100% found this document useful (1 vote)
179 views242 pages

Quantum Mechanics Problems and Solutions

Quantum mechanics
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
100% found this document useful (1 vote)
179 views242 pages

Quantum Mechanics Problems and Solutions

Quantum mechanics
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/ 242

Quantum Mechanics

Quantum Mechanics
Problems and Solutions

K. Kong Wan
June 23, 2020 10:49 JSP Book - 9in x 6in 00-Solution˙Manual-Prelims

Published by
Jenny Stanford Publishing Pte. Ltd.
Level 34, Centennial Tower
3 Temasek Avenue
Singapore 039190

Email: [email protected]
Web: www.jennystanford.com

British Library Cataloguing-in-Publication Data


A catalogue record for this book is available from the British Library.

Quantum Mechanics: Problems and Solutions


c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright �

All rights reserved. This book, or parts thereof, may not be reproduced in any
form or by any means, electronic or mechanical, including photocopying,
recording or any information storage and retrieval system now known or to
be invented, without written permission from the publisher.

For photocopying of material in this volume, please pay a copying


fee through the Copyright Clearance Center, Inc., 222 Rosewood Drive,
Danvers, MA 01923, USA. In this case permission to photocopy is not
required from the publisher.

ISBN 978-981-4800-72-3 (Paperback)


ISBN 978-0-429-29647-5 (eBook)
June 23, 2020 10:49 JSP Book - 9in x 6in 00-Solution˙Manual-Prelims

To my beautiful granddaughter Orly Rose,


whose new arrival brings
infinite joy and jubilation.
June 23, 2020 10:49 JSP Book - 9in x 6in 00-Solution˙Manual-Prelims

Contents

Preface xi

1 Structure of Physical Theories 1

2 Classical Systems 3

3 Probability Theory for Discrete Variables 5

4 Probability Theory for Continuous Variables 9

5 Quantum Mechanical Systems 17

6 Three-Dimensional Real Vectors 21

7 Matrices and Their Relations with Vectors 27


3

8 Operations on Vectors in IE 35

9 Special Operators on IE 3 41

10 Probability, Selfadjoint Operators, Unit Vectors and the


Need for Complexness 51

11 Complex Vectors 55

12 N-Dimensional Complex Vectors 59

13 Operators on N-Dimensional Complex Vectors 65

14 Model Theories Based on Complex Vector Spaces 81


June 23, 2020 10:49 JSP Book - 9in x 6in 00-Solution˙Manual-Prelims

viii Contents

15 Spectral Theory in Terms of Stieltjes Integrals 89

16 Infinite-Dimensional Complex Vectors and Hilbert Spaces 93


17 Operators in a Hilbert Space H 99


18 Bounded Operators on H 107


19 Symmetric and Selfadjoint Operators in H 115


20 Spectral Theory of Selfadjoint Operators in H 121


21 Spectral Theory of Unitary Operators on H 127

22 Selfadjoint Operators, Unit Vectors and Probability


Distributions 129

23 Physics of Unitary Transformations 133

24 Direct Sums and Tensor Products of Hilbert Spaces and


Operators 135

25 Pure States 143

26 Observables and Their Values 145

27 Canonical Quantisation 149

28 States, Observables and Probability Distributions 161

29 Time Evolution 167

30 On States after Measurement 175

31 Pure and Mixed States 177

32 Superselection Rules 181

33 Many-Particle Systems 185


June 23, 2020 10:49 JSP Book - 9in x 6in 00-Solution˙Manual-Prelims

Contents ix

34 Conceptual Issues 187

35 Harmonic and Isotropic Oscillators 189

36 Angular Momenta 201

37 Particles in Static Magnetic Fields 225

Bibliography 229
June 23, 2020 10:49 JSP Book - 9in x 6in 00-Solution˙Manual-Prelims

Preface

This is a solutions manual to accompany the book Quantum


Mechanics: A Fundamental Approach by the author published in
2019 by Jenny Stanford Publishing, Singapore. It provides detailed
solutions to all the questions listed at the end of each chapter of the
book, except for the introductory Chapters 1 and 2. These questions
are reproduced here chapter by chapter, followed by their solutions,
which are labelled to correspond to the questions. For example,
SQ3(1) is the solution to question Q3(1), which is the first question
listed in Exercises and Problems at the end of Chapter 3 of the book.
The solutions presented make full use of the materials in the
book. All the theorems, definitions, examples, comments, properties,
postulates and equations in the book are referred to by their chapter
or section numbers. For instance, Theorem 13.3.2(2) refers to the
second theorem in section 13.3.2, Eq. (4.18) refers to equation
(4.18) in Chapter 4, P15.1(5) refers to property (5) in section
15.1, and C28.2(3) refers to comment (3) in section 28.2. Equation
labelling in terms of (∗), (∗∗), (∗∗∗) and (∗∗∗∗) is introduced here
in some questions when they are needed for reference later in their
solutions.

