(ξ) + hodd(ξ) (24.27)

heven = a0 + a2ξ2 + a4ξ

4 + · · · (24.28)

hodd = a+ 0ξ + a3ξ3 + a5ξ

5 + · · · (24.29)

These are called Hermite functions. However, notice a problem: consider therecursion relation from equation (23.24) for large j

aj+2 ≈2

jaj (24.30)

This is an algebraic equation with solution

aj ≈C(j2

)!

(24.31)

for any constant C.Now consider the behavior of the Hermite functions atlarge ξ. There, the higher powers of ξ dominate, so

h(ξ)→ C∑ 1(

j2

)!ξj ≈ C

∑ 1

j!ξ2j ≈ Ceξ

2

(24.32)

Thenψ(ξ) = e−

ξ2

2 h(ξ) limξ→∞

⇒ Ceξ2

2 (24.33)

which diverges. Hence, neither Hermite function is a useful solution for us!How do we wiggle out of this problem? This leads to a loophole worthy ofany attorney: the power series must terminate as polynomials!

There is some highest j for which aj 6= 0. Then the recursion relationmust lead to aj+2 = 0! Now, the recursion relation was

aj+2 =2j + 1−K

(j + 1)(j + 2)aj (24.34)

so we require2j + 1−K = 0 (24.35)

207

Page 209: PHY 416, Quantum Mechanics - SIUE

for j = n (highest nonzero j). This leads to

K = 2n+ 1 (24.36)

but, K ≡ 2Ehω , so

2E

hω= 2n+ 1 (24.37)

and by solving for the energy leads to

En =

(n+

1

2

)hω n = 1, 2, 3, · · · (24.38)

so, the allowed energies (energies of the eigenstates) are given by this formula!

24.2 Harmonic Oscillator Potential, Continued

We saw that for the potential energy profile V (x) = 12kx

2. the energy valuesassociated with the eigenstates are given by

En =

(n+

1

2

)hω (24.39)

This is very basic and very important result in quantum mechanics. We seethat the energy levels are equally spaced in units of h times the classicalfrequency, and that the ground state does not have zero energy, but ratherenergy 1

2hω ( “zero-point” energy) (n=0).fig 7

Since so many potentials have minima, which are therefore locally quadratic(by Taylor expansion), “nothing is still” in quantum mechanics is a oft-repeated rule of thumb. Now let us look at our eigenfunctions. The recursionrelation is

aj+2 =2j + 1−K

(j + 1)(j + 2)aj (24.40)

and the series termination condition is

K =2E

hω= 2n+ 1 (24.41)

leading to

aj+2 =2j + 1− 2n− 1

(j + 1)(j + 2)aj = − 2(n− j)

(j + 1)(j + 2)aj (24.42)

We consider successive values of n in turn:

208

Page 210: PHY 416, Quantum Mechanics - SIUE

• n = 0: Then a2 = 0. To kill the series hodd, we choose a1 = 0, so

h0(ξ) = a0 (24.43)

andψ0(ξ) = a0e

− ξ2

2 (24.44)

which meansψ0(x) = a0e

−mω2h x2

(24.45)

We see that the ground state is Gaussian in x (or in ξ). The normalizedeigenfunction is

ψ0(x) =(mωπh

) 14

e−mω2h x

2

(24.46)

since ∫ +∞

−∞emωh x2

dx =

√πh

mω(24.47)

Note that, as we expect, the ground state eigenfunction has no nodes.Note also there is significant penetration into the classically forbiddenregion.

fig 8

• n = 1: We choose a0 = 0 to kill off heven. Then the recursion relationwith n = 1, leads to a3 = 0, so ξ,

h1(ξ) = a1ξ (24.48)

ψ1(ξ) = a1ξe− ξ

2

2 (24.49)

From this point on, it is easier to leave ψ in terms of ξ. We see thatψ1(ξ) is an odd function with one node (at ξ = 0); both of these featureswe expect from our experience.

• n = 2: As you can easily show,

h2(ξ) = a0(1− 2ξ2) (24.50)

ψ2(ξ) = a0(1− 2ξ2) e−

ξ2

2 (24.51)

which is even and has two nodes.

209

Page 211: PHY 416, Quantum Mechanics - SIUE

We see that, for any n, hn(ξ) is an nth order polynomial in ξ with only evenor odd powers of ξ, according to whether n is even or odd. These polynomialsare well-known in mathematics and in physics, and when their normalizationis chosen to be such that the coefficient of the highest poser is 2n, they arecalled Hermite polynomials. Some of these famous polynomials are listedbelow

H0(ξ) = 1

H1(ξ) = 2ξ

H2(ξ) = 4ξ2 − 2

H3(ξ) = 8ξ3 − 12ξ

H4(ξ) = 16ξ4 − 48ξ2 + 12

H5(ξ) = 32ξ5 − 160ξ3 + 120ξ

In terms of these, the ψn(ξ), properly normalized, are

ψn(x) =(mωπh

) 14 1√

2nn!Hn(ξ)e

− ξ2

2 (24.52)

where ξ =√

mωh x. The first five of these are plotted in figure 23.9.

fig 9We see that these look more or less as we expect (smaller amplitude where

E − V is biggest, etc.).

210

Page 212: PHY 416, Quantum Mechanics - SIUE

Chapter 25

Twenty-Fifth Class: FurtherProperties of Harmonic OscillatorEigenfunctions

Thursday, November 20, 2008Recall that the eigenfunctions for the harmonic oscillator potential energy

V (x) = 12kx

2 are, once normalized,

ψn(x) =(mωπh

) 14 1√

2nn!Hn(ξ)e

− ξ2

2 (25.1)

where ξ ≡√

mωh x and ω ≡

√km . As we mentioned, the harmonic oscilla-

tor eigenfunctions, being eigenfunctions of Hermitian Hamiltonian, form anorthonormal set, i.e., ∫ +∞

−∞ψn(ξ)ψm(ξ) dξ = δm,n (25.2)

An interesting mathematical consequence that follows easily is that the Her-mite polynomials themselves obey

1√π2nn!

∫ +∞

−∞Hn(ξ)Hm(ξ)e−ξ

2

dξ = δm,n (25.3)

We say that that the Hermite polynomials are orthogonal “with respect tothe ‘weight function’ e−ξ

2

”.

211

Page 213: PHY 416, Quantum Mechanics - SIUE

25.1 Expansion Postulate for Harmonic Oscillator

We saw quite a while ago that the eigenfunctionsψn(x) = An sin

(nπa x)

forthe infinite square well potential form a complete set for the well boundaryconditions − i.e. any “reasonable” f(x) vanishing at x = 0 and at x = a canbe expanded as

f(x) =∑

anψn(x) (25.4)

this being nothing more than a Fourier sine expansion. This is then allowedsolution of the “initial value problem” for the infinite square well potential:given an arbitrary Ψ(x, t = 0), to find Ψ(x, t), one technique is to expand

Ψ(x, t = 0) =∑n

anψn(x) (25.5)

thenΨ(x, t) =

∑n

anψn(x)e−iEnh t (25.6)

The question is whether a similar result is true for the harmonic oscillatorpotential. You will also recall that we stated, some time ago, the expansionpostulate of quantum mechanics, or at least, a special (but still very general)case of it − namely that:

Postulate: For any continuous potential energy function V (x), the set ofeigenfunctions is a complete set, i.e., and “reasonable” function f(x) obeyingthe same boundary conditions as the eigenfunctions can be expanded as

f(x) =∞∑n=1

anψn(x) (25.7)

Thus, the set of harmonic oscillator eigenfunctionsψn(x) = CnHn(ξ)e

− ξ2

2

, Cn ≡

(mωπh

) 14

, ξ ≡√mω

hx (25.8)

is a complete orthonormal set. Thus, any square-integrable f(x) on (−∞,∞)can be expanded as

f(x) =∞∑n=1

anψn(x) (25.9)

212

Page 214: PHY 416, Quantum Mechanics - SIUE

for this set. Assuming this, we can find the expansion coefficients: Multiplyboth sides of the expansion by ψ∗k and inegrate∫ +∞

−∞ψ∗k(x)f(x) dx =

∞∑n=1

an

∫ +∞

−∞ψ∗k(x)ψn(x) dx∫ +∞

−∞ψ∗kf(x) dx =

∞∑n=1

anδk,n = ak

ak =

∫ +∞

−∞ψ∗k(x)f(x) dx (25.10)

which is to say

ak = Cn

∫ +∞

−∞f(x)e−

mω2h x

2

Hk

(√mω

hx

)dx (25.11)

The condition that f(x) be “reasonable” is the condition that the integralfrom of this equation for ak exists. If F (x) does not diverge more rapidlythen e±

mω4h x

2

as x→ ±∞, we do not need to worry about this.

25.1.1 Mathematical Aside

Since the Hermite polynomials themselves obey the orthogonality realtion

1√π2nn!

∫ +∞

−∞Hn(x)Hm(x)e−x

2

dx = δm,n (25.12)

it should be very plausible to you (and is, indeed true) that any squareintegrable f(x) on (−∞,∞) can be expanded in a Hermite series

f(x) =∑n

anHn(x) (25.13)

with, as you can easily show

an =1

2nn!√π

∫ +∞

−∞e−x

2

Hn(x)f(x) dx (25.14)

such a series expansion is sometimes useful in mathematical physics.

213

Page 215: PHY 416, Quantum Mechanics - SIUE

25.2 Further Important Results

Returning to our wave function expansion, we have

f(x) = Ψ(x, t = 0) =∑n

anψn (25.15)

with

an =

∫ +∞

−∞ψ∗n(x)Ψ(x, t = 0) dx (25.16)

Since this is formally the same as what we had for the infinite square well,the same results follow:

1. ∞∑n=0

|an|2 = 1 (25.17)

(proof same as on page 37 of Griffiths and in our previous notes)

2.

〈H〉 =∞∑n=0

|an|2En (25.18)

3. ⟨√E⟩

=∞∑n=0

|an|2√En (25.19)

and

4.

Ψ(x, t) =∞∑n=0

anψn(x)e−iEnh t (25.20)

where

En =

(n+

1

2

)hω (25.21)

with

an =

∫ +∞

−∞ψ∗n(x)Ψ(x, t = 0) dx (25.22)

This formally solves the initial value problem.

214

Page 216: PHY 416, Quantum Mechanics - SIUE

Example: Suppose that V (x) = 12kx

2 and that

Ψ(x, t = 0) =

√β√π

e−β2(x−b)2

2 , β ≡√mω

h, ω ≡

√k

m(25.23)

(this function is just like the harmonic oscillator ground state, except that itis centered at x = b rather than at x = 0)

fig 1If the energy is measured, what is the probability of finding hω

2 ? 3hω2

What is Ψ(x, t)?This may be a homework problem.

25.3 The Classical Limit in the Harmonic Potential −A First Look

We can make a first pass at trying understand the classical limit by looking,still at eigenfunctions only, but those for very large n. Let us first ask whatwe expect in the classical limit. Classically, the oscillator should spend moretime where it moves more slowly − near the edges of the classically allowedregion. Also, the “leakage trail” into the forbidden region should becomenegligible.

fig 2We note that even for very low n (n = 4), the amplitude of ψn is smaller

near the center x = 0, and indeed, this is what we expect from our observa-tions last week in class. At large n, this behavior persists − the situation for“low intermediate” n (n = 100) is shown in figure 25.3.

fig 3Let us calculate the classical probability distribution. Let the classical

period τ = 2πω . Then

Pclassical(x)∆x =∆t

τ=

ω

2∆x

v(x)(25.24)

where v is the classical speed. Now,

xclassical = A cos(ωt) (25.25)

215

Page 217: PHY 416, Quantum Mechanics - SIUE

where we ignore the phase, and where A =√

2Emω2 where E is the energy.

Thus,

v(x) = −ωA sin(ωt)

= −ω[√

1− cos2(ωt)]

v(x) = ωA

(1− x2

A2

) 12

(25.26)

which gives us v as a function of x. Thus,

Pclassical∆x =1

πA

1(1− x2

A2

) 12

∆x (25.27)

Pclassical =1

πA

1(1− x2

A2

) 12

(25.28)

and this is plotted as the dashed line in figure 25.3. So, in the limit oflarge n, the agreements of the probability distributions is very good. Now,what about leakage into the forbidden region? We note that the exponential

factor e−ξ2

2 is the same for all n; however, the classically allowed region getwider (why?) as n increases, so the probability leakage fraction does get less.Of course, we would like to see a probability lump moving back and forthharmonically in the classical limit, and, of course, we cannot get that justconsidering solely eigenfunctions (why not?). So, we need to understand themotion of a peaked wave packet in a harmonic oscillator potential.

