0% found this document useful (0 votes)
228 views218 pages

James W. Rohlf - Guide To Modern Physics. Using Mathematica For Calculations and Visualizations-CRC Press (2024)

This document provides a guide to performing calculations in modern physics using Mathematica. It is aimed at second year college physical science and engineering students. Key features include a concise summary of physics concepts, over 300 worked examples in Mathematica with available notebooks, and a tutorial for beginners to produce fast results. The guide stresses graphical visualization, units, and numerical answers to help students learn physics without getting stuck on mathematics. The author is James Rohlf, a professor at Boston University with extensive experience in particle physics experiments.

Uploaded by

malossi
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
228 views218 pages

James W. Rohlf - Guide To Modern Physics. Using Mathematica For Calculations and Visualizations-CRC Press (2024)

This document provides a guide to performing calculations in modern physics using Mathematica. It is aimed at second year college physical science and engineering students. Key features include a concise summary of physics concepts, over 300 worked examples in Mathematica with available notebooks, and a tutorial for beginners to produce fast results. The guide stresses graphical visualization, units, and numerical answers to help students learn physics without getting stuck on mathematics. The author is James Rohlf, a professor at Boston University with extensive experience in particle physics experiments.

Uploaded by

malossi
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 218

Guide to Modern Physics

This is a “how to guide” for making beginning calculations in modern physics. The academic
level is second year college physical science and engineering students. The calculations are
performed in Mathematica, and stress graphical visualization, units, and numerical answers.
The techniques show the student how to learn the physics without being hung up on the math.
There is a continuing movement to introduce more advanced computational methods into
lower-level physics courses. Mathematica is a unique tool in that code is written as “human
readable” much like one writes a traditional equation on the board.

Key Features:
• Concise summary of the physics concepts.
• Over 300 worked examples in Mathematica with available notebooks.
• Tutorial to allow a beginner to produce fast results.

James Rohlf is a Professor at Boston University. As a graduate student, he worked on the first
experiment to trigger on hadron jets with a calorimeter, Fermilab E260. His thesis (G. C. Fox,
advisor, C. Barnes, R. P. Feynman, R. Gomez) used the model of Field and Feynman to compare
observed jets from hadron collisions to that from electron-positron collisions and made detailed
acceptance corrections to arrive at first the measurement of quark-quark scattering cross sec-
tions. His thesis is published in Nuclear Physics B171 (1980) 1. At the Cornell Electron Storage
Rings, he worked on the discovery of the Upsilon (4S) resonance and using novel event shape
variables developed by Steven Wolfram and his thesis advisor, Geoffrey Fox. He performed par-
ticle identification of kaons and charmed mesons to establish the quark decay sequence, b –>
c. At CERN, he worked on the discovery of the W and Z bosons and measurement of their
properties. Presently, he is working on the Compact Muon Solenoid (CMS) experiment at the
CERN Large Hadron Collider (LHC) which discovered the Higgs boson and is searching for new
phenomena beyond the standard model.
Guide to Modern Physics
Using Mathematica for Calculations
and Visualizations

James W. Rohlf
Designed cover image: James W. Rohlf

First edition published 2024


by CRC Press
2385 NW Executive Center Drive, Suite 320, Boca Raton FL 33431

and by CRC Press


4 Park Square, Milton Park, Abingdon, Oxon, OX14 4RN

CRC Press is an imprint of Taylor & Francis Group, LLC

© 2024 James W. Rohlf

Reasonable efforts have been made to publish reliable data and information, but the author and
publisher cannot assume responsibility for the validity of all materials or the consequences of
their use. The authors and publishers have attempted to trace the copyright holders of all material
reproduced in this publication and apologize to copyright holders if permission to publish in this
form has not been obtained. If any copyright material has not been acknowledged please write and
let us know so we may rectify in any future reprint.

Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced,
transmitted, or utilized in any form by any electronic, mechanical, or other means, now known or
hereafter invented, including photocopying, microfilming, and recording, or in any information
storage or retrieval system, without written permission from the publishers.

For permission to photocopy or use material electronically from this work, access www.copyright.
com or contact the Copyright Clearance Center, Inc. (CCC), 222 Rosewood Drive, Danvers, MA
01923, 978-750-8400. For works that are not available on CCC please contact mpkbookspermissions@
tandf.co.uk

Trademark notice: Product or corporate names may be trademarks or registered trademarks and are
used only for identification and explanation without intent to infringe.

ISBN: 978-1-032-49801-0 (hbk)


ISBN: 978-1-032-49686-3 (pbk)
ISBN: 978-1-003-39551-5 (ebk)

DOI: 10.1201/9781003395515

Typeset in Nimbus Roman


by KnowledgeWorks Global Ltd.

Publisher’s note: This book has been prepared from camera-ready copy provided by the authors.
Contents

Preface xiii

Chapter 1  Basis of Modern Physics 1


1.1 CHARGE AND THE ELECTRONVOLT 1
1.1.1 Elementary Charge 1
1.1.2 Strength of Electromagnetism 3
1.2 PLANCK’S CONSTANT 4
1.2.1 The Combination hc 4
1.2.2 Reduced Planck’s Constant (h̄) 5
1.2.3 Fine Structure Constant 6
1.2.4 Strength of Gravity 7
1.3 ENERGY 8
1.3.1 Mass and Momentum Units 8
1.3.2 Kinetic and Mass Energy 9
1.3.3 Momentum 10
1.3.4 Potential Energy 12
1.3.5 Nuclear Binding Energy 13
1.3.6 Q Value 14
1.4 THE PHOTON 16
1.4.1 Wavelength and Energy 16
1.4.2 Speed and Frequency 17
1.5 DE BROGLIE WAVELENGTH 18
1.5.1 Particle Wave Duality 18
1.5.2 Wavelength and Kinetic Energy 20
1.5.3 Classical Regime 20

v
vi  Contents

1.6 UNCERTAINTY PRINCIPLE 22


1.6.1 Position and Momentum 22
1.6.2 Energy and Time 23

Chapter 2  Thermal Radiation 25


2.1 RAYLEIGH-JEANS FORMULA 25
2.1.1 Average Oscillator Energy 25
2.1.2 Number of Modes 27
2.1.3 Power per Area 27
2.2 PLANCK FORMULA 29
2.2.1 Comparison with Rayleigh-Jeans 31
2.2.2 Planck Formula vs. Temperature 32
2.2.3 Radiation Peak 32
2.2.4 Planck Formula vs. Frequency 34
2.2.5 Number of Photons 34
2.3 STEFAN-BOLTZMANN LAW 36
2.4 WEIN’S LAW 40

Chapter 3  Key Processes 42


3.1 RADIOACTIVE DECAY 42
3.1.1 Decay Types 42
3.1.2 Decay Probability 43
3.1.3 Carbon Dating 44
3.2 MOTION OF A CHARGED PARTICLE IN ELECTRIC
AND MAGNETIC FIELDS 45
3.2.1 Charged Particle in an Electric Field 45
3.2.2 Charged Particle in a Magnetic Field 45
3.2.3 Electron Charge-to-Mass Ratio 46
3.2.4 Electron Charge Measurement 48
3.3 PHOTOELECTRIC EFFECT 49
3.4 ELECTRON DIFFRACTION 52
3.4.1 Scattering Off a Crystal 52
3.4.2 Neutrons in Thermal Equilibrium 53
3.4.3 Quarks Inside a Proton 54
Contents  vii

3.5 COMPTON SCATTERING 54


3.5.1 Compton Formula in Terms of Energy 55
3.5.2 Limiting Cases 56
3.5.3 Compton Formula in Terms of Wavelength 57
3.5.4 Relationship of λc to Other Quantities 57
3.6 RUTHERFORD SCATTERING 59
3.6.1 Effect of the Electrons 60
3.6.2 Scattering from a Nucleus 60
3.6.3 Cross Section 62
3.6.4 Scattering from a Thin Foil 65
3.7 THE WEAK INTERACTION 66
3.7.1 Weak Coupling 66
3.7.2 Neutrino Cross Section 66
3.7.3 Neutrino Scattering Rate 67
3.7.4 Neutrino Mean Free Path 67

Chapter 4  Special Relativity 69


4.1 BETA AND GAMMA 69
4.2 SPACE AND TIME 70
4.2.1 Time Dilation 70
4.2.2 Lorentz Contraction 73
4.3 ENERGY AND MOMENTUM 73
4.4 4-VECTORS 74
4.4.1 Invariant Mass 75
4.4.2 Center of Mass 75
4.5 LORENTZ TRANSFORMATION 76
4.5.1 Transformation of Time-Space 76
4.5.2 Transformation of Energy-Momentum 77
4.6 DOPPLER EFFECT 79
4.6.1 Colinear Light Source 79
4.6.2 Redshift 80
4.6.3 Observation at an Angle 80
viii  Contents

Chapter 5  Bohr Model 82


5.1 QUANTIZATION OF ANGULAR MOMENTUM 82
5.2 GROUND STATE 83
5.2.1 Bohr radius 83
5.2.2 Energy 83
5.3 EXCITED STATES 84
5.3.1 Orbits 84
5.3.2 Speeds 84
5.3.3 Energies 85
5.4 TRANSITIONS BETWEEN ENERGY LEVELS 85
5.4.1 Lyman Series 85
5.4.2 Balmer Series 86
5.4.3 Paschen Series 86
5.5 RYDBERG CONSTANT 87
5.6 REDUCED MASS 87
5.7 COLLAPSE OF THE BOHR ATOM 88
5.8 CORRESPONDENCE PRINCIPLE 89
5.8.1 Orbit and Radiation Frequency 89
5.8.2 Earth’s Orbit 90
5.8.3 LHC Proton 91

Chapter 6  Particle in a Box 92


6.1 THE POTENTIAL 92
6.2 THE SCHRÖDINGER EQUATION 93
6.3 SOLUTION 93
6.3.1 Wave Functions 94
6.3.2 Electron in a Box 94
6.3.3 Proton in a Box 96
6.4 COMPARISON WITH THE DE BROGLIE WAVELENGTH
AND THE UNCERTAINTY PRINCIPLE 96
6.5 EXPECTATION VALUES 97
6.5.1 Position 97
6.5.2 Momentum 98
6.6 CONSISTENCY WITH THE UNCERTAINTY PRINCIPLE 98
Contents  ix

6.7 CORRESPONDENCE PRINCIPLE 99


6.7.1 Phone in Box 99
6.7.2 Classical Probability 100
6.8 THREE DIMENSIONS 102
6.9 FINITE POTENTIAL 103
6.9.1 Solution Technique 103
6.9.2 Energy Condition 104
6.9.3 Solving for the Energy 106
6.9.4 Wave Functions 109

Chapter 7  Quantum Harmonic Oscillator 112


7.1 GROUND STATE 112
7.1.1 Wave Function 112
7.1.2 Energy 113
7.1.3 Normalization 114
7.1.4 Quantum Tunneling 114
7.1.5 Uncertainty Principle 115
7.2 FIRST EXCITED STATE 116
7.2.1 Wave Function 116
7.2.2 Energy 117
7.2.3 Normalization 117
7.2.4 Quantum Tunneling 117
7.2.5 Uncertainty Principle 118
7.3 GENERAL SOLUTION 119
7.3.1 Hermite Polynomials 119
7.3.2 Normalization 121
7.3.3 Energy 121
7.3.4 Wave Functions 122
7.4 CORRESPONDENCE PRINCIPLE 123
7.4.1 Large Quantum Numbers 123
7.4.2 Classical Probability 124
x  Contents

Chapter 8  Hydrogen Atom 127


8.1 GROUND STATE 128
8.1.1 Solution 128
8.1.2 Normalized Wave Functions 129
8.1.3 Radial Probability 130
8.1.4 Consistency with the Uncertainty Principle 132
8.2 FIRST EXCITED STATES 133
8.2.1 Wavefunctions 133
8.2.2 Energies 134
8.2.3 Radial Probability Distributions 134
8.3 MORE EXCITED STATES 136
8.4 CORRESPONDENCE PRINCIPLE 138
8.5 TRANSITIONS BETWEEN LEVELS 138
8.6 ELECTRON INTRINSIC ANGULAR MOMENTUM 141
8.6.1 Addition of Angular Momentum 141
8.6.2 Magnetic Moment 141
8.6.3 Zeeman Effect 142
8.6.4 Spin-Orbit Interaction 142
8.6.5 Hyperfine Splitting 144
8.6.6 Lamb Shift 148
8.6.7 The Electron g-Factor 150

Chapter 9  Statistical Physics 152


9.1 PROBABILITY DISTRIBUTIONS 152
9.1.1 Binomial Distribution 152
9.1.2 Poisson Distribution 155
9.1.3 Gaussian Distribution 156
9.2 MAXWELL-BOLTZMANN DISTRIBUTION 157
9.2.1 The Ideal Gas 157
9.2.2 Gas Pressure 158
9.2.3 Mean Free Path 158
9.2.4 Velocity Distribution 160
Contents  xi

9.2.5 Speed Distribution 161


9.2.6 Energy Distribution 164
9.3 QUANTUM DISTRIBUTIONS 165
9.3.1 Bose-Einstein Distribution 165
9.3.2 Fermi-Dirac Distribution 166
9.3.3 Comparison of the Distribution Functions 166
9.3.4 Density of States 167
9.3.5 Photon Gas 168
9.3.6 Electron Gas 169
9.3.7 Superfluid Helium 170

Chapter 10  Astrophysics 173


10.1 THE SUN 173
10.1.1 Proton Cycle 173
10.1.2 Distance to the Sun 174
10.1.3 Solar Constant 174
10.1.4 Temperature of Sun 175
10.1.5 Neutrino Flux from the Sun 175
10.2 MAGNITUDE SCALE FOR SKY OBJECTS 175
10.2.1 Apparent Magnitude 175
10.2.2 Absolute Magnitude 176
10.3 THE MILKY WAY 176
10.4 WHITE DWARF 177
10.5 NEUTRON STAR 178
10.5.1 Density 178
10.5.2 Fermi Energy 178
10.5.3 Binding Energy 179
10.6 BLACK HOLES 180
10.6.1 Schwarzschild Radius 180
xii  Contents

10.6.2 Hawking Radiation 180


10.7 THE DARK NIGHT SKY 181
10.8 HUBBLE’S LAW 182
10.8.1 Hubble Constant 182
10.8.2 Cosmic Redshift 182
10.9 COSMIC BACKGROUND RADIATION 183
10.10 COSMIC NEUTRINO BACKGROUND 184
10.11 CRITICAL MASS DENSITY 185
10.12 PLANCK MASS 185
10.12.1 Planck Length and Planck Time 186
10.12.2 Relationship to Schwarzschild Radius 187

Appendix A  Mathematica Starter 188


A.1 CELLS 188
A.2 PALETTES 189
A.3 FUNCTIONS 189
A.4 RESERVED NAMES 190
A.5 PHYSICAL CONSTANTS AND THEIR UNITS 191
A.6 INTEGRATION 193
A.6.1 Indefinite Integrals 193
A.6.2 Definite Integrals 193
A.6.3 Numerical Integration 194
A.6.4 Assumptions 194
A.7 RESOURCE FUNCTIONS 195
A.8 SERIES EXPANSION 195
A.9 SOLVING AN EQUATION 195
A.10 PLOTTING A FUNCTION 196

Appendix B  Physical Constants 197

Index 201
Preface

Modern physics uses combinations of physical constants which are unwieldy


in both magnitude and units. The beginning student can easily be over-
whelmed in the details of calculations that obscure the underlying physics
at precisely the time when getting a numerical answer is the key to learning.
This book is a guide to how to perform these calculations in Mathematica and
includes many examples that will be encountered in a first course in modern
physics. Mathematica is used as a data base for physical constants, a calcula-
tor with units, an algebraic manipulator with math functions, an integrator, a
differentiator, an equation solver, a series expander, a plotter, and more.
Appendix A is intended to help a beginner immediately get started in
Mathematica. One of the most important things to know at the outset is how
physical constants are stored and evaluated. The Mathematica names, sym-
bols, and numerical values of important quantities are given in App. B.
The body of the text puts the calculations in physics context. The cal-
culations are displayed as numbered examples, followed by executable code
labeled as In[ ]:=. The output is labeled Out[ ].

Example 5.3 Calculate the numerical value of the Bohr radius (n = 1).

The beauty of Mathematica is that the code is human readable because


the core part of the calculation looks like what you would see printed in a
book. No calculation should ever be done in physics without checking the
units. Mathematica has the extremely useful built in property that code will
not execute if the units do not match. In the example above, the expression,

4n2 πh̄2 0
,
e2 m

xiii
xiv  Preface

will not evaluate if its units are incompatible with the requested unit of
nanometers (nm). If the unit is not specified, the output will be in the in-
ternational system of units (SI), no matter what units were assigned to the
input. Thus, Mathematica is a powerful unit converter. This can be a huge
time saver.
The book contains a number of figures that are all drawn in Mathematica.
The chapters are largely independent of one another, except that the ma-
terial in Chap. 1 should be digested first, and Chap. 6 (Particle in a Box)
should come before Chaps. 7 (Quantum Harmonic Oscillator) or 8 (Hydro-
gen Atom).
CHAPTER 1

Basis of Modern
Physics

1.1 CHARGE AND THE ELECTRONVOLT


Charge is an intrinsic property of the electron. There is no such thing as an
electron without its charge. No free particles have ever been observed with
fractional electron charge.

1.1.1 Elementary Charge


The function Quantity[“ElementaryCharge”] is used to acquire the funda-
mental electric charge in Mathematica. The charge is elegantly displayed as
e and has units assigned. The quantity e is positive, so the proton charge is e
and the electron charge is −e. Details for how to get a physical constant with
units using the alternate natural language box are given in A.5 and a list of
constants is given in App. B.

Example 1.1 Get the elementary charge.

The output of Ex. 1.1 is written in a slightly different shade e to distin-


guish its representation of a unit from an italic e.
The unit of charge is the coulomb (C) which is an amp (A) times a second
(s). The coulomb unit is acquired with the function Quantity[“Coulombs”].

DOI: 10.1201/9781003395515-1 1
2  Guide to Modern Physics

Example 1.2 Get the unit coulomb.

The function UnitConvert[ ] converts the symbolic unit to a number in


a specified unit. The function N[ ] displays that number in decimal form to
the specified number of digits. It is important to note that the first argument
of UnitConvert [e,C] displayed in the input cell of ex. 1.3, is a stored unit as
generated in ex. 1.1. It is the abbreviation for Quantity[“ElementaryCharge”].
The second argument is the coulomb unit obtained in ex. 1.2. It is the abbre-
viation for Quantity[“Coulombs”].

Example 1.3 Get the numerical value of the elementary charge in C to 4


figures.

The preferred unit of energy in modern physics is the electronvolt (eV),


defined as the kinetic energy gained by an electron when accelerated through
a potential difference of 1 volt (V). The eV is the energy scale of outer elec-
trons in atoms. Similarly, keV is the scale of inner electrons in atoms, MeV is
the scale of protons in the nucleus, and GeV is the scale of energetic quarks
in the proton. There is a subtle and significantly important property of this
unit in that the elementary charge (e) × 1 V is equal to 1 eV. One may take
advantage of this in many numerical calculations.
For the input cell of ex. 1.4, V is the unit from Quantity[“Volts”] and eV
is the unit from Quantity[“ElectronVolts”].

Example 1.4 Compare e times 1 V to 1 eV (the “==” perfoms a logical


comparison).

One e per coulomb (C) is equal to 1 eV per joule (J). Thus, the elementary
charge of 1.602×10−19 C gives the eV to be 1.602×10−19 J.
Basis of Modern Physics  3

e eV
Example 1.5 Compare the units C to J and calculate the numerical value
of the ratio.

1.1.2 Strength of Electromagnetism


The strength of the electric force between 2 elementary charges (e) is written
e2
F= ,
4πε0 r2
where r is the separation distance and ε0 is the electric constant.
Example 1.6 Get the electric force constant.

e2
The preferred unit of 4πε0 in modern physics is eV·nm.
e2
Example 1.7 Calculate the value of 4πε0 in units eV·nm.

The units are always output in the same order (mass, length, time, etc.)
no matter what order they are input.
The unit eV/nm is a unit of force that is practical to use at the atomic
scale. The attractive force between an electron and proton separated by a dis-
tance of 0.1 nm is
e2 1.44eV · nm 14.4 eV
F= 2
= = .
4πε0 r (0.1 nm)2 0.1 nm
The strength of the electric force in these units can be identified with the
atomic energy and distance scales.
4  Guide to Modern Physics

1.2 PLANCK’S CONSTANT


At the heart of modern physics lies the fact that measured quantities such as
energy and angular momentum are not continuous. Instead, they take on dis-
crete values, a concept referred to as quantization. Planck’s constant (h) sets
the scale for this granularity which is not observable in classical physics. The
unit of Planck’s constant is that of momentum × distance or energy × time,
 
kg · m
(m) = (J · s).
s

Planck’s constant is extremely tiny compared to 1 kg · m2 /s. Even compared


to a microscopic scale, 1 µg · (µm)2 /s, Planck’s constant is still very tiny
which is why the quantization goes unnoticed in the classical regime.

Example 1.8 Get Planck’s constant.

The function NumberForm[ ] is one of several ways to output a numerical


value.

Example 1.9 Get the numerical value of Planck’s constant in SI units.

A more suitable unit is eV·s.

Example 1.10 Get the numerical value of Planck’s constant in eV · s.

1.2.1 The Combination hc


Planck’s constant appears naturally together with the speed of light (c).
Basis of Modern Physics  5

Example 1.11 Get the speed of light.

Example 1.12 Get the numerical value of the speed of light in vacuum.

The product hc is relatively easy to remember and is extremely useful in


calculations involving h.

Example 1.13 Get Planck’s constant times the speed of light.

For higher energies of the nuclear scale, one may note that

1240 eV · nm = 1240 MeV · fm.

The numerical value of hc is revealing that smaller distances are associated


with higher energies. This is the central theme of modern physics.

1.2.2 Reduced Planck’s Constant (h̄)


The quantization is commonly written in terms of the “reduced” Planck’s
constant (h̄), called “h-bar”,
h
h̄ = .

Example 1.14 Get the reduced Planck’s constant.