K. Kong Wan
St Andrews
Scotland
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 1

Structure of Physical Theories

This introductory chapter sets out a general structure of physical


theories which is applicable to both classical and quantum me­
chanics. We start with measurable properties of a given physical
system, be it classical or quantum. These properties are called
observables. We then introduce a definition of the state of a physical
system in terms of measured values of a sufficiently large set of
observables. A theory to describe the system should consists of four
basic components:
1. Basic mathematical framework This comprises a set of elements
endowed with some specific mathematical structure and properties.
In mathematics such a set is generally known as a space.
2. Description of states States are described by elements of the
space in the chosen mathematical framework. For this reason the
space is called the state space of the system.
3. Description of observables Observables are to be described
by quantities defined on the state space. The description should
yield all possible values of observables. The relationship between
observables and states should be explicitly stated. The following two
cases are of particular interest:

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

2 Structure of Physical Theories

(1) For a deterministic theory like classical mechanics a state


should determine the values of all observables.
(2) For a probabilistic theory like quantum mechanics a state
should determine the probability distribution of the values of
all observables.
4. Description of time evolution (dynamics).
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 2

Classical Systems

This chapter sets out some general physical properties of classical


systems which are divided into discrete and continuous:
1. Discrete systems These are systems of discrete point particles.
The specific structure of classical mechanics is presented with
position, linear and angular momenta serving as basic observables.
2. Continuous systems These systems are illustrated by a vibrat­
ing string. Continuous systems have different kinds of properties
and observables, e.g., wave properties. In particular we have
discussed:
(1) Description of states by solutions of the classical wave equation.
(2) The concept of eigenfunctions with orthonormality property
and their superposition and interference.
(3) The concept of a complete set of states.
These discussions are given specifically to provide an intuition
to help a better understanding of similar properties of quantum
systems.

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 3

Probability Theory for Discrete Variables

Q3(1) Prove Theorem 3.4(1).


SQ3(1) Theorem 3.4(1) can be proved using properties PM3.4(1),
PM3.4(2) and PM3.4(3) in Definition 3.4(2).

(1) To prove Eq. (3.25) let E be any event. Then E ∅ ∈ =


E and E ⊥ ∈ = ∈. By PM3.4(3) we have,
    
M p  E ∅ ∈ = M p  E   
  � M p ∈ = 0.
Mp E ∅ ∈ = Mp E + Mp ∈
 
(2) To prove Eq. (3.26) we start with M p Sam = 1 by PM3.4(2).
Since E ⊥ E c = ∈ and E ∅ E c = Sam we have
       
M p Sam = M p E ∅ E c = M p E + M p E c = 1
   
� Mp E c = 1 − Mp E .
(3) To
 prove Eq. (3.27) we first observe  that E 1 ≥ E 2 � E 2 =
E 2 − E 1 ∅ E 1 . Since (E 2 − E 1 and E 1 are disjoint, i.e., (E 2 −
E 1 ⊥ E 1 = ∈, we have
    
Mp E2 = Mp E2 − E1 ∅ E1
     
= Mp E2 − E1 + Mp E1 ⊕ Mp E1 .

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

6 Probability Theory for Discrete Variables

(4) The proof of Eq. (3.28) is based on the decomposition, obvious


from the Venn diagram,
   
E1 = E1 − E2 ∅ E1 ⊥ E2 ,
   
where E 1 − E 2 and E 1 ⊥ E 2 are disjoint. Then PM3.4(3)
implies
     
Mp E1 = Mp E1 − E2 + Mp E1 ⊥ E2 ,

which is Eq. (3.28).


Two cases are noteworthy:
(a) E 1 and E 2 are disjoint, i.e., E 1 ⊥ E 2 = ∈. Then E 1 − E 2 = E 1 ,
we have
M p (E 1 ⊥ E 2 ) = 0 and M p (E 1 − E 2 ) = M p (E 1 ).
Equation (3.28) is again satisfied.
(b) E 1 is a subset of E 2 . Then

E1 ⊥ E2 = E1 and E 1 − E 2 = ∈.

Equation (3.28) is again satisfied in these two cases.


(5) To prove Eq. (3.29) we first note that if E 1 and E 2 are disjoint
the equation is obviously true, i.e.,
     
Mp E1 ∅ E2 = Mp E1 + Mp E2

by PM3.4(3). If E 1 and E 2 are not disjoint, then the Venn


diagram tells us that
 
E 1 ∅ E 2 = E 1 − E 2 ∅ E 2,
 
where E 1 − E 2 and E 2 are disjoint. We have, by PM3.4(3),
     
Mp E1 ∅ E2 = Mp E1 − E2 + Mp E2 .

Using Eq. (3.28) we immediately get


       
Mp E1 ∅ E2 = Mp E1 + Mp E2 − Mp E1 ⊥ E2 ,

which is Eq. (3.29).


We have assumed E 2 is a subset of E 1 in the above proof. When
E 1 is a subset of E 2 we can similarly establish the result.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Probability Theory for Discrete Variables 7

Q3(2) Prove Theorem 3.5(1).


SQ3(2) Equation (3.36) is proved as follows:
n 
 2
Var (℘ A ) = aℓ − E(℘ A ) ℘ A (aℓ )
ℓ=1
n 
 
= aℓ2 − 2aℓ E(℘ A ) + E(℘ A )2 ℘ A (aℓ )
ℓ=1
 n

= aℓ2 ℘ A (aℓ ) − E(℘ A )2 ,
ℓ=1
using the fact that
n   n

− 2aℓ E(℘ A ) ℘ A (aℓ ) = −2E(℘ A )2 and ℘ A (aℓ ) = 1.
ℓ=1 ℓ=1

Q3(3) What is the value F(a4 ) − F(a3 ) of the probability


distribution function in Eq. (3.38)?
SQ3(3) The value F(a4 ) − F(a3 ) is equal to ℘(a4 ).