25.4 Behavior of Harmonic Oscillator Superposition States

As a start toward this, we can consider the behavior over time of an equalsuperposition o the ground and first excited states

Ψ(x, t = 0) =1√2

[ψ0(x) + ψ1(x)] (25.29)

Notice that these are the two must non-classical states! However, as you willshow in the homework, a plot of the probability density as a function of timeoscillates as shown in figure 25.4.

216

Page 218: PHY 416, Quantum Mechanics - SIUE

fig 4We note, in fact, that any packet or superposition state in a harmonic

oscillator potential (only) is periodic (and will therefore be in the same placeand look the same after one or an integral number of periods). To see this,note that, by the completeness theorem, any initial state Ψ(x, t = 0) = f(x)can be expanded as

Ψ(x, t = 0) =∞∑n=0

cnψn(x) (25.30)

so

Ψ(x, t) =∞∑n=0

cnψn(x)e−iEnh t (25.31)

Then the probability density is

P (x, t) = |Ψ(x, t)|2 =∞∑k=0

∞∑n=0

c∗kcnψ∗k(x)ψn(x)e−(k−n)ωt (25.32)

where ω ≡√

km . Each term in this sum oscillates in time at an integral

multiple (k − n) of the classical frequency ω. Hence the whole series isperiodic in time at the classical frequency. Of course, this doesn’t meanthat the superposition state moves in simple harmonic motion. So, we mustlook into this more deeply.

25.4.1 Motion of Gaussian Wave Packet in Harmonic Potential

We consider starting with a packet state initially centered at the point x = 0and having 〈p〉 positive. A good choice would be a Gaussian in momentum

φ(p) ∼ e−p−p0

2σ2p (25.33)

where 〈p〉 = p0. As you know, a Gaussian φ(p) leads to a Gaussian ψ(x) inposition; so

Ψ(x, t = 0) =1√√πσx

e− x2

2σ2x e

ip0h x (25.34)

fig 5

217

Page 219: PHY 416, Quantum Mechanics - SIUE

This packet has 〈p〉 = p0. If it were in free space (i.e. V = 0), such aspacket would move to positive x continuously, always spreading. but, thepacket is not in free space. To find out what happens over the course of time,we use the completeness theorem:

Ψ(x, t) =∞∑n=0

cnψn(x)e−iEnh t (25.35)

where

cn =

∫ +∞

−∞ψ∗n(x

′)Ψ(x′, 0) dx′ (25.36)

Recall that we rewrote this expansion in the form

Ψ(x, t) =

∫ +∞

−∞Ψ(x′, t = 0)K(x′, x, t) dx′ (25.37)

where K is the “propagator” is iven by

K(x′, x, t) =∞∑n=0

ψ∗n(x′)ψn(x)e−

iEnh t (25.38)

For the harmonic potential, as we know

En =

(n+

1

2

)hω

ψn(x) =(mωπh

) 14 1√

2nn!Hn

(√mω

hx

)e−

mω2h x

2

thus, in this case. the propagator is very complicated. However, this is one ofthe very few cases in which it has been evaluated in closed form! The result,after much work, is for the harmonic oscillator,

K(x′, x, t) =

√mω

2iπh sin(ωt)eimωh

(x2+x′2) cos(ωt)−2xx′2 sin(ωt) (25.39)

which itself is not at all simple. By solving for Ψ(x, t), it has been found(again, after much work) that, for an initial state as in equation (25.34),

|Ψ(x, t)|2 =1√π∆(t)

e− [x− p0

mω sin(ωt)]∆2(t) (25.40)

218

Page 220: PHY 416, Quantum Mechanics - SIUE

where

∆(t) ≡

√σ2x cos2(ωt) +

(h

mωσx

)2

sin2(ωt) (25.41)

This result is remarkably simple: We see that the probability packet remainsGaussian in shape, although the center moves, at time t it is apparentlycentered at coordinate xc given by

xc −p0

mωsin(ωt) = 0 (25.42)

xc =p0

mωsin(ωt) (25.43)

Thus, we see that the center of the packet oscillates harmonically in positionat exactly the classical frequency! Let us attempt a more detailed comparisonwith a classical mass-spring oscillator: Classically, the energy and amplitudeare related by

E =1

2kA2 ⇒ A =

√2E

mω2 (25.44)

using the classical relation pmax =√

2Em , we have

Aclassical =p

mω(25.45)

therefore,

x(t) =pmax

mωsin(ωt) (25.46)

In the quantum expression, we recall that p0 = 〈p〉, so

xclass,quantum =〈p〉mω

sin(ωt) (25.47)

We see that close correspondence. Note also that the width of the packet∆(t) also oscillates in time. As yo can easily show, the frequency for this istwice the classical frequency. Thus, the packet itself moves back and forth, itbreathes!. An interesting special case occurs when σx is chosen for the initial

packet as σx =√

hmω ; then, the packet width, ∆(t), is independent of time.

This special packet, called a “coherent state” oscillates rigidly back and forthat frequency ω much like a classical particle.

219

Page 221: PHY 416, Quantum Mechanics - SIUE

25.5 The Time Dependence of Expectation Values

We will see that an understanding of how expectation values change in timewill greatly help our understanding of how classical physics comes from quan-tum mechanics. We are interested in the general case −i.e., not just the caseof a harmonic potential or “particle in a box”, etc. With this in mind, we

now consider the time dependence of⟨Q⟩

, where Q is any operator (x, p, L,

etc.) We keep in mind that⟨Q⟩

may have derivatives with respect to x (e.g.

p = hi∂∂x), In any case, Q acts on functions of x to yield ther functions of x.

Now ⟨Q⟩t

=

∫ +∞

−∞Ψ∗(x, t)QΨ(x, t) dx (25.48)

We seek ddt

⟨Q⟩

. Then

d

dt

⟨Q⟩

=d

dt

∫ +∞

−∞Ψ∗(x, t)QΨ dx

=

∫ +∞

−∞

(∂Ψ∗

∂tQΨ + Ψ∗Q

∂Ψ

∂t

)dx

Now, the Schrodinger equation is

HΨ = ih∂Ψ

∂t⇒ ∂Ψ

∂t=

1

ihHΨ = − i

hHΨ (25.49)

and its conjugate is

HΨ∗ = −ih∂Ψ∗

∂t⇒ ∂Ψ∗

∂t=i

hHΨ∗ (25.50)

sod

dt

⟨Q⟩

=i

h

∫ +∞

−∞

(HΨ

)∗QΨ dx (25.51)

Now, H is a Hermitian operator. Recall that, for any Hermitian operator, O,by definition, for any two allowed state functions Ψ1 and Ψ2,∫ +∞

−∞Ψ∗1

[OΨ2

]dx =

∫ +∞

−∞

[OΨ1

]∗Ψ2 dx (25.52)

220

Page 222: PHY 416, Quantum Mechanics - SIUE

Thus, ∫ +∞

−∞

(HΨ

)∗ (QΨ)dx =

∫ +∞

−∞Ψ∗H

(QΨ)dx (25.53)

Thus, equation (25.51) is

d

dt

⟨Q⟩

=i

h

∫ +∞

−∞Ψ∗[HQ− QH

]Ψ dx (25.54)

Now the combination of operators AB − BA is very common in quantummechanics; it is denoted as [

A, B]≡ AB − BA (25.55)

and given a special name − it is called “the commutator of A and B”. So

d

dt

⟨Q⟩

=i

h

∫ +∞

−∞Ψ∗[H, Q

]Ψ dx =

i

h

⟨[H, Q

]⟩(25.56)

d

dt

⟨Q⟩

=i

h

⟨[H, Q

]⟩(25.57)

As we will shortly get at least a first idea of this is a very important relationin quantum mechanics.

Note: In rare cases, an operator may also have an explicit time depen-dence. An example might be a harmonic potential energy in which the “springconstant” (and hence the classical frequency) changes in time − then H for

this system changes in time (H = p2

2m + 12k(t)x2 = H(t)). In such cases, the

time derivative of the expectation value is modified to

d

dt=∂Q

∂t+⟨[H, Q

]⟩(25.58)

In this course, we do not deal with such cases. We now reap our first resultfrom this theorem: Conservation of Energy Let Q be the Hamiltonian H.Then [

H, Q]

=[H, H

]= HH − HH = 0 (25.59)

Thend

dt

⟨H⟩

= 0 ⇒ d

dt〈E〉 = 0 (25.60)

221

Page 223: PHY 416, Quantum Mechanics - SIUE

Thus, we have derived conservation of energy of energy as a statistical result.Relation between x and px We now consider the very fundamental

commutator:[x, px] (25.61)

To find out what this is, we operate with it is a function f(x):

[x, px] f(x) =

= xpxf(x)− pxxf(x)

= xh

i

∂f(x)

∂x− h

i

∂x[xf(x)]

= xh

i

∂f(x)

∂x− h

i

(f(x) + x

df

dx

)=

h

ix∂f

∂x− h

if(x)− h

ixdf

dx= ihf(x)

since this is true for any f(x),

[x, px] = ih (25.62)

As we will see later, this result is very fundamental in quantum mechanics.Relation between 〈x〉 and 〈px〉: We had, for the time development of

expectation values,

d

dt〈Q〉 =

i

h

∫ +∞

−∞Ψ∗[H, Q

]Ψ dx ≡ i

h

⟨[H, Q

]⟩(25.63)

Suppose Q = x. Now [H, x

]=

[(p2x

2m+ V (x)

), x

]=

1

2m

[p2x, x]

=1

2m[pxpx, x]

Now, as you will show in the homework, for any three operators A, B, andC, [

AB, C]

= A[B, C

]+[A, C

]B

222

Page 224: PHY 416, Quantum Mechanics - SIUE

also, [A, BC

]=[A, B

]C + B

[A, C

]so [

p2x, x]

= px [px, x] + [px, x] px (25.64)

now [A, B

]= −

[B, A

][p2x, x]

= −ihpx − ihpx =2h

ipx (25.65)

so

md

dt〈x〉 = 〈px〉 (25.66)

which shows that the classical connection between momentum and positionis true, on average, even though both can’t be known exactly together!

Newton’s Second Law In classical mechanics, this is

dpxdt

= −∂V∂x

(25.67)

So we consider ddt 〈px〉. This is

d

dt〈px〉 =

i

h

⟨[H, px

]⟩(25.68)

Now, [H, px

]=

[p2

2m+ V, px

]=

1

2m

[p2x, px

]+[V , px

](25.69)

now [p2x, px

]= [pxpx, px] = 0 (25.70)

To find [V (x), px], consider

[V (x), px]ψ(x) = V (x)h

i

∂Ψ

∂x− h

i

∂VΨ

∂x

=h

i

[V∂Ψ

∂x− ∂V

∂xΨ− V ∂Ψ

∂x

]= ih

∂V

∂xΨ

so

[V (x), px] = ih∂V

∂x(25.71)

223

Page 225: PHY 416, Quantum Mechanics - SIUE

Thus, [H, px

]= ih

∂V

∂x(25.72)

Henced

dt〈px〉 = −

⟨dV (x)

dx

⟩(25.73)

Equations (25.66) and (25.73) together are called Ehrenfest’s theorem (1927).Note that they are true for any continuous V (x) and for any superposition(or eigen-) state. We note, of course, that equation (25.73) looks a lot likeNewton’s second law for expectation values. Thus, it looks like our longsought general connection between quantum and classical physics. It is notquite that. For “Newton’s Second Law for the expectation values”, we wouldlike

d

dt〈px〉 = − d

dxV (〈x〉) (25.74)

since

− d

dxV (〈x〉) = F (〈x〉) (25.75)

but, equations (25.73) and (25.74) are generally not the same since, in general,⟨dV (x)

dx

⟩6= d

d 〈x〉V (〈x〉) (25.76)

or〈F (x)〉 6= F (〈x〉) (25.77)

Proof: ⟨dV (x)

dx

⟩=

∫ +∞

−∞Ψ∗(x, t)

∂V

∂xΨ(x, t) dx (25.78)

d

dxV (〈x〉) =

d

dx

V

(∫ +∞

−∞Ψ∗(x, t)xΨ(x, t) dx

)(25.79)

and these two are not equal in general. Of course, this somewhat deepens (ormuddies, depending on your point of view) to gap (or connection) betweenquantum and classical physics.