The quantity h̄c appears frequently and is also easy to remember.


6  Guide to Modern Physics

Example 1.15 Get the reduced Planck’s constant times the speed of light.

Again,
197 eV · nm = 197 MeV · fm.

1.2.3 Fine Structure Constant


The strength of the electromagnetic force is made dimensionless by dividing
by h̄c, giving the famous fine structure constant (α),
 2 
e
4πε0
α= .
h̄c
Example 1.16 Get the fine structure constant.

The fine structure constant is historically written as the ratio of integers.


The function Rationalize [ ] will do this.

Example 1.17 Calculate α, rationalized to order 10−5 .

The fine structure constant of Ex. 1.16 can be compared with e2 /4πε0 h̄c.

e2
Example 1.18 Compare the quantity α to 4πε0 h̄c .
Basis of Modern Physics  7

The fine structure constant appears frequently in atomic physics which is


described by the electromagnetic force strength combined with the quantiza-
tion in terms of Planck’s constant. For example, the root-mean-square (rms)
speed v of an electron in the hydrogen atom is

v = αc.

Example 1.19 Calculate the rms speed and kinetic energy of an electron in
hydrogen.

The atomic electron is not relativistic.

1.2.4 Strength of Gravity


The gravitational force between an electron and a proton can be written
Gme mp
F=
r2
Example 1.20 Get the universal gravational constant G.

The strength of gravity between and electron and proton can be made di-
Gm m
mensionless with the expression h̄ce p which gives the gravitational equiva-
lent of α. The particle masses may be obtained with the natural language box
(A.5). They are discussed further in 1.3.1.
8  Guide to Modern Physics

Example 1.21 Calculate the dimensionless strength of gravity.

Gravity is 40 orders of magnitude weaker than the electric force.

1.3 ENERGY
1.3.1 Mass and Momentum Units
A convenient mass unit is MeV/c2 , where c is the speed of light in vacuum.
Masses may be obtained with either the function Quantity[ ] or the natural
language box.

Example 1.22 Get the electron mass with both the Quantity[ ] function and
the natural language box.

Example 1.23 Compare the 2 methods of acquiring the electron mass.

These representations of the mass are used interchangeably in this work.


Basis of Modern Physics  9

Example 1.24 Get the proton mass with both the Quantity[ ] function and
the natural language box.

The preferred momentum unit is eV/c (or keV/c, MeV/c, etc.) depending
on the scale. This is convenient whether or not the electron is relativistic. The
formula for relativistic momentum is given in 1.3.3.
Example 1.25 Calculate the momentum in eV/c for an electron with speed
v = 104 m/s.

The relationship between momentum (p) and nonrelativistic kinetic en-


ergy (K) is
p2
K= .
2m
Example 1.26 Calculate the momentum in MeV/c of an electron whose ki-
netic energy is 10 keV.

1.3.2 Kinetic and Mass Energy


There are two forms of energy: energy due to motion (kinetic energy) and
energy stored as mass (mass energy). The mass energy (E0 ) is

E0 = mc2 .
10  Guide to Modern Physics

The total energy (E) is


E = mc2 + K.

1.3.3 Momentum
If the particle is relativistic, the rule for momentum (as observed to be con-
served in the lab) is
mv
p=  .
2
1 − vc2
and the total energy is
 
2 p2
E = (mc2 )2 + (pc)2 = mc 1+ 2 2.
m c
The second term inside the square root is a small number if v << c, and the
kinetic energy reduces to the familiar nonrelativistic form.

Example 1.27 Get the nonrelativistic kinetic energy from the relativistic en-
ergy equation.

In Ex. 1.27, the variables are user defined (not units).

Example 1.28 Calculate the fractional error in the nonrelativistic formula


for momentum when v = 0.1 c .
Basis of Modern Physics  11

The expression is good to 0.5% when v = 0.1 c, so that is a good approxi-


mate number for the boundary at which one can get a realistic answer without
using the relativistic expression.
The exact expression for the kinetic energy is

K = (mc2 )2 + (pc)2 − mc2

# !()*"

!"%

!",

!"$

!"+

!"#

!"'

!" !()*"
!"# !"$ !"% !"& '"!

Figure 1.1 The plot shows the kinetic energy of an electron vs. momentum
times c. At low momentum the expression is quadratic but at high momentum
it is linear.

Example 1.29 Calculate the kinetic energy and speed of an electron with
momentum 1 MeV/c.
12  Guide to Modern Physics

1.3.4 Potential Energy


There is another energy that is called the potential energy U. It is not a new
type of energy but is a binding energy related to the mass and kinetic ener-
gies. In the description of the hydrogen atom, the mass energy is left out and
the total energy is written as the sum of kinetic (Ex. 1.19) plus potential,

K + U = −13.6 eV,

where the potential energy has been calculated from

e2
U =− .
4πε0 r
(In the language of quantum mechanics, both the position r and the momen-
tum p are smeared out and the above energies become averages.) The neg-
ative sign means that the force is attractive and the electron is bound. An
energy of 13.6 eV must be added to free the electron.

Example 1.30 Get the energy needed to ionize a hydrogen atom.

This means that


mH c2 + 13.6 eV = me c2 + mp c2
and using eV/c2 as a mass unit,

mH = me + mp − 13.6 eV/c2

The mass of the hydrogen atom is smaller than the sum of the masses of its
parts. The value of U is negative.

U = mH c2 − (me c2 + mp c2 ) − K

In the lowest energy (most bound) state of the hydrogen atom, the (average)
electron kinetic energy is
K = 13.6 eV
and the (average) potential energy is

U = −27.2 eV
Basis of Modern Physics  13

In the describing atom, one doesn’t worry about the constant mass ener-
gies of the electron and proton and they are left out of the expression for total
energy,
E = U + K,
This is understood to mean
U = −27.2 eV + me c2 + mp c2 ,
and
E = U + K = m H c2 .
One has to know from the context (GeV vs. eV scale), if the shortcut of leav-
ing out the mass part has been made.

1.3.5 Nuclear Binding Energy


The binding energy of an atomic nucleus is calculated by summing the mass
energies of all the neutrons and protons and subtracting the mass energy of
the nucleus.
Example 1.31 Get the neutron mass.

The neutron mass is larger than the proton mass (Ex. 1.24) by about 1.3
MeV/c2 .
The alpha particle is the nucleus of helium and is composed of 2 neutrons
and 2 protons. It is very stable.
Example 1.32 Calculate the binding energy of the alpha particle.

The uranium-238 nucleus has 92 protons and 146 neutrons.


14  Guide to Modern Physics

Example 1.33 Calculate the binding energy of 238 U.

Notice that Mathematica reports the atomic mass of 238 U so the electrons
need to be subtracted from the atomic mass to get the nuclear mass.
A neutron is added to an oxygen-16 nucleus to make an oxygen-17 nu-
cleus,
n +16 O →17 O + energy.

Example 1.34 Calculate the binding energy of the neutron.

1.3.6 Q Value
The energy released in a decay (Q) is the mass energy of the decaying paricle
minus the sum of the mass energies of the decay products. The neutron decays

n → p + e + ν̄e .

The neutrino (ν̄e ) is essentially massless. This process is called β− decay.

Example 1.35 Calculate the Q value for neutron decay.


Basis of Modern Physics  15

The decay cannot occur if Q is negative, but it is possible for a bound


proton to convert into a neutron inside the nucleus,

p → n + e+ + νe .

This process is called β+ decay. The positron (e+ ) has the same mass as the
electron. The β+ process cannot happen for a free proton because the proton
mass is smaller than the neutron mass. An example is the decay of boron-8,
8
B →8 Be + e+ + νe .

Example 1.36 Calculate Q for β+ decay of 8 B.

Another example of the same process is the fusion of 2 protons to make


a proton-neutron bound state called the deuteron (d),

p + p → d + e+ + νe .

Example 1.37 Calculate Q for the fusion of 2 protons.


16  Guide to Modern Physics

Note that Q of the process is subtly different from the binding energy of
the deuteron because the positron mass must be included. The positron will
annihilate with an electron and produce energy in the form of 2 photons.
The fusion process generally occurs for elements lighter than iron, but
does not occur for heavier elements.

Example 1.38 Show that the decay 230 U →229 Pa + p + e + ν̄e cannot occur.

The Q value is negative so the decay cannot occur.

Example 1.39 Show that the decay U230 → U229 + n + e+ + νe cannot occur.

The Q value is negative so the decay cannot occur.

1.4 THE PHOTON


1.4.1 Wavelength and Energy
A photon is a particle (quantum) of light. It has zero mass so its energy is
pure kinetic. The photon energy as a function of wavelength (λ) is
hc
E=
λ
Example 1.40 Calculate the energy range of visible photons (wavelengths
400 nm to 700 nm).
Basis of Modern Physics  17

The wavelength of a 10 keV x ray is


hc 1240 eV · nm
λ= = = 0.124 nm
E 104 eV
and a 1 MeV γ ray has
1240 eV · nm
λ= = 1240 fm
106 eV
Example 1.41 Calculate the energy in meV of a 3 cm wavelength microwave
photon.

1.4.2 Speed and Frequency


The relationship between wave speed (c), wavelength (λ), and frequency ( f )
is
c = λf
Photon frequency is usually given in hertz (Hz) which is just an inverse sec-
ond.

Example 1.42 Calculate the frequency range of visible photons.

The relationship between energy and frequency is

E = hf
18  Guide to Modern Physics

For example, the energy of photons from a 100 MHz radio broadcast is

E = h f = (4.14 × 10−15 eV · s)(100 × 106 s−1 ) = 4.14 × 10−7 eV

1.5 DE BROGLIE WAVELENGTH


1.5.1 Particle Wave Duality
The light wave has a particle (photon) interpretation, and the electron has a
wave behavior. This is the single most important concept in modern physics.
It is the small mass of the electron that makes its wave properties observable.
The de Broglie wavelength of any particle is defined in terms of its momen-
tum as
h hc
λ= = .
p pc
The expression works for relativistic particles, provided one uses the rela-
tivistic expression for momentum. Note that for a photon, E = pc, resulting
in the usual formula, E = hc
λ.

λ (nm)

10-6

10-7

10-8

10-9

10-10

10-11

p (eV/c)
10 100 1000 104 105 106

Figure 1.2 The wavelength is shown vs. momentum on a log-log plot.

Example 1.43 Calculate the de Broglie wavelength of a an electron whose


speed is αc.
Basis of Modern Physics  19

The de Broglie wavelength of an outer electron in an atom gives the size


of the atom. Since outer electrons have a limited range energy of a few eV
(and therefore, a limited momentum range), all atoms are roughly the same
size.

#&

#!

"&
02

"!

&

!
! "! #! $! %! &!
'()*+, -.*/01

Figure 1.3 The atomic ionization energy is shown vs. atomic number.

Example 1.44 Calculate the de Broglie wavelength in fm of a neutron whose


kinetic energy is 10 MeV.

The de Broglie wavelength of a nucleon inside the nucleus gives the size of
the nucleus.
20  Guide to Modern Physics

Example 1.45 Calculate the de Broglie wavelength of a quark whose mo-


mentum is 200 MeV/c.

The de Broglie wavelength of a quark inside a proton gives the size of the
proton.

1.5.2 Wavelength and Kinetic Energy


The relationship between momentum and kinetic energy ,

K = E − mc2 = (pc)2 + (mc2 )2 − mc2 ,

gives 
pc = K 2 + 2mc2 K,
and
hc hc
λ= = √
pc K + 2mc2 K
Example 1.46 Calculate the wavelength of an electron that has a kinetic en-
ergy of 2 MeV.

1.5.3 Classical Regime


In the regime of classical physics, de Broglie wavelengths are unmeasurably
small.
Basis of Modern Physics  21

!+,"

#!%

#!!!

!"#

#!#*

#!#)
& !'("
!"!# #! #!$ #!% #!#!

Figure 1.4The de Broglie wavelengths of a photon, electron, and proton are


shown vs. kinetic energy.

Example 1.47 Calculate the de Broglie wavelength of the earth in its orbit
about the sun.

Example 1.48 Calculate the de Broglie wavelength of a 5 gram snail moving


at 0.03 miles per hour.

Even if we assign a ridiculously small speed to the snail, its wavelength is


tiny.
22  Guide to Modern Physics

Example 1.49 Calculate the de Broglie wavelength of a 5 gram snail moving


at a speed of one atomic distance per year.

1.6 UNCERTAINTY PRINCIPLE


1.6.1 Position and Momentum
The wave nature of particles makes it impossible to simultaneously know
both the location (∆x) and momentum (∆p) of a particle with infinite preci-
sion. The limit is

∆x∆p ≥ .
2
The conceptual reasoning is that the act of observing the position necessitates
an interaction with at least one photon of wavelength equal to the position ac-
curacy (λ ∼ ∆x), and the collision of that photon produces an uncertainty in
momentum (∆p ∼ λh ), resulting in

∆x∆p ∼ h.

Example 1.50 Calculate the minimum uncertainty in momentum for a par-


ticle whose position is known to 0.1 nm.

Note that the calculation of momentum above does not depend on the parti-
cle mass or whether or not the particle was relativistic. Suppose the confined
particle is an electron.

Example 1.51 Calculate the minimum kinetic energy of an electron whose


momentum is equal to ∆p as calculated above, corresponding to ∆x = 0.1 nm.
Basis of Modern Physics  23

The use of the nonrelativistic form of kinetic energy is seen to be justified


since K << mc2 (Ex. 1.22).

1.6.2 Energy and Time


A second form of the uncertainty principle says that

∆E∆t ≥ .
2
Suppose (as once thought) that the strong force was mediated by exchange
of a massive particle m whose range is governed by the uncertainty principle.
The minimum energy of the particle is

∆E = mc2

and the maximum range is


c∆t ≈ 1 fm.

Example 1.52 Use the uncertainty principle to estimate the mass energy of
the exchanged particle.

In beta decay, weak force is mediated by the W particle whose large mass
gives the force an extremely short range, even compared to that of the strong
force.

Example 1.53 Get the W boson mass.


24  Guide to Modern Physics

Example 1.54 Use the uncertainty principle to estimate the range of the
weak force.
CHAPTER 2

Thermal Radiation

2.1 RAYLEIGH-JEANS FORMULA


In its classical interpretation, radiation comes from atomic oscillators having
a continuous probability distribution, which is distributed exponentially. The
exponential distribution has a number of remarkale properties.
Example 2.1 Calculate the integral, average, and root-mean-square (rms)
deviation from the average for the exponential distribution e−x .

2.1.1 Average Oscillator Energy


To get the total radiation energy density, one needs to count the number of os-
cillators and then multiply by the average energy per oscillator. The average
energy comes from an exponential probability distribution. For equilibrium
at temoperature T , the probability per energy f (E) is written
E
f (E) = Ce− kT dE.
where k is the Boltzmann constant that converts temperature into energy, and
C is a normalization constant making unit probability after summing over all
energies.
Example 2.2 Get the Boltzmann constant.

DOI: 10.1201/9781003395515-2 25
26  Guide to Modern Physics

!!"

!('

'(*

'()

'($

'("

&
! " # $ %

Figure 2.1 The exponential distribution, e−x , has the remarkable property the
the integral, average, and rms deviation from the average are all unity.

Example 2.3 Calculate the average energy per atomic oscillator.

In Ex. 2.3, the Greek letter E is used to avoid a reserved name (A.4),
and k is entered as a user-defined variable (not the unit of Ex. 2.2) to avoid
substitution of a number in the output.
An exponential distribution of oscillators gives an average energy of kT .

Example 2.4 Calculate the approximate value of kT at 290 K.

It is handy to remember that at room temperature,


1
kT ≈ eV.
40
In Ex. 2.4, k is used as the unit.
Thermal Radiation  27

2.1.2 Number of Modes


The next step is to count the number of oscillation modes. Consider a stand-
ing wave in one dimension of length L and wavelength λ. For 3 dimensions,
consider a spherical volume 34 πL3 and divide by the volume of one mode, λ3 .
This gives
4
πL3
N=23 3 ,
λ
where the factor of 2 accounts for the 2 polarizations of electromagnetic
waves. The number of states per L3 , ns (λ), is
N 8π
ns = 3
= 3.
L 3λ
The number per volume per unit wavelength, referred to as the density of
states n(λ), is
dns 8π
n(λ) = − = .
dλ λ4
The energy per volume per wavelength stored in the cavity is the density of
states times the average energy.
8πkT
u(λ) = n(λ)kT = .
λ4

2.1.3 Power per Area


There is one more thing to be done which is to uncover the relationship be-
tween energy per volume in a cavity and power per area radiated from its
surface. Consider radiation at an angle θ to some surface (Fig. 2.2). The
power radiated per wavelength (R) is related to the energy per volume per
wavelength (u) by
1 uV 1 uLA cos θ 1
R= = = cuA cos2 θ),
2 ∆t 2 L/(c cos θ) 2
where the 1/2 comes from only one-half of the radiation in the cavity moving
in a direction away from the wall. Averaging over all angles gives another
factor of 1/2. Thus, the fundamental relationship between volume energy and
power radiated from the surface is
c
R = u.
4
28  Guide to Modern Physics

!
θ

Figure 2.2 Radiation at an angle θ from a cavity wall. The time for radiation to
cross the cavity is L/(c cos θ) and the effective area from which the radiation
originates is reduced by a factor cos θ (all the radiation makes it across when
θ = 0 and none makes it across when θ = π2 ).

Example 2.5 Calculate the average value of cos2 θ from − π2 to π2 .

The power per area per wavelength radiated from the surface is
c 2πckT
R(λ) = u(λ) = .
4 λ4
This is the famous Rayleigh-Jeans formula (Fig. 2.3). It only works in the
limit of long wavelengths, suffering from what is referred to as the “ultravio-
let catastrophe”, as it blows up when λ → 0. Solution of this problem was the
birth of modern physics at a time when it was once widely thought that that
all of fundamental physics was known!
The power per area due to photon wavelengths in an interval

λ1 < λ < λ2

may be obtained by direct integration (area under the curve of Fig. 2.3).

Example 2.6 Calculate the power per area radiated for 1 m < λ < 2 m at
300 K.
Thermal Radiation  29

*$!" +,- %

!&#)

!&#(

!&#'

!&#!&

!&#!!

!&#!"

!&#!#
!%"
! " # $

Figure 2.3Radiated power per area per wavelength vs. λ predicted by the
Rayleigh-Jeans formula at 300 K.

This result is reasonable but at smaller wavelengths something goes badly


wrong.

Example 2.7 Calculate the power per area radiated for 400 nm < λ < 700 nm
at 300 K.

This is clearly not correct: objects are not glowing from radiation in the vis-
ible spectrum at room temperature!

2.2 PLANCK FORMULA


The small-wavelength correction to the Rayleigh-Jeans formula was obtained
by Planck who assumed that the atomic oscillators that produce the radiation
do not have a continuous distribution of energy levels, but that they were
30  Guide to Modern Physics

quantized with energies


En = nh̄ω.
A photon emitted from the transition between adjacent levels has an energy
hc
Ep = En+1 − En = h̄ω = .
λ
The levels still have an exponential drop with energy, but are quantized with
a spacing
hc
∆E = En+1 − En = .
λ
Figure 2.4 shows a quantized probability distribution with energy spacing of
kT
2 . The average energy has been reduced to 0.77 kT compared to kT for
a continuous distribution. The average energy drops rapidly with increased
spacing.

()*+,+-.-/0 1-2/)-+3/-*4

%'&

&'$

&'#

&'"

&'!

! !"#"!#
! " # $ %&

Figure 2.4Exponential energy distribution for quantized energy levels with


spacing kT/2.

Example 2.8 Calculate the average energy per oscillator for a quantized dis-
tribution with kT
2 energy spacing.
Thermal Radiation  31

Example 2.9 Calculate the average energy per oscillator for a quantized dis-
tribution with 10kT energy spacing.

At room temperature, this corresponds to about a factor of 2000 lower at


λ = 5 × 10−6 m.

Example 2.10 Calculate the average energy per oscillator in terms of λ for
energy spacing hc/λ.

Note that Mathematica chooses the order of the symbols in the output.
The famous universal thermal (blackbody) radiation formula derived by
Planck is obtained by replacing kT in the Rayleigh-Jeans formula (2.1.3) with
the average energy per oscillator from Ex. 2.10, giving

2πh̄c2
R= hc
.
λ5 (e λkT − 1)
This is the correct answer for the observed power per area per wavelength
radiated by an object in thermal equilibrium at a temperature T .

2.2.1 Comparison with Rayleigh-Jeans


At large λ the blackbody formula reduces to the Rayleigh-Jeans for-
mula.

Example 2.11 Calculate the leading term in the blackbody formula at large
λ.
32  Guide to Modern Physics

($!' )*+ &

!"!$

!"!#

!"!'#

"&#
!"!# !"!$ "%"! !

Figure 2.5 The Rayleigh-Jeans (upper) is compared to the Planck blackbody


distribution (lower) at 300 K.

2.2.2 Planck Formula vs. Temperature


At higher temperatures, the total power radiated increases rapidly and peaks
at lower wavelengths. The temerature 300 K (5800 K) is comparable to the
surface of the earth (sun).

2.2.3 Radiation Peak


Plotting the blackbody radiation on a linear scale makes it easier to visualize
the location of the peak.

The location of the peak of the distribution is found by setting the derivative
equal to zero. The calculation is done over the domain of real numbers.
Thermal Radiation  33

*$!) +,- &


!"!)

!""

!"!(

!"!!(

!"!)(

!"!'(
"&#
!"!# !"!$ "%""! "%!""

Figure 2.6 The blackbody distribution is shown vs. λ at 300 K and 5800 K.

)#!$ *+, #! (

%!!

$'!

$!!

#'!

#!!

'!

!("
!"!!!!# !"!!!!$ !"!!!!% !"!!!!& !"!!!!'

Figure 2.7 The blackbody distribution is shown vs. λ at 300 K on a linear


scale.

Example 2.12 Get the symbol for the real domain.