Q3(4) In an experiment of tossing a fair die a number from 1 to 6


will be obtained with equal probabilities.

(a) Write down the probability mass function ℘ and evaluate the
expectation value and the uncertainty.
(b) Write down the corresponding probability distribution function
F(τ ) and sketch a plot of F(τ ) versus τ . What are the values
F(τ ) at τ = 0.9, 1, 2.5, 6 and 6.1?

SQ3(4)(a) For a fair die every number is equally likely to appear in


a toss. The probability mass function is a function ℘ on the sample
space Sam := {a1 = 1, a2 = 2, a3 = 3, a4 = 4, a5 = 5, a6 = 6}
defined by ℘(aℓ ) = 1/6 for all aℓ ⇒ Sam . The expectation value is
1 1 1 1 1 1
1 × + 2 × + 3 × + 4 × + 5 × + 6 × = 3.5.
6 6 6 6 6 6
The variance is given by Theorem 3.5(1) to be
1 1 1 1 1 1
1× + 4 × + 9 × + 16 × + 25 × + 36 × − 3.52
6 6 6 6 6 6
≤ 2.9.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

8 Probability Theory for Discrete Variables

� �
The uncertainty = variance with an approximate value 2.9.
SQ3(4)(b) The probability distribution function F(τ ) is
piecewise-constant with discontinuous steps occurring at τ =
1, 2, 3, 4, 5, 6. Explicitly F(τ ) is related to ℘(aℓ ) by


⎪ 0, τ <1



⎪ 1/6, 1 ≤ τ <2



⎨ 2/6, 2 ≤ τ <3
F(τ ) = 3/6, 3 ≤ τ < 4 .


⎪ 4/6, 4 ≤ τ < 5




⎪ 5/6, 5 ≤ τ < 6


1, 6≤τ

The function F(τ ) is continuous from the right for all τ . A plot of
F(τ ) against τ is given below:

F(τ )

0 1 2 3 4 5 6 τ
We have F(0.9) = 0, F (1) = 1/6, F(2.5) = 2/6, F(6) = 1,
F(6.1) = 1.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 4

Probability Theory for Continuous


Variables

Q4(1) For the characteristic function χ  show that the inverse


image of every Borel set is a Borel set.
SQ4(1) Let � be a Borel set in the codomain of the characteristic
function χ  (τ ) which is be taken as IR . Then:

(1) If � does not contain the value 0 or 1 the inverse image


χ −1 �
 ( ) = ∈.
(2) If � contain the value 0 but not 1 the inverse image χ −1 �
 ( ) =
c
 , the complement of .
(3) If � contain the value 1 but not 0 the inverse image χ −1 �
 ( ) =
.
(4) If � contain both the value 0 and 1 the inverse image χ −1 �
 ( ) =
IR .
All the inverse images are Borel sets � χ  (τ ) is a Borel function.
Q4(2) Prove Eq. (4.12).
SQ4(2) Use the additive and the defining properties of Lebesgue
measure, i.e., the Lebesgue measure of a singleton set is zero and
that the measure of a half-open interval (τ1 , τ2 ] is equal to τ2 − τ1 .

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

10 Probability Theory for Continuous Variables

Since [τ1 , τ2 ] = {τ1 } ∅ (τ1 , τ2 ] and {τ1 } ⊥ (τ1 , τ2 ] = ∈ we get


   
Ml [τ1 , τ2 ] = Ml {τ1 } ∅ (τ1 , τ2 ]
     
= Ml {τ1 } + Ml (τ1 , τ2 ] = Ml (τ1 , τ2 ]
 
� Ml [τ1 , τ2 ] = τ2 − τ1 . (�)
We can prove two similar cases:
 
(1) To prove Ml [τ1 , τ2 ) = τ2 − τ1 we have
   
Ml [τ1 , τ2 ] = Ml [τ1 , τ2 ) ∅ {τ2 }
     
= Ml [τ1 , τ2 ) + Ml {τ2 } = Ml [τ1 , τ2 )
 
� Ml [τ1 , τ2 ) = τ2 − τ1 by Eq. (�) above.
 
(2) To prove Ml (τ1 , τ2 ) = τ2 − τ1 we have
   
Ml (τ1 , τ2 ] = Ml (τ1 , τ2 ) ∅ {τ2 }
     
= Ml (τ1 , τ2 ) + Ml {τ2 } = Ml (τ1 , τ2 )
 
� Ml (τ1 , τ2 ) = τ2 − τ1 by Eq. (�) above.

Q4(3) Show that the Lebesgue measure of the set of irrational


numbers in [a, b ] is equal to b − a.
SQ4(3) Let the set of rational and irrational numbers in [τ1 , τ2 ]
be denoted by [τ1 , τ2 ]ra and [τ1 , τ2 ]ir respectively. Then we have
[τ1 , τ2 ] = [τ1 , τ2 ]ra ∅[τ1 , τ2 ]ir and [τ1 , τ2 ]ra ⊥[τ1 , τ2 ]ir = ∈ . It follows
that
   