224

Page 226: PHY 416, Quantum Mechanics - SIUE

25.6 When Does Classical Mechanics Apply?

When, then does classical mechanics apply? Must the system be macro-scopic? The answer is no − for example, motion of electrons in an acceleratorbeam are well described by the classical law

med2~x

dt2= q ~E + q~V × ~B (25.80)

in accelerator guide fields. One key is that V (x) is a “slowly varying” (com-pared with the mean deBroglie wavelength in the packet under consideration)

function of x. To see this, we expand f(x) = −dV (x)dx around x = 〈x〉:

F (x) = F (〈x〉) + (x− 〈x〉) dF (〈x〉)dx

+(x− 〈x〉)2

2!

d2F (〈x〉)dx2 + · · · (25.81)

If (x− 〈x〉) dF (〈x〉)dx F (〈x〉) for all (x− 〈x〉) occupied by the packet at all

times t under consideration, then we can drop all terms but the first and

F (x) ≈ F (〈x〉) ⇒ dV

dx≈ dV (〈x〉)

d 〈x〉(25.82)

Under these conditions,

md2 〈x〉dt2

→ −∂V (〈x〉)∂ 〈x〉

(25.83)

25.6.1 Exceptions

Consider the harmonic potential V (x) = 12kx

2. Then

dV

dx= kx

⟨dV

dx

⟩= 〈kx〉 = k 〈x〉 (25.84)

Then, Newton’s second law can be written as

md2 〈x〉dt2

= −k 〈x〉 (25.85)

Which has as solution

〈x〉 = A cos

(√k

mt+ φ

)(25.86)

225

Page 227: PHY 416, Quantum Mechanics - SIUE

Note that this is true for any state for any value of k (which controls howfast V varies with x). Thus, for the harmonic potential, 〈x〉 always obeysequation (25.85), which leads to “classical behavior”.

Question: What about for an eigenstate in this potential? is equation(25.86) obeyed? Is equation (25.85) obeyed?

As you can show, 〈x〉 always obeys a classical equation of motion of V (x) =Cxn as long as n is a positive integer greater than 2.

Question: Why doesn’t this work if n = 1?

25.7 Another Example of an Operator − Angular Momentum

In classical physics, the orbital angular momentum of a particle around anorigin is

~L = ~r × ~p (25.87)

In cartesian coordinates, this is

Lx = ypz − zpy (25.88)

Ly = zpx − xpz (25.89)

Lz = xpy − ypx (25.90)

The normal procedure is now the following: For every dynamical variableof classical physics, we have an operator in quantum mechanics constructedaccording to the prescription px → h

i∂∂x , etc. Thus, in quantum mechanics,

we have

Lx = −ih(y∂

∂z− z ∂

∂y

)(25.91)

Ly = −ih(z∂

∂x− x ∂

∂z

)(25.92)

Lz = −ih(x∂

∂y− y ∂

∂x

)(25.93)

(25.94)

As we will see, these three operators are very important. However, in nature,quantum mechanics does not “come from” classical physics, rather, we believethat it is somewhat the other way around. Thus, really, we believe that the

226

Page 228: PHY 416, Quantum Mechanics - SIUE

classical quantities (“momentum”, “position”, “angular momentum”, etc.)do not exist as quantities except in a certain limit. Rather, there are onlyoperators and states, and in the “classical limit” certain superposition statesbehave as if certain quantities are conserved.

227

Page 229: PHY 416, Quantum Mechanics - SIUE

Chapter 26

Twenty-Sixth Class: QuantumMechanics and the Hydrogen Atom

Tuesday, December 2, 2008

26.1 An Introduction

We ow begin application of quantum mechanics to atomic physics − specifi-cally, to the simplest atom − that of hydrogen. To prepare, let us raise threequestions:

1. What determines the size of an atom? What sets the scale? Classicalorbits (e.g., planets around the sun) can occur with any radius. Is thistrue in quantum mechanics?

2. Why is the atom stable? Why does it not decay, as predicted by classicalmechanics?

3. Associated with atoms are their atomic spectra. For example, for hy-drogen, the Balmer series wavelengths are well-known. But, why shouldatom emit photons at all? Is it possible to derive, from “first principles”,the Balmer formula for the wavelengths?

These are excellent questions. To refresh our memories, let us recall somefacets of the semiclassical Bohr model. In his original model, Bohr assumedcircular, classical orbits. He also assumed, ad-hoc, that the orbital angularmomentum of the electron is restricted to values

L = nh, n = 1, 2, 3, · · · (26.1)

228

Page 230: PHY 416, Quantum Mechanics - SIUE

This led to “results”:

1. That the ground state orbit has radius

a0 = 4πε0h2

mee2 ≈ 0.529× 10−10 m (26.2)

(“Bohr radius”)

2. That the energy of the ground state is

E1 = −(

1

4πε0

)2mee

4

2h2 = −13.6 eV (26.3)

in agreement with the experimental value of the ionization energy.

3. That the energy of state n is

En =E1

n2 (26.4)

so, En ∝ 1n2 , which is the assumption of photon emission accompanying

“transitions” between energy levels, reproduces the Balmer series for-mula. Still, due to the ad-hoc nature of its assumption, the Bohr modelis unsatisfactory.

Now we see what the Schrodinger equation quantum mechanics has to say.We consider an electron, still not in fully objective reality, but in the Coulombfield of the proton− the latter considered as a classical object that is infinitelyheavy and sets stationary at the origin. figure 26.1 shows one possible pointthat the electron could materialize at upon position measurement.

fig 1Thus, the electron has a state function that depends on all three of its

coordinates and the time

Ψ(z, y, z, t) ≡ Ψ(~r, t) (26.5)

As we know, the time dependent Schrodinger equationfor this is

ih∂Ψ(~r, t)

∂t= − h2

2m∇2Ψ(~r, t) + V (~r)Ψ~r, t) (26.6)

229

Page 231: PHY 416, Quantum Mechanics - SIUE

We look for the normal-mode or standing wave or eigenstates:

Ψ(~r, t) = ψ(~r)e−iEh t (26.7)

where

− h2

2m∇2ψ(~r) + V (~r)ψ(~r) = Eψ(~r) (26.8)

Here, V (x, y, z) is the Coulomb potential due to the proton at each “potentialposition” (x, y, x) of the electron

V (~r) = V (x, y, z) = − e2

4πε0√x2 + y2 + z2

= − e2

4πε0r(26.9)

Clearly, spherical coordinates are more convenient for us here than the Carte-sian coordinates. So, we re-express

ψ(x, y, z) = ψ(r, θ, φ) (26.10)

Now we need the Schrodinger equation (which we have in Cartesian coordi-nates) in spherical coordinates. We need to transform

∇2 ≡ ∂2

∂x2 +∂2

∂y2 +∂2

∂z2 (26.11)

to spherical coordinates. This is done in multivariable calculus, and the resultis

∇2 =1

r2

∂r

(r2 ∂

∂r

)+

1

r2 sin θ

∂θ

(sin θ

∂θ

)+

1

r2 sin θ

∂2

∂φ2 (26.12)

With this, the time independent Schrodinger equationis1

− h2

2m

∂2ψ(r, θ, φ)

∂r2 +2

r

∂ψ(r, θ, φ)

∂r+

1

r2 sin θ

∂θ

(sin θ

∂ψ(r, θ, φ)

∂θ

)

+1

r2 sinθ∂2ψ(r, θ, φ)

∂φ2

− e2

4πε0rψ(r, θ, φ) = Eψ(r, θ, φ) (26.13)

We seek solutions of this monstrous-looking equation.1where I’ve expanded out the term 1

r2∂∂r

`r2 ∂

∂r

´using the product rule for derivatives.

230

Page 232: PHY 416, Quantum Mechanics - SIUE

26.2 Spherically Symmetric Solutions

Before getting involved in the more complicated business of trying to findrather general solution of equation (26.13), let us first see if we can somehowfind (or guess) a few simple solutions. For example, let’s see if we can findany solutions for which ψ depends only or r (and not on θ or φ).2. Then,equation (26.13) becomes

d2ψ(r)

dr2 +2

r

dψ(r)

dr+

2mke2

h2rψ(r) = −2m

h2 Eψ(r) (26.14)

where k ≡ 14πε0

. This ordinary differential equation still looks formidable.Can we guess a solution? We know that the probability that the electron willmaterialize at points for r →∞ must go to zero. So, we look for a ψ(r) thatdecreases when r is large and increasing. This still seems like a formidable“guessing problem”. However, notice an interesting feature of the differentialequation. Let us rewrite it as

d2ψ(r)

dr2 +2

r

dψ(r)

dr= −2m

h2

(E +

ke2

r

)ψ(r) (26.15)

Now, the two pieces that look troublesome (one on the left and one on theright) both have 1

r factors. How nice it would be if they were equal − thenwe could just cancel them! So, we can’t resist trying − just trying

2

r

dψ(r)

dr= −2m

h2ke2

rψ(r) (26.16)

Thendψ(r)

dr= −mke

2

h2 ψ(r) (26.17)

But this we can solve − the solution is

ψ(r) = Ae−br (26.18)

where b = ke2mh2 . Note incidentally that this “play function” at least agrees

with our observation that ψ(r) decreases at large r. Of course, if life were so2Such solutions would then be spherically symmetric.

231

Page 233: PHY 416, Quantum Mechanics - SIUE

simple, then this same function,ψ(r) = Ae−br, would also have the solve therest of equation (26.14), namely,

d2ψ(r)

dr2 = −2m

h2 Eψ(r) (26.19)

This looks a bit familiar. Is it an oscillator equation? No − since E < 0(why?) the right hand side is positive, not negative. That means that tissolution is also a decaying exponential

ψ(r) = Be−( 2mE

h2 )1/2r (26.20)

so

ψ(r) = Ae+√−2mE

h2 r+ Be

−√−2mE

h2 r(26.21)

but we must have A = 0 (why?). Of course, now we have two differentfunctions ψ(r), each solving an additive piece of the differential equation.Let’s just try to equate them − is this possible? It is if

e−(−2mE

h2 )1/2r = e−

ke2m

h2 r (26.22)(−2mE

h2

)1/2

=ke2m

h2 (26.23)

−2mE

h2 =

(ke2m

)2

h4 (26.24)

E = −k2me4

2h2 (26.25)

which, amazingly, is exactly the ground state Bohr model energy (and it isequal to −13.6 eV)! So let us look at our solution:

ψ(r) = e−br (26.26)

where b = ke2mh2 This is a simple exponential fall off with “1

e” distance of 1b .

fig 2But,

1

b=

1

k

h2

me2 = 4πε0h2

mee2 (26.27)

232

Page 234: PHY 416, Quantum Mechanics - SIUE

which is nothing but the Bohr radius a0!

a0 = 4πε0h2

mee2 (26.28)

So our solution isψ(r) = Be−

ra0 (26.29)

which we expect top be the ground-state eigenfunction. This turns out to becorrect.

26.3 Other Spherically Symmetric Solutions

As you can show for yourself, another spherically symmetric solution to equa-tion (26.15) is

ψ2(r) =

(2− r

a0

)e−

r2a0 (26.30)

fig 3If you go through the algebra, you will find tat his is a solution only if

E = E2 =E1

4= −13.6 eV

4(26.31)

again agreeing with the Bohr theory. In fact, there is an entire series ofsolutions to equation (26.15)

ψ1(r), ψ2(r), ψ3(r), ψ4(r), · · ·

ψ3(r) =

(1− 2r

3a0+

2r2

27a20

)e−

r3a0 (26.32)

fig 4with energies

En =−13.6 eV

n2 (26.33)

This is a fantastic result! It shows that, although the Bohr model is reallywrong, the Bohr formula for the energies

En = −(mk2e4

2h2

)1

n2 (26.34)

233

Page 235: PHY 416, Quantum Mechanics - SIUE

is apparently correct in Schrodinger quantum mechanics! From this point ofview, it is interesting to contemplate the far from obvious fact that equation(26.15) is apparently equivalent to the Balmer formula.

Of course, there is one aspect of the procedure we used above to findsolutions that is, perhaps, a bit unsatisfying − we guessed we guessed thesolutions beyond ψ1. This raises the question − is there a systematic methodof finding these and also non-spherical symmetric eigenfunctions? The answeris yes − and in the following we use methods of differential equations to showthis and to find them.