Example 2.13 Calculate the wavelength at which the blackbody distribution


peaks at T = 300 K.
34  Guide to Modern Physics

2.2.4 Planck Formula vs. Frequency


To get the radiation formula as a function of photon frequency ( f ), make a
change of variables to the entire Planck formula, not forgetting the differen-
tial,
c
λ→ ,
f
c
dλ = − 2 d f,
f
and
2πh f 3
R( f ) = hf
c2 (e kT − 1).
The negative sign in the differential, which just says f increases as λ de-
creases, is ignored.
Example 2.14 Calculate the frequency at which the blackbody distribution
peaks at T = 300 K.

2.2.5 Number of Photons


To get the photon flux (number per s per m2 ), convert the freqeuncy scale to
energy (E = h f ) and divide by the photon energy.
1 2πE 2
F(E) = R(E) = E
E (hc)2 (e kT − 1)
Thermal Radiation  35

W/m2 per 1013 s-1

150

100

50

f (s-1 )
2 × 1013 4 × 1013 6 × 1013 8 × 1013

Figure 2.8 The blackbody distribution is shown vs. f at 300 K on a linear


scale.

photons per (m2 s) per eV

4 × 1065

3 × 1065

2 × 1065

1 × 1065

E (eV)
0.05 0.10 0.15 0.20

Figure 2.9 The photon flux is shown vs. energy at 300 K.

Example 2.15 Calculate the flux of visible photons at T = 300 K. A variable


kT is defined to facilitate the numerical integration.
36  Guide to Modern Physics

Example 2.16 Calculate the flux of visible photons at T = 2000 K.

2.3 STEFAN-BOLTZMANN LAW


If we integrate the power per area per wavelength over all wavelengths, the
answer depends on only one parameter, the temperature, which appears to the
fourth power.
Thermal Radiation  37

Example 2.17 Integrate the thermal radiation formula over all wavelengths
to get the total radiated power per area.

)#"#

!"'

!"""

! !("
!""" #""" $""" %""" &""" '"""

Figure 2.10 The total radiated power per area is shown vs. temperature.

Example 2.18 Find the temperature where the radiated power per area is
1 MW/m2 .

Example 2.19 Find the value of kT where the radiated power per area is
1 MW/m2 .
38  Guide to Modern Physics

The Stefan-Boltzmann law for the radiated power per area is written
P
= σT 4
A
where the constant
2π5 k4
σ=
15h3 c2
is called the Stefan-Boltzmann constant.

Example 2.20 Get the Stefan-Boltzmann constant.

Example 2.21 Calculate the Stefan-Boltzmann constant.

It is may be useful to express this in eV, for example, if we are counting


photons.

Example 2.22 Calculate the Stefan-Boltzmann constant using eV for energy


units.
Thermal Radiation  39

It may be useful to factor out k4 so we have something (σ ) times (kT )4 .


P
= σ (kT )4
A
Example 2.23 Calculate σ .

1
Example 2.24 Calculate the radiated power per area for kT = 40 eV.

Example 2.25 Find the total radiated power per area from an object at room
temperature.

Example 2.26 Get the surface temperature of the sun.

Example 2.27 Calculate the power per area radiated from sun.
40  Guide to Modern Physics

Example 2.28 Calculate the total power radiated from sun.

2.4 WEIN’S LAW


The peak wavelength value of the thermal spectrum is found by tak-
ing its derivative with respect to λ and setting it equal to 0. This re-
sults in a transcendental equation which can be solved numerically with
the use of the function NSolve [ ]. The peak value occurs at about
hc
5kT .

Example 2.29 Find λ that gives the peak radiated power per area.

This is expression commonly written as λT equals a constant.

Example 2.30 Find the value of λT which gives the wavelength for maxi-
mum power.

This is Wein’s law. The peak wavelength is given by

λmax T = 0.002898 m · K.
Thermal Radiation  41

Example 2.31 Find the wavelength for maximum power from the sun.

The human eye has evolved to have sensitivity at the wavelength where
the sun puts out its peak power.
CHAPTER 3

Key Processes

3.1 RADIOACTIVE DECAY


Radioactive decay produced the mystery particles α, β, and γ (He nucleus,
electron, photon) and played a central role at the beginning of the modern
physics era. Study of these decays revealed the structure of the atom at a time
when α, β, and γ were still unknown radiation phenomena.

3.1.1 Decay Types


Decay always occur if they are not forbidden. Alpha decay is a common de-
cay of a heavy nucleus. The process is

Z−2 Y + α,
A A−2
ZX →

where the decaying nucleus has shed both 2 neutrons and 2 protons that are
bound into the α particle.
Beta decay has two types, where either a neutron turns into a proton in-
side the nucleus,
+
Z X → Z+1 Y + β + νe ,
A A

or a proton turns into a neutron inside the nucleus,

Z−1 Y + β + ν̄e ,
A A −
ZX →

where the β− (β+ ) particle is an electron (positron).


Related to β decay is the process of electron capture,

ZX +β Z−1 Y + νe ,
A − A

where a nucleus has captured an inner atomic electron and turned a proton
into a neutron.

DOI: 10.1201/9781003395515-3 42
Key Processes  43

Gamma decay occurs when a nucleus has been left in an excited state
(X ∗ ),
A ∗ A
Z X →Z X + γ.

3.1.2 Decay Probability


Radioactive decays are governed by an exponential distribution (see 2.1),

N(t) = N0 e−t/τ ,

where N(t) represents the size of the sample as a function of time, N0 is the
sample size at time t = 0, and τ is the average lifetime.

Example 3.1 Calculate the average lifetime.

The exponential distribution has the special property that the rms deviation
from the mean is equal to the mean.

Example 3.2 Calculate the rms deviation from the mean lifetime.

The half-life (t1/2 ) is defined by

1
= e−t1/2 /τ
2
which gives
t1/2 = τ ln 2.
Uranium-238 decays by α emission,
238
U →234 Th + α.
44  Guide to Modern Physics

Example 3.3 Calculate the initial decay rate in a 1 g sample of U-238.

Thorium-234 decays by β emission,


234
Th →234 Pa + e− + ν̄e .

Example 3.4 Calculate the initial decay rate of Th-234.

3.1.3 Carbon Dating


The isotope carbon-14 and normal carbon-12 are in atmospheric equilibrium
in the ratio 1.2 × 10−12 . Living organisms contain this natural mix of 12 C and
14 C, but upon death no longer take in any more 14 C which then decays away

with a half-life of 5730 years.

Example 3.5 An old bone has a 14 C fraction that has dropped by 20% from
atmospheric equilibrium to 0.96 × 10−12 . How old is the bone?
Key Processes  45

3.2 MOTION OF A CHARGED PARTICLE IN ELECTRIC AND


MAGNETIC FIELDS
3.2.1 Charged Particle in an Electric Field
Newton’s 2nd law for a charged particle in an electric field (E) is
q∆V
F = qE = ,
d
where the field is given by a potential difference (∆V) across a distance (d).
This force is most conveniently expressed in eV/m (1.1.2).

Example 3.6 Calculate the force on an electron in an electric field of


100 V/cm in both eV/m and N.

3.2.2 Charged Particle in a Magnetic Field


Newton’s 2nd law for the circular motion of radius r for a particle with mass
m and charge q in a magnetic field B is
v2
F=m = qvB,
r
46  Guide to Modern Physics

which gives
mv = p = qrB.
If the electron is relativistic, the same equation, p = qrB, holds provided that
the relativistic expression for momentum (1.3.3) is used.

Example 3.7 Find the radius of curvature of a 10 keV/c electron in a 1 mT


magnetic field.

3.2.3 Electron Charge-to-Mass Ratio


The charge-to-mass ratio of the electron was measured by J. J. Thompson
who built the first mass spectrometer consisting of parallel plates having per-
pendicular electric and magnetic fields. The idea of the spectrometer is that
an electron velocity is very difficult to measure but its path is relatively easy.
With no magnetic field, the electron gets an upward acceleration (a) of
F qE
a= = .
m m
The upward component of the electron velocity (vy ) is

aL
vy = at = ,
vx
where (L) is the distance traveled between the plates, t is the travel time, and
v x is the initial electron velocity. Thus,
q v x vy
= .
m EL
Key Processes  47

When the magnetic field is also applied with a strength that balances the elec-
tric force,
qE = qv x B.
Thus, using
E
vx =
B
and vy
tan θ = ,
vx
the result may be written
q E tan θ
= .
m LB2
In the above expression, E is the electric field that causes a deflection θ and
B is the magnetic field that causes no deflection.

( )!'*+ $))
" " !!
#
!"#$%!"& ' ( )!'*+ $"

Figure 3.1 A region of space has an electric field in the downward direction
and a magnetic field into the page. When the electric and magnetic forces can-
cel, an incoming electron is undeflected. When the magnetic field is turned
off, the electron is deflected upward.

Example 3.8 Calculate the electron speed for a θ = 0.1 (radian) deflection
from an electric field of 104 V/m and L = 5 cm.
48  Guide to Modern Physics

This is an atomic speed (1.2.3); the electrons used for this measurement came
from atoms.

3.2.4 Electron Charge Measurement


The electron charge was measured by R. Millikan in his famous oil-drop ex-
periment. A charged droplet of mass m suspended in air reaches a terminal
speed vT given by the force balance

mg = bvT = 6πηRvT ,

where the constant b is proportional to both the viscosity of air (η) and the
droplet radius (R) through a relationship called Stokes’s Law. Using the for-
mula for density (ρ),
m
ρ= 4 ,
3 πR
3

the droplet radius can be calculated as


s
9vT η
R= .
2gρ

If an electric field is now applied in a direction to make the droplet move


upward, the force balance is

qE = mg + bvE ,

where vE is the electron velocity with the field on. This can be used to elim-
inate b to get
mg(1 + vvET )
q= .
E

Example 3.9 For vT = 1.3 mm/s and ρ = 0.9g/cm3 , determine b, R, and m.


Key Processes  49

The electron charge measurement combined with the charge-to-mass ra-


tio measurement (3.2.3) gives the electron mass.

3.3 PHOTOELECTRIC EFFECT


The photoelectric effect (PE) occurs when a photon has sufficient energy to
eject an electron from the metal. The electron which was originally bound
has absorbed the photon and become free. The electron binding energy is re-
ferred to as the work function. The PE cannot occur if the photon energy is
less that the binding energy. If the photon energy is greater than the binding
energy, the freed electron gets the excess kinetic energy. The process is

γ + e (in metal) → e (free)

Note that a free electron cannot absorb a photon. The metal is needed in this
case to conserve momentum, but it is so heavy that it gets no energy.

Traditionally, the electron kinetic energy is measured by the voltage (V0 )


needed to stop all emitted electrons, the so-called stopping voltage. The min-
imum binding energy is called the work function (φ) and an electron with
50  Guide to Modern Physics

!"#$%&'(
+&## ,%-%#,
#(#&)*
ϕ
!23%* 0'1(/ ,%-%#,
+.""#/ 0'1(/ ,%-%#,

Figure 3.2 The energy diagram for the photoelectric effect shows empty and
filled states.

binding energy φ gets the maximum allowed kinetic energy. Conservation of


energy reads
eV0 = Kmax = h f − φ

Example 3.10 Get the work function for lead.

Example 3.11 Get the work functions for elements 28 to 32.


Key Processes  51

The threshold frequency ( f0 ) below which there are no electrons emitted is


given by
φ
f0 =
h

  

 
        

Figure 3.3 The stopping voltage is shown vs. radiation frequency.

Example 3.12 Calculate the threshold wavelength for the photoelectric ef-
fect in potassium.

The PE occurs instantaneously. Classically, it could not occur until enough


energy was absorbed.
Example 3.13 Photons with intensity 10−8 W/m2 are incident on potassium.
Estimate the amount of time classically that it would take an atom to absorb
enough energy to eject an electron.
52  Guide to Modern Physics

3.4 ELECTRON DIFFRACTION


3.4.1 Scattering Off a Crystal
When electrons are incident on a regular array of atoms, there are interference
effects and the wave properties of the electrons are observable. In Fig. 3.4, θ
is the angle between the incident electron direction and the planes of atoms
indicated by the dashed lines. The condition for constructive interference is

nλ = 2d sin θ,

where d is the distance between layers of atoms (distance between the dashed
lines in Fig. 3.4).

incident e

scattered e
θ ϕ
          
          
          
          
          
Figure 3.4 Electron scattering geometry from a crystal array of atoms.

The incident electrons are accelerated by a voltage V and have kinetic


energy
1
K = eV = mv2 .
2
In the original experiment of Davisson-Germer, peaks were observed at
equally spaced intervals when plotted as the square-root of the acceleration
voltage,

r
m
V =v .
2e
Key Processes  53

The interference condition gives


2d sin θ
n= ,
λ
thus, the data were suggesting that
1
∼ v.
λ
In fact, precise agreement is given by the de Broglie formula (1.5),
h h h
λ= = = √ .
p mv 2meV
Example 3.14 Calculate the wavelength of an electron accelerated through
100 V.

Direct observation of electron wave properties proved that the concept of


de Broglie wavelength applies to the particle world.

3.4.2 Neutrons in Thermal Equilibrium


Nonrelativistic particles in thermal equilibrium have average kinetic energy

p2 3
K= = kT.
2m 2
This gives √
p= 3mkT ,
and
h
λ= √ .
3mkT
54  Guide to Modern Physics

Example 3.15 Calculate the wavelength of thermal neutrons at 300 K and


77 K (liquid nitrogen temperature).

3.4.3 Quarks Inside a Proton


The concept of de Broglie wavelength applies to quarks inside the protron,
even though they are relativistic.

Example 3.16 Estimate the speed of a charm quark inside the J/ψ particle.

3.5 COMPTON SCATTERING


Compton scattering originally referred to the scattering of a photon from an
electron at rest. More generally it can refer to the scattering of a photon with
any charged particle, whether moving or not. The electron gains kinetic en-
ergy in the collision and the photon loses kinetic energy.
Key Processes  55

!"#$%" &$''()($* /#-"% &$''()($*

+,$-$*

+,$-$* θ
.'"&-%$*
ϕ
.'"&-%$*

Figure 3.5 When a photon scatters off an electron at rest, the photojn and
electron emerge at different angles.

3.5.1 Compton Formula in Terms of Energy


Let E1 ( E2 ) be the incoming (outgoing) photon energy and p be the outgoing
electron momentum. Conservation of energy gives

E1 + mc2 = E2 + (pc)2 + (mc2 )2

Conservation of momentum gives


E1 E2
= p cos φ + cos θ
c c
E2
p sin φ = sin θ
c
Solve by squaring the square root in the energy equation and substituting for
pc using the momentum equations.

Example 3.17 Solve for the photon energy.


56  Guide to Modern Physics

3.5.2 Limiting Cases


When the incident photon energy is much smaller than the electron mass
energy, it is like scattering off a “brick wall”. Momentum is transferred but
essentially no energy is lost by the incoming photon.

E1 = E2

Example 3.18 Find the scattered photon energy at small values (compared
to the electron rest energy) of incident energy.

For forward scattering (θ = 0), the photon energy is unchanged.

Example 3.19 Evaluate the Compton formula at θ = 0.

For backward scattering (θ = π) in the limit of large incoming photon energy


(greater than the electron mass energy), the outgoing photon gets 1/2 of the
electron mass energy.

Example 3.20 Evaluate the Compton formula at θ = π and large incoming


photon energy.
Key Processes  57

3.5.3 Compton Formula in Terms of Wavelength


The standard way to describe the collision is to calculate the change in wave-
length of the photon, the so-called Compton formula by using the formula
for photon wavelength ( hc
λ ).

hc
∆λ = λ1 − λ2 = (1 − cos θ)
mc2
Example 3.21 Solve for the photon wavelength difference.

The quantity,
hc
λC = ,
mc2
is known as the Compton wavelength.

Example 3.22 Calculate the numerical value of the Compton wavelength.

3.5.4 Relationship of λc to Other Quantities


The Bohr radius and the Rydberg constant are discussed in ?? and 5.5. The
Compton wavelength compared to the Bohr radius (a) is

λc = 2παa
58  Guide to Modern Physics

Δλ (m)
5. × 10-12

4. × 10-12

3. × 10-12

2. × 10-12

1. × 10-12

θ
0.5 1.0 1.5 2.0 2.5 3.0

Figure 3.6 The change in photon wavelength depends on the scattering angle.
! !'()"

!"#

!"%
θ
!"! !"# $"! $"# %"! %"# &"!

Figure 3.7 The energy of the scattered photon is shown vs. scattering angle
for 1, 2, and 5 MeV incident photon energies.

Example 3.23 Calculate the Compton wavelength from the Bohr radius.
Key Processes  59

Example 3.24 Calculate the Compton wavelength from the Rydberg con-
stant.

The classical electron radius (re )) is given by setting

e2
= mc2
4πε0 re
which is the radius where the potential energy of an electron charge distribu-
tion would equal to its mass energy. It turns out that there is nothing special
about this distance. The electron has no measurable structure (to date) and
experiments have been performed at distances 5 orders of magnitude smaller
than re . The Compton wavelength compared to the classical electron radius is
re
λc = 2π
a
Example 3.25 Calculate the Compton wavelength from the classical elec-
tron radius.

Thus, we have
α3
re ∼ αλc ∼ α2 a ∼
R∞

3.6 RUTHERFORD SCATTERING


Rutherford scattering refers to a hard-scattering process with a r−2 force. The
original experiment was scattering α particles (helium nucleus) off of a thin
metal foil, scattering of nuclei by the electric force at a time before the ex-
istence of the nucleus was known. The α particles came from nuclear decay
with kinetic energies of about 6 MeV and they are not relativistic.
60  Guide to Modern Physics

3.6.1 Effect of the Electrons


Since the α particles are massive, they cannot transfer much energy to an
electron. The speed of an α particle with kinetic energy Kα is

2Kα
vα = .

From momentum conservation, the maximum speed of the electron after the
collision is 
2Kα
v ≈ 2vα = 2 .

Example 3.26 Calculate the maximum speed of the electron after the colli-
sion with an energetic but still non-relativistic α particle.

Example 3.27 In a collision with a 6-MeV α particle with an electron at rest,


calculate the maximum kinetic energy of the electron.

This energy is much larger than the energy of outer electrons in atoms (eV
scale) but much smaller than the kinetic energy of the α particle. Collisions
with electrons are negligible in Rutherford scattering.

3.6.2 Scattering from a Nucleus


The concept of Rutherford scattering is that the scattering angle can be very
large if the force is large which can only happen if the α particle comes very
close to the nucleus, which can in-turn only happen if the nucleus is tiny on
the atomic scale. Suppose the α particle (charge qα ) has an incoming speed
Key Processes  61

vα and passes approximately a distance b from the nucleus (charge Q). The
momentum transferred (∆p) to the α particle is
qα Q 2b
∆p = F∆t = ,
4πε0 b2 vα
where the time (∆t) for the α partiple to pass the nucleus (the time it spends
feeling the large part of the force) is estimated to be 2b/vα . The factor of 2
is not important in the estimate; the important part is that the time experienc-
ing the force is proportional to the distance scale divided by the speed. The
scattering angle (θ) is
∆p qα Q 2b 1 qα Q
θ= = 2
= ,
p 4πε0 b vα mα vα 4πε0 bKα
where mα is the α-particle mass and Kα is its kinetic energy.

Example 3.28 Calculate the scattering angle for a 6-MeV α particle that
passes a distance 100 fm from a gold nucleus.

The above answer is in the dimensionless unit, radians. Even for a rel-
atively large approach distance of 100 fm (the nuclear scale is 1 fm), the
scattering angle is quite large, about 22◦ .
Notice that a 6-MeV α particle cannot penetrate the gold nucleus, which
would require a kinetic energy exceeding that of the Coulomb repulsion at a
distance equal to the nuclear radius,
qα Q
Kα > .
4πε0 r

Example 3.29 Calculate the α particle kinetic energy needed to penetrate the
gold nucleus of radius 7 fm.
62  Guide to Modern Physics

3.6.3 Cross Section


Cross section (σ) is a geometrically intuitive quantity that is useful in spec-
ifying collision probabilities. It is defined as the transition or scattering rate
(Rs ), the number of times that something predefined happens per second, di-
vided by the incident flux (Φ),
Rs
σ= .
Φ
Since the numerator has units of s−1 and the denominator has units of m−2 s−1 ,
the unit of cross section is area. Cross section represents the effective geomet-
ric size of the target that corresponds to the specified transition.
To get the exact solution for Rutherford scattering, the differential cross
section may be written in terms of the so-called impact paramater (b) of Fig.
3.8,
dσ = 2πbdb,
the area of a ring with radius b and thickness db. Smaller (larger) values of b
correspond to larger (smaller) scattering angles. The differential cross section
in terms of x = cos θ is
dσ dσ dσ db db
= = = 2πb .
dx d cos θ db d cos θ d cos θ
Thus, the problem is reduced to deducing the relationship between impact
parameter and scattering angle. This can be done by deducing two expres-
sions for the momentum transfer, one in terms of θ and the other in terms
of b.

Scattered α

ϕ
Incident α ϕ θ

b
Nucleus

Figure 3.8 An α particle comes within a distance r of a nucleus and scatters


at an angle θ. The angle φ indicates the axis of symmetry.
Key Processes  63

Momentum is exchanged in the collision but very little energy (like a ball
bouncing off a brick wall) because the nucleus is so massive. Therefore, the
direction of the alpha particle changes but not its magnitude and
(∆p)2 = p2 + p2 − 2p2 cos θ,
because the initial and find vectors each of length p and ∆p make a triangle.
The momentum transfer in the collision is
∆p = mα vα 2(1 − cos θ.
p

p2 Δp

p1

Figure 3.9The momentum vector before scattering (p1 ), after scattering (p2 )
and the momentum transfer vector (∆p) form a triangle.

A second expression for the momentum transfer comes from integrating


the component of force in the direction of momentum transfer,
cos φ
Z Z
qα Q
∆p = dtF cos φ = dt 2 ,
4πε0 r
and applying conservation of angular momentum (L),
dφ0
L = mα r2 = mα vα b,
dt
to get
Zφ r
qα Q 2qα Q 2qα Q 1 + cos θ
∆p = dφ cos φ =
0 0
sin φ = .
4πε0 vα b 4πε0 vα b 4πε0 vα b 2
−φ

where the last step comes from the relationship between θ and φ,
π θ
φ= − .
2 2
64  Guide to Modern Physics

Example 3.30 Verify the relationship between φ and θ.