Ml [τ1 , τ2 ] = Ml [τ1 , τ2 ]ra ∅ [τ1 , τ2 ]ir
     
= Ml [τ1 , τ2 ]ra + Ml [τ1 , τ2 ]ir = Ml [τ1 , τ2 ]ir ,
since [τ1 , τ2 ]ra is a countable set which
 has a measure zero. The
desired result follows from Ml [τ1 , τ2 ] = τ1 − τ1 .
Q4(4) Prove Eqs. (4.18) to (4.23).
SQ4(4) Use the additive property of measures and the defining
properties of a Lebesgue-Stieltjes measure Mls, g in Eqs. (4.16) and
(4.17), i.e., the Lebesgue-Stieltjes measure of a half-open interval
(τ1 , τ2 ] is equal to g(τ2 ) − g(τ1 ) and of a singleton set {τ } is equal
to is g(τ + 0) − g(τ − 0).
(1) The proof of Eq. (4.18) is the same as the proof of Eq. (3.25) in
SQ3(1).
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Probability Theory for Continuous Variables 11

(2) Eq. (4.19) is part of the defining properties of the measure given
in Eq. (4.17), bearing in mind that g(τ + 0) = g(τ ).
(3) Eq. (4.20) is part of the defining properties of the measure given
in Eq. (4.16).
(4) To prove of Eq. (4.21) we first observe that
   
Mls, g (τ1 , τ2 ] = Mls, g (τ1 , τ2 ) ∅ {τ2 }
   
= Mls, g (τ1 , τ2 ) + Mls, g {τ2 }
 
= Mls, g (τ1 , τ2 ) + g(τ2 ) − g(τ2 − 0).
It follows that
     
Mls, g (τ1 , τ2 ) = Mls, g (τ1 , τ2 ] − g(τ2 ) − g(τ2 − 0)
= g(τ2 − 0) − g(τ1 ).

(5) We can prove of Eq. (4.22) as follows:


   
Mls, g [τ1 , τ2 ] = Mls, g {τ1 } ∅ (τ1 , τ2 ]
   
= Mls, g {τ1 } + Mls, g (τ1 , τ2 ]
= g(τ1 ) − g(τ1 − 0) + g(τ2 ) − g(τ1 )
= g(τ2 ) − g(τ1 − 0).

(6) We can prove of Eq. (4.23) we first observe that:


   
Mls, g [τ1 , τ2 ] = Mls, g [τ1 , τ2 ) ∅ {τ2 }
   
= Mls, g [τ1 , τ2 ) + Mls, g {τ2 }
   
= Mls, g [τ1 , τ2 ) + g(τ2 ) − g(τ2 − 0) .
It follows that
     
Mls, g [τ1 , τ2 ) = Mls, g [τ1 , τ2 ] − g(τ2 ) − g(τ2 − 0)
= g(τ2 − 0) − g(τ1 − 0).
Q4(5) Explain the main differences between Riemann and
Lebesgue integrals.
SQ4(5) The main difference lies in the way the area under the
curve is partitioned into a sum of rectangular areas. For Riemann
integrals we first partition the domain of the function (integrand)
into subintervals ℓ and then partition the area under the curve
into a sum of rectangles with ℓ as the base. For Lebesgue integrals
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

12 Probability Theory for Continuous Variables

we partition the range of the function (integrand) into subintervals


�ℓ and take the inverse images ℓ = f −1 (�ℓ ) under the function.
For Borel functions these inverse images are necessarily Borel sets
which may not be intervals. We then take the Lebesgue measure of
each of these inverse images Mℓ (�ℓ ) as the base to form an area
which is not necessarily of the form of a standard rectangle. The area
under the curve is approximated by the sum of these areas. The two
integrals agree for some functions but not all functions. Lebesgue
integrals are applicable to a wider class of functions.
Q4(6) Explain why Lebesgue integrals are only defined for Borel
functions.
SQ4(6) By dividing the range of the function (integrand) into
subintervals we need to take inverse images of the subintervals
under the function (integrand) and we have to give each of these
inverse images a length. If these inverses are Borel sets then we can
take the Lebesgue measure of these inverse images to give each of
these inverse images a length. The integrand being a Borel function
ensures that all inverses are Borel sets. Otherwise we will be unable
to give each of these inverse images a length, i.e., we are unable to
obtain a area associated with each of these inverse images and we
are unable to obtain an approximation to the area under the curve.
Q4(7) Explain the main differences between Riemann–Stieltjes
and Lebesgue–Stieltjes integrals.
SQ4(7) The Riemann-Stieltjes integrals are extensions of Riemann
integrals by integrating with respect to a function rather than an
independent variable. The Lebesgue-Stieltjes integrals are exten­
sions of Lebesgue integrals by using a Lebesgue-Stieltjes measure
to construct the integrals.
Q4(8) Explain the meaning of Eq. (4.68) which expresses the Dirac
delta function δ(τ ) formally as the derivative of the unit step function
gus (τ ).
SQ4(8) The meaning of Eq. (4.68) which expresses the Dirac delta
function δ(τ ) formally as the derivative of the unit step function
gus (τ ), i.e., δ(τ ) = dgus (τ )/dτ , lies in Eq. (4.69) on the Riemann­
Stieltjes integral with respect to the unit step function. Indeed the
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Probability Theory for Continuous Variables 13

Dirac delta function is meaningful only within the context of such an


integral.
Q4(9) Show that the coefficients c1 , c2 , c3 in Eq. (4.79) must
satisfy the condition c1 + c2 + c3 = 1.
SQ4(9) Using the limiting value of distribution functions F(τ ) at
τ = → in Definition 3.6(2), i.e., property MP3.6(2), we get

F (→) = Fd (→) = Fac (→) = Fsc (→) = 1.