26.4 The Road to More General Hydrogen Atom Eigenfunctions

The hydrogen atom still holds more secrets. Thus, we return to equation(26.13)

1

r2

∂r

(r2∂ψ(r, θ, φ)

∂r

)+

1

r2 sin θ

∂θ

sin θ

(∂ψ(r, θ, φ)

∂θ

)+

1

r2 sin2 θ

∂ψ(r, θ, φ)

∂φ+

2m

h2

[e2

4πε0r− E

]ψ(r, θ, φ) = 0 (26.35)

We try separation of variables: Search for solutions of the form

ψ(r, θ, φ) = R(r)Θ(θ)Φ(φ) (26.36)

we plug this into the Schrodinger equation. This yields

− h2

2m

[1

r2

∂r

(r2∂ [RΘΦ]

∂r

)+

1

r2 sin θ

∂θ

(sin θ

∂ [RΘΦ]

∂θ

)+

1

r2 sin2 θ

∂2 [RΘΦ]

∂φ2

]+ V (r)RΘΦ = ERΘΦ (26.37)

Now,

∂ [RΘΦ]

∂r= ΘΦ

dR

dr(26.38)

∂ [RΘΦ]

∂θ= RΦ

dθ(26.39)

∂2 [RΘΦ]

∂φ2 = RΘd2Φ

dφ2 (26.40)

234

Page 236: PHY 416, Quantum Mechanics - SIUE

so the Schrodinger equation becomes

− h2

2m

[ΘΦ

r2

d

dr

(r2dR

dr

)+

r2 sin θ

d

(sin θ

)+

r2 sin2 θ

d2Φ

dφ2

]+ V (r)RΘΦ = ERΘΦ (26.41)

Now we divide by RΘΦ and multiply by −2µr2 sinθ (from now on, we let µstand for me)

1

Φ

d2Φ

dφ2 =sin2 θ

R

d

dr

(r2dR

dr

)− sin θ

Θ

d

(sin θ

)−

h2 r2 sin2 θ [E − V (r)] (26.42)

In spite of the complication, the above echoes a by now familiar situation −the left hand side depends only on one variable (φ while the right hand sidedoes not depend on this variable at all. Therefore, each side must equal aconstant, which for later connivence, we denote as “−m2

`”3. (At this stage,

this is completely general − so far “−m2`” can be any complex number). So

we now have two simpler equations.

1

Φ(φ)

d2Φ

dφ2 = −m2` (26.43)

and

− 1

R

d

dr

(r2dR

dr

)=

1

Θ sin θ

d

(sin θ

)− 2µ

h2 r2 [E − V (r)] =

− m2`

sin2 θ(26.44)

which is

1

R

d

dr

(r2dR

dr

)+

2µr2

h2 [E − V (r)] =m2`

sin2 θ− 1

Θ sin θ

d

(sin θ

)(26.45)

Again, the same sort of thing has happened − one side depends only on onevariable (θ) and the other only on another (r). So again, each side is equal to

3The notation for the constant is conventional; the reason for the “`” subscript will become apparent in thesubsequent lectures.

235

Page 237: PHY 416, Quantum Mechanics - SIUE

a constant, call it “α”. So now we have three ordinary differential equations,each of which is much simpler than the original partial differential equation:

1

r2

d

dr

(r2dR

dr

)+

h2 [E − V (r)] = αR

r2 (26.46)

− 1

sin θ

d

(sin θ

)+m2`Θ

sin2 θ= αΘ (26.47)

d2Φ

dφ2 = −m2`Φ (26.48)

In these equations lie locked the secrets of the hydrogen atom!

26.5 Hydrogen Atom Eigenstate Functions

We assumed separation of variables for the time independent Schrodingerequation

ψ(r, θφ) = R(r)Θ(θ),Φ(φ) (26.49)

and broke the time independent Schrodinger equationup into three ordinarydifferential equations:

1

r2

d

dr

(r2dR

dr

)+

h2 [E − V (r)] = αR

r2

− 1

sin θ

d

(sin θ

)+m2`Θ

sin2 θ= αΘ

d2Φ

dφ2 = −m2`Φ

While we have no guarantee yet that we can cover all solutions of the timeindependent Schrodinger equationby this procedure, we proceed anyway. Theφ-equation is easiest; it is an “oscillator equation”, with general solutionexpressible as

Φ(φ) = A cos (m`φ) + B sin (m`φ) (26.50)

or asΦ(φ) = Ceim`φ +De−im`φ (26.51)

236

Page 238: PHY 416, Quantum Mechanics - SIUE

To remove any lingering doubt as to the equivalence of these solutions (bearin mind that A,B, C,D are all allowed to range over all complex numbers),we expand that last form by Euler’s theorem:

Φ(φ) = C [cos (m`φ) + i sin (m`φ)] +D [cos (m`φ)− i sin (m`φ)]

= (C +D) cos (m`φ) + i (C − D) sin (m`φ)

where we define (C +D) to be A and i (C − D) to be B. We can allow forboth terms by allowing m` to be either positive or negative, so we take

Φ(φ) = Ceim`φ (26.52)

26.5.1 Boundary Conditions on Φ(φ)

This is not all with this equation, however. Since changing an azimuthalangle by 2π brings one back to the same point in space, we should require

Φ (φ+ 2π) = Φ (φ) (26.53)

This means

eim`(φ+2π) = eim`φ

e−m`φe2πim` = eim`φ

e2πim` = 1

cos (2πm`) + i sin (2πm`) = 1 + 0i

cos (2πm`) = 1

sin (2πm`) = 0

m` = 0,±1,±2,±3, · · · (26.54)

We see that m` is restricted to be zero or a positive or negative integer, andthat

Φm`(φ) = e−m`φ (26.55)

No other values of m` produce acceptable solutions.

237

Page 239: PHY 416, Quantum Mechanics - SIUE

26.6 The Θ Equation

We have

− 1

sin θ

d

(sin θ

)+m2`Θ

sin2 θ= αΘ (26.56)

This is an equation that is well known to physicists and mathematicians. Itcvan be solved by standard power series methods4; as with the harmonic os-cillator, the solutions that do blow up at θ = 0 and/or θ = π are polynomialstimes well behaved functions of θ; these exist for integer values of of twoindices (`,m`) where α = `(`+ 1)

Θ`,m`(θ) = APm

` (cos θ) (26.57)

where Pm` (cos θ) is the “associated Legendre function”, is

Pm` (cos θ) ≡

(1− cos2 θ

)|m`|/2(d

dx

)|m`|P`(cos θ) (26.58)

where P`(cos θ) is the `th Legendre polynomial, given by

P`(x) =1

2``!

(d

dx

)` (x2 − 1

)`(26.59)

where the first few Legendre polynomials are

P0(x) = 1

P1(x) = x

P2(x) =1

2

(3x2 − 1

)P3(x) =

1

2

(5x3 − 3x

)P4(x) =

1

8

(35x4 − 30x2 + 3

)...

Some mathematical properties they obey: The Legendre polynomials areorthogonal on the interval [−1, 1]:∫ 1

−1Pm(x)Pn(x) dx =

0 m 6= n

22n+1 m = n

(26.60)

4see any decent differential equations book.

238

Page 240: PHY 416, Quantum Mechanics - SIUE

They also form a complete set on this interval: Any “arbitrary” function f(x)on [−1, 1] can be expressed as

f(x) =∞∑n=0

anPn(x) (26.61)

where, from the orthogonality relation above,

an =

(n+

1

2

)∫ 1

−1f(x)Pn(x) dx (26.62)

You can verify that equation (26.57)is a solution of equation (26.56) by simply“plugging it in”. Note that for equation (26.59) to exist, ` has to be positiveinteger. Further, if |m`| > `, |m`| ≤ `. Thus acceptable solutions exist onlyfor

` = 0, 1, 2, 3, 4, · · · (26.63)

m` = −`,−`+ 1,−`+ 2, · · · ,−1, 0, 1, · · · , `− 1, ` (26.64)

Now, as a second order differential equation (26.56) has two linearly indepen-dent solutions for any ` and any m`, integer or not; the point is that thesediverge at θ = 0 and/or θ = π unless equation (26.63) and equation (26.64)are satisfied.

26.7 Spherical Harmonics

Putting these together, we define (dropping the subscript on m`)

Y m` (θ, φ) = NΘ(θ)Φ(φ) (26.65)

where N is normalization constant. The determine N , we note that thevolume element in spherical coordinates is d3~r = r2 sin θ dr dθ dφ; thus, tonormalize Y m

` , we require∫ 2π

φ=0

∫ π

θ=0|Y m` |

2 sin θ dθ dφ = 1 (26.66)

This works out to be

Y m` (θ, φ) = ε

√(2`+ 1)

(`− |m|)!(`+ |m|)!

eimφPm` (cos θ) (26.67)

239

Page 241: PHY 416, Quantum Mechanics - SIUE

where ε = (−1)m for m ≥ 0, ε = 1 for m ≤ 0. These are orthogonal; theyare listed (for low `, m) in the following tables. The orthogonality of thespherical harmonics requires that∫ 2π

0

∫ π

0[Y m` (θ, φ)]∗

[Y m′

`′ (θ, φ)]

sin θ dθ dφ = δ`,`′δm,m′ (26.68)

Y 00 =

(1

)1/2

Y 10 =

(3

)1/2

cos θ

Y ±11 = ∓

(3

)1/2

sin θe±iφ

Y 02 =

(5

16π

)1/2 (3 cos2 θ − 1

)Y ±1

2 = ∓(

15

)1/2

sin θ cos θe±iθ

Y ±22 =

(15

32π

)1/2

sin2 θe±2iφ

Y 30 =

(7

16π

)1/2 (5 cos3 θ − 3 cos θ

)Y ±1

3 = ∓(

21

64π

)1/2

sin θ(5 cos2 θ − 1

)e±iπ

Y ±23 =

(105

32π

)1/2

sin2 θ cos θe±2iφ

Y ±33 = ∓

(35

64π

)1/2

sin3 θe±3iφ

Essentially due to their orthogonality,the spherical harmonics are also com-plete − any behaved function of θ and φ, f(θ, φ) can be expanded as

f(θ, φ) =∞∑`=0

∑m=−`

a`,mYm` (θ, φ) (26.69)

240

Page 242: PHY 416, Quantum Mechanics - SIUE

while we do not prove this here, it is not terribly surprising since

Y m` (θ, φ) ∝ e−mφPm

` (cos θ) (26.70)

and since the set

eimφ

is complete for functions of φ (it is just the complexFourier series set) and the set P`(cos θ) is complete for functions of θ. (Thesame turns out to be true for associated Legendre functions Pm

` (cos θ)).The completeness of the Y m

` is very useful in slightly more advancedcourses, however, we mention it now as a theoretical point.

241

Page 243: PHY 416, Quantum Mechanics - SIUE

Chapter 27

Twenty-Seventh Class: More on theHydrogen Atom

Thursday, December 4, 2008

27.1 The Radial Equation

The radial equation for the hydrogen atom is

1

r2

d

dr

(r2dR

dr

)+

h2 [E − V (r)] = αR

r2 (27.1)

Since α = `(`+ 1),this is

d

dr

(r2dR(r)

dr

)− 2mr2

h2 [V (r)− E]R(r) = `(`+ 1)R(r) (27.2)

We have already some experience with this for the case ` = 0; in any case,before jumping into mathematics, let’s try to further our intuition. In dealingwith the semi-infinite square well in three dimensions (for the binding energyof the deuteron problem), I mentioned that the three dimensional Schrodingerequation for spherically symmetric states (ψ depends only on r) looks likethe one-dimensional Schrodinger equation if we write it in term of the newvariable u(r) ≡ rψ(r). So we try the same trick here. We define

u(r) ≡ rR(r) (27.3)

then the radial equation becomes

− h2

2m

d2u(r)

dr2 +

[V (r) +

h2

2m

`(`+ 1)

r2

]u(r) = Eu(r) (27.4)

242

Page 244: PHY 416, Quantum Mechanics - SIUE

Note that this is the same as a one-dimensional time independent Schrodingerequationexcept for the appearance of a “new piece” h2

2m`(`+1)r2 added to the

potential energy. This “new piece” is called the centrifugal potential; itshould be familiar from classical mechanics. To connect with this, considerthe energy of, say, an orbiting planet in classical mechanics.

fig 1The energy of this situation is

E =1

2mv2 + V (r) (27.5)

Now, in polar coordinates (since the orbit is in a plane)

~v = rr + rθθ (27.6)

sov2 = r2 + r2θ2 (27.7)

so, the kinetic energy is

K.E. =1

2mv2 =

1

2mr2 +

1

2mr2θ2 (27.8)

The 12mr

2 term is the radial term and the 12mr

2θ2 term is the centrifugalterm. Note that the kinetic energy is the “one-dimensional” ”radial kineticenergy” plus a “centrifugal” piece. Thus

E =1

2mr2 + V (r) +

1

2mr2θ2 (27.9)

is the equivalent one-dimensional form of the energy. We now rewrite thecentrifugal term in terms of the orbital angular momentum L. We have,from ~L = ~r × ~p,

L = m |~v × ~r| = mrvθ = mr · rθ = mr2θ (27.10)

soL2

2mr2 =m2r4θ2

2mr2 =1

2mr2θ2 (27.11)

Comparing equations (27.1) and (27.11), we get

E =1

2mr2 + V (r) +

L2

2mr2 (27.12)

243

Page 245: PHY 416, Quantum Mechanics - SIUE

Written in the last form, the centrifugal term L2

2mr2 looks like an addition to thepotential energy; for this reason it is often called the “centrifugal potential”1.This makes us suspect that our term in the Schrodinger equation

h2

2m

`(`+ 1)

r2 (27.13)

is somehow the value of L2

2mr2 (i.e., L2 = `(` + 1)h2) for the eigenstates wewill find with “quantum numbers” ` and m`. As we will see next week, thisconjecture is exactly correct. Notice that the centrifugal potential acts like aclassical “force” that is repulsive from the origin and is very strong at small r− this is like the centrifugal force in a rotating frame of reference in classicalmechanics

−∂Vcent

∂r= +

h2

2m

`(`+ 1)

r3 (27.14)

the + sign shows that this force is repulsive to + r and the 1r3 term grows

strongly with decreasing r.