Equating the two expressions for momentum transfer gives the relation-
ship between impact parameter and scattering angle,

qα Q 1 + cos θ
b= .
4πε0 mα vα 1 − cos θ
2

Example 3.31 Calculate the angular dependance of the cross section, dσ/dx
where x = cos θ.

The whole expression is


 2
dσ db π qα Q 1
= 2πb = .
d cos θ d cos θ 2 4πε0 Kα (1 − cos θ)2
The cross section for scattering is some specified range comes from inte-
gration. This may be easier accomplished using b,
b
max

σ= db(2πb) = π(b2max − b2min ).


bmin

Note that b = 0 corresponds to cos θ = −1.


Example 3.32 Calculate the cross section for scattering of a 12 MeV α par-
ticle scattering off silver (Z = 47) at angles greater than 90◦ . Give the answer
in barns (1 b = 10−28 m).
Key Processes  65

3.6.4 Scattering from a Thin Foil


In performing a scattering experiment, one does not have the liberty to choose
the impact parameter. When the α particle passes through a foil, it will en-
counter nuclei at random values of b. A good way to think of the scattering is
that of a flux of nuclei from the target streaming by the α particle. Consider
a foil with atomic mass A and density ρ. The number of target nuclei per
volume is
N N0 ρ
= ,
V A(10−3 kg)
where N0 is Avogadro’s number.

Example 3.33 Get Avogadro’s number.

The number of target nuclei per area is dN/V where d is the target thick-
ness. If the rate of incoming particles is Ri , then the flux is
Ri dN Ri dN0 ρ
Φ= = .
V A(10−3 kg)
The cross section is
Rs Rs A(10−3 kg)
σ= = ,
Φ Ri dN0 ρ
where Rs is the scattering rate.

Example 3.34 Alpha particles of 6 MeV are incident at normal angle on a 1


µm thick gold foil at a rate of 1000 per s. Calculate the rate that α particles
scattered at angles greater than 1 radian.
66  Guide to Modern Physics

3.7 THE WEAK INTERACTION


3.7.1 Weak Coupling
The strength of the weak interaction is traditionally written in terms of the
Fermi constant (GF ). The best value of GF is given by measurement of the
muon lifetime. The muon decay probability goes as G2F , so the lifetime goes
as the inverse square. The result is

192π3 h̄
GF = .
(mc2 )5 τ

Example 3.35 Calculate GF .

The Fermi constant is stored as Quantity[“ReducedFermiConstant”] (G0F ).

Example 3.36 Get the weak coupling constant.

The dimension of GF is inverse energy squared. The dimensionless weak


coupling at the energy scale E may be written as

αw = GF E 2 .

At 1 GeV,
αw ≈ 10−5 .

3.7.2 Neutrino Cross Section


Event probabilities are measured in particle physics by the interaction cross
section (3.6.3). The neutrino has size (cross sectional area) is given by its
Key Processes  67

wave properties as ( h̄c 2


E ) . We may estimate the cross section per nucleon (σN )
as  2
2 h̄c
σN ≈ αw .
E

Example 3.37 Estimate the interaction cross section per nucleon for the in-
teraction of a 1 GeV neutrino.

3.7.3 Neutrino Scattering Rate


The cross section that was calculated in 3.7.2 is per nucleon. When calculat-
ing the neutrino scattering rate in a target, one must us the same technique
derived for Rutherford scattering in a foil 3.6.4.

Example 3.38 Estimate the interaction rate for 1 GeV neutrinos incident on
a 104 kg target with a flux of 108 m−2 s−1 .

3.7.4 Neutrino Mean Free Path


In passing a distance d through the matter that has n nucleons per volume, a
neutrino would encounter nd nucleons per cross-sectional area. For σN nd = 1,
one interaction is expected, on the average. The mean free path (9.2.3) is
1
d= .
nσN
The number of nucleons per volume in the earth is its mass density divided
by the mass of a nucleon (0.94 GeV/c2 ).
68  Guide to Modern Physics

Example 3.39 Estimate the mean free path for a 1 GeV neutrino in the earth.

The 1 GeV neutrino mean free path is about 5000 times greater than the
earth’s diameter (1.3 × 107 m). A 1 GeV neutrino passing through the full
diameter of the earth will only have a probability of 1 in 5000 of interacting.
CHAPTER 4

Special Relativity

The first postulate of special relativity states that the laws of physics are iden-
tical in all inertial (non-accelerated) frames of reference. The second postu-
late states that the speed of light in vacuum (c) is the same for all observers.

4.1 BETA AND GAMMA


Calculations in relativity involve the the relative velocity (v) of 2 frames of
reference. This is expressed in terms of a dimensionless quantity
v
β= ,
c
which has a range from −1 to 1. Note that β can be negative. Sometimes β is
written as a vector,
v
β= .
c
The Lorentz gamma factor (γ) appears frequently:
1
γ= p ,
1 − β2
which has a range of 1 to ∞.
The inverse relationship is
s
1
β= 1− .
γ2
The γ and β factors are the ingredients of the Lorentz transformation de-
scribed in 4.5.
As γ gets large, the particle speed is essentially the speed of light. For a
γ-factor in the 104 range, the difference between particle and light speed is in
the m/s range.

DOI: 10.1201/9781003395515-4 69
70  Guide to Modern Physics

2.5

2.0

1.5

β
0.2 0.4 0.6 0.8 1.0

Figure 4.1 As β approaches 1, γ goes rapidly to ∞.

c-v (m/s)

108

106

104

100

0.01
γ
1 10 100 1000 104 105

Figure 4.2 The speed of light in vacuum minus particle speed is shown vs. γ.

4.2 SPACE AND TIME


4.2.1 Time Dilation
Consider a frame of reference S wherelight moves a distance d from points
A to B in the y-direction.
Now consider the motion as viewed from a frame S0 that moves with a
speed v in the −x direction. In this frame both points A and B move with
speed v in the x direction. In the time that light is traveling, point B is mov-
ing and the light has a longer distance to travel to get there. Its speed is fixed
so it takes a longer time to get from A to B. Since the speed of light is fixed,
the two frames are related by the velocity triangle of Fig. 4.4.
Special Relativity  71

Frame ℬ

Figure 4.3 In frame S, light moves from A to B.

Frame ' ℬ Light arrived

c c2 - v 2

v
Light left

Figure 4.4 In frame S0 , in the time that light moves from A to B, points A
and B have both moved to the right. The speed of light and the speed v make
a right triangle as shown.
.

In a frame where a clock is moving, time intervals are measured to be


longer by a factor of γ longer compared to their measurement in a frame
where the clock is at rest,
∆t0 = γ∆t.
An example is the cosmic ray muon that is produced in the upper atmo-
sphere by cosmic ray protons which make energetic short-lived pions which
in turn decay into energetic muons. The muons have a lifetime of about
2.2 × 10−6 s in their rest frame. The energetic muons are moving at nearly
the speed of light in our frame, having a typical γ of about 10.
72  Guide to Modern Physics

Example 4.1 Calculate the speed of a muon that has γ = 10.

The muon is its own clock ticking away until its decay. The mean lifetime
for the muon in its rest frame is about 2.2 µs. In a frame in which it is moving,
the lifetime is a factor of γ longer.

Example 4.2 Calculate the lifetime of the muon in the frame where it moves
corresponding to γ = 10.

Example 4.3 Calculate the average distance the muon can travel before de-
caying.

The muons are observed to reach the surface of the earth because of time
dilation.
An atomic clock was flown on the space shuttle at a speed v for a time
∆T as measured on earth. The time interval measured on the space shuttle is
shorter by a factor of γ. The difference of the time intervals due to special
relativity is     
1  v 2 
∆T − ∆T = T 1 − 1 − .
γ c 

Example 4.4 Calculate the time interval difference for v = 7710 m/s and T
= 7 days.
Special Relativity  73

4.2.2 Lorentz Contraction


The concept of length contraction appears together with time dilation.
Lengths are shorter when measured in a frame where they are moving. There
is one special frame where the length is at rest and in that frame it is the
longest.
Consider a stick moving with speed v measured with a stationary clock
(frame with no primes). The time for the stick to pass by the clock is L/v.
In the frame where the stick is stationary (frame with primes), the clock is
moving with speed v. Time intervals are longer in this frame,
L L0
∆t0 = γ∆t = γ = ,
v v
or
L0
L= .
γ
Of course, it does not matter in what frame “prime” is used. The physics says
the moving clock runs slower and the moving stick is measured to be shorter,
as seen in the example of the cosmic-ray muon (4.2).

4.3 ENERGY AND MOMENTUM


Many problems in relativity are more easily visualized in terms of energy-
momentum rather than space-time. A moving proton with corresponding β
and γ has momentum
mv
p = q   = γβmc.
2
1 − vc

The energy is
q q q
E = (mc2 )2 + (pc)2 = (mc2 )2 + (γβmc2 )2 = (mc2 )2 ((1 + γ2 β2 ).

Using
β2 1
1 + γ2 β2 = 1 + = = γ2 ,
1−β 2 1 − β2
gives
E = γmc2
and
pc
β= .
E
74  Guide to Modern Physics

In 2022, the CERN Large Hadron Collider (LHC) increased the energy
of the protons to 6.8 TeV per beam. One TeV is 1012 eV.

Example 4.5 Calculate the gamma factor for LHC protons.

The gamma factor is so large that β is very nearly equal to 1.

Example 4.6 Calculate 1 − β for LHC protons.

Example 4.7 Calculate how much the speed of LHC protons differ from the
speed of light.

4.4 4-VECTORS
A 4-vector is a 4-component vector with one scalar “time” part and an or-
dinary vector “space part whose length does not depend on the frame of
reference. The length-squared of a 4-vector is defined as the square of the
time part minus the square of the space part. The combination (ct, r) makes a
4-vector of length (L) squared:

L2 = (ct)2 − x2 + y2 + z2 ).

The matrix that flips the sign of the space part,


 
 1 0 0 0 
 0 1 0 0 
g =  ,
 0 0 1 0 
0 0 0 1
is useful for calculating 4-vector lengths.
Special Relativity  75

Example 4.8 Calculate the length-squared of the space-time 4-vector.

Other common 4-vectors are energy-momentum (E, pc), charge-current


density (ρc, J), and electromagnetic scalar-vector potential (V, cA).

4.4.1 Invariant Mass


For any particle, the total energy, momentum, and mass are related by (1.3)

(mc2 )2 = E 2 − (pc)2 ,

which is precisely the definition of 4-vector length. The addition of 2 4-


vectors is also a 4-vector, and for energy-momentum this length is called
the invariant mass. For two 4-vectors (E1 , p1 c) and (E2 , p2 c) , the invariant
mass squared (s) is

s = (E1 + E2 )2 − (p1 c + p2 c)2 .

4.4.2 Center of Mass


There is a special frame where the two particles have the same magnitude of
momentum, in opposite directions. This is commonly referred to as the center
of mass (CM) frame. In this case

s = (E1∗ + E2∗ )2 ,

where E1∗ and E2∗ are the energies in the CM frame. Thus, the total energy
squared in the CM frame is the invariant mass of the system. The collision
of two particles is simplest when viewed in the CM frame because the total
momentum is zero both before and after the collision. If the two particles have
the same mass, then the energies E ∗ are also equal in the CM frame, and

s = (E ∗ + E ∗ )2 = 4E ∗2 .

Example 4.9 A proton of momentum p = 200 GeV/c strikes another proton


at rest. Calculate the energy of each proton in the CM frame.
76  Guide to Modern Physics

The minimum reaction to make an antiproton in a pp collision is

p + p → p + p + p + p̄

in order to conserve baryon number. The antiproton has the same mass as the
proton.

Example 4.10 Calculate the (total) energy that a proton needs to hit another
proton at rest and make an antiproton.

4.5 LORENTZ TRANSFORMATION


The Lorentz transformation can be written as a 4×4 matrix (Λ) that multiplies
a 4-vector. The matrix that corresponds to speed v = βc in the x-direction is
 
 γ −βγ 0 0 
 −βγ γ 0 0 
Λ =  .
 0 0 1 0 

0 0 0 1

4.5.1 Transformation of Time-Space


Example 4.11 Transform the time-space 4-vector.
Special Relativity  77

Example 4.12 Calculate the length-squared of the transformed space-time


4-vector.

The Lorentz transformation preserves the length of any 4-vector. The un-
derlying physics states whether something transforms like a 4-vector or not.
Time-space makes a 4-vector because the speed of light does not depend on
the frame of reference.
The Lorentz transformation accounts for time dilation. and length con-
traction. Consider the example of the clock. In the frame where the clock
is at rest, a time interval (t2 − t1 ) is measured at the same location. Make a
Lorentz transformation to the frame where the clock is moving. In this frame
the time interval is longer by γ.

Example 4.13 Calculate the time interval c(t2 − t1 ) in the moving frame.

4.5.2 Transformation of Energy-Momentum


Energy-momentum makes a 4-vector because the mass of a particle is the
same in all frames of reference. The length of energy-momentum 4-vector is
 
E − [(p x c) + (py c) + (pz c) ] = E 2 − (pc)2 = mc2 .
2 2 2 2

If we transform a mass m from rest, its speed divided by c in the moving


frame is seen to be that of the moving frame but in the opposite direction.
Its energy is γmc2 and its momentum is −γmvx̂, where the unit vector is the
direction of the transformation.
78  Guide to Modern Physics

Example 4.14 Transform a mass m from rest and calculate its β-factor.

When we transform (E, pc), the transformed energy is smaller (greater)


than the original energy if there is a component of momentum in the same
(opposite) direction as the transformation.
Example 4.15 Transform the energy-momentum 4-vector.

The relative velocity formula can be readily determined from a Lorentz


transformation. Consider 2 particles with velocities v and u as shown in Fig.
4.5. The relative speed of the particles is the speed one of the particles in
the frame where the other particle is at rest. We can get to this frame with a
Lorentz transformation in the x-direction with speed v. The relative speed is
then the speed of the moving particle which is obtained by dividing momen-
tum times c by energy.

In frame S , particles have velocites vx̂ and −ux̂. In frame S  which


Figure 4.5
moves with speed v in the x-direction, one of the particles is at rest.
Special Relativity  79

Example 4.16 Calculate the relative speeds of the particles in Fig. 4.5.

4.6 DOPPLER EFFECT


4.6.1 Colinear Light Source
Consider a light source that is moving toward an observer. Let the photon
energy be E in the frame where the source is at rest. In the frame where ths
source is moving, the photon energy E  is given by the Lorentz trnasforma-
tion: 
1 + β 1+β
E  = γE + βγE = E  =E .
1 − β2 1−β
Since the frequency is proportional to the energy by E = h f (see 1.4.2), we
have 
1+β
f = f ,
1−β
where f  is the frequency in the frame where the source moves and f is the
frequency in the frame where the source is at rest. Negative values of β (or
switching the sign of β) correspond to the source moving away from the ob-
server.
Example 4.17 The Lyman alpha line (5.4.1) is at λ = 122 nm for a hydrogen
atom at rest. This line is observed to be at λ = 362 nm when emitted from a
moving star. Calculate the speed of the star.

The minus sign indicates the star is moving away.


80  Guide to Modern Physics

4.6.2 Redshift
Astrophysicists define redshift (z) to be a fractional change in frequency for
stars that are receding from us,

f − f 1+β
z= 
= −1
f 1−β
Example 4.18 Find β corresponding to redshift z.

'(*

'()

'($

'("

&
! " # $ %

Figure 4.6 Beta vs. redshift is plotted for a receding star.

4.6.3 Observation at an Angle


Suppose the light is observed at an angle θ in the lab. Let the photon energy
in the lab be E  and its x-component of momentum (along the boost direc-
tion) be E  cos θ. The energy of the photon (E) in its rest frame is given by
the Lorentz transformation,
E = γE  − βγE  cos θ,
and
E
E = .
γ(1 − β cos θ)
Special Relativity  81

The same relationship holds for the frequencies,


f
f = .
γ(1 − β cos θ)
This result is referred to as the “transverse Doppler effect”. Note that this
result reduces to that of 4.6.1 for approaching source (θ = 0) and receding
source (θ = π).
The condition of zero redshift is given by f  = f . This is the boundary
between Doppler redshift and blueshift (increase in frequency)..

Example 4.19 Find cos θ for the boundary between redshift and blueshift as
a function of β.

()*θ

'"!

!"&

!"%

!"$

!"#

β
!"# !"$ !"% !"& '"!

Figure 4.7 Cosine of the boundary angle between redshift and blueshift is
plotted as a function of β. Note that at very large values of β there will be a
redshift unless the source is headed sharply toward the observer.
CHAPTER 5

Bohr Model

The Bohr model is historically important in the understanding of quantum


physics. It gives many important results for the hydrogen atom including
atomic size and ionization energy. It is sometimes referred to as a semi-
classical calculation, because it includes the physics of angular momentum
quantization without the full quantum mechanical description of an electron
wave. It is tremendously useful for gaining physical intuition of the quantum
nature of the atom.

5.1 QUANTIZATION OF ANGULAR MOMENTUM


In the Bohr model of the hydrogen atom, an electron has a circular orbit with
speed v and radius r about the proton. The Bohr quantization condition is that
an integer number of electron (de Broglie) wavelengths must fit to the atomic
circumference,
h
nλ = n = 2πr.
p
Written in terms of h̄ with p = mv, angular momentum quantization is

mvr = nh̄,

where n is a positive integer. Newton’s 2nd law gives

e2 mv2
= .
4πε0 r2 r
The possible values of the orbit radius as a function of the quantum number
n are
4n2 πh̄2 ε0
r= .
e2 m

DOI: 10.1201/9781003395515-5 82
Bohr Model  83

5.2 GROUND STATE


The lowest energy (ground state) corresponds to n = 1.

5.2.1 Bohr radius


The value of r corresponding to n = 1 is called the Bohr radius (a0 ),
4πh̄2 ε0
a0 = .
e2 m
Example 5.1 Get the Bohr radius.

4πh̄2 ε0
Example 5.2 Compare a0 with e2 m
.

Example 5.3 Calculate the numerical value of the Bohr radius.

5.2.2 Energy
The kinetic energy is one-half the potential energy with the opposite sign.
1 1 e2 1
K = mv2 = = − U.
2 2 4πε0 r 2
The total energy (E) is
V 1 e2
E = K +U = =−
2 2 4πε0 r
with r given in 5.1 as
r = n 2 a0 .
84  Guide to Modern Physics

Example 5.4 Calculate the total energy for the n = 1 orbit.

The electron speed is


nh̄ e2 αc
v= = =
mr 4πε0 nh̄ n

Example 5.5 Calculate the electron speed for n = 1 orbit.

The electron has an interesting speed. It is extremely fast compared to


speeds of macroscopic objects, but yet it is (just barely) nonrelativistic.
Since K is one-half U with the opposite sign, the total energy is −K.
Therefore, a convenient alternate way to write the total energy is

1  αc 2
E=− m .
2 n

5.3 EXCITED STATES


The excited states are given by n = 2, 3, 4...

5.3.1 Orbits
The size of the orbits in the Bohr model scale by the square of the quantum
number, n2 .

Example 5.6 Calculate the orbit radii for n = 2 to 5.

5.3.2 Speeds
The orbit speeds are inversely proportional to the quantum number n.
Bohr Model  85

Example 5.7 Calculate the orbit speeds for n = 2 to 5.

5.3.3 Energies
The energies scale as −13.6 eV/n2 . Increasing values of the quantum number
n give larger energies.

Example 5.8 Calculate the orbit energies for n = 2 to 5.

5.4 TRANSITIONS BETWEEN ENERGY LEVELS


When an electron moves from a higher energy orbit to a lower energy orbit, a
photon is emitted with an energy equal to the energy difference between the
orbits.

5.4.1 Lyman Series


Transistons to the n = 1 (ground) state are called the Lyman series. In Ex. 5.9,
a function is defined (A.9) to evaluate the energy E(n). A trick is played to
avoid use of the reserved E (A.4) by writing it as a Greek letter which looks
nearly identical in Mathematica.

Example 5.9 Get the first 4 Lyman series energies.

These photons are in the ultraviolet region.


86  Guide to Modern Physics

Example 5.10 Get the first 4 Lyman series wavelengths.

5.4.2 Balmer Series


Transitions to the n = 2 (1st excited) state are called the Balmer series.

Example 5.11 Get the first 4 Balmer series energies.

These photons are in the visible region.

Example 5.12 Get the first 4 Balmer series wavelengths.

5.4.3 Paschen Series


Transitions to the n = 3 (2nd excited) state are called the Paschen series.

Example 5.13 Get the first 4 Paschen series energies.

These photons are in the infrared region.

Example 5.14 Get the first 4 Paschen series wavelengths.


Bohr Model  87

5.5 RYDBERG CONSTANT


The Rydberg constant (R∞ ) is defined to be the inverse wavelength of a pho-
ton corresponding to a transiton from n = ∞ → n = 1.
E1
R∞ = − .
hc
Example 5.15 Calculate the Rydberg constant.

Example 5.16 Get the Rydberg constant.

In terms of R∞ , the wavelength of a photon from the transition j → i is given


by  
1  1 1 
= R∞  2 − 2  .
λ ni n j

Example 5.17 Calculate the energy of the first Lyman transition using the
Rydberg constant.

5.6 REDUCED MASS


The Bohr model assumes that the proton is stationary. We can correct for this
by working in the center-of-mass. The electron mass is then replaced by
mm p
µ= .
m + mp
88  Guide to Modern Physics

Example 5.18 Calculate the reduced mass divided by the electron mass.

The deuteron has a slightly bigger reduced mass than the proton. Con-
sider the Balmer series n = 3 to n = 2 transition. This photon from this tran-
sition was calculated to have a wavelength of 656 nm (5.4.2). The photon
energy is  
1 2 1 1
E = µ(αc) − ,
2 4 9
and the wavelength is
hc
λ= .
E
Example 5.19 Calculate the wavelength difference between the n = 3 to
n = 2 transition in hydrogen vs. deuterium.

5.7 COLLAPSE OF THE BOHR ATOM


The power (P) radiated by a charged particle is proportional to the square of
the acceleration (a). For a non-relativistic charge e, the result is

e2 a2
P= .
6πε0 c3
For an atom in a circular orbit, the acceleration is

v2
a= .
r
Bohr Model  89

Example 5.20 Calculate the power radiated for a classical electron in the
Bohr orbit.