For a general probability distribution function F(τ ) in Eq. (4.79) we


have

F(→) = c1 Fd (→) + c2 Fac (→) + c3 Fsc (→) = c1 + c2 + c3 .

The fact that we must also have F(→) = 1 implies c1 + c2 + c3 = 1.


Q4(10) Show that expectation value is additive in the sense that
for the distribution function in Eq. (4.79) we have

E(c1 Fd + c2 Fac + c3 Fsc )


= c1 E(Fd ) + c2 E(Fac ) + c3 E(Fsc ).
 
SQ4(10)1 The expectation value E F for the distribution func-
tion F in Eq. (4.79) is given by
 →
 
E F = τ dF(τ )
−→
 →  
= τ d c1 Fd (τ ) + c2 Fac (τ ) + c3 Fsc (τ )
−→
 →  →  →
= c1 τ dFd (τ ) + c2 τ dFac (τ ) + c3 τ dFsc (τ )
−→    −→   −→
= c1 E Fd + c2 E Fac + c3 E Fsc .

Q4(11) Show that the Gaussian probability distribution function


given by the density function in Eq. (4.80) by
 τ
FG (τ ) = wG (τ ) dτ
−→

1 Papoulis p. 143.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

14 Probability Theory for Continuous Variables

satisfies the defining properties of probability distribution func­


tions. Verify that the expectation value and the uncertainty are given
respectively by
E(FG ) = b and (FG ) = a.

SQ4(11) The defining properties of probability distribution func­


tions are given by Definition 3.6(2). The function FG (τ ), being
an integral of the Gaussian probability density function, is clearly
nondecreasing, continuous from the right and having the limiting
value FG (−→) = 0. For the limiting value at → we have
 →
1 2 2
� e−(τ −b) /2a dτ = 1,
2
2πa −→
due to the well-known integral
 → 
2 π
e−λ(x−c) dx = .
−→ λ

The expectation value is calculated by replacing τ by x = τ − b,


i.e.,
 →
1 2
/2a2
E(FG ) = � τ e−(τ −b) dτ
2πa2 −→
 →
1 2
/2a2
= � (x + b) e−x dx
2πa2 −→
 →  →
1 −x 2 /2a2 1 2
/2a2
= � xe dx + � b e−x dx
2πa2 −→ 2πa2 −→

= b.
The first integral vanishes since the integrand is an odd function of x.
The variance is calculated using Eq. (4.92), i.e.,
 →
1 2 2
Var(FG ) = � τ 2 e(τ −b) /2a dτ − E(FG )2
2πa2 −→
 →
1 2 2
= � τ 2 e−(τ −b) /2a dτ − b2
2
2πa −→
 →
1  2 2 2
= � x + b e−x /2a dx − b2
2
2πa −→
 →
1  2  2 2
= � x + 2xb + b2 e−x /2a dx − b2 (�)
2
2πa −→
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Probability Theory for Continuous Variables 15

 →
1 2
/2a2
= � x 2 e−x dx
2πa2 −→
= a2 ,
due to the integral
 →

2 −λx 2 1 π
x e dx = .
−→ 2λ λ

The uncertainty is then (FG ) = Var(FG ) = a. Note that there
are three integrals in the expression (�) above. The middle integral
is zero on account of an odd integrand. The last integral is equal to
b2 since the integral of FG (τ ) is equal to 1.
Q4(12) Find the probability distribution function FU (τ ) given by
the density function wU (τ ) in Eq. (4.81).
SQ4(12) The uniform probability distribution function is given by

 τ ⎨0 τ ≤a
FU (τ ) = wU (τ � ) dτ � = (τ − a)/(b − a) x ⇒ (a, b) .
−→ ⎩
1 x⊕b

Q4(13) Discuss the mathematical and physical differences be­


tween discrete and continuous probability distributions.
SQ4(13) Mathematically, a discrete probability distribution func­
tion is piecewise-constant with many discontinuities while a con­
tinuous probability distribution function is absolutely continuous
(we ignore singularly continuous distributions). Physically, the
probability of obtaining a single precise value in a statistical
experiment with a continuous probability distribution is zero. Any
measurement result has to be specified in terms of a Borel set or
interval of non-zero measure. This is not the case for a discrete
distribution.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 5

Quantum Mechanical Systems

Q5(1) Explain why the values of a classical observable are deemed


to be objective while the values of a quantum observable are
generally regarded as non-objective.
SQ5(1) As stated in §2.1.3 and §2.1.4 a kinematic observable
of a classical mechanical system is defined to be a numerical
function of the state. It follows that a kinematic observable of a
classical mechanical system possesses a value in an arbitrary state
at any instance of time. If we wish to know what this value is
we can perform a measurement which will reveal the value of the
observable at that time. A measurement can be executed within a
short period of time without disturbing the system significantly so
that the value revealed by a measurement is the same as the value
the observable has before, during and after the measurement with
arbitrary accuracy. This is why we say that classical observables
have objective values independent of measurement. On account of
QMP5.3(3) quantum observables do not have the above properties.
Generally the values of a quantum observable are not objective
in the following sense. An observable of a quantum system do
not possesses a value in an arbitrary state. A measurement would
yield a value but this value cannot be assumed to be the value