27.2 More on the Hydrogen atom Radial Equation

With u(r) ≡ rR(r), our radial equation (from any spherically symmetricpotential V (r)) is

− h2

2m

d2u

dr2 +

[V (r) +

h2

2m

`(`+ 1)

r2

]u = Eu (27.15)

For the hydrogen atom this becomes

− h2

2m

d2u

dr2 +

[− e2

4πε0

1

r+h2

2m

`(`+ 1)

r2

]u = Eu (27.16)

To handle this, first we tidy it up (following Griffith’s notation) by puttingit in terms of a dimensionless dependent variable. We define

κ =

√−2mE

h, ρ ≡ κr, ρ0 ≡

me2

2πε0h2κ

(27.17)

1In classical physics, the “force” − ∂Vcent∂r

associated with Vcent is called the “inertial” or “fictitious” force.

244

Page 246: PHY 416, Quantum Mechanics - SIUE

then the equation is

d2u(ρ)

dρ2 =

[1− ρ0

ρ+`(`+ 1)

ρ2

]u(ρ) (27.18)

It is tempting to try a direct power series solution of this,

u(ρ) =∞∑n=0

anρn

but this does not work. From your reading for today, you know that a methodaround this is to extract the behavior of u(ρ) in the limits ρ→ 0 and ρ→∞,factoring these, then, we define a function v(ρ) by

u(ρ) ≡ ρ`+1e−ρv(ρ) (27.19)

then

ρd2v

dρ2 + 2(`+ 1− ρ)dv

dρ+ [ρ0 − 2(`+ 1)] v = 0 (27.20)

A simple power series trial

v(ρ) =∞∑j=0

cjρj

as a solution of equation (27.20) does not obviously fail, it leads to the re-cursion relation

cj+1 =

[2(j + `+ 1)− ρ0

(j + 1)(j + 2`+ 2)

]cj (27.21)

As you know from your reading, here we again encounter the same kind ofproblem we had for the harmonic oscillator problem − namely, the seriessolution represented by equation (27.21) diverges as ρ → ∞ − and it badlyenough to make u(ρ) = ρ`+1e−ρv(ρ) also diverge as ρ → ∞. Interestingly,“the same problem has the same solution” − our loophole (for the the atomto exist) is the terminate the series into polynomials. Then, equation (27.21)says for a given ρ0, there must be a “jmax” such that

2 (jmax + `+ 1)− ρ = 0 (27.22)

Letting n ≡ jmax + `+ 1, then the termination condition is

2n = ρ0 (27.23)

245

Page 247: PHY 416, Quantum Mechanics - SIUE

which is

2n =me2

2πεoh2

h√−2mE

(27.24)

or√−2mE =

mee2

4πε0nh(27.25)

or

En = −(

1

4πε0

)2me4

2n2h2 (27.26)

therefore,

En =E1

n2 , E1 = −(

1

4πε0

)2mee

4

2h2 (27.27)

these are exactly the Bohr energies! Further,

κ =

√−2mE

h=

(1

4πε0

m2e

h2

)1

n=

1

a0n(27.28)

where a0 = 4πε0h2

mee2 is the Bohr radius.

27.3 A check

Let us pause before going on to state our result in generality and check thatwhat we have makes sense. We consider the ground state. the question is −how do we identify it? Recall our termination condition

n = jmax + `+ 1 (27.29)

The lowest energy is the minimum n, n = 1, and this means that jmax = ` = 0.Recalling that this only allows only m` = 0 (m` ranges from −` to `), we seethat our set of quantum numbers for the ground state is n = 1, ` = 0,m` = 0.So, we are looking at

ψ100(r, θ, φ) = R1,0(r)Y0

0 (θφ) (27.30)

Since Y 00 is a constant, all we need is R1,0(r). Now equation (27.21) with

j = 0

c1 =2(0 + 0 + 1− 1)

(0 + 1)(0 + 0 + 2)= 0

246

Page 248: PHY 416, Quantum Mechanics - SIUE

so for R1,0 only c0 is not zero. Then from equation (27.16),

R1,0(r) =u1,0

r=ρ0+1e−ρv1,0(ρ)

r(27.31)

with v1,0(ρ) = c0 (a constant). Now, ρ ≡ κr, and recalling that κ = 1a0n

, here,n = 1

R1,0(r) = c0r

a0re−

ra0 =

c0

a0e−

ra0 = constant× e−

ra0 (27.32)

which is the same as we had by our original “easier method” early last week.Properly normalized (we’ll see how to do this shortly),

ψ1,0,0(rθ, φ) =1√πa3

0

e−ra0 (27.33)

Now consider the next energy level − n = 2. Referring to our definition ofn, n ≡ jmax + `+ 1, we see that there two possible values of `

` = 0 ⇒ jmax = 1 (27.34)

and` = 1 ⇒ jmax = 0 (27.35)

This means that there are actually four different states with this energy: for` = 1, m` can be −`, 0, or +1! If ` = 0 equation (27.21) says c1 = −c0,c2 = 0, so all higher cn = 0. Then, v(ρ) = c0 + c1ρ = c0(1− ρ), so

R2,0(r) =c0

2a0

(1− r

2a0

)e−

r2a0 (27.36)

If ` = 1, jmax = 0 ⇒ v(ρ) = c0

R2,1(r) =c0

4a20re−

r2a0 (27.37)

so, you see how this thing works: for any n, n = jmax + `+ 1 tell us that thepossible values of ` are

` = 0, 1, 2, 3, · · · , n− 1 (27.38)

and we already know that for any `,

m` = −`,−`+ 1,−`+ 2, · · · ,−1, 0, 1, · · · , `− 1, ` (27.39)

247

Page 249: PHY 416, Quantum Mechanics - SIUE

(there are 2`+ 1 values). Thus, level n has degeneracy

n−1∑`=0

(2`+ 1) = n2 (27.40)

In general,

ψn,`,m`(r, θ, φ) = Rn,`(r)Y

m`

` (θ, φ) (27.41)

=1

rρ`+1e−ρvn,`(ρ)Y m`

` (θ, φ) (27.42)

=1

r

(r

a0n

)`+1

e−r/(a0n)vn,`

(r

a0n

)Y m`

` (θ, φ) (27.43)

27.4 Hydrogen Atom − Eigenstates, Completeness, Probabilities

For the hydrogen atom,

ψn,`,m`(r, θ, φ) = Rn,`(r)Y

m`

` (θ, φ) (27.44)

=1

rρ`+1e−ρvn,`(ρ)Y m`

` (θ, φ) (27.45)

=1

r

(r

a0n

)`+1

e−r/(a0n)vn,`

(r

a0n

)Y m`

` (θ, φ) (27.46)

Vn,`(ρ) is a polynomial of order (n− `− 1):

nn,`(ρ) ≡ L2`+1n−`−1(2ρ) (27.47)

where

Lpq−p(x) ≡ (−1)p(d

dx

)pLq(x) (27.48)

where

Lq(x) ≡ ex(d

dx

)q (e−xxq

)(27.49)

Lq(x) is called the “qth Laguerre polynomial”, and Lpq−p(x) is called the “as-sociated Laguerre polynomial”. Some of the Laguerre polynomials (Lq) are

248

Page 250: PHY 416, Quantum Mechanics - SIUE

listed below.

L0 = 1

L1 = −x+ 1

L2 = x2 − 4x+ 2

L3 = −x3 + 9x2 − 18x+ 6

L4 = x64− 16x3 + 72x2 − 96x+ 24

L5 = −x5 + 25x4 − 200x3 + 600x2 − 600x+ 120

L6 = x6 − 36x5 + 450x4 − 2400x3 + 5400x2 − 4320x+ 720

Some of the associated Laguerre polynomials (Lpq−p) are listed below

L00 = 1

L01 = −x+ 1

L02 = x2 − 4x+ 2

L10 = 1

L11 = −2x+ 4

L12 = 3x2 − 18x+ 18

L20 = 2

L21 = −6x+ 18

L22 = 12x2 − 96x+ 144

L30 = 6

L31 = −24x+ 96

L32 = 60x2 − 600x+ 1200

With the normalization, the wave functions are

ψ(n, `,m`) =√(2

na0

)3(n− `− 1)!

2n [(n+ `)!]3e−r/(na0)

(2r

na0

)` [L2`+1n−`−1

(2r

na0

)]Y m` (θ, φ) (27.50)

which is very complicated looking. Combining the orthonormality of theYell

m’s with that of the associated Laguerre polynomials yields (we don’t

249

Page 251: PHY 416, Quantum Mechanics - SIUE

show this) the result that the ψn,`,m is a complete orthonormal set; theorthonormality relation is∫

ψ∗n,`,m(r, θ, φ)ψn′,`′,m′(r, θ, φ)r2 sin θ dr dθ dφ = δn,n′δ`,`′δm,m′ (27.51)

The completeness property means that we can be assured that we have not“left out” any solution of the Schrodinger equation by looking for productsolutions; any behaved function f(r, θ, φ) can be expanded as

f(r, θ, φ) =∞∑n=1

∞∑`=0

∑m`=−`

an,`,m`ψn,`,m`

(r, θ, φ) (27.52)

In the next classes, we will try to make physical sense of our complicatedlooking results.

250

Page 252: PHY 416, Quantum Mechanics - SIUE

Chapter 28

Twenty-Eigth Class: Resolutions ofParadoxes

Tuesday, December 9, 2008

28.1 The Location of the Electrons in the Hydrogen Atom

We have our eigenstates for the Schrodinger hydrogen atom, but let’s nowlook again at the most basic solution − the ground state. According tothe Bohr model, the electron orbits at definite radius a0 ≈ 0.5 × 10−11 m.However, in the Schrodinger theory

ψ1,0,0 = C1,0,0e−r/a0 (28.1)

This then means that the probability density of ψ∗ψ,

ψ∗ψ = |C|2 e−2r/a0 (28.2)

peaks at r = 0 (inside the nucleus!)fig 1This seems to make no sense! Let us then, recall the volume element in

spherical coordinates.fig 2Thus, probability of materializing in dV is

ProbindV (r,θ,φ) = ψ∗(r, θ, φ)ψ(r, θ, φ)r2 sin θ dr dθ dφ (28.3)

Let us ask, then for the total probability that the electron will materialize“between r and r + dr.” Since we are dealing here with a spherically sym-metric state (ψ∗ψ is independent of θ and φ depends only on r), all we have

251

Page 253: PHY 416, Quantum Mechanics - SIUE

to do in this case is multiply ψ∗ψ by differential volume between r and r+drdV = 4πr2 dr. (hatched volume in figure 28.3.

fig 3Thus

P (r) dr = ψ∗(r)ψ(r) · 4πr2 dr (28.4)

this is for s-states only1. We now see the resolution of our paradox:

P1,0,0(r) dr = 4πe−2r/a0r2 dr ∝ r2e−2r/a0 (28.5)

The factor r2 vanishes at the nucleus; thus the plot of P1,0,0)(r) dr is shownin figure 28.4.

fig 4Let us ash where this function peaks. To do this, we set

dP (r)

dr= 0

d

dr

(r2e−2r/a0

)= 0

r2(− 2

a0

)· e−2r/a0 + 2re−2r/a0 = 0

r2

a0= r

therefore,r = a0 (28.6)

Thus, for the ground state (ψ1,0,0), the most likely small range of r that theelectron can materialize in is centered on r = a0!