The energy radiated per second is enormous compared to the kinetic energy
of the electron and the Bohr atom would collapse on a time scale of 10−11
s.

5.8 CORRESPONDENCE PRINCIPLE


5.8.1 Orbit and Radiation Frequency
In the Bohr model, the orbit frequency from which an electron radiates when
moving to the next lower state,
αc/n
f= ,
2πn2 a0
is smaller than the frequency of the emitted photon
En+1 − En
f= .
h
For large values of the quantum number n, we must get agreement with
classical physics in which the orbit frequency ( f ) of the charge gives radia-
tion of the same frequency.

Example 5.21 Calculate the orbit frequency.


90  Guide to Modern Physics

Example 5.22 Calculate the radiation frequency and take the limit as n → ∞.

0.005

0.002
% difference

0.001

5.× 10-4

2.× 10-4

0 2000 4000 6000 8000 10 000


n

Figure 5.1 The percentage difference between the radiation and orbit frequen-
cies drops vs. the quantum number n.

5.8.2 Earth’s Orbit


We may apply the Bohr model to the earth’s orbit about the sun by replac-
e2
ing the electric force strength ( 4πε 0
) with the gravitational force strength
(Gme Ms ) in the expression for r calculated in 5.1, giving
n2 h̄2
r= .
Gm2e Ms
Example 5.23 Calculate the quantum number for the earth’s orbit about the
sun.
Bohr Model  91

5.8.3 LHC Proton


Protons orbit the Large Hadron Collider (LHC), which has a circumference
of 27 km, with a momentum of 13.6 TeV/c. Quantization of angular momen-
tum gives
pr = nh̄.

Example 5.24 Calculate the quantum number for LHC protons.


CHAPTER 6

Particle in a Box

The particle in a box, also referred to as “infinite square well” is an impor-


tant example because it illustrates the many features of the solution to the
Schrödinger equation in an example requiring minimal math. The problem
will be worked first in one dimension.

6.1 THE POTENTIAL


The box is defined by a potential that is zero in some region (0 < x < L) and
infinite at the boundaries. Thus, the particle is absolutely confined and the
wave function is zero beyond the boundaries.
2.0

1.5

1.0
U

0.5

0.0
-1.0 -0.5 0.0 0.5 1.0 1.5 2.0
x/L

Figure 6.1 The potential for the one-dimensional particle in a box becomes
infinite at the boundaries.

DOI: 10.1201/9781003395515-6 92
Particle in a Box  93

6.2 THE SCHRÖDINGER EQUATION


The Schrödinger equation inside the box is

d2 ψ 2mE
2
= − 2 ψ.
dx h̄
This is the most important differential equation that appears in physics. It
is of the same form that describes mechanical oscillations: two derivatives
of a function gives the same function back with a negative sign and a con-
stant multiplier. The only functions that satisfy this are the sine and cosine.
The sine and cosine have the same shape and differ only by placement of the
origin.

6.3 SOLUTION
There are multiple solutions (in this case, an infinite number) corresponding
to the number of oscillations that fit inside the box. The general solution that
satisfies the boundary conditions (ψ → 0. at at x = 0 and x = L) is
 nπx 
ψ = A sin ,
L
where A is a constant that is determined by normalization to unit
probability and n is a positive integer that tells us how many peaks
the sine function makes inside the box. It is a general feature in
the quantum world that the lowest energy cannot be zero because
that would require ψ = 0 which would correspond to no particle at
all.

Example 6.1 Solve the Schrödinger equation for a particle in a box.

The energy levels increase with n, but the percentage of increase de-
creases with n, in such a manner that we arrive at the classical limit at large n.
The interpretation of the wave function is that its square is the probability
per distance x (or per volume for 3 dimensions) of finding the particle at lo-
cation x. The integral of the wave function squared over all possible positions
94  Guide to Modern Physics

'()*+, !π # ℏ# "#-.# #

&"

%"

$"

#"

!"

π2 h̄2
Figure 6.2 The energy levels are indicated in units of 2mL2
.

is normalized to unity.

Example 6.2 Find the normalization constant A.

6.3.1 Wave Functions


The wave functions for the 3 lowest energy states are shown in fig. 6.3. The
square of the wave functions is shown in fig. 6.4.

6.3.2 Electron in a Box


For an electron in a box of the atomic size (0.2 nm), we expect to get a ground
state energy on the eV scale.
Particle in a Box  95

#"! # #

'"( ' '

'"! ! !

ψ
ψ

!"( ! "' ! "'


!"! "# "#
!"! !"# !"$ !"% !"& '"! !"! !"# !"$ !"% !"& '"! !"! !"# !"$ !"% !"& '"!
)!* )!* )!*

Figure 6.3 Particle-in-a-box wave functions for n = 1, 2, and 3.

$ $ $

( ( (

# # #
ψ#

ψ#

ψ#
' ! ' ! '

! ! !
!"! !"# !"$ !"% !"& '"! !"! !"# !"$ !"% !"& '"! !"! !"# !"$ !"% !"& '"!
)!* )!* )!*

Figure 6.4 Particle-in-a-box wave functions squared for n = 1, 2, and 3.

Example 6.3 Calculate the ground state energy for an electron in a box of
width 0.2 nm.

We may make a comparison with the Bohr model (Chap. 5) which re-
quires
λ = 2πa0 .
This is the n = 2 state for a box of size L = 2πa0 .

Example 6.4 Calculate the n = 2 energy for an electron in a box of width


L = 2πa0 .
96  Guide to Modern Physics

6.3.3 Proton in a Box


For a proton in a box of the nucleus size, we expect to get a ground state
energy on the MeV scale.

Example 6.5 Calculate the energy for a proton in a box of width 3 fm.

6.4 COMPARISON WITH THE DE BROGLIE WAVELENGTH


AND THE UNCERTAINTY PRINCIPLE
The minimum energy from the particle in a box calculation is

π2 h̄2 h2
∆E > = .
2mL2 8mL2
In the de Broglie picture, this can be thought of as a standing wave where
1/2 wavelength fits inside the box,
λ
= L.
2
Thus gives
h h
∆p = = ,
λ 2L
corresponding to a minimum kinetic energy,

(∆p)2 h2
∆E > = .
2m 8mL2
The de Broglie wavelength interpretation gives the same minimum energy as
the Schrödinger equation.
To apply the uncertainty principle, the uncertainty ∆x must be evaluated.
This may be taken to be the rms of a flat distribution.
Particle in a Box  97

Example 6.6 Calculate the rms position if a flat probability from 0 to L.

For
L
∆x = √ ,
2 3
the uncertainty principle gives

h̄ h 3
∆p = = .
2∆x 2πL
The minimum allowed kinetic energy
  2
(∆p)2 3 h
∆E > = 2 .
2m π 8mL2
In this case, the minimum kinetic energy is greater than required by the un-
certainty principle by about a factor of 3. To meet the uncertainty principle
minimum, a Gaussian distribution is required.

6.5 EXPECTATION VALUES


6.5.1 Position
The wave function squared is the probability distribution and its average
value is referred to as the “expectation value”. The probability distributions
look very different for different values of n but they all have an average value
of L/2. There is an equal probability of the particle being found on the left-
or right-hand half of the box. The wave function squared is symmetric about
x = L/2.
Example 6.7 Calculate the average value of x for n = 1, 2, and 3.
98  Guide to Modern Physics

Example 6.8 Calculate the average value of x2 for n = 1, 2, and 3.

6.5.2 Momentum
The momentum operator is given by
d
p = −ih̄ .
dx
Example 6.9 Calculate the average value of momentum for n = 1, 2, and 3.

Example 6.10 Calculate the average value of momentum squared for n = 1,


2, and 3.

The average value of momentum squared is seen to give the energy,

(p2 )ave
E= .
2m

6.6 CONSISTENCY WITH THE UNCERTAINTY PRINCIPLE


We may estimate the uncertainty in the knowledge of the position and mo-
mentum from the root-mean-square values which have been calculated in
6.5.
Particle in a Box  99

Example 6.11 Check that the product of the root-mean-square values of po-
sition and momentum satisfy the uncertainty principle.

This value is larger than h̄/2 as required by the uncertainty principle.

6.7 CORRESPONDENCE PRINCIPLE


When the quantum number n is large, the energy levels become close together
and we get so many wiggles in the wave function that we get agreement with
classical physics.

0.025

0.020

0.015
ΔE/E

0.010

0.005

0 200 400 600 800 1000


n

Figure 6.5 The fractional energy difference between adjacent levels is indi-
cated vs. n.

6.7.1 Phone in Box


For a macroscopic object we expect the ground state energy will be zero and
also that any measurable energy corresponds to a huge quantum number by
100  Guide to Modern Physics

the correspondence principle. The iPhone 14 has a mass of 6.07 oz (ounce is


a mass unit in Mathematica).

Example 6.12 Calculate the ground state energy of a mobile phone confined
to be in a room of width 15 feet.

Example 6.13 Calculate the quantum number n if the iPhone had a kinetic
energy of a thermal photon (1/40 eV).

6.7.2 Classical Probability


Consider the probability that the particle is at the center of the box in the tiny
interval
L L L L
− <x< +
2 200 2 200
The classical answer is 0.01 independent of the size of the box because the
interval has been chosen to be a fraction (1/100) of the length. The n = 1
wave function has a much larger probability of being found near there, in
fact, nearly twice as large.

L L L L
Example 6.14 Integrate ψ2 from 2 − 200 <x< 2 + 200 .
Particle in a Box  101

The n = 2 wave function has a node there and has nearly zero probability of
being found near there.

Example 6.15 Calculate the probability that the particle is found at the cen-
ter of the box within 1% for n = 1 and 2.

The correspondence principle tells us that at large n, when we get


enough wiggles, the probability will go to the classical value of 0.01.

0.020

0.015
∫ ψ 2 dx

0.010

0.005

0.000
0 200 400 600 800 1000
n

The probability that the particle is found at the center of the box
Figure 6.6
within 1% is shown for n = 1 to 1000.

The over and undershoot gets progressively smaller with increasing n. From
symmetry of the wave function, we always get the exact answer when n
is a multiple of 100 and the worst case when n is 50 plus a multiple
of 100. How close is the solution to the classical result when n = 106 +
50?
102  Guide to Modern Physics

Example 6.16 Calculate the probability that the particle is found at the cen-
ter of the box within 1% for n = 106 + 50.

6.8 THREE DIMENSIONS


For three dimensions, the wave function is ψ(x, y, x). The Schrödinger equa-
tion inside the box is
d2 ψ d2 ψ d2 ψ 2mE
2
+ 2 + 2 = − 2 ψ.
dx dy dz h̄
Choosing the size of the box to be L1 × L2 × L3 , the solution is obtained by
the technique of separation of variables to be of the form
     
n1 πx n2 πx n3 πx
ψ = A sin sin sin ,
L1 L2 L3
There are now 3 quantum numbers (n1 , n2 , n3 ) corresponding to the boundary
conditions on x, y, and z.

Example 6.17 Solve the Schrödinger equation for a particle in a box in 3D.

The ground-state energy is obtained when by

n1 = n2 = n3 = 1.

The ordering of the energies after that will depend on the relative sizes of
L1 , L2 , and L3 . The energy levels can be degenerate. The greatest degeneracy
will happen when L1 = L2 = L3 . In that case (n1 , n2 , n3 ) being (1,1,2), (1,2,1),
or (2,1,1) all have the same energy.
Particle in a Box  103

6.9 FINITE POTENTIAL


For a box where the sides are not infinitely high (“finite square well”) the
Schrödinger equation becomes

d2 ψ 2m(V − E)
= ψ,
dx2 h̄2
where V is the value of the potential outside the box. The principle difference
in the solution, compared to the infinite potential, is that now the wave func-
tion extends outside the box. It must go to 0 at ±∞ instead of at the boundary
of the box.
2.0

1.5

1.0
U

0.5

0.0
-1.0 -0.5 0.0 0.5 1.0 1.5 2.0
x/L

Figure 6.7 The potential for a finite square well is constant outside the bound-
ary.

6.9.1 Solution Technique


Defining s
2m(V − E)
α= ,
h̄2
the solution for x < 0 becomes

ψ = Aeαx ,

and for x > L,


ψ = Be−αx .
104  Guide to Modern Physics

Inside the box, the Schrödinger equation is identical to the infinite potential.

d2 ψ 2mE
2
= − 2 ψ.
dx h̄
Defining s
2mE
k= ,
h̄2
the solution may be written

ψ = Ceikx + De−ikx .

The remaining task is to match the boundary conditions at x = 0 and x = L.


Continuity of ψ at x = 0 gives

A = C + D,

and at x = L gives
Be−αL = CeikL + De−ikL .
The derivative dψdx must also be continuous in order for the second derivative
to exist. This gives at x = 0,

αA = ikC − ikD,

and at x = L
−αBe−αL = ikCeikL − ikDe−ikL .

6.9.2 Energy Condition


The relationship between α and k gives the allowed energies. It is use-
ful to separate these into 2 classes according to the symmetry of the wave
function. The even functions have ψ(x) = ψ(−x). The ground state will be
even. The excited states, if they exist, will alternate between even and odd,
ψ(x) = −ψ(−x).

The even states are obtained eliminating the constants making use of
ψ(x) = ψ(−x). The conditions at x = 0 give
−α − ik
C= D,
α − ik
Particle in a Box  105

and combining this with the condition at x = L gives the relationship between
α and k to be
−α − ik ikL −α − ik ikL
−α e − αe−ikL = ik e − ike−ikL .
α − ik α − ik
The even functions have kL < π and the wave function for 0 < x < L has the
form   L
ψ ∼ cos k x − .
2
Example 6.18 Solve for the relationship between α and k for the even func-
tions.

The equation for the energy is


   
2m(V − E) 2mE  2mE L 
= tan    .
2 2 2 2 
h̄ h̄ h̄
This is a transcendental equation which must be solved numerically.
The odd functions have π < kL < 2π and the wave function for 0 < x < L
has the form   L
ψ ∼ sin k x − .
2
Example 6.19 Solve for the relationship between α and k for the odd func-
tions.

The equation for the energy is


   
2m(V − E) 2mE  2mE L 
=− cot   .

h̄2 h̄2 h̄2 2 
106  Guide to Modern Physics

This is another transcendental equation which must be solved numerically.

6.9.3 Solving for the Energy


Define the dimensionless variables
mV L2
W=
2h̄2
and √
L 2mE
ξ=
2h̄
to write the transcendental equation in dimensionless form for the even states
as
W − ξ2
p
tan ξ = ,
ξ
and the odd states as
ξ
tan ξ = − p .
W − ξ2
Given numerical values of V and L we can solve for the allowed ener-
gies. The dimensionless transcendental equation allows visualization of the
allowed solutions graphically. There is always at least 1 bound state but there
could be any number in general.

Consider an electron with V = 100 eV and L = 0.2 nm. √


Figure 6.8 shows
W−ξ2
that there are two solutions for the energy where tan ξ and ξ are equal.

Example 6.20 Find the energies of the even states.


Particle in a Box  107

'!

&
! ! ξ"

ξ
()*ξ )*+

!&

! " # $ %
ξ


W−ξ2
Figure 6.8 Plot of tan ξ and ξ . vs. ξ. The solution for the even energies
are the points where the curves meet. In this case there are 2 solutions for the
even energies, corresponding to the ground and 2nd excited states.
108  Guide to Modern Physics

Figure
√ 6.9 shows that there are two solutions for the energy where tan ξ
W−ξ2
and ξ are equal.

&

$
! ! ξ"

"
ξ

)
*+,ξ +,- !

!"

!$

!&
! " # $ % & ' (
ξ

ξ
Figure 6.9 Plot of tan ξ and - √ . vs. ξ. The solution for the odd energies
W−ξ2
are the points where the curves meet. In this case there are 2 solutions for the
odd energies, corresponding to the 1st and 3rd excited states.

Example 6.21 Find the energies of the odd states.


Particle in a Box  109

6.9.4 Wave Functions


The wave functions are plotted in fig. 6.10 and the squares are plotted in fig.
6.11.
100 000

80 000

50 000
60 000
ψ (m-12 )

ψ (m-12 )
0
40 000

20 000 -50 000

0
-100 000
-1. × 10-10 0 1. × 10-10 2. × 10-10 3. × 10-10 -1. × 10-10 0 1. × 10-10 2. × 10-10 3. × 10-10

x (m) x (m)

50 000
50 000
ψ (m-12 )

ψ (m-12 )

0 0

-50 000
-50 000

-1. × 10-10 0 1. × 10-10 2. × 10-10 3. × 10-10 -1. × 10-10 0 1. × 10-10 2. × 10-10 3. × 10-10
x (m) x (m)

Figure 6.10 Wave functions for the 4 bound states.

8 × 109 8 × 109

6 × 109 6 × 109
ψ2 (m-1 )

ψ2 (m-1 )

4 × 109 4 × 109

2 × 109 2 × 109

0 0

-1. × 10-10 0 1. × 10-10 2. × 10-10 3. × 10-10 -1. × 10-10 0 1. × 10-10 2. × 10-10 3. × 10-10
x (m) x (m)
8 × 109

5 × 109

6 × 109
4 × 109
ψ2 (m-1 )

ψ2 (m-1 )

4 × 109 3 × 109

2 × 109
2 × 109
1 × 109

0 0

-1. × 10-10 0 1. × 10-10 2. × 10-10 3. × 10-10 -1. × 10-10 0 1. × 10-10 2. × 10-10 3. × 10-10
x (m) x (m)

Figure 6.11 Wave functions squared for the 4 bound states.

The wave function squared is normalized to unity.

Example 6.22 Calculate the normalization constant for the ground state and
110  Guide to Modern Physics

evaluate the probability for the electron to be outside the box. The ground
state energy has been saved as Eg (see Ex. 6.20).
Particle in a Box  111

Since the wave function extends beyond the potential boundary with expo-
nential decay, the ground state energy may be approximated as that due to an
infinite potential of a box of size L + α2 .

Example 6.23 Estimate the ground state energy from the infinite potential
result,

The estimation is very good as long as E << V.


CHAPTER 7

Quantum Harmonic
Oscillator

The harmonic oscillator is defined by a quadratic potential,


1
U = mω2 x2 ,
2
where ω is the angular frequency of an oscillating mass m.

The Schrödinger equation is

d2 ψ 2mE 1
2
= − 2 ψ + mω2 x2 ψ.
dx h̄ 2
Many important problems in physics reduce to that of a quantum har-
monic oscillator because the leading term of a potential near its minimum
is quadratic.

7.1 GROUND STATE


7.1.1 Wave Function
The ground state wave function is of the form
2
ψ = Ae−αx .

where A and α are constants. This is a Gaussian distribution with the parame-
ter α giving the width of the wave function. It is the simplest function whose
second derivative gives the same function back again with a minus sign plus

DOI: 10.1201/9781003395515-7 112


Quantum Harmonic Oscillator  113

2.0

1.5

1.0
U

0.5

-3 -2 -1
0.0
0 1 2 3

x ( m ω)

Figure 7.1 The potential for the harmonic oscillator is quadratic.

a term which has x2 times the function. The Schrödinger equation may be
written
1 d2 ψ 2mE 1
= − 2 + mω2 x2 .
ψ dx 2
h̄ 2

1 d2 ψ
Example 7.1 Calculate ψ dx2 .

Comparing separately both x2 and x terms gives α and E.

Example 7.2 Solve for E and α.

7.1.2 Energy
Knowing the value of α, one can directly substitute ψ into the Schrödinger
equation, verifying the solution.
114  Guide to Modern Physics

Example 7.3 Solve the Schrödinger equation to get the ground state energy.

It is a general and noteworthy feature of quantum systems that the lowest


energy is not zero. Compare to a particle in a box (6.3).

7.1.3 Normalization
The constant A is determined from the normalization condition,
 ∞
ψ2 dx = 1.
0

Example 7.4 Find the normalization constant A.

7.1.4 Quantum Tunneling


A classical oscillator has maximum position (xmax ) given by conservation of
energy,
1
mω2 xmax
2
= E.
2
For the ground state energy (h̄ω/2) this would be


xmax =

(see fig. 7.2).
Quantum Harmonic Oscillator  115

#$'

#$&
)ω # ! ℏ $

#$%
ψ! "

#$!

#$#
!! !" # " !

( " ! ℏ # )ω $

Figure 7.2Wave function squared for the ground state with the classical limits
indicated by vertical lines (left and right). The square of a Gaussian is also a
Gaussian.

Example 7.5 Calculate the probability that the quantum oscillator is found
beyond the classical limit (on either side).

Examination of a plot of ψ2 (fig. 7.2) shows that the probability peaks at


x = 0. This is opposite to that of a classical oscillator that spends the least
time at x = 0 where it is moving the fastest.

Example 7.6 Calculate the probability that the quantum oscillator is found
within the center half of the classical range.

In 7.4.2 the classical probability is calculated to be 13 .

7.1.5 Uncertainty Principle


Agreement with the uncertainty principle may be checked by calculating the
expectation values of x2 and p2 and then taking the square root of the product.
116  Guide to Modern Physics

Example 7.7 Calculate the expectation value of x2 .

Example 7.8 Calculate the expectation value of p2 .

Example 7.9 Calculate the product ∆x∆p.

The absolute limit of the uncertainty ∆x∆p has been achieved because the
wave function squared is a Gaussian distribution.

7.2 FIRST EXCITED STATE


7.2.1 Wave Function
The first excited state wave function is of the form
2
ψ = Axe−αx .
The normalization constant A will differ from that of the ground state. This
function is a solution because it’s second derivative again gives the same
function back again with a minus sign plus a term which has x2 times the
function.
1 d2 ψ
Example 7.10 Calculate ψ dx2 .
Quantum Harmonic Oscillator  117

2
The x2 term of ψ1 ddxψ2 is seen to be identical to that of the ground state, there-
fore, the solution for α is the same. The constant term is 3 times that in the
ground state, giving 3 times the energy.

7.2.2 Energy
Example 7.11 Solve the Schrödinger equation to get the energy of the first
excited state.

7.2.3 Normalization
Example 7.12 Find the normalization constant A.

7.2.4 Quantum Tunneling


The classical limit of xmax for the first excited state (E = 3h̄ω/2) is

3h̄
xmax = .