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

18 Quantum Mechanical Systems

possessed by the system before the measurement. A repetition of


the experiment to measure the observable in the same state may
well yield a different value. This is why we say the values of quantum
observables are generally non-objective.
Q5(2) Give a brief account of the relationship between states and
possessed values of discrete observables of a quantum system.
SQ5(2) The relationship between states and possessed values of
observables is contained in QMP5.3(2) which says that there are
states in which a given discrete observable has a definite value which
can be revealed by measurement. Such values are by definition
the value possessed by the observable in those states. A quantum
observable does not possess a definite value in an arbitrary state.
Q5(3) Explain why quantum observables cannot be related to the
state in the same way kinematic observables of a classical system are
related to the state.
SQ5(3) As stated in §2.1.4 a kinematic observable of a classical
mechanical system is defined to be a numerical function of the state.
As a result a kinematic observable of a classical mechanical system
possesses a value in an arbitrary state at any instance of time. A
given state automatically determines simultaneously the values of
all these observables. QMP5.3(1) tells us that the values of quantum
observables are not simultaneously determinable and QMP5.3(3)
informs us that a state cannot determine the values of an arbitrary
observable. It follows that quantum observables cannot be related
to the state in the same way classical kinematic observables are
related to the state, i.e., quantum observables cannot be defined as
numerical functions of the state.
Q5(4) Discuss the effect of QMP5.3(1) on the specification of
states.
SQ5(4) QMP5.3(1) implies that we cannot determine the state of
a quantum system by the simultaneous values of all its observables
(obtained by simultaneous measurements of all the observables).
Instead we have to determine a state by the simultaneous values of
an appropriate set of compatible observables.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Quantum Mechanical Systems 19

Q5(5) Explain what is meant by the behaviour of quantum


systems being intrinsically probabilistic.
SQ5(5) For classical systems the probabilistic behaviour is due
to a lack of a maximum knowledge of the system. Quantum
systems behave probabilistically despite a maximum knowledge
of the system, as stated in QMP5.3(3). In particular a state
cannot determine the values of all observables. Instead a state
can only determine of the probability distributions of the values
of an arbitrary observable. We call such behaviour intrinsically
probabilistic.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 6

Three-Dimensional Real Vectors

Q6(1) Prove Eqs. (6.1) to (6.3).

SQ6(1)

(1) To prove Eq. (6.1) we start with the premise v∀1 + u∀ = v∀2 + u.

Adding u∀ −1 on both side this equation we get

v∀1 + u∀ + u∀ −1 = v∀2 + u∀ + u∀ −1
� v∀1 + 0∀ = v∀2 + 0∀ � v∀1 = v∀2 .

(2) To prove the first equation in Eq. (6.2) we start with

au∀ = (a + 0)u∀ = au∀ + 0u∀.


∀ = 0u∀.
Adding (au∀ )−1 on both side of the equation we get 0
To prove the second equation in Eq. (6.2) we start with
∀ + au∀ = a(0
a0 ∀ + u) ∀ + au∀.
∀ = au∀ = 0
∀ = 0.
It then follows from Eq. (6.1) that a0 ∀
(3) Eq. (6.3) is proved as follows:
 
∀ � (−1)u∀ = u∀ −1 .
u∀ + (−1)u∀ = 1 + (−1) u∀ = 0u∀ = 0

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

22 Three-Dimensional Real Vectors

Q6(2) Prove Theorem 6.2.1(1).


SQ6(2) Let cℓ� be a different set of coefficients such that
n

v∀ = cℓ� u∀ℓ .
ℓ=1

Subtract Eq. (6.9) from the above equation we get


n

∀=
0 (cℓ − cℓ� )u∀ℓ .
ℓ=1

On account of the linear independence of u∀ℓ the above equation


implies that cℓ − cℓ� = 0 ∞ℓ. In other words we have cℓ� = cℓ , i.e.,
there is just one set of coefficients cℓ for any given v∀.

Q6(3) Verify that the expression for �u∀ | v∀ ∪ in Eq. (6.13) satisfies
properties SP6.3.1(1), SP6.3.1(2) and SP6.3.1(3) of scalar product.

SQ6(3)

(1) To prove SP6.3.1(1) we note that changing the order of u∀ and v∀


would change the angle between the two vectors from θ to −θ.
But this does not change the value of the scalar product since
cos θ = cos(−θ), i.e., property SP6.3.1(1) is satisfied.
(2) The distributive property SP6.3.1(2) can be verified diagram­
matically. Draw a diagram on the x-y plane with the vector u∀
lying along the x-axis and then draw a1 v∀1 , a2 v∀2 and their sum
a1 v∀1 + a2 v∀2 . Then draw a diagram of projections of a1 v∀1 , a2 v∀2
and a1 v∀1 + a2 v∀2 , onto u∀, i.e., onto the x-axis. It will then become
obvious that the component of a1 v∀1 + a2 v∀2 onto u∀ is equal to the
sum of the components of a1 v∀1 and a2 v∀2 onto u∀. In other words
the distributive property SP6.3.1(2) is satisfied.
(3) Property SP6.3.1(3) is satisfied since �u∀ | u∀ ∪ = u2 ⊕ 0, and
�u∀ | u∀ ∪ = 0 � u = 0 � u∀ = 0. ∀ Formally this is because u∀ is
equal to its norm ||u∀ || times its unit directional vector u∀(u) , i.e.,
u∀ = ||u∀ || u∀(u) . Then ||u|| ∀
∀ = 0 implies u∀ = 0 u∀(u) = 0.