28.2 Probability Distributions for Excited States

Now, unlike the ground state, generally the excited state functions dependon θ and φ (unless, of course, ` = 0, in which cases the procedure is similarto that for the ground state.

ψ∗n,`,m`(r, θ, φ)ψn,`,m`

(r, θ, φ) = R∗n,`Θ∗`,m`

Φ∗m`Rn,`Θ`,m`

Φm (28.7)

1“s-states” are eigenstates with ` = 0. The notation “s” is historical, but still widely used.

252

Page 254: PHY 416, Quantum Mechanics - SIUE

subject to the normalization condition∫ ∞0

∫ π

0

∫ 2π

0ψ∗n,`,m`

(r, θ, φ)ψn,`,m`(r, θ, φ)r2 sin θ dr dθ dφ = 1 (28.8)

often,however, we are not interested in keeping track of the angular depen-dence − we often only want to know about the radial probability distribution− e.g., about dP

dr dr, the probability of materialization in dr centered on rirrespective of θ and φ, or the total probability to materialize between dis-tances r1 and r2 from the nucleus, irrespective of θ and φ, etc. If we areinterested in the radial probability density, we want2.

Pn,`(r) dr =

∫ π

0

∫ 2π

0ψ∗ψr2 sin θ dr dθ dφ

= r2R∗n,`(r)Rn,`(r) dr

∫ π

0Θ∗`,mΘ`,m sin θ dθ

∫ 2π

0Φ∗Φ dφ

Now the Θ and Φ functions are separately normalized so that the integralsabove involving them are each equal to one. Thus,

Pn,`(r) dr = r2R∗n,`(r)Rn,`(r) dr (28.9)

We note that this is independent of m`. Of course∫ ∞0

Pn,`(r) dr (28.10)

for any n and `. Thus, for example, in state n, `,m 〈r〉 is found from

〈rn,`〉 =

∫ ∞0

rPn,`(r) dr (28.11)

Here is a way of remembering the r2 factor in Pn,`(r) : P (r) is the probabilityof finding the electron anywhere in the spherical shell of thickness dr centeredon r − this is ψ∗ψ × shell volume.

= ψ∗ψ · surface area× shell thickness

= ψ∗ψ · 4πr2 dr2Pn,`(r) dr ≡ dP

drdr for the state n, `,m`

253

Page 255: PHY 416, Quantum Mechanics - SIUE

The 4π is absorbed in the conversion from ψ to R and the angular integra-tions. To see this, consider, e.g., the 1− s state

ψ1,0,0(r) =1√4πR(r) =

1√π

1

a3/20

e−r/a0 (28.12)

with probability

P (r) dr = |ψ|2 · 4πr2 dr (28.13)

=1

4π|R(r)|2 · 4πr2 dr (28.14)

= r2 |R(r)|2 dr (28.15)

In any case, equation (28.9) on page 253 is correct for any eigenstate ψn,`,m`.

The factor r2 in r2 |Rn,`|2 is important for any radial probability calculationand must not be omitted. Figure 28.5 shows plots for the n = 1, 2 and 3states.

fig 5The small vertical arrows indicate the position of 〈r〉 in units of a0; note

that, in general 〈r〉 6= rpeak of dP/dr. The expectation value is, of course, cal-culated from

〈rn,`〉 =

∫ ∞0

rPn,`(r) dr

=

∫ ∞0

r · r2 |Rn,`|2 dr

=

∫ ∞0

∫ π

0

∫ 2π

0ψ∗n,`,mrψn,`,mr

2 sin θ dr dθ dφ

These expectation values have been evaluated; they are

〈rn,`〉 = n2a0

1 +

1

2

[1− `(`+ 1)

n2

](28.16)

we see from this that the `-dependence is suppressed by the factor 12n2 ; a

qualitatively similar conclusion holds for the peaks. Thus, we see that thepeaks for n = 2 are at larger r than that for n = 1, and for n = 3 atstill larger r. Further, all the n = 2 functions peak near r = 4a0 (mostexact for n = 2, ` = 1); all the n = 3 functions peak at r ∼ 9a0 (most

254

Page 256: PHY 416, Quantum Mechanics - SIUE

exact for n = 3, ` = 2), etc. This is roughly in agreement with the Bohrmodel (Rn = n2a0) and provides support for a shell model of the atom. Inorder order to make much sense of the angular probability distributions, it isgood to first discuss the connection with the angular momentum in quantummechanics, and that we turn to next.

28.3 Angular Momentum and Hydrogen Atom States

We have, in cartesian coordinates,

Lx = ypz − zpy (28.17)

Ly = zpx − xpz (28.18)

Lz = xpy − ypx (28.19)

As we have discussed, in quantum mechanics, classical quantities “do notexist”, rather they are represented by operators. Since px = − − h ∂

∂x , etc.,we have

Lx = −ih(y∂

∂z− z ∂

∂y

)(28.20)

Ly = −ih(z∂

∂x− x ∂

∂z

)(28.21)

Lz = −ih(x∂

∂y− y ∂

∂z

)(28.22)

As we are working in spherical coordinates, we need to transform these. Ifwe do this, one finds

Lx = −h(

sinφ∂

∂θ+ cot θ cosφ

∂φ

)(28.23)

Ly = −h(− cosφ

∂θ+ cot θ sinφ

∂φ

)(28.24)

Lz = −ih ∂∂φ

(28.25)

255

Page 257: PHY 416, Quantum Mechanics - SIUE

We will also be interested in L2 = L2x + L2

y + L2z; in spherical coordinates this

works out to be

L2 = −h[

1

sin θ

∂θ

(sin θ

∂θ

)+

1

sin2 θ

∂2

∂φ2

](28.26)

Let us consider finding the eigenfunctions of Lz and L2. Recall that, for anyoperator Q, the function f(~r) is an eigenfunction of Q if and only if

Qf(~r) = constant× f(~r) (28.27)

the constant is then the eigenvalue corresponding to the eigenfunction f(~r)of Q. Consider, then, the eigenfunction problem for Lz:

−ih ∂∂φf(φ) = cf(φ)⇒ df(φ)

dφ=ic

hf(φ)

therefore,f(φ) = Ae

ich φ (28.28)

any function f(φ) of this form is an eigenfunction of Lz. Now consider thefunctions e−m`φ from the hydrogen atom. Clearly

Lzeim`φ = m`heim`φ (28.29)

sinceY m` (θ, φ) ∝ Pm

` (cos θ) eim`φ (28.30)

LzYm` (θ, φ) = m`hY

m` (θ, φ) (28.31)

Further since

Lzψn,`,m`(r, θ, φ) = −ih ∂

∂φ

= R(r)

[−ih ∂

∂φY m` (θ, φ)

]= m`h [R(r)Y m

` (θ, φ)]

= m`hψn,`,m(r, θ, φ)

thus, ψn,`,m(R, θ, φ) is an eigenfunction of Lz with eigenvalue m`h. Nowconsider equation (28.3)

L2 = −h[

1

sin θ

∂θ

(sin θ

∂θ

)+

1

sin2 θ

∂2

∂φ2

]256

Page 258: PHY 416, Quantum Mechanics - SIUE

The right hand side of this looks kind of familiar. Recall the “Θ-equation”for the hydrogen atom eigenstates; it is[

− 1

sin θ

∂θ

(sin θ

∂θ

)+

m2`

sin2 θ

]= `(`+ 1)Θ(θ) (28.32)

This is solved by Θ`,m`(θ). Consider, then

L2ψm,`,m`= −h2

[1

sin θ

∂θ

(sin θ

∂θ

)]ψn,`,m`

− h2

sin2 θ

∂2

∂φ2ψn,`,m`(28.33)

since∂2

∂φ2ψn,`.m`= R(r)Θ(θ)

∂2

∂φ2 eim`φ = (im`)2R(r)Θ(θ)Φ(φ)

L2ψn,`,m`= −h2

[1

sin θ

∂θ

(sin θ

∂θ

)]ψn,`,m`

− m2`

sin2 θψn,`,m`

(28.34)

therefore,

L2ψn,`,m`= −h2Rn,`(r)Φm`

(φ)

[1

sin θ

∂θ

(sin θ

∂θ

)− m2

`

sin2 θ

]Θ(θ) (28.35)

by using equation (28.32), this is

L2ψn,`,m`= h2`(ell + 1)ψn,`,m`

(28.36)

i.e., ψn,`,m`(r, θ, φ) is an eigenfunction of L2 with eigenvalue `(`+ 1)h2

Thus, in addition to to being an eigenfunction of the hydrogen atom Hamil-tonian with eigenvalue En

Hψn,`,m`= Enψn,`,m`

(28.37)

ψn,`,m`is simultaneously also an eigenfunction of L2 and Lz

As we will see, the consequences of this are interesting are profound. How-ever, already we see the physical meaning of the quantum numbers of ` andm`. Thus, if the state function is ψn,`,m`

If we measure the energy, we def-initely get the value En. If we measure the magnitude of the “orbital” an-gular momentum, we definitely get the value

√`(`+ 1)h. If we measure the

z-component of the angular momentum, we definitely get the value m`h. Ifthe state function is a superposition state, then the probabilities follow theusual statistical rules we’ve developed.

257

Page 259: PHY 416, Quantum Mechanics - SIUE

28.4 Compatible and Incompatible Quantities

We have established3 that the following three operators commute with eachother: [

H, L2]

=[H, Lz

]=[H, Lz

](28.38)

Further, we have seen that the ψn,`,m`are simultaneously, eigenfunctions of

all three of these operators. This means that the values of of E, L2, and Lzare sharp for any state ψn,`,m`

. These results illustrate a general theorem inquantum mechanics

Theorem: If observable operators A and B commute ([A, B

]= 0), then

there exists a complete set of functions ψA,B that are simultaneously eigen-functions of A and eigenfunctions of B − i.e.,

AψA,B = AψA,B

BψA,B = BψA,B

thus, in such a state, the quantities corresponding to both A and B are sharp.We will not have time to prove this theorem in general, but, we note that the

operators H, L2, and Lz for the hydrogen atom provide an example.On the other hand, the inverse of the theorem is also true − i.e., If A and

B do not commute ([A, B

]6== 0), then other than the zero function, there

exists no function that is simultaneously an eigenfunction of A and also an

eigenfunction of B. Thus, if[A, B

]6= 0, there exists no state (other than

ψ = 0) for which A and B are both sharp.Example: Consider x and pxWe know that [x, px] = ih 6==. We also already know that there does not

exist any states for which x and px are simultaneously sharp − the HeisenbergUncertainty Principle forbids this. Now, we begin to see why.

In fact, in any state, the product of the statistical uncertainties σAσB ofA and B is determined by the extent of the noncommutation:

σ2Aσ

2B ≥

(1

2i

⟨[A, B

]⟩)2

(28.39)

3You can easily show these

258

Page 260: PHY 416, Quantum Mechanics - SIUE

where 〈[ ]〉 is the expectation value of the commutator[A, B

]. Equation

(28.39)is the generalized uncertainty principle. It is proved in the appendixto this lecture.

Example: [x, px] according to equation (28.39)

σ2xσ

2p ≥

(1

2i· ih)2

=

(h

2

)2

⇒ σxσp ≥h

2(28.40)

Now, we know, e.g., that (similarly, one can show that)[Lx, Ly

]= ihLz[

Ly, Lz

]= ihLx[

Lz, Lx

]= ihLy

(28.41)

This means that there is no state (other than ψ = 0) that is sharp in morethan one cartesian component of L.

Example: ψn,`,m is sharp in Lz (Lz = m`h).Therefore, it is not sharp in Lx and not sharp in Ly.Question: Suppose then that the state function is ψn,`,m and we measure

Lx. What possible results can you obtain? What is the probability of eachresult?.

Answer:To answer this, one must expand the state function ψn,`,m over the com-

plete set of eigenfunctions of the operator Lx. (Since Lx is a Hermitianoperator, like all other Hermitian operators it possesses a complete set ofeigenfunctions). Let us call these functions fLx. Then our expansion is

ψn,`,m =∑k

ckfLx,k (28.42)

whereLxfLx,k = LxfLx,k

The probability of obtaining the result Lx,k is |ck|2.We will now inquire as to the nature of the eigenfunctions of Lx. We

will note, however, (without proof) that the eigenvalues of Lx are integer

259

Page 261: PHY 416, Quantum Mechanics - SIUE

multiples of h, as are the eigenvalues of Lz (and Ly). For this reason, h is thefundamental unit of angular momentum

28.5 Angular Dependence of States and Probabilities

We now return to the angular dependence of probabilities. We note that

ψ∗n,`,m`ψn,`,m`

= R∗n,`Rn,`Θ∗`,m`

Θ`,m`Φ∗m`

Φm`(28.43)

but,Φ∗m`

Φm`= eim`φeim`φ = 1 (28.44)

consequently, for state ψn,`,m`, the probability density does not depend of φ.