Example 7.13 Calculate the probability that the quantum oscillator is found
beyond the classical limit.
118  Guide to Modern Physics

#$(

#$'

#$&
*ω # ! ℏ $

#$%
ψ! "

#$!

#$"

#$#
!! !" # " !

) " ! ℏ # *ω $

Figure 7.3 The wave function squared for the 1st excited state is shown with
the classical limits indicated by vertical lines (left and right).

7.2.5 Uncertainty Principle


Example 7.14 Calculate the expectation value of x2 .

Example 7.15 Calculate the expectation value of p2 .

Example 7.16 Calculate the product ∆x∆p.


Quantum Harmonic Oscillator  119

7.3 GENERAL SOLUTION


The normalized wave functions of the quantum harmonic oscillator may be
generally expressed as
1 2α √ 2
ψn = √ ( )1/4 Hn ( 2αx)e−ax ,
2 n! π
n


where Hn ( 2αx) is a Hermite polynomial.

7.3.1 Hermite Polynomials


1 d2 ψ
Example 7.17 Get the Hermite polynomial and calculate ψ dx2 for n = 0.

1 d2 ψ
Example 7.18 Get the Hermite polynomial and calculate ψ dx2 for n = 1.
120  Guide to Modern Physics

1 d2 ψ
Example 7.19 Get the Hermite polynomial and calculate ψ dx2 for n = 2.

1 d2 ψ
Example 7.20 Get the Hermite polynomial and calculate ψ dx2 for n = 3.

1 d2 ψ
Example 7.21 Get the Hermite polynomial and calculate ψ dx2 for n = 4.
Quantum Harmonic Oscillator  121

7.3.2 Normalization
Example 7.22 Check the normalization of the first 20 states.

7.3.3 Energy
The oscillator energy is given by
 
1
E = n + h̄ω
2

for n = 0, 1, 2, etc.

Example 7.23 Calculate the energies of the first 20 states.


122  Guide to Modern Physics

Energy (ℏ ω)

Figure 7.4 The energy levels of the quantum harmonic oscillator are equally
spaced.

7.3.4 Wave Functions


The ground state wave function has a single “bump” and each excited state
has one bump more than the state below it.

Figure 7.5 The wave functions for the first 9 states are shown.
Quantum Harmonic Oscillator  123

Figure 7.6 The wave functions squared for the first 9 states are shown.

7.4 CORRESPONDENCE PRINCIPLE


7.4.1 Large Quantum Numbers
For large quantum numbers, the solution must approach that of the classical
harmonic oscillator. The (angular) frequency of a simple pendulum consist-
ing of a mass m on a string of length L is

g
ω= ,
L
124  Guide to Modern Physics

where g is the acceleration of gravity. If the pendulum swings to a height H


above its lowest point, its energy is

E = mgH.

Example 7.24 Calculate the quantum number of a simple pendulum with


m = 0.1 kg, L = 1 m, and H = 1 mm.

7.4.2 Classical Probability


In one-half period (T/2), a classical particle traverses the entire allowed
range. The classical probability distribution (P) is
1
P= T
,
2 v(x)

where the speed is given by


    
2K 2 2 1 1
v(x) = = (E − U) = mω xmax − mω x
2 2 2 2
m m m 2 2

Using T = 2π/ω,
1
P=  .
π 2 − x2
xmax
The classical maximum displacement corresponding to energy with quantum
number n is 
(2n + 1)mω
xmax = .

Quantum Harmonic Oscillator  125

!$%!

!$#"

∫ψ% '(

!$#!

!$!"

!$!!
! " #! #"
&

Figure 7.7 The probability for a quantum oscillator to be found beyond the
classical limit is shown vs. n.

Example 7.25 Calculate the classical probability that the particle is in the
center half of the oscillation range.

Example 7.26 Calculate the probability that a quantum oscillator is in the


center half of the classical oscillation range for n = 50.
126  Guide to Modern Physics

0.20

0.15
mω / 2 ℏ )

0.10
ψ2 (

0.05

0.00

-10 -5 0 5 10

x ( 2 ℏ / mω )

Figure 7.8 The wave function squared for n=50 is shown together with the
classical probability distribution.
CHAPTER 8

Hydrogen Atom

The Bohr model (5) is highly successful in explaining many features of the
hydrogen atom. The next level of sophistication is to solve the Schrödinger
equation,
h̄2 e2
− ∇2 ψ − ψ = Eψ.
2m 4πε0 r
The solution may be obtained by the technique of separation of variables,

ψ(r, θφ) = R(r)Y(θ, φ),

resulting in radial

1 d 2 dR 2m h̄2 `(` + 1)
(r ) + (E − V − )R = 0
r2 dr dr h̄2 2mr2
and angular
1 d dY m2`
(sin θ + [`(` + 1) − 2 ]Y = 0
sin θ dθ dθ sin θ
equations. The integers ` and m` appear as part of the separation process. The
radial solutions are of the form
r
R = f (r)e− na ,

where f (r) is a polynomial and n is a positive integer. The angular solution is


the spherical harmonic. The quantity

L = h̄2 `(` + 1)

has an interpretation as orbital angular momentum with

Lz = m` h̄

DOI: 10.1201/9781003395515-8 127


128  Guide to Modern Physics

being its “z component”, the possible value of L along any given direction.
The integer n appears in the solution for allowed energies from the second
derivative in the radial equation.
E0
E= ,
n2
where E0 is the ground state energy.

8.1 GROUND STATE


The ground state wave function does not have any angular dependence and
may be written
ψ = Ae−r/a .
It is the simplest function whose Laplacian gives the terms with function re-
produced and 1/r times the function. Its quantum numbers are (n, , m ) =
(1,0,0). In spectroscopic notation, this is referred to as the 1s state.

8.1.1 Solution
The Schrödinger equation is
1 2 2me2 2mE
∇ ψ=− − 2 .
ψ 2
4πε0 h̄ r h̄
Example 8.1 Calculate ψ1 ∇2 ψ.

Comparing, separately, the terms containing r and the constant terms gives a
and E.

Example 8.2 Solve for E and a.


Hydrogen Atom  129

The parameter a appearing in the exponential of the wave function is seen to


be the Bohr radius as calculated in 5.1.

8.1.2 Normalized Wave Functions


The normalized hydrogen atom wave functions are directly accessible in
Mathematica (see A.7) and can be readily verified as solutions of the
Schrödinger equation.

Example 8.3 Get the hydrogen atom ground state wave function in terms of
the Bohr radius a.

Example 8.4 Show that this function satisfies the Schrödinger equation and
calculate the energy.

Example 8.5 Evaluate the ground state energy numerically.


130  Guide to Modern Physics

Example 8.6 Check the normalization by direct integration.

8.1.3 Radial Probability


The probability of finding the electron per volume at a given location is given
by the square of the wave function,
dP
= |ψ|2 .
dV
The integration over the angular variables gives 4π since there is no angular
dependence. This leads to a radial probability distribution,

dP 4r2
= 4πr2 |ψ|2 = 3 e−2r/a .
dr a

+(,-(. /'01(1-.-23

)*%

)*$

)*#

)*"

)*!

'!(
! " # $ % &

Figure 8.1 Radial probability for the ground state of hydrogen is plotted in
units of the Bohr radius.

The most-probable radius is found by setting the derivative of the proba-


bility function equal to zero and then solving for r. It turns out to be the Bohr
radius as can be seen from Fig. 8.1.
Hydrogen Atom  131

Example 8.7 Calculate the most probable distance for the electron to be
from the proton.

Example 8.8 Calculate the average distance.

Example 8.9 Calculate the root-mean-square distance.

It is seen that the most probable (rmp ), average (rave ), and rms (rrms ) dis-
tances are related by
rmp < rave < rrms .
Figure 8.1 shows that there is a substantial probability of finding the elec-
tron beyond the Bohr radius.

Example 8.10 Calculate probability that the electron is found beyond the
Bohr radius.

This is about 68%.


The electron has only a very tiny probability of being found inside the
proton, whose size is about 1 fm. At such a small distance, the exponential
132  Guide to Modern Physics

part of ψ2 is essentially unity, and it can either be expanded (or just set to
unity and pulled outside the integral).

Example 8.11 Estimate the probability that the electron may be found inside
the proton.

8.1.4 Consistency with the Uncertainty Principle


The expectation value for the kinetic energy can be evaluated with use of the
momentum operator,
 ∞  2
2 ∗ h̄
K = 4πr ψ − ∇2 ψdr.
0 2m

Example 8.12 Calculate the average kinetic energy and give the numerical
value.

To check the uncertainty principle, one may assign a value



∆p = 2mK
Hydrogen Atom  133

due to lack of knowledge of the direction or sign of p, as we only know its


average square. The rms radial position may be used to give

∆x = 3a.

Example 8.13 Check the uncertainty principle for the ground state of hydro-
gen.

8.2 FIRST EXCITED STATES


The first excited state has n = 2 and 2 possibilities for the quantum number
, 0 or 1, referred to as the 2s and 2p states, respectively. The = 1 state has
three possibilities for m , −1, 0, or 1. Thus, the Schrödinger equation gives 4
solutions.

8.2.1 Wavefunctions
Example 8.14 Get the n=2 wave functions.
134  Guide to Modern Physics

8.2.2 Energies
Example 8.15 Show that the 2s wave function satisfies the Schrödinger
equation and calculate the energy.

Example 8.16 Show that the 2p wave functions satisfy the Schrödinger
equation and calculate the energies.

All 4 states are seen to have the same energy which is the ground state
energy divided by n2 = 4.

8.2.3 Radial Probability Distributions


All of the quantities calculated for the 1s state may be easily reproduced for
the 2s and 2p states. The radial probability distribution for the 2s state is
shown in Fig. 8.2.
Example 8.17 Calculate the average electron distance for the 2s state.
Hydrogen Atom  135

+(,-(. /'01(1-.-23
&)!&

&)%*

&)%&

&)&*

'!(
! " # $ %& %! %"

Figure 8.2Radial probability for the 2s state of hydrogen is plotted in units


of the Bohr radius.

The 2p wave functions have angular dependence in both θ and φ. Note


also that the φ parts are represented as complex numbers. Thus, in calculating
the radial probability, we must integrate over all angles and also use

ψ2 = ψ∗ ψ.

The general form is


 2π  π
dP
= dφ dθ sin θ r2 ψ∗ ψ.
dr 0 0

Example 8.18 Compare the radial probabilities for the 2p states.

The 2p radial distribution is shown in Fig. 8.3.


136  Guide to Modern Physics

+(,-(. /'01(1-.-23

&)!&

&)%*

&)%&

&)&*

'!(
! " # $ %& %! %"

Figure 8.3Radial probability for the 2p state of hydrogen is plotted in units


of the Bohr radius.

Example 8.19 Calculate the average distance for the 2p states.

8.3 MORE EXCITED STATES


The letters used in the spectroscopic notation are s, p, d, f, g... for =
0, 1, 2, 3, 4.... After the letter f , they are just alphabetical order. A schematic
of the allowed energy levels looks like

4s 4p 4d 4 f
3s 3p 3d
2s 2p
1s

with n on the vertical and horizontal.


Hydrogen Atom  137

Example 8.20 Get the wave function for the 3d state.

Example 8.21 Calculate the average electron distance.

+&,-&. /%01&1-.-23

#'"#

#'#*

#'#)

#'#(

#'#$

%!&
! "# "! $# $!

Figure 8.4 The radial probability for the 3d state is shown.

Example 8.22 Calculate probability that the electron is found beyond 9


times the Bohr radius.
138  Guide to Modern Physics

8.4 CORRESPONDENCE PRINCIPLE


At large values of the quantum numbers n and , one expects a wave function
that is sharply peaked at a distance of n2 times the Bohr radius as the wave
function approaches a classical circular orbit.

Example 8.23 Calculate the radial probability for the (100,99,0) state of hy-
drogen.

The integer in front of a201 is a very long number (2.45... × 10716 ) whose
output has been iconized.
Figure 8.5 shows the radial probability for n = 100 and = 99.

8.5 TRANSITIONS BETWEEN LEVELS


Consider the possible transition between two states of hydrogen. The expec-
tation value of the dipole moment (er) sandwiched between the wave func-
tions of the initial and final states,

ψ∗f (−er)ψi dV,

is a measure of whether or not the transition can occur. The semi-classical


intuitive mechanism for the radiation is an oscillating electric dipole.
Hydrogen Atom  139

*&+,&- .%/0&0,-,12

"'"""!

"'""")

"'"""(

"'"""$

"'"""#

%!&
!""" #" """ #! """ $" """

Figure 8.5 Radial probability for n = 100 and = 99.

Example 8.24 Calculate the expectation value of the dipole vector connect-
ing the 2p and 1s states.

The lifetime (τ) of the state is proportional to the inverse square


 −2
µ 0 ω3
τ= ψ∗f (−er)ψi dV ,
3πh̄c

where ω is the angular frequency of the photon emitted in the transition. It is


the photon energy the transition divided by the power radiated from a classi-
cal oscillating electric dipole.
140  Guide to Modern Physics

Example 8.25 Calculate the lifetime of the 2p → 1s transition.

Electric dipole transitions are forbidden when the dipole integral is zero
(corresponding to infinite lifetime). Transitions are also forbidden when the
quantum number l does not change by one unit. This is a result of angular
momentum conservation with the photon carrying one unit of angular mo-
mentum.
Example 8.26 Calculate the dipole factor for the 2s → 1s transition (∆ =
0).

Example 8.27 Calculate the dipole factor for the 3d → 1s transition (∆ =


2).
Hydrogen Atom  141

8.6 ELECTRON INTRINSIC ANGULAR MOMENTUM


There is one more quantum number for an electron in the hydrogen atom, the
physics of which is not contained in the Schrödinger equation. The electron
has an intrinsic angular momentum, commonly referred to as spin, which can
take on 2 possible values. The spin quantum number (s) can be 12 or − 12 . This
corresponds to a contribution to the angular momentum (S ) of

3
S = s(s + 1)h̄ =
p
h̄.
2
The total angular momentum of the electron (J) is the vector addition

J = L + S.

8.6.1 Addition of Angular Momentum


The rules for adding angular momentum quantum numbers is that the sum
can take on all values from |` − s| to ` + s|, Thus,
1 1
0+ → ,
2 2
and
1 1 3
1+ → or .
2 2 2
For the n = 2 levels of hydrogen these states are labeled 1s1/2 , 2p1/2 , 2p3/2 ,
where the subscript denotes the value of the quantum number j and

J = j( j + 1)h̄.
p

8.6.2 Magnetic Moment


Compared to classical orbital motion of a charge whose magnetic moment is
q
µ= L,
2m
the electron has a gyromagnetic factor (g) that accounts for the spin quantum
number being 1/2,
e
µ = g S.
2m
The value of g for the electron is very nearly equal to −2. (Note that the
electron’s negative charge is contained in g when it is written this way.) With
142  Guide to Modern Physics

g = −2, 1/2 unit of intrinsic angular momentum gives the same magnetic con-
tribution as one unit of orbital angular momentum.
The Bohr magneton is the dipole strength of the electron’s permanent
magnet. It is the projection of the magnetic moment in some direction (S → s)
and defined with g = 2 so that the spin 12 factor exactly cancels, giving
eh̄
µB =
2m
Example 8.28 Get the Bohr magneton.

Example 8.29 Get the value of the Bohr magneton.

8.6.3 Zeeman Effect


A magnet has lower (higher) energy when aligned (anti-aligned) with an ex-
ternal magnetic field, giving an energy shift of
e eh̄
∆E = µ · B = µz Bz = (Lz + 2S z )Bz = (m + 2m s )Bz
2m 2m
compared to zero field. Here B = Bz ẑ.

Example 8.30 Calculate the magnetic field need to make a fractional energy
shift of 10−4 on the 3d5/2 state of hydrogen.

8.6.4 Spin-Orbit Interaction


From the viewpoint of the electron in a p ( = 1) state, the orbiting proton
makes a magnetic field with which the electron’s magnet moment interacts,
causing an energy shift between the 2p1/2 and 2p3/2 states, which are other-
wise degenerate in the solution of Schrödinger’s equation (8.2). The magnetic
Hydrogen Atom  143

field (B) may be estimated as that at the center of of a tiny current loop with
radius (r). From Ampère’s law,

µ0 (Idl) × r̂ µ0 2πr µ0 I
B= = = ,
4π r2 4π r2 2r
where µ0 is the magnetic constant and I is the current from a proton charge
v
with orbit frequency ( 2πr ),
ev
I= .
2πr
This gives
µ0 ev
B= .
4πr2
Example 8.31 Get the magnetic constant.

The magnetic constant, is related to the electric constant by


1
µ0 = .
c2 ε0
Example 8.32 Check the relationship between the electric and magnetic
constants.

The energy shift caused by the electron magnetic moment is


e  e   h̄   µ ev 
0
∆E = µ · B = g S · B = 2 .
2m 2m 2 4πr2
For the first excited state in the Bohr model (5.2.2),

v= ,
2
and
4h̄
r= .
αmc
This gives
α4 mc2
∆E = .
64
144  Guide to Modern Physics

Example 8.33 Solve for ∆E.

In solving for ∆E in Ex. 8.33, the input quantities are user-defined vari-
2
ables (not units) to allow the substitutions µ0 → c21ε and h̄ → 4πεe0 αc . The
0
energy levels are themselves proportional to α2 and the correction to them is
proportional to α4 . This is how α got its name as the fine structure constant.
Note that the correction to spin up (down) is positive (negative) so the energy
difference between the 2p3/2 and 2p1/2 states is 2∆E.

Example 8.34 Calculate the energy difference between the 2p3/2 and 2p1/2
states.

Despite using a simple Bohr orbit for getting the magnetic field, the en-
ergy splitting turns out to be correct. Such semi-classical classical calcula-
tions are often extremely physically intuitive compared to a full calculation
with wave functions.

8.6.5 Hyperfine Splitting


An electron in the ground hydrogen atom has zero orbital angular momen-
tum, however, it will still feel the magnetic field caused by the internal magnet
of the proton (due to intrinsic angular momentum). There will be an energy
difference between an electron having spin up or spin down (Fig. 8.6).
The energy difference between spin orientations may be estimated by
evaluating the magnetic field (Fig. 8.7) from the proton at the location of the
electron. The proton magnetic moment is
eh̄
µ p = gp ,
2mp
Hydrogen Atom  145

e p p

Figure 8.6 Visualization of the spin-flip transition in hydrogen. The state with
spins aligned is higher energy than spins opposite.

1.0

0.5

0.0
z

-0.5

-1.0

-1.0 -0.5 0.0 0.5 1.0


y

Figure 8.7 Dipole field from the proton magnetic moment.

where analogous to the electron case, gp is the proton g-factor to account for
the charge-to-mass ratio and mp is the proton mass.
The expression for the field can be calculated from

µ0 ẑ
A= µp ∇ × ,
4π r
and
B = ∇ × A.
146  Guide to Modern Physics

Example 8.35 Calculate the magnetic dipole field.

The magnetic field from the proton dipole at the location of the electron (Bohr
radius, a) is
µ0
B= µp .
4πa3
Notice that electron spin down is aligned with the magnetic field from the
proton (Fig. 8.7) and spin up is anti-aligned. Therefore, spin down is lower
energy than spin up. The energy difference is
 
µ0 eh̄
∆E = 2 gp ge
4πa3 mp

Example 8.36 Calculate the difference in energy levels for electron and pro-
ton spins aligned or anti-aligned.
Hydrogen Atom  147

This is the correct order of magnitude for the observed energy difference.
To get a more accurate answer, on needs to use the 1s hydrogen wave func-
tion (8.1.2) and evaluate its square over the magnetic field produced by the
proton. The only contribution is at the origin where the magnetic field is a
delta function. The magnetic field at the proton location is
2
B = µδ(r).
3
The delta function has the property that its value is zero everywhere except
at the origin where it is infinite, such that its integral gives unity. It can be
thought of as a Gaussian with unit area under the curve in the limit where its
width goes to zero.

Example 8.37 Calculate the energy difference using the wave function.

This answer is correct to order α.

Example 8.38 Calculate the wavelength of the radiation emitted when the
electron flips its spin.

This is the famous 21-cm line from hydrogen that allowed the mapping
of the spiral nature of the Milky Way galaxy due to its ability to penetrate
interstellar dust.
The lifetime of the hyperfine transition can be calculated from the classi-
cal formula for radiated power,
µ2p ω4
P = µ0 .
12πc
The lifetime is
∆E
τ= .
P
148  Guide to Modern Physics

Example 8.39 Calculate the lifetime of the hyperfine transition.

8.6.6 Lamb Shift


In the solution to the Schrödinger equation, even after accounting for elec-
tron spin and spin-orbit interaction, the 2s1/2 and 2p1/2 states have identical
energy (8.2.2). The wave functions (Figs. 8.2 and 8.3) of these 2 states, how-
ever, are very different. The electron has a self-interaction in which it can
emit and absorb virtual photons. (A virtual photon is one that never enters
the real world, but can “exist” on borrowed energy for a short time allowed
by the uncertainty principle.) The effect of the self-interaction spreads out
the electron on a distance scale of about 0.1 fm. The 2s wave function has a
little piece that brings the electron close to the nucleus. Without the electron
self-interaction, the energies are identical, but when it is included, the 2s1/2
energy becomes slightly higher because the smearing results in the electron
feeling a smaller force. A smaller force translates to less binding energy and
a larger total energy.
The energy difference between the 2s1/2 and 2p1/2 states is called the
Lamb shift after Willis lamb who first measured the energy difference by
measuring the transitions

2s1/2 → 2p1/2 → 1s1/2 .

(Note that the direct transition 2s → 1s is forbidden.) The frequency of the


photons in the transition 2s1/2 → 2p1/2 is 1058 Hz. The Lamb shift is the
same order of magnitude as the hyperfine splitting (8.6.5) and is proportional
to the wave function squared at the origin.
Hydrogen Atom  149

Example 8.40 Calculate the wave function squared at the origin for the 2s1/2
and 2p1/2 states.