Q6(4) Show that two orthogonal vectors are linearly independent.


June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Three-Dimensional Real Vectors 23

SQ6(4) Let v∀1 and v∀2 be two (non-zero) orthogonal vectors and a1
∀ Then:
and a2 be two real numbers such that a1 v∀1 + a2 v∀2 = 0.
�∀
v 1 | a1 v∀1 + a2 v∀2 ∪ = 0
v 1 | a1 v∀1 + a2 v∀2 ∪ = a1 �∀
� �∀ v 1 | v∀1 ∪ + a2 �∀
v 1 | v∀2 ∪
= a1 �∀
v 1 | v∀1 ∪ = 0
� a1 = 0 � a2 = 0.
The two vectors are therefore linearly independent.
One can further check this by assuming linear dependence, i.e.,
v∀1 = a v∀2 . We will then have
v 1 | v∀1 ∪ = a �∀
�∀ v 1 | v∀2 ∪ = 0,
v 1 | v∀1 ∪ = || v∀1 ||2 �= 0.
which is a contradiction since �∀
Q6(5) Prove Eq. (6.19).
SQ6(5) Take the scalar product of e∀ℓ and v∀ we get
3
 3

�e∀ℓ | v∀ ∪ = �e∀ℓ | cℓ� e∀ℓ� ∪ = cℓ� �e∀ℓ | e∀ℓ� ∪
ℓ� =1 ℓ� =1
3

= cℓ� δℓℓ� = cℓ .
ℓ� =1

Note that we have rewritten Eq. (6.19) as 3ℓ� =1 cℓ� e∀ℓ� rather than
3
the original expression ℓ=1 cℓ e∀ℓ which will cause confusion when
taking the scalar product with e∀ℓ . If one wishes to use the original
expression one should take the scalar product with e∀ℓ� which will
show that cℓ� = �e∀ℓ� | v∀ ∪.
Q6(6) Prove the Pythagoras theorem in the forms of Eqs. (6.22)
and (6.27).
SQ6(6) For the Pythagoras theorem in Eq. (6.22) we have
v | v∀ ∪ = v x v x � i∀ | i∀ ∪ + v x v y � i∀ | ∀j ∪ + · · · .
v ||2 = �∀
||∀
The orthonormality of the basis vectors i∀, ∀j , k∀ immediately leads to
the desired result.
For the Pythagoras theorem in Eq. (6.27) we have
3
 3
 3

�u∀ | v∀ ∪ = � uℓ e∀ℓ | vℓ� e∀ℓ� ∪ = uℓ vℓ� �e∀ℓ | e∀ℓ� ∪.
ℓ=1 ℓ� =1 ℓ, ℓ� =1
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

24 Three-Dimensional Real Vectors

The orthonormality of the basis vectors, i.e., �e∀ℓ | e∀ℓ� ∪ = δℓℓ� ,


immediately leads to the desired result.
Q6(7) Prove Eqs. (6.30) and (6.31).
SQ6(7) For Eq. (6.30) we first express v∀ in terms of e∀ℓ as in Eq.

(6.19), i.e., v∀ = 3ℓ=1 vℓ e∀ℓ with vℓ = �e∀ℓ | v∀ ∪. Then:
vℓ = �e∀ℓ | v∀ ∪ = 0 ∞ℓ � v∀ = 0. ∀
For Eq. (6.31) we observe that
�e∀ℓ | u∀ ∪ = �e∀ℓ | v∀ ∪ ∞ ℓ � �e∀ℓ | u∀ − v∀ ∪ = 0 ∞ ℓ.
Then Eq. (6.30) proved above implies u∀ − v∀ = 0, ∀ i.e., u∀ = v.

Q6(8) Verify the Gram-Schmidt orthogonalisation procedure
given in §6.3.5.
SQ6(8) For the Gram-Schmidt orthogonalisation we have:
v 1 | v∀2 ∪
�∀
�u∀ �1 | u∀ �2 ∪ = �∀
v 1 | v∀2 ∪ − v 1 | v∀1 ∪ = 0,
�∀
�∀
v 1 | v∀1 ∪
v 1 | v∀3 ∪
�∀
�u∀ �1 | u∀ �3 ∪ = �∀
v 1 | v∀3 ∪ − �∀
v 1 | v∀1 ∪
�∀
v 1 | v∀1 ∪
�u∀ �2 | v∀3 ∪ �
− � �u∀ 1 | u∀ �2 ∪ = 0,
�u∀ 2 | u∀ �2 ∪
�u∀ �1 | v∀3 ∪ �
�u∀ �2 | u∀ �3 ∪ = �u∀ �2 | v∀3 ∪ − � �u∀ 2 | u∀ �1 ∪
�u∀ 1 | u∀ �1 ∪
�u∀ �2 | v∀3 ∪ �
− � �u∀ 2 | u∀ �2 ∪ = 0.
�u∀ 2 | u∀ �2 ∪
We have used the results �u∀ �1 | u∀ �2 ∪ = 0 and �u∀ �2 | u∀ �1 ∪ = 0.
Q6(9) Prove triangle inequalities (6.38) and (6.39).
SQ6(9) For the triangle inequality in expression (6.38) we have,
using the Schwarz inequality, i.e., |�u∀ | v∀ ∪| ≤ ||u∀ || || v∀ ||,
||u∀ + v∀ ||2 = �u∀ + v∀ | u∀ + v∀ ∪ = �u∀ | u∀ ∪ + �∀v | v∀ ∪ + �u∀ | v∀ ∪ + �∀
v | u∀ ∪
2 2
 2
≤ || u∀ || + || v∀ || + 2 ||u∀ || || v∀ || = || u∀ || + || u∀ ||
� || u∀ + v∀ || ≤ || u∀ || + || u∀ ||.
For the triangle inequality in expression (6.39) we have, using the
above inequality,
||u∀ || = ||u∀ − v∀ + v∀ || ≤ ||u∀ − v∀ || + ||∀
v ||
� || u∀ || − || v∀ || ≤ ||u∀ − v∀ ||.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Three-Dimensional Real Vectors 25