Thus, to study the angular dependence of probabilities for the eigenstates,we need only study Θ∗`,m`

Θ`,m`. A simplified way of studying this involves a

polar diagram (see figure 28.6); the origin is at the nucleus, and the z-axis istaken alone the direction from which θ is measured. The distance from theorigin to the curve, at angle θ is proportional to Θ∗`,m`

(θ)Θ`,m`(θ).

fig 6Let is look at a few of the states in more detail. The 1s and 2s states are,

of course, spherically symmetric. However, consider the 2p states.

ψ2,1,0 = C2,1,0

(r

a0

)e

r2a0 cos θ (28.45)

ψ2,1,±1 = C2,1,1

(r

a0

)e−

r2a0 sin θe±iφ (28.46)

The angular probability densities Θ∗Θ behaves as

Θ∗1,0Θ1,0 ∝ cos2 θ (28.47)

Θ∗1,±1Θ1,±1 ∝ sin2 θ (28.48)

(where θ is measured from the z-axis). Computer plots in which the densityof the dots gives an indication of the r-dependence of the probability densityare shown in figures 28.7 through 28.12.

fig 7-12

260

Page 262: PHY 416, Quantum Mechanics - SIUE

Appendix

28.6 Schwarz Inequality

to proceed we need to establish a result called the “Schwarz Inequality”. Thissay that, for any two normalizable functions f(x) and g(x), and any interval[a, b], (∫ b

a

|f |2 dx

)(∫ b

a

|g|2 dx

)≥∣∣∣∣∫ b

a

f ∗(x)g(x) dx

∣∣∣∣2 (28.49)

In our more compact notation, this is

〈f | f〉 〈g | g〉 ≥ |〈f | g〉|2 (28.50)

Before we prove this, first let’s get some intuition on it in ordinary threedimensional space. SUppose that, in ordinary three dimensional space, wehave two vectors, ~A and ~B. Then

textitfig 13

〈A |B〉 ≡ ~A · ~B = AB cos θ ≤ AB (28.51)

therefore,|〈A |B〉|2 ≤ |A|2 |B|2 (28.52)

The Schwartz inequality say “the same thing” on the infinite dimensionalfunction space.

Proof

1. Lemma: For any normalizable function f(x), 〈f | f〉 ≥ 0.

Proof : Using completeness, expand in the complete set of eigenfunctionsof any hermitian observable

f(x) =∑n

cnψn (28.53)

261

Page 263: PHY 416, Quantum Mechanics - SIUE

Then,

〈f | f〉 =

⟨N∑i=1

ciψi |N∑j=1

cjψj

=∑i=1N

N∑j=1

c∗i cj 〈ψi |ψj〉

=N∑i=1

N∑j=1

c∗i cjδi,j

=N∑i=1

|ci|2 ≥ 0

2. Now let ψ = f + bg where b is an arbitrary complex number.

By our lemma,0 ≤ 〈|ψ〉 = 〈f + bg | f + bg〉 (28.54)

Therefore,

0 ≤ 〈f | f〉+ b 〈f | g〉+ b∗ 〈g | f〉+ bb∗ 〈g | g〉 (28.55)

To minimize the right hand side, we differentiate with respect to b andset the result equal to zero.

0 =d

db〈ψ |ψ〉 = 〈f | g〉+ b∗ 〈g | g〉 (28.56)

therefore,

b∗ = −〈f | g〉〈g | g〉

(28.57)

and

b = −〈g | f〉〈g | g〉

(28.58)

By putting these into equation (28.55), we get

0 ≤ 〈f | f〉 − 〈g | f〉〈g | g〉

〈f | g〉 − 〈f | g〉〈g | g〉

+〈g | f〉 〈f | g〉|〈g | g〉|2

(28.59)

262

Page 264: PHY 416, Quantum Mechanics - SIUE

Therefore,

0 ≤ 〈f | f〉 − |〈f | g〉|2

〈g | g〉+|〈f | g〉|2

〈g | g〉(28.60)

Therefore,〈f | f〉 〈g | g〉 ≥ |〈f | g〉|2 (28.61)

28.7 The Uncertainty Principle − Proved

Suppose we have an ensemble of systems, all in the same state Ψ, and weconsider measurements of two observables, A and B. From our previousdiscussions, the variances are

σ2A =

⟨(A− 〈A〉

)2⟩

=⟨

Ψ |(A− 〈A〉

)(A− 〈A〉

)Ψ⟩

=⟨(A− 〈A〉

)Ψ |(A− 〈A〉

)Ψ⟩

Therefore,σ2A = 〈f | f〉 (28.62)

where f(x) ≡(A− 〈A〉

)Ψ. Likewise,

σ2B = 〈g | g〉 (28.63)

where g(x) ≡(B − 〈B〉

Consider nowσ2Aσ

2B = 〈f | f〉 〈g | g〉 (28.64)

By the Schwarz inequality, the right hand side of this is greater than |〈f | g〉|2,hence

σ2Aσ

2B ≥ |〈f | g〉|

2 (28.65)

Now 〈f | g〉 is some complex number; call it z. Since for any complex number,

|z|2 = z∗z = [Re(z)− iIm(z)] [Re(z) + iIm(z)]

= [Re(z)]2 + [Im(z)]2 ≥ [Im(z)]2

=

[z − z∗

2i

]2

263

Page 265: PHY 416, Quantum Mechanics - SIUE

We have

σ2Aσ

2B ≥

[〈f | g〉 − 〈g | f〉

2i

]2

(28.66)

Let us work this out in terms of A and B

〈f | g〉 =⟨(A− 〈A〉

)Ψ |(B − 〈B〉

)Ψ⟩

=⟨

Ψ |(A− 〈A〉

)(B − 〈B〉

)Ψ⟩

=⟨

Ψ |AB − 〈A〉 B − A 〈B〉+ 〈A〉 〈B〉

Ψ⟩

=⟨

Ψ |(AB)

Ψ⟩− 〈A〉

⟨Ψ | BΨ

⟩− 〈B〉

⟨Ψ | AΨ

⟩+ 〈A〉 〈B〉 〈Ψ |Ψ〉

= < 〈AB〉 − 〈A〉 〈B〉 − 〈B〉 〈A〉+ 〈A〉 〈B〉= 〈AB〉 − 〈A〉 〈B〉

Likewise,〈g | f〉 = 〈f | g〉∗ = 〈BA〉 − 〈A〉 〈B〉

Therefore, equation (28.66) says

σ2Aσ

2B ≥

[〈AB〉 − 〈BA〉

2i

]2

=

[〈AB −BA〉

2i

]2

⟨[A, B

]⟩2i

2

(28.67)

Thus, the generalized uncertainty principle is

σ2Aσ

2B ≥

⟨[A, B

]⟩2i

2

(28.68)

this gives a quantitative expression to “how incompatible” A and B are whenthey don’t commute.

Example: Suppose A = x and B = px = hi∂∂x .

As you know, [x, px] = −h. Therefore, equation (28.68) says that, for anystate Ψ

σ2xσ

2px≥(ih

2i

)2

=

(h

2

)2

(28.69)

264

Page 266: PHY 416, Quantum Mechanics - SIUE

or

σxσpx ≥h

2(28.70)

Which you recognize as the original Heisenberg Uncertainty Principle , nowproved. However, we now see that this is just one case a much more general(and deeper) principle. We will discuss one consequence of this in the nextclass.

265

Page 267: PHY 416, Quantum Mechanics - SIUE

Chapter 29

Twenty-Ninth Class: Four LastThings

Friday, December 2, 2008

29.1 What’s Special About the z-Axis?

Look back at our polar plots of the angular dependence of the hydrogen atomeigenstate probability densities. We see that in these plots that, for a givenstate with ` 6= 0, there are conical “nodal surfaces” on which the electroncan never materialize. This is very strange, you might think. You might say− prepare a million different hydrogen atoms in the same state, then mea-sure the position of the electron for each. This would then locate the nodalsurfaces, from which one could locate the z-axis. But, this should be im-possible − since V (r) is spherically symmetric, this would amount to findinga preferred direction in space, which is a violation of a very basic physicalprinciple. Thus, it appears that we have a serious paradox undermining theentire Schrodinger theory of the hydrogen atom on our hands. Let us thinkon this.

Consider any one of the n = 2 states, for example, ψ2,1,1. How could one seta system into this state? We need to perform three measurements − energy(to segregate for further measurements later all atoms materializing the n = 2energy level), magnitude of angular momentum (to segregate all atoms with` = 1 in addition to n = 2), and Lz (to finally define Lz = 1h). Consider,for example, the least measurement. To accomplish it requires an externalmagnetic field set up to point in the direction we wish to call “z”. thus, to

266

Page 268: PHY 416, Quantum Mechanics - SIUE

make the measurement requires the positive setting up of a special directionin space, thus destroying the isotropy. Without destroying the isotropy ofspace by forcing a “special direction” it is not possible to measure Lz. Thus,without destroying the isotropy of space, all we an say is that an atom couldbe in any of ψ2,1,−1, ψ2,1,0, ψ2,1,1, or any superposition of these. Thus, onaverage, a large ensemble would correspond to an equal weight superpositionof these states; thus

ψ∗ψeffective =1

3

[ψ∗2,1,−1ψ2,1,−1 + ψ∗2,1,1ψ2,1,1 + ψ∗2,1,0ψ2,1,0

](29.1)

Two of these terms are proportional to 12 sin2 θ, and one is proportional to

cos2 θ. Therefore, the sum is spherically symmetric (sin2 θ + cos2 θ = no θ

dependence).This logic shows that at least, we cannot “force” the paradox to appear

experimentally. To further resolve the paradox conceptually, we must realizeagain that the “wave function” ψ is not a real thing that exists in real space.(In fact, this paradox demonstrates one of the major problems with such anaive “realism” interpretation of ψ.

29.2 A Brief Basis for Chemical Bonding − Atomic “Orbital” inDifferent Directions

We have, for p-states for any n > 1,

ψn,1,1)r, θ, φ) = −√

3

8πRn,1(r) sin θeiφ (29.2)

ψn,1,0(r, θ, φ) =

√3

4πRn,1(r) cos θ =

√3

4πRn,1(r)

z

r(29.3)

ψn,1,−1(r, θ, φ) =

√3

8πRn,1(r) sin θe−iφ (29.4)

These three states are degenerate (correspond to the same energy). Now,chemical bonding in a molecule is greatly enhanced when the electron’s wavefunction overlap from the atom’s is maximally dense. The state ψn,1,0 issuited for this since its “lobes” stick out purely along z; however, the highdensity region for the other two states is distributed all around the equator,

267

Page 269: PHY 416, Quantum Mechanics - SIUE

and hence is “diluted”. With this in mind, we investigate the possibility offorming new eigenstates that have a more “centered” probability density thando ψ2,1,1 and ψ2,1,−1. to begin to understand how this works, we consider thefollowing statement: Any linear combination of eigenstates corresponding tothe same energy for a given Hamiltonian H is also an eigenstate of H for thatsame energy.