There is no shift for the 2p1/2 state. The 2s1/2 state is shifted up by an
amount proportional to
1 (mc2 )3
ψ2s (0)2 = = .
8πa3 8π(4πh̄2 ε0 )3
The expression to leading order is,
2
 
5 mc 1
∆E = α ln .
6π 8.9α2
Example 8.41 Calculate the Lamb shift and transition frequency.

This result is good to about 2%.


150  Guide to Modern Physics

Example 8.42 Use the measured value of the 2s1/2 Lamb shift to estimate
the Lamb shift frequency for the 3s1/2 state.

The measurement of the Lamb shift and the calculation that followed ush-
ered in the development of quantum electrodynamics which is the foundation
on which the standard model of particle physics was constructed.

8.6.7 The Electron g-Factor


Due to the self-interaction of the electron, the spin component (g/2) con-
tributes slightly more to the total magnetic moment than one unit of orbital
angular momentum. If the g-factor were exactly 2, there would be no differ-
ence in the energies of the states. The self-interaction alters the value of g by
about 0.1% (8.6.2). Thus, the electron in a 2s1/2 state is a slightly stronger
magnet than in the 2p1/2 state.

Example 8.43 Get the g-factor of the electron.

The g-factor of the electron has been calculated to 10 decimal places (or-
der α5 ) using the theory of quantum electrodynamics. The leading corrections
are  
|g| − 2 α  α 2 197 π2 2
= + + − 2π ln 2 + 3ζ(3) ,
2 2π 2π 36 3
where  ∞
1 x2
ζ(3) = ≈ 1.20206.
2 0 ex − 1
Hydrogen Atom  151

|g|−2
Example 8.44 Calculate 2 to order α2 .
CHAPTER 9

Statistical Physics

9.1 PROBABILITY DISTRIBUTIONS


Probability distributions are at the heart of modern physics and have been
encountered in thermal radiation (2.2), radioactive decay (3.1) as well as ψ2
from the solution to the Schrödinger equation (6.5). In general, a probability
distribution may be written as P(x) which represents some probability per
unit of x. The function is usually normalized to unity,

P(x)dx = 1.

9.1.1 Binomial Distribution


Some distributions like flipping a coin are not continuous but have discreet
outcomes. The binomial distribution fb (x) represents the probability of get-
ting an integer result x from n trials when the probability per trial is p,

n!p x (1 − p)n−x
fb (x) = .
x!(n − x)!
The normalization condition is
n
fb (x) = 1.
x=0

Example 9.1 Sum all the binomial terms.

DOI: 10.1201/9781003395515-9 152


Statistical Physics  153

Example 9.2 Calculate the probability of getting a coin toss in your favor
exactly 5 times out of 10 tosses, assuming the probability of the coin landing
either way is 1/2.

&'!(

&'!&

&'%(

&'%&

&'&(

! " # $ %&

Figure 9.1 The binomial distribution is shown for n = 10 and p = 1/2.

The average is
x = np,
and the rms is 
xrms = np(1 − p).

Example 9.3 Calculate the average of a binomial distribution.

Example 9.4 Calculate the rms of a binomial distribution.


154  Guide to Modern Physics

In the limit where n → ∞, the variable x becomes continuous and the bi-
nomial distribution because a Gaussian, provided that the average np is not
too small. In the case where p << 1, one gets a Poisson distribution.

0.025

0.020

0.015

0.010

0.005

200 400 600 800 1000

Figure 9.2 The binomial distribution is shown for n = 1000 and p = 1/2. The
distribution has become Gaussian.

0.3

0.2

0.1

2 4 6 8 10

Figure 9.3 The binomial distribution is shown for n = 1000 and p = 0.001.
The distribution has become Poisson.
Statistical Physics  155

9.1.2 Poisson Distribution


The Poisson distribution corresponds to large n and small x and is obtained
from the binomial distribution with the approximations

(1 − p)n = en ln(1−p) ≈ e−np ,

and
n!
= (n)(n − 1)(n − 2)...(n − x) ≈ n x .
(n − x)!
With both p << 1 and x << n, the binomial distribution reduces to the Poisson
distribution
e−a a x
fP (x) = ,
x!
where a = np, the average.

Example 9.5 Expand ln(1 − p) about p = 0.

The Poission distribution is normalized to unity.

Example 9.6 Sum all the Poisson terms.

A major-use case of the Poisson distribution comes about when one wants
to know an upper limit on the average value (a) when a small number of oc-
currences are observed. For example, suppose 0 events are observed. What is
the upper limit on a at 90% confidence level (CL)? The answer is given by

e−a a0
fP (0) = = e−a = 0.1.
0!
Solving for a,
a = − ln 0.05 = 2.3.
One says that a is less than 2.3 at 90% CL.
156  Guide to Modern Physics

Example 9.7 Find the 90% CL limits for 0 to 3 events observed.

Example 9.8 Find the 95% CL limits for 0 to 3 events observed.

9.1.3 Gaussian Distribution


The Gaussian distribution is the result of the sum of random processes. As
pointed out, it is obtained from the binomial distribution in the limit of large n
and np that is not small. Figure 9.4 shows a histogram of the sum 10 random
numbers between 0 and 1, done 106 times. The result is a Gaussian.

$) )))

#) )))

") )))

!) )))

)
! " # $ % & ' (

Figure 9.4 For a histogram of the sum of 10 random numbers between 0 and
1, the average is 5.00 and the rms is 0.913.
Statistical Physics  157

The Gaussian distribution may be written


2 /2σ2
fG = Ce−(x−a) ,
where a is the average, σ is the standard deviation, and C is the normalization
constant. When the average is zero,
2 /2σ2
fG = Ce−x ,
Example 9.9 Calculate the normalization constant.

Example 9.10 Calculate the Gaussian rms.

About 32% of the distribution is outside 1 σ (on either side).


Example 9.11 Calculate the probability to exceed n sigma (on either side),
for n from 1 to 5.

9.2 MAXWELL-BOLTZMANN DISTRIBUTION


The Maxwell-Boltzmann (MB) distribution applies to a large number of par-
ticles that are in thermal equilibrium and are distinguishable because they are
far apart and their wave properties do not overlap. It is the classical regime.

9.2.1 The Ideal Gas


Temperature on an absolute scale is a measure of average kinetic energy. For
an ideal gas, that scale may be defined as
3 1
K ≡ kT = m v2 .
2 2
158  Guide to Modern Physics

Thus, temperature is also a measure of the rms speed.

Example 9.12 Calculate the rms speed of nitrogen molecules at T = 300 K.

9.2.2 Gas Pressure


The gas pressure is proportional to the average speed squared. Consider gas
molecules hitting a walls of a container. The pressure gets one power of v
from momentum transfer during a collision and another power of v for the
probability per time of hitting the wall. This gives P ∼ v2 ∼ T and is written
as the ideal gas law,
N
P = kT.
V
Example 9.13 Calculate the volume of 1 mole (Avogadro’s number) of
molecules at standard temperature and pressure (STP) of 273 K and 1 atm.

9.2.3 Mean Free Path


A cylinder traced out by a molecule as it moves between collisions may be
equated to the volume that contains one molecule (see Fig. 9.5),

πa2 l ≈ d3 ,

where a is the molecular size, d is the average distance between molecules,


and l is the mean free path between collisions. This gives

d3
l= .
πa2
Statistical Physics  159

Writing n = d−3 is the number of particles per volume,


1
l= ,

where σ = πa2 as the collision cross section as defined in 3.6.3. The mean
free path is inversely proportional to both the number of particles per volume
and the collision cross section.
#

"

Figure 9.5 Schematic to indicate the mean free path (l) and its qualitative
relationship to molecular size (a) and average distance between molecules
(d).

Example 9.14 Estimate the mean free path of a molecule at STP.

Example 9.15 Estimate the average distance between molecules at STP.

The average kinetic energy may be written

p2 3
= kT.
2m 2
160  Guide to Modern Physics

Therefore, the rms momentum is



 2 √
p = 3kT .

Example 9.16 Calculate the de Broglie wavelength of nitrogen molecules at


STP.

The idea gas is characterized by

a << d << l.

9.2.4 Velocity Distribution


The velocity distribution for each of the x, y, x components is Gaussian as a
result of random collisions and may be written

m −mv2x /2kT
f (v x ) = e ,
2πkT

m −mv2y /2kT
f (vy ) = e ,
2πkT

m −mv2z /2kT
f (vz ) = e .
2πkT
Each component is symmetric about zero and has zero average.

Example 9.17 Check the normalization of the velocity distribution.


Statistical Physics  161

"$""!&

"$""!%

"$""!"
")#*$!!

"$"""(

"$"""'
# "!" $

"$"""&

"$"""%

"$""""
!!""" !#"" " #"" !"""
!" ")#*$

The Maxwell-Boltzmann velocity distribution is shown for nitro-


Figure 9.6
gen molecules at STP.

Example 9.18 Calculate the rms of the velocity distribution.

It is seen that
1 1 1 1
m v2x = m v2y = m v2z = kT.
2 2 2 2
Each component of the velocity contributes 12 kT to the kinetic energy. This
result is known as the equipartition theorem. Together, the 3 translational
components contribute 32 kT to the kinetic energy in agreement with the def-
inition of temperature (9.2.1).

9.2.5 Speed Distribution


The speed distribution depends only on the magnitude, not direction. The
speed distribution is obtained by integrating over all angles,

dv x dvy dvz = v2 sin θdvdθdφ,


162  Guide to Modern Physics

which produces a factor of v2 , The resulting distribution of particle speeds in


thermal equilibrium is
 m 3/2 2
f (v) = 4π v2 e−mv /2kT
2πkT
The distribution is normalized to unity.

Example 9.19 Integrate the speed distribution from 0 to ∞.

The most probable speed is found by setting the derivative of the distribution
equal to 0.

Example 9.20 Calculate the most probable speed.

The dimensionless speed distribution N(x) in units of most probable speed is


4
N(x) = √ x2 e−x
π

Example 9.21 Calculate fraction of particles that are within 10% of the most
probable speed.

The average speed is calculated from


 ∞
vave = v f (v)dv
0
Statistical Physics  163

!&(

!&'
"#"
# "#

!&%
!
$ !!"

!&#

!&!
! " # $ %

! "
# "#
!
!

Figure 9.7The Maxwell speed distribution is shown in units of the most prob-
able speed.

Example 9.22 Calculate the average speed.

The root-mean-square (rms) speed is calculated from




vrms = v2 f (v)dv
0

Example 9.23 Calculate the rms speed.


164  Guide to Modern Physics

Example 9.24 Calculate the most probable, average, and rms speeds for ni-
trogen molecules at T = 300 K.

9.2.6 Energy Distribution


The energy distribution may be obtained from the speed distribution by a
change in variables,
1
E = mv2 ,
2
where E is the kinetic energy. The differential is
dE = mvdv,
and 
2 v 1 2E
v dv = dE = dE.
m m m

Thus, the energy distribution goes as Ee−E/kT . The normalized distribution
is
2 √
f (E) = √ (kT )−3/2 Ee−E/kT
π
Example 9.25 Verify that the energy distribution is normalized to unity.

Example 9.26 Calculate average energy.


Statistical Physics  165

The average kinetic energy from the Maxwell-Boltzmann distribution is


seen to be in agreement with the definition of temperature (9.2.1).

9.3 QUANTUM DISTRIBUTIONS


The number density of particles with energy E may be written as

n(E) = ρ(E) fMB ,

where ρ(E) is a function called the density of states (discussed further in


9.3.4) that gives the number of states per energy and fMB is the probability
factor that the state is occupied,
1
fMB = .
eE/kT
It is exponentially rarer to be in a state of higher energy. This holds true at
lower energies (longer wavelengths), only if the particles are distinguishable,
i.e., the average distance between particles is larger than their wavelength as
calculated for an ideal gas in 9.2.3.
In the quantum world where the wave properties are important, the parti-
cles are no longer distinguishable by their position and there are two classes:
particles like the electron that obey the exclusion principle, and particles like
the photon that do not.

9.3.1 Bose-Einstein Distribution


The Bose-Einstein (BE) distribution function applies to particles with integer
spin, namely the photon. Its properties are derived from the fact that, unlike
electrons, there is no exclusion principle preventing a large number of pho-
tons from occupying the same state. The BE distribution is subtly different
from the MB distribution at low energies ( kTE
<< 1) where it becomes much
larger. It is written
1
fBE = E/kT .
e −1
The main example of this is thermal radiation where the average energy per
quantized oscillator was calculated (2.2) to be
Ep
hEosc i = E /kT
,
e p −1
where Ep = hc/λ is the photon energy. The −1 in the denominator is crucial
for understanding the distribution of photons at low energy.
166  Guide to Modern Physics

9.3.2 Fermi-Dirac Distribution


The Fermi-Dirac (FD) distribution applies to particles with half-integer spin,
namely the electron. No two electrons are allowed to be in the same state. The
FD distribution is subtly different from the MB distribution at low energies,
and is written
1
fFD = E/kT
Ae +1
where A is a parameter that is temperature dependent and is conventionally
written as
A = e−EF /kT .
The constant EF is called the Fermi energy, and it denotes the boundary be-
tween filled and open states at low temperature (EF >> kT ). This FD distri-
bution,
1
fFD = (E−E )/kT ,
e F +1
gives approximately 1 when E < EF and 0 when E > EF (Fig. 9.8).

1.0

0.8

0.6
fFD

0.4

0.2

0.0

0.0 0.2 0.4 0.6 0.8 1.0 1.2


E (EF )

Figure 9.8 Fermi-Dirac distribution function. The energy is plotted in units of


EF

9.3.3 Comparison of the Distribution Functions


Figure 9.9 shows a comparison of the distribution functions at low energies.
The FD distribution is unity (one electron per state) up to energy EF (not
shown), after which it drops rapidly to zero (see Fig. 9.8). The BE distribu-
tion goes to ∞ as the energy goes to 0.
Statistical Physics  167

2.0

1.5
f

1.0

0.5

0.0
0.0 0.5 1.0 1.5 2.0
E (kT)

Figure 9.9 Comparison of the distribution functions at low energy. The FD


distribution drops to zero quickly above the Fermi energy and the BE and
MB distributions go to zero exponentially at large energy.

9.3.4 Density of States


The distribution functions give occupancy probability but do not give the
number of states which is energy dependent. For that we need another func-
tion called the density of states. For an √
ideal gas, the density of states is
proportional to the number density and E. Using the result of 9.2.6, the
density of states for the ideal gas is
N 2 √
ρ(E) = √ (kT )−3/2 E.
V π

For distributions of electrons or photons, one can get the density of states
using the concept of phase space, which is counting the number of waves that
can fit inside a given volume in the same way as done in 2.1 with λ replaced
by the de Broglie wavelength h/p. The number of states per volume is

3 πp
4 3
ns (p) = g 3 ,
h
where the factor g accounts for the degeneracy of the states due to spin. The
density of states is
p
dns dns d p 4πp2 d( E 2 − (mc2 )2 /c)
ρ(E) = = =g 3 .
dE d p dE h dE
168  Guide to Modern Physics

Taking the derivative


4πp2 E 4πp2
ρ(E) = g = g ,
h3 pc2 h3 v
since the speed is given by v = pc/E. Thus, the density of states is propor-
tional to the momentum squared divided by the speed.

9.3.5 Photon Gas


Applying this to photons, one gets
g=2
for 2 possible polarizations, and
p2 E 2
= 3.
v c
The photon distribution becomes
8πE 2 1
n(E) = ρ(E) fBE = .
3
(hc) e E/kT −1
This is the number of photons per volume per energy that corresponds to
thermal radiation. The energy per volume is E times this.
The energy density (u) of thermal radiation is (2.1.3)
4
u = σT 4 .
c
The average photon energy is the ratio of total photon energy per volume
divided by the number of photons per volume. The number per volume is
obtained by integrating n(E),
4
σT 4
E =  ∞c .
0
n(E)dE
Example 9.27 Calculate the average energy of thermal photons as a function
of T .
Statistical Physics  169

9.3.6 Electron Gas


For a nonrelativistic electron,
g=2
for two spin states, and
4πp2 4π 2mE 4π(2m)3/2 √
ρ(E) = g = 2 √ = E.
h3 v h3 2E/m h3
The electron distribution becomes
4π(2m)3/2 √ 1
n(E) = 3
E (E−E )/kT .
h e F +1
An expression for the Fermi energy may be obtained by integrating n(E)
to get the number density (N/V).
 ∞  ∞
N 4π(2m)3/2 √ 1
= n(E)dE = 3
E (E−E )/kT dE.
V 0 0 h e F +1
The integration may be done at T = 0 because the Fermi energy does not
depend on temperature as long as EF >> kT .

N 4π(2m)3/2 EF √ 4π(2m)3/2 2 3/2
= EdE = E .
V h3 0 h3 3 F
Solving for the Fermi energy,
 2/3  2/3
h2 3 N
EF = .
8m π V
For a metal with one conduction electron, one may estimate
N 1
≈ .
V (0.3 nm)3
Example 9.28 Estimate the Fermi energy for a metal with one conduction
electron per atom.

This is the correct order of magnitude. The Fermi energy for metals is a few
eV. To calculate the Fermi energy more accurately, one may use the known
density of the metal to get N/V.
170  Guide to Modern Physics

Example 9.29 Calculate the Fermi energy of copper which has one conduc-
tion electron per atom.

Example 9.30 Calculate the average energy of a conduction electron.

9.3.7 Superfluid Helium


Helium is a boson with spin zero. Therefore, it obeys BE statistics. The den-
sity of states is similar to nonrelativistic electrons, except there is no factor
of 2 for spin states, giving

2π(2m)3/2 √ 1
n(E) = E E/kT .
h3 e −1
The number per volume is
 ∞ √
N 2π(2m)3/2 1
= E dE.
V h3 0 eE/kT −1
Making a change of variables, x = E/kT ,
 √
N 2π(2mkT )3/2 ∞ x
= dx.
0 e −1
V h3 x
Statistical Physics  171

Example 9.31 Evaluate the integral above.

Solving for T,
 2/3
h2 (N/V)
T= .
2mk 2π 2.315
which is the temperature at which liquid helium would obey BE statistics, i.e.,
become a superfluid. This process is called a Bose-Einstein condensation.

Example 9.32 Evaluate the temperature at which He becomes a superfluid.


Take the mass density at low temperature to be 140 kg/m3 .

The answer above is very close to the true value of 2.17 K. To get a more
precise answer, one would need knowledge of a constant A which needs to be
added to the BE probability (compare to the FD case where the Fermi energy
was introduced),
1
fBE = E/kT .
Ae −1
which has been set to 1 as in the case of the thermal distribution where it gave
the correct answer.
The next candidate, after helium, to be a superfluid might be neon, but its
melting point is too high , i.e., it is a solid at the temperature it would need to
reach to condense to a superfluid.
172  Guide to Modern Physics

Example 9.33 Evaluate the temperature at which Ne would need to reach


to become a superfluid and compare it to the melting point. Take the mass
density at low temperature to be 1200 kg/m3 .
CHAPTER 10

Astrophysics

10.1 THE SUN


10.1.1 Proton Cycle
The primary mechanism by which energy is produced in the sun is by the
proton cycle. This is a multistep process in which the first step is the fusion
of 2 protons. Two protons cannot make a stable bound state because of their
electrical repulsion, but the bound state of a proton and a neutron, a deuteron
(d), is stable. The reaction is

p + p → d + e+ + νe .

The neutrino (νe ) is produced to conserve lepton number. The kinetic energy
released is the sum of the mass energies on the left minus that on the right.
the neutrino is essentially massless. This may be followed by

p + d → He3 + γ,

and
He3 + He3 → p + p + α,
The net reaction is
4p → α + 2e+ + 2νe + 3γ.
There are other variances that have the same net reaction. The positron anni-
hilates with an electron to give its mass energy back,

e+ + e− → γ + γ,

DOI: 10.1201/9781003395515-10 173


174  Guide to Modern Physics

Example 10.1 Calculate the amount of energy released in the proton cycle.

10.1.2 Distance to the Sun


The distance from the earth to the sun is called 1 astronomical unit (au).

Example 10.2 Convert au to m.

The parsec (pc) is a length unit defined by the distance at which the mean
radius of the earth’s orbit subtends an angle of one second of arc.

Example 10.3 Convert pc to m.

The Mpc is in common use in astrophysics.

10.1.3 Solar Constant


The amount of solar energy reaching the earth, the solar constant ( f ) is mea-
sured to be
f = 1.36 × 103 W/m2 .
The total power output if the sun (luminosity) is

Ls = 4π(1 au)2 f.

Example 10.4 Calculate the solar luminosity.


Astrophysics  175

10.1.4 Temperature of Sun


We may use the solar luminosity together with the Stefan-Boltzmann law to
calculate the surface (blackbody) temperature of the sun.
Ls
R = σT 4 = ,
4πrs2
where rs is the radius of the sun.

Example 10.5 Calculate the temperature of the sun.

10.1.5 Neutrino Flux from the Sun


The sun produces energy by fusing protons (see 10.1.1). Each proton cycle
releases 27 MeV of energy and produces 2 neutrinos.

Example 10.6 Calculate the solar neutrino flux at the surface of the earth.

That is an enormous number of neutrinos, often pointed out to be about 1011


through a thumbnail every second!

10.2 MAGNITUDE SCALE FOR SKY OBJECTS


10.2.1 Apparent Magnitude
Apparent magnitude is a measure of how bright a star or other sky object
appears. The brightness depends on the luminosity, distance, and how much
light is absorbed by material between the earth and the object. The bright-
ness is characterized by the apparent magnitude parameter m on a reverse
logarithmic scale such that brighter objects have smaller numbers. Apparent
176  Guide to Modern Physics

magnitude is scaled such that 5 units of m correspond to a factor of 100 in


brightness. The brightest stars in the night sky are approximately m = 1 and
the weakest to the naked eye are about m = 6. By comparison, the sun has
m = −26.7.

10.2.2 Absolute Magnitude


Example 10.7 Convert Mpc to light years.

The absolute magnitude (M) is defined as the apparent magnitude an ob-


ject would have if viewed from a distance of 10 pc. For the The equation
that relates apparent magnitude m and absolute magnitude M of a star at a
distance d is called the distance modulus:
d
M − m = −5 log10 .
10 pc
Example 10.8 Calculate the absolute magnitude of the sun.

10.3 THE MILKY WAY


The distance of the sun from the center of the galaxy is measured by motion
of the stars. The answer is about 8 kpc. The rotation speed of the sun about
the galactic center is 2.3 × 105 m/s. The age of the sun, estimated by radioac-
tive dating of the oldest meteorites, is about 4.6 By.

Example 10.9 Calculate the number of rotations the sun has made about the
Milky Way center.
Astrophysics  177

Newton’s law for uniform circular motion gives

GMg ms ms v2
= .
R2 R
This gives the galactic mass (MG ) in terms of the solar mass (ms ) to be

MG Rv2
= .
ms Gms

Example 10.10 Calculate the galactic mass in terms of the solar mass.

10.4 WHITE DWARF


A star that has consumed all its fusion energy may collapse to
become a white dwarf which is stable against further gravitational
collapse by electron pressure caused by the Pauli exclusion princi-
ple. A white dwarf is very dense having a solar mass in an earth
size.

Example 10.11 Estimate the density of a white dwarf is solar masses.

There is a maximum mass to the white dwarf beyond which it will col-
lapse further to form a neutron star or black hole. This maximum mass is
called the Chandrasekhar limit, the order-of-magnitude of which is given by
 3/2
1 h̄c
M∼ 2 .
mH G
178  Guide to Modern Physics

Example 10.12 Estimate the maximum size of a white dwarf.

A more precise calculation gives about 1.4 solar masses for the maximum
white dwarf mass.

10.5 NEUTRON STAR


A neutron star is formed from the gravitational collapse of a massive star.
Further gravitational collapse is halted by the strong force when the neutron
star density reaches that of nuclear density.

10.5.1 Density
Example 10.13 Estimate the neutron star density in kg/m3 and (short) tons
per tsp.

That’s a whopping 2 billion tons per teaspoon!

10.5.2 Fermi Energy


Example 10.14 Estimate the Fermi energy (see 9.3.6) for a neutron star.
Astrophysics  179

10.5.3 Binding Energy


The gravitational potential energy for a differential mass dm separated a dis-
tance r from. a mass m is
Gmdm
dU = .
r
The mass is related to the density by
4
m = πr3 ρ,
3
so
G 2/3
dU =  1/3 m dm.
3
4πρ

Assuming constant density which can be written in terms of the neutron star
mass M and radius R (ρ = 4 M 3 ), the binding energy is
3 πR

 M  
G 2/3 G 3 3GM 2
Eb =  1/3 m dm =  1/3 M 5/3 = .
3 0 3 5 5R
4πρ 4πρ

Example 10.15 Calculate the binding energy of a neutron star with 1.4 solar
masses and a radius of 10 km.
180  Guide to Modern Physics

10.6 BLACK HOLES


10.6.1 Schwarzschild Radius
For a mass M, the distance beyond which no particle, including photons, can
escape. may be calculated in general relativity and is called the Schwarzschild
radius (rs ).
2GM
rs = 2 .
c
This gives the formula for the size of a black hole vs. mass.

Example 10.16 Calculate the Schwarzschild radius of the sun.

10.6.2 Hawking Radiation


A black hole can radiate when pair production near the event horizon allows
one particle to escape with the other falling into the black hole. The energy
for the process comes from the gravitational field of the black hole. The re-
sulting radiation follows the blackbody formula with temperature

h̄c3
T= .
8πkGM
Example 10.17 Find the blackbody temperature of a black hole of mass 3
times that of the sun
Astrophysics  181

The radiated power per area is given by the Stefan-Boltzmann law, where the
area comes from the Schwarzschild radius. The equation is
d(Mc2 ) h̄c3 4 c8 h̄4 σ
= −4πσrs2 ( ) =− ,
dt 8πkGM 256π3 k4G2 M 2
or  0  τ
2 c8 h̄4 σ
M dM = − dt,
M0 256π3 k4G2 0
where M0 is the initial mass. The solution for the blackhole lifetime is
256π3 k4G2 M03
τ= .
3c6 h̄4 σ
Example 10.18 Calculate the lifetime of a black hole having a mass of 3 so-
lar masses.

10.7 THE DARK NIGHT SKY


Suppose stars with radius (R) and luminosity of the sun were distributed ran-
domly with number density n in an infinite universe. The mean free path (see
3.7.4) of light before being absorbed and reradiated by a star is
1
d= .
nπR2
Example 10.19 Calculate the mean free path of starlight in an infinite uni-
verse. Take the visible mass density to be 3 × 10−28 kg/m3 .
182  Guide to Modern Physics

It would take 1031 s for light to travel from a star at distance d. The night
sky would be bright because in any direction out to far enough distance there
would be star in the line of sight. However, the universe is evolving on a time
scale that is much faster than 1031 s and the night sky is dark. In other words
there are no stars to be seen at such a distance.

10.8 HUBBLE’S LAW


Hubbles law gives the relationship between velocity, as determined by ob-
served cosmological redshift, and distance
v = H0 r,
where H0 is called the Hubble constant.

10.8.1 Hubble Constant


The measured value of the Hubble constant is
km/s
H0 = 73 .
Mpc
The inverse of the Hubble constant is called the Hubble lifetime which
is the expansion time of the universe, excluding the effect of gravitational
pull.

Example 10.20 Calculate the Hubble lifetime.

10.8.2 Cosmic Redshift


The relationship between redshift (z)and β (v/c) is

1+β
1+z = ,
1−β
or
(1 + z)2 − 1
β= .
(1 + z)2 + 1
Astrophysics  183

Example 10.21 Calculate the distance to a galaxy which has a redshift of 1.

10.9 COSMIC BACKGROUND RADIATION


The cosmic background radiation (CMB) is one of the cornerstones of cos-
mology. The CMB is left over from the early universe when radiation was
in equilibrium with a hot, dense plasma of electrons, protons, and neutrons.
When the plasma cooled, the first hydrogen and helium atoms were formed
and the radiation decoupled from the charges that were no longer free. As the
universe expanded, the CMB cooled to it present temperature of about 2.73 K.
The CMB spectrum fits that of a blackbody, and has been measured very
accurately. The CMB temperature together with the Stefan-Boltzmann law
gives the energy density (u) to be
4
u = σT 4 .
c
Example 10.22 Calculate the energy density of cosmic photons.

In 2.2.5 the photon flux from blackbody radiation was calculated. The
photon number density (n) in a cavity (in this case the universe) is given by
dn 1 du
= ,
dE E dE
with   
du 4  2πE 3 
 .
= 
dE c h3 c2 (e kT − 1) 
E
184  Guide to Modern Physics

E
One integrates over all energies to get n. With the change invariables, x = kT ,
 3  ∞
kT x2
n = 8π .
hc 0 ex − 1

Example 10.23 Calculate the CMB photon number density.

10.10 COSMIC NEUTRINO BACKGROUND


At a temperature corresponding to kT = 2.5 MeV, neutrinos would decouple
from matter. After expansion from that extremely early time to now, the uni-
verse is predicted to be left with a cosmic neutrino background (CNB) at a
temperature of 1.95 K. The CNB is a black body distribution analogous to the
CMB and the same technique (10.9) can be used to calculate their numbers.
The CNB has not been observed due to the low neutrino interaction proba-
bility.

Example 10.24 Calculate the energy density of cosmic neutrinos.

Example 10.25 Calculate the CNB photon number density.


Astrophysics  185

10.11 CRITICAL MASS DENSITY


The critical mass density (ρc ) is defined to be the mass density of the universe
beyond which expansion cannot continue forever. It is given by the condition
on escape velocity
1 2 GMm
mv = ,
2 R
Hubble’s law
v = H0 R,
and definition of mass density
M
ρc = 4
.
3
3 πR

The result is
3H02
ρc = .
8πG
Example 10.26 Solve for the critical density.

Example 10.27 Calculate the critical mass density.

This is about 5 hydrogen atoms per m3 .

10.12 PLANCK MASS


The Planck scale is defined by the energy that gives a dimensionless strength
(analogous to electromagnetic α) of unity for gravity. It is an energy scale
186  Guide to Modern Physics

where the physics is unknown. The condition is


h̄c
= 1.
GmP c2

Example 10.28 Calculate the Planck mass.

10.12.1 Planck Length and Planck Time


Apart from a factor of 2π, the Planck length is the wavelength of a photon
that has an energy of MP c2 .

Example 10.29 Calculate the Planck length.

The Planck time is the time it would take a photon of energy MP c2 to travel
a distance equal to the Planck length.

Example 10.30 Calculate the Planck time.


Astrophysics  187

10.12.2 Relationship to Schwarzschild Radius


At the Planck mass, the Compton wavelength becomes

hc hc 2πhcG
λc = 2
= q = ,
MP c hc c2
c2 2πG

and the Schwarzschild radius is


q q
hc 2
2GMP 2G 2πG π hcG
rs = = = .
c2 c2 c2
Thus, they are the same order of magnitude.
APPENDIX A

Mathematica Starter

A.1 CELLS
Mathematica notebooks have 2 types of cells: “text” cells that can only dis-
play what is written and “input” cells that are executable. Input cells are ex-
ecuted by typing (simultaneously)

SHIFT RETURN

which generates the label In[ ]:= with the output going to another input cell
with the label Out[ ]= (but it is still an input cell and can be executed).

Example A.1 Calculate 1+1.

A semicolon after a line of code means the code will still execute but the
output will be suppressed. This is a useful feature for debugging code.

Example A.2 Set x = 7 and y = 2 and output x + y.

When a variable is set, its value may be used in other cells until cleared.

DOI: 10.1201/9781003395515-A 188


Mathematica Starter  189

A.2 PALETTES
The menu has several extensive palettes that are useful in formatting the in-
put. For example, there is a Writing Assistant that manages cells and fonts.
This is useful for quick access to Greek letters. There is a Math Assistant
that has templates for operations like division, raising to a power, summa-
tion, integration, etc. This makes it easy to enter something like a summation
of squares into an input cell in a very clean format.

Example A.3 Sum the squares of integers from 0 to 10.

A.3 FUNCTIONS
Mathematica functions always begin with a capital letter. When typing in
a cell, Mathematica will give autocomplete options for existing functions.
Mousing over a function gives extensive documentation for the function’s use
with examples. The function Clear [ ] clears a variable. It produces no output.

Example A.4 Set x = 1 and clear x.

The function ClearAll[ ] is extremely useful to perform a global clear of


everything.

Example A.5 Clear all variables.

The function Simplify[ ] reduces the result algebraically.


190  Guide to Modern Physics

Example A.6 Calculate sin2 x + cos2 x.

It can also be written as //Simplify, placed after the code.

x2 −4x+4
Example A.7 Simplify x−2 .

The function D[ ] gives the dervative.


d cos x
Example A.8 Calculate dx .

A user may define a function by placing an underscore after the argument,


f [x ]. This allows the function to be evaluated for any value of the argument.

Example A.9 Define the function f (x) = x2 and evaluate it for x = 2.5.

A.4 RESERVED NAMES


There are a few names that are reserved and may not be user defined. One of
them is D which is reserved for differentiation. The reserved names always
begin with a capital letters. Others include, E , I, and Pi, which stand for the
exponential e , imaginary i, and π.

Example A.10 Calculate eiπ + 1.

A double equal sign makes a logical comparison.


Mathematica Starter  191

Example A.11 Compare Pi to π .

A.5 PHYSICAL CONSTANTS AND THEIR UNITS


You can get physical constants in Mathematica by typing simultaneously (in
input cell)

CTRL +

and then typing into the natural language box that appears, for example,
“speed of light”:

Figure A.1 Typing “speed of light” into the natural language box.

Clicking outside the box gives you (hopefully) what you were looking for,
displayed in standard physics notation.

Figure A.2 Successful procurement of the speed of light.

The physical constant c is stored as a “unit” and it appears in italics with a


different shading so you can recognize the difference between a unit and a
user-defined variable with the same name. The numerical value is displayed
together with units using the function UnitConvert[ ]. The default units will
be SI.
192  Guide to Modern Physics

Example A.12 Get the numerical value of the speed of light.

The units to be displayed may be specified. There are 2 ways to get a unit 1)
typing into the natural language box, and 2) using the function Quantity[ ].

Example A.13 Get the unit miles per second.

Example A.14 Get the numerical value of the speed of light in miles per
second.

The function N[ ] will calculate the numerical value to the specified num-
ber of significant figures. The value has been stored in the variable x and it
remains so until cleared.

Example A.15 Get the numerical value of the previous calculation of the
speed of light to 3 significant figures.

Example A.16 Get π to 50 figures.

The functions NumberForm [ ], and ScientificForm[ ] can also be used to


display a decimal answer with specified number of digits.
Mathematica Starter  193

Other useful physical constants are similarly obtained by typing the fol-
lowing into the natural language box: elementary charge, epsilon 0, planck’s
constant, hbar, electron mass, proton mass, boltzmann constant, etc. Physical
constants and their names are given in App. B.
Mathematica is extremely useful as a calculator because it will automat-
ically check the units of a calculation and report errors.
Example A.17 Try to get the speed of light in kg.

A.6 INTEGRATION
Integration is performed with the function Integrate[ ].

A.6.1 Indefinite Integrals



Example A.18 Calculate x3 e−x dx.

Mathematica can convert your input cell into a “standard form” which
looks very much like you would see it typed in a book. This is equivalent code
that executes identically. The computer code has become human readable!

Example A.19 Calculate x3 e−x dx with the input in standard form.

A.6.2 Definite Integrals


∞
Example A.20 Calculate x3 e−x dx.
0
194  Guide to Modern Physics

∞
Example A.21 Calculate x3 e−x dx with the input in standard form.
0

A.6.3 Numerical Integration


Numerical integration is performed with the function NIntegrate[ ].

∞ x2
Example A.22 Numerically integrate e x −1 dx.
0

A.6.4 Assumptions
Mathematica will not assume that a variable is real.

Example A.23 Calculate the average value of e−λx for 0 < x < ∞.

Assumptions about variables can be made with the global command $As-
sumption. The can also make the code run faster for involved calculations.

Example A.24 Calculate the average value of e−λx for 0 < x < ∞ with the
assumption that λ > 0.
Mathematica Starter  195

A.7 RESOURCE FUNCTIONS


Mathematica has a large library of resource functions. These are called with
the function ResourceFunction[ ] with the resource name in quotes, for ex-
ample ResourceFunction[“HydrogenWavefunction”].

Example A.25 Get the 2p, m = 0 wave function of hydrogen with Bohr ra-
dius a.

A.8 SERIES EXPANSION


Example A.26 Get the first 4 terms (order x3 ) of sin x expanded about x = 0.

Example A.27 Get the first 3 terms of the binomial expansion.

A.9 SOLVING AN EQUATION


Example A.28 Solve x2 − 4x + 2 = 0.
196  Guide to Modern Physics

Solve produces a list. It is useful sometimes to put the results from Solve
into a variable.

Example A.29 Extract the results from Solve above for the solution to x
stored in the variable y.

A.10 PLOTTING A FUNCTION

Plot[Sin[x], {x, 0, 6π}, PlotStyle→GrayLevel[.5],

AxesLabel → {"x", "sin(x)"}]








  





Figure A.3 Plot of sin x vs. x.


APPENDIX B

Physical Constants

Physical constants may be called in two ways. One way is to use the function
Quantity[ ] with the argument equal to the name of the constant. A second
way which makes a much cleaner look to the code is to use the defined sym-
bol obtained from the natural language box as described in A.5. Executing
Quantity[ ] produces the an output identical to that of the natural language
box.

Example B.1 Compare the elementary charge as obtained from the Quan-
tity[ ] function and the natural language box.

The names of the fundamental constants used in this book with their sym-
bols and values are shown in B.2 and derived combinations are shown in
B.1.

Particle masses may be acquired with either the function Quantity[ ] or


the natural language box as described in A.5.

Example B.2 Get the numerical value of the electron mass using the natural
language box.

Particle masses are displayed in B.3.

DOI: 10.1201/9781003395515-B 197


198  Guide to Modern Physics

Table B.1 Mathematica names (symbol) and numerical values for physical
constants.

Table B.2 Mathematica names (symbol) and numerical values for derived
constants.
Physical Constants  199

Table B.3 Mathematica names (symbol) and numerical values for particle
masses.

Table B.4 Common names (symbol) and numerical values for sky objects.
Index

alpha, 6 Boltzman constant, 25


alpha decay, 42 Bose-Einstein distribution, 165
alpha particles, 59
angular momentum addition, 141 carbon dating, 44
astronomical unit, 174 cells, Mathematica, 188
atomic clock, 72 center of mass, 75
atomic collapse, 88 Chandrasekhar limit, 177
atomic mass, 14 classical electron radius, 59
average distance, 131 classical probability, 100, 124
average energy, 31, 164 Compton scattering, 54
average kinetic energy, 132, 159 correspondence principle, 89, 99,
average oscillator energy, 25 123, 138
average speed, 162 cosmic background radiation, 183
Cosmic neutrino background, 184
Balmer series, 86 cosmic ray muon, 71
beta decay, 23, 42 Coulomb unit, 1
binding energy, alpha particle, critical mass density, 185
13 cross section, 62
binding energy, neutron star, 179 cross section, neutrino, 66
binding energy, nuclear, 13
binomial distribution, 152 dark night sky, 181
black hole, 180 de Broglie wavelength, 18, 82, 96
black hole, lifetime, 181 decay probability, 43
black hole, temperature, 180 density of states, 167
blackbody radiation, 31 deuteron, 88
blackbody radiation peak, 32 Doppler effect, 79
Bohr atom, ground state, 83 duality, particle wave, 18
Bohr electron speed, 84 earth’s orbit, 90
Bohr excited states, 84 electric charge, 1
Bohr magneton, 142 electric dipole transition, 140
Bohr model, 81, 82 electric force constant, 3
Bohr orbit radii, 84 electron g-factor, 150
Bohr radius, 83, 129 electron charge, 48
Bohr total energy, 84 electron charge to mass ratio, 46

201
202  Index

electron diffraction, 52 hydrogen atom, 127


electron gas, 169 hyperfine splitting, 144
electron in a box, 94
electron speed, 7 ideal gas, 157
electron spin, 141, 144 infinite square well, 92
electronvolt, 1 intrinsic angular momentum,
Elementary charge, 1 141
energy, 8 invariant mass, 75
energy and momentum, 73 ionization energy, 12
energy density, cosmic neutrinos,
kinetic energy, 2, 9, 20, 49, 83
184
energy distribution, 164 Lamb shift, 148
energy scale, 2 Large Hadron Collider, 74, 91
energy-momentum transformation, lifetime, 140
77 lifetime, average, 43
even states, 106 Lorentz γ factor, 69
excited states, hydrogen, 136 Lorentz contraction, 73
expectation values, 97 Lorentz transformation, 76
exponential distribution, 43 Lyman alpha line, 79
Lyman series, 85
fermi constant, 66
Fermi energy, 169, 178 magnetic constant, 143
Fermi-Dirac distribution, 166 magnetic dipole field, 146
fine structure constant, 6 magnetic moment, 141
finite square well, 103 magnitude, absolute, 176
four vectors, 74 magnitude, apparent, 175
frequency, 17, 34 mass energy, 9
mass, electron, 8
galaxy, 176
Maxwell-Boltzmann
gamma decay, 42
distribution, 157
Gaussian distribution, 154, 156
mean free path, 158, 181
gravitational constant, 7
mean free path, neutrino,
ground state, 112
67
ground state energy, 129
Milky Way, 147, 176
Hawking radiation, 180 Milky Way, mass, 177
Hermite polynomial, 119 minimum kinetic energy, 22
Hubble constant, 182 momentum, 10
Hubble lifetime, 182 momentum unit, 9
Hubble’s law, 182 most probable distance, 131
hydrogen 21 cm line, 147 most probable speed, 162
Index  203

neutron star, 178 Quantity, Mathematica, 2


Newton’s 2nd law, 45, 82 quantization of angular momentum,
normalization, 94, 114, 130 82
nuclear decay, Q value, 14 quantization of energy, 29
number density, cosmic neutrinos, quantum distributions, 165
184 quantum harmonic
number of modes, 27 oscillator, 111, 112
number of photons, 34 quantum tunneling, 114
NumberForm, Mathematica, 4 quarks, 54

odd states, 108 radial probability, 130, 134


orbit frequency, 89 radioactive decay, 42
Rayleigh-Jeans formula, 25
palettes. Mathematica, 189 redshift, 80
parsec, 174 redshift, cosmic, 182
particle in a box, 92 reduced mass, 87
particle in a box, 3 dimensions, root mean square, 161
102 root mean square speed, 163
Paschen series, 86 root-mean-square distance, 131
photoelectric effect, 49 Rutherford Scattering, 59
photon, 16 Rydberg constant, 87
photon energy density, 183
photon flux, 36 scattering rate, neutrino, 67
photon gas, 168 Schrödinger equation, 93, 102,
photon number density, 184 112, 128
physical constants, 197 Schwarzschild radius, 180, 187
Planck length, 186 solar constant, 174
Planck mass, 185 solar distance, 174
Planck radiation formula, 29 solar luminosity, 174
Planck time, 186 solar neutrino flux, 175
Planck’s constant, 4 solar power, 40
Planck’s constant, reduced, 5 solar temperature, 175
Poisson distribution, 155 space and time, 70
potential energy, 12, 83, 92 special relativity, 69
power per area, 27, 37 speed distribution, 161
pressure, 158 speed of light, 4, 5, 143
probability distributions, 152 spin-orbit interaction, 142
proton cycle, 173 Stefan-Boltzmann constant, 38
proton fusion, 15 Stefan-Boltzmann law, 36, 180
proton in a box, 96 strength, electric force, 3
204  Index

superfluid helium, 170 visible photon, 16

temperature, 32 W boson, 23
thermal neutrons, 53 wave function, 94, 109, 116, 122
thermal radiation, 25 wave function, hydrogen atom, 133
threshold frequency, 51 wave speed, 17
time dilation, 72 wavelength, 17, 20
transitions, 85 weak force, range, 24
transitions in hydrogen, 138 weak interaction, 66
Wein’s law, 40
uncertainty principle, 22, 23, white dwarf, 177
96, 98, 115, 132 work function, 50
UnitConvert, Mathematica, 2
Zeeman effect, 142
velocity distribution, 160

You might also like