Proof : Suppose Hψi = Eψ1 and Hψ2 = Eψ2. Let c1 and c2 be anycomplex numbers. Then

c1Hψ1 = H(c1ψ1) = c1Eψ1 = Ec1ψ1 (29.5)

c2Hψ2 = H(c2ψ2) = c2Eψ2 = Ec2ψ2 (29.6)

(29.7)

(since H is linear). Adding these

H (c1ψ1 + c2ψ2) = E (c1ψ1 + c2ψ2) (29.8)

(again using the linearity of H). Therefore, c1ψ1+c2ψ2 is also an eigenfunctionof H with eigenvalue E.

with this in mind, let us consider the following three linear combinationsof ψn,1,0, ψn,1,1, and ψn,1,−1

pz = ψn,1,0(~r) =

√3

4πRn,1(r)

z

r(29.9)

px = − 1√2

[ψn,1,1(~r)− ψn,1,−1(~r)] =

√3

4πRn,1(r)

x

r(29.10)

py =i√2

[ψn,1,1(~r) + ψn,1,−1(~r)] =

√3

4πRn,1(r)

y

r(29.11)

These new eigenfunctions (“orbitals” to chemists) are shown in figure 29.1.fig 1In each of these we have two regions of relatively high electron probability

density, which is god for chemical bonding. An example wherein nature“makes use of this” is in the crucially important H2O molecule. In the oxygenatom (at the “center” (total 8 electrons), the 1s state is filled (uses twoelectrons, spin up and spin down). the 2s state is also filled (uses 2 moreelectrons), one 2p orbital is fully occupied (uses 2 more electrons) and the

268

Page 270: PHY 416, Quantum Mechanics - SIUE

remaining two 2p orbitals are each only half full (1 electron each) and so eachcan join with a 1s orbital from a hydrogen atom in its ground state to formwhat are “spσ′’ bonding orbitals (shown in figure 29.2). In this, essentiallythe 2px and 2py orbitals of the oxygen are used.

fig 2

29.3 The Necessary “Bohr-Limit”

We wish to further investigate the classical limit of the quantum mechanicsof the hydrogen atom. Ultimately, we would like to be able to constructtight wave packets that circle the nucleus in orbitals,part of this is impliedby Ehrenfest’s theorem, which tells us that 〈~r〉 for the electron follows theclassical path. As a conceptual step along the way, we could ask − to whatextent are certain of the eigenfunctions ψn,`,m`

Bohr-like? Now we’re alreadyseen that the radial probability distributions peak around n2a0 (with sub-peaks in case for which ` < n− 1); as a reminder, a figure corresponding tothis is shown figure 29.3.

fig 3The states with the “single bump in r” are the most Bohr-like. It is

interesting that these occur for the highest possible `-values for each n (i.e.,` = n− 1). This is explained in the following

textClassical Limit: One can imagine “setting up” a superposition state

function Ψ(x, t = 0) that is the sum of large n and ` = n− 1 eigenfunctions,and one that is localized to a small azimuthal range ∆φ. (For large n and` = n−1, Ψ(x, t = 0) is already very localized around θ = π

2 and in r ), If thetime dependence for each term is included, then, presumably, the followingis what would unfold over over the course of time: the packet would “split”into two pieces, each of which orbits the point at the proton in the oppositedirection. This, presumably represents the classical limit.

269

Page 271: PHY 416, Quantum Mechanics - SIUE

29.4 If an Eigenstate is Forever, Why Do Atoms Radiate?

Consider an electron in a hydrogen atom; say in a 2p state. After a meanlifetime τ ∼ 10−8 seconds in this state, the atom “decays” to the groundstate, with the emission of a photon of frequency corresponding to the energydifference between 2p and 1s. But why? 2−p is an eigenstate, and accordingto our work so far, eigenstates should last forever.

The classical theory, on the other hand, seems to present the oppositeproblem. According to Maxwell;’s equations, whenever a charge accelerates,the vacuum is obliged to emit electromagnetic radiation; since electrons inclassical orbits accelerate, they should always radiate. We will sketch theanswer to the answer to the “eigenstates should live forever” question shortly.In the meantime, to get a physical picture (again, we’ll justify this shortly)we will assume that “during the transition”, the state function is a timedependent superposition of the initial and final atomic eigenstates

Ψ(~r, t) = ai(t)ψi(~r)e−Eit/h + af(t)ψf(~r)e

−iEf t/h (29.12)

Note that the coefficients ai and af are time dependent where ai(t) = 1before the transition and zero after the transmission and af(t) = 0 beforethe transition and one after the transition. Now consider 〈~r〉 for the electronduring the transition

〈~r〉t =

∫Ψ∗(~r, t)~rΨ(~r, t) d3~r

= a∗1ai

∫~r |ψi|2 d3~r + a∗faf

∫~r |ψf |2 d3~r

+ a∗iaf

(∫ψ∗i~rψf d

3~r

)e+i(Ei−Ef )t/h

+ afa∗f

(∫ψ∗f~rψi d

3~r

)e−i(Ei−Ef )t/h

Now, since ψi and ψf are eigenstates ψn,`,m`, the first two integrals vanish.

The third and fourth terms are complex conjugates, so their sum is

〈~r〉t = 2Re

(a∗fai

)(∫ψ∗f~rψi d

3~r

)e−i(Ei−Ef )t/h

(29.13)

270

Page 272: PHY 416, Quantum Mechanics - SIUE

This expression has three factors. The middle factor is called “the matrixelement of ~r”; it is an overlap integral involving the final and initial states;it is a measure of the amplitude of the oscillation defined by the real part of

the last factor, cos(Ei−Efh t

). Thus, in this picture, during the transition, the

position of the electron expectation value undergoes simple harmonic motionof frequency

Ei−Efh = E2−E1

h ; semi-classically, this would require the emissionof electromagnetic radiation of this frequency.

Now, why should the state function during the transition be a superpo-sition of hydrogen atom eigenstates? The answer is deep, and have timeonly to sketch it here. The point is that, according to quantum field theory,the Hamiltonian of the electron is the atom is not the whole story! Rather,the real Hamiltonian has three pieces, the Hamiltonian for the electron inthe hydrogen atom, the Hamiltonian for the vacuum radiation filed, and theHamiltonian for the interaction between the electron and the vacuum radia-tion field. Thus, the total Hamiltonian for the hydrogen atom is the sum ofthese

He− ≡ H1 + H2 + H3 (29.14)

The eigenstates of H1 are the ψn,`,m`’s. The eigenstates of H2 are the existence

of various numbers and frequencies of photons in the vacuum.But, what can be said about H3? Electrodynamics shows that it is of the

formHint = − e

2me

(~p · ~A+ ~A · ~p

)(29.15)

where ~p is the electron momentum operator and ~A is the operator for theelectromagnetic vector potential.

At this point,you might think that, in “darkness” (atom in vacuum), ~A =0, so nothing will happen, but quantum field theory shows this to be wrong!The background quantum vector potential field, whose excited modes arephotons, looks like an infinite collection of quantum oscillators. Accordingto quantum mechanics, each oscillator mode (photon) has a nonzero “zero-point energy” (recall that this is 1

2hω for a quantum oscillator), so even in

“darkness”, Hint is always nonzero.Now consider the atom in the 2p state. this is an eigenstate of H1, but

it is not an eigenstate of Hfull = H1 + H2 + H3. Therefore, its real time

271

Page 273: PHY 416, Quantum Mechanics - SIUE

dependence is not e−iE2t/h, where E2 is the energy of 2p. Rather, its best tojust write it as Ψ(~r, t).

So, what is Ψ(~r, t)? According to the completeness theorem (again!), what-ever the eigenstates, ψn,`,m`

are a complete set, so

Ψ(~r, t) =∑

an,`,m`(t)ψn,`,m`

e−iEnt/h (29.16)

where the expansion coefficients an,`,m`(t) depend on time. We note that

equation (29.16) is of the form of equation (29.12). Detailed analysis showsthat only the initial 2p and the ground state 1s states contribute to equation(29.16), so equation (29.12) does indeed result.

29.5 Transition Rates and Spin Angular Momentum of Photons

We have seen that the transition rate (probability of decay from excitedstate to ground state with emission of electromagnetic radiation, per second)should be proportional to the square of the “matrix element” between theinitial and final electron states, i.e.,

R ∝ |Pfi|2 , Pfi ≡∫ψ∗fe~rψi dτ (29.17)

where dτ ≡ d3~r. A rigorous treatment from quantum electrodynamics cor-roborates this conjecture. Now, Pfi can be evaluated for every pair of eigen-states for the hydrogen atom. When this is done, it is found that Pfi = 0unless

∆` = ±1

∆m` = 0,±1(29.18)

for the transition. These are known as the “selection rules”. Thus, unlessthese “selection rules” are obeyed, the rate for the transition is much sup-pressed (in fact, in our approximate treatment the rate would be zero.) Suchtransitions are called “dipole forbidden”. We can get some insight into the∆`±1 rule by noting that ψn,`,m`

∝ Y m`

` (θ, φ) and the space reflection on par-ity of Y m`

` is (−1)! That is, if ~r → −~r, then Y m`

` (θ, φ)→ (−1)`Y m`

` (θ′, φ′)⇒ψn,`,m`

→ (−1)`ψn,`,m`(r,−θ,−φ). Now consider

Pfi = e

∫ψ∗f~rψi dτ (29.19)

272

Page 274: PHY 416, Quantum Mechanics - SIUE

Under reflection ~r → −~r. Thus, if ψ∗fψi → +ψ∗fψi, the integrand is off and

Pfi = 0. But ψ∗fψi → (−1)`f (−1)`i ⇒ ψ∗fψi is even if ∆` = 0, 2 and ψ∗fψi isodd if ∆` = ±1. If ∆` = 1, it is fair to ask what happens to the “disappeared”angular momentum. The theory of relativity plus quantum mechanics showsthat the photon carries a “spin” or intrinsic angular momentum of either+1h or −h in the propagation direction, these are “right circularly” and “leftcircularly” polarized photons.

a slightly more detailed discussion follows1

1From R. Eisberg and R. Resnick, Quantum Physics of Atoms, Molecules, Solids, Nuclei, and Particles(Wiley), 2nd Edition.

273

Page 275: PHY 416, Quantum Mechanics - SIUE

Index

Heisenberg Uncertainty Principle , 88,89, 123, 134, 135

Schrodinger equation, 111, 116, 118

Airy Integral, 198Albert Einstein, 6, 24, 27, 36Angular Momentum, 226angular momentum, 255associated Legendre function’, 238

Balmer series, 228bandwidth theorem, 100Bohr model of the atom, 228Boltzmann’s Constant, 7Born Probability Rule, 26Born probability rule, 23, 44Box Normalization, 106

Charles Hermite, 206classical wave equation, 55coherence time, 155coherent state, 219commutator relation, 221completeness postulate, 137, 138complex conjugate, 43, 44Compton scattering, 75Copenhagen interpretation, 24, 25, 27,

28, 31, 37correspondence limit’, 144

deBroglie, 35, 36, 38, 41, 42

DeBroglie wavelength, 147deBroglie wavelength, 30–33Dirac delta function, 92Dirac Normalization, 106dispersion relation, 35, 38

Ehrenfest’s theorem, 200, 224eigenfunction, 115eigenstate, 114eigenstate , 115Eigenstate Property, 111eigenstates, 112, 114, 117eigenvalue, 113, 114, 118eigenvalue equation, 112eigenvalue’, 112Euler’s formula, 42expectation value, 70

Fourier, 35, 51, 63, 66Fourier analysis, 76Fourier Bandwidth Theorem, 86Fourier Series, 79Fourier series, 78, 136Fourier transform, 84Fourier transforms, 81–83, 89, 91, 93,

94, 96free particle, 111Fresnel, 12

Gamow’s approximation, 168

274

Page 276: PHY 416, Quantum Mechanics - SIUE

Garnow’s approximation, 167Gaussian, 32, 50, 52generalized uncertainty principle, 259,

264ground state, 123

Hamiltonian, 112, 114, 116, 119hamiltonian, 111Heisenberg Uncertainty Principle, 66–

69, 74, 76Hermite functions, 207Hermite polynomials, 210Hermite’s equation, 206Hermitian operator, 112, 114Hermitian operators, 104Huygen’s principle, 8, 13

infinite square well, 125

Laguerre polynomial, 248Laurent Schwartz, 92Legendre polynomial, 238

normalization, 26, 43, 45, 47, 123, 127

orthogonality property, 136orthonormal set, 77, 78

parity, 203Parseval’s theorem, 89, 100Paul Dirac, 92phasors, 14–16potential

barrier potential, 160finite square well, 130, 132, 190hydrogen atom, 228, 242infinite square well, 45, 47, 121,

123, 132, 143

infintie square well, 147simple harmonic oscillator, 202, 208,

211step-potential, 155, 160

potentialFinite square well, 182

Poynting vector, 10Probability Current, 57probability flux, 58propagator, 218

quantization of energy, 46, 122, 128quantum initial value problem, 50Quantum Mechanics Convention, 59quantum tunneling, 160

recursion relation, 206reflection coefficient, 151reflection coefficient’, 159

Schrodinger, 47–52, 54, 55, 72Schrodinger equation, 40, 43, 45, 46Schrodinger equation, 45Schwarz inequality, 261selection rules, 272sinc(x), 17state function, 23, 25, 31–33, 36, 37,

41, 42, 49, 50, 112, 125state functions, 43stationary states’, 117, 143

Thomas Young, 8Time Dependent Schrodinger Equa-

tion, 119–121time dependent Schrodinger equation,

135

275

Page 277: PHY 416, Quantum Mechanics - SIUE

Time Dependent Schrodinger Equa-tion, 120

Time Independent Schrodinger Equa-tion, 116, 119, 121

time independent Schrodinger equa-tion, 125

transmission coefficient, 162transmission coefficient’, 159

wave function, 23, 42WKB Approximation, 167

zero-point energy, 123, 134, 208

276