Similarly we have
||∀ v − u∀ + u∀ || ≤ ||∀
v || = ||∀ v − u∀ || + ||u∀ ||
v − u∀ || = ||u∀ − v∀ ||.
� || v∀ || − || u∀ || ≤ ||∀
It follows that
 
|| u∀ || − || v∀ || ≤ ||u∀ − v∀ ||.

Q6(10) Let {e∀ℓ } be an orthonormal basis. Show that any vector v∀ is


expressible as a sum of the projections v∀e∀ℓ of v∀ onto the basis vectors
e∀ℓ , i.e., v∀ = v∀e∀1 + v∀e∀2 + v∀e∀3 .
SQ6(10) Since e∀ℓ form an orthonormal basis we have, by Eqs.
(6.19) and (6.61),
v∀ = v1 e∀1 + v2 e∀2 + v3 e∀3 = v∀1 + v∀2 + v∀3 ,
where v∀1 = v1 e∀1 , v∀2 = v2 e∀2 and v∀3 = v1 e∀3 are the projections of v∀
onto the basis vectors e∀ℓ in accordance with Eq. (6.46).
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 7

Matrices and Their Relations with


Vectors

Q7(1) Verify Eq. (7.37) on the trace of square matrices.


SQ7(1) Since
  
M·N mn
= M mk N kn
k
   
� tr M · N = M nk N kn .
n k

The
 sameresult is obtained
 if we interchange M and N , i.e., we get
tr M · N = tr N · M .
Q7(2) Show that the inverse of a square invertible matrix is
unique.
SQ7(2) Let M −1 be an inverse of M and let N be another inverse
of M , i.e., we have
M −1 · M = M · M −1 = I and N · M = M · N = I .
Then
   
M −1 = M −1 · I = M −1 · M · N = M −1 · M · N = N .
In other words there is no other inverse.

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

28 Matrices and Their Relations with Vectors

Q7(3) Prove Eq. (7.71).


SQ7(3) Suppose a matrix N exists such that N · M = I . Let C be
a column matrix such that M · C = 0. Then
 
N · M · C = 0.
 
We get, because of N · M = I ,
   
N · M · C = N · M · C = C = 0.

In other words we have

N · M = I and M · C = 0 � C = 0.

This implies M is invertible


 by property P7.4(3) in §7.4. Now
multiplying N · M = I by the inverse M −1 from the left we
get N = M −1 . This result implies M · N = M · M −1 = I .
Q7(4) What is the inverse of a diagonal matrix with non-zero
diagonal elements M j j ?
SQ7(4) The inverse is also a diagonal matrix with diagonal
element M−1
jj .

Q7(5) Show that the orthogonal matrix R z (θz ) in Eq. (7.149) is not
selfadjoint.
SQ7(5) The matrix R z (θz ) in Eq. (7.149) has real elements. This
means that its transpose R zT (θz ) is equal to its adjoint R†z (θz ). But
R zT (θz ) not equal to R z (θz ). It follows that its adjoint R†z (θz ) is
different from R z (θz ), i.e., R z (θz ) is not selfadjoint.
Q7(6) Prove Eq. (7.160).
SQ7(6) By Eq. (7.81) we get

�U C | U C ∪ = �U † U C | C ∪, �U C 1 | U C 2 ∪ = �U † U C 1 | C 2 ∪.

Since U † U = I by Eq. (7.158) we get

�U † U C | C ∪ = �C | C ∪, �U † U C 1 | C 2 ∪ = �C 1 | C 2 ∪.

Q7(7) Verify Eq. (7.162) on simultaneous unitary transforma­


tions.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Matrices and Their Relations with Vectors 29

SQ7(7) Using Eqs. (7.81) and (7.158) we get


  
�C � | M � C � ∪ = �U C | U MU † U C ∪
 
= �U C | U M U † U C ∪
= �U C | U MC ∪ = �U † U C | MC ∪
= �C | MC ∪.
Q7(8) Pauli matrices are denoted by σx , σ y , σz .

(a) Verify the six properties of Pauli matrices shown in Eqs.


(7.41) to (7.49).
 2  
(b) Show that ax σx + ay σ y + az σz = ax2 + a2y + az2 I 2×2 ,
where I 2×2 is the 2 × 2 identity matrix and ax , ay , az are
real numbers.
(c) What are the determinant and trace of each of the Pauli
matrices?
(d) What is the inverse of each of Pauli matrices?
(e) Verify that Pauli matrix σx satisfies the selfadjointness
condition in Eq. (7.163) for the vectors C α∀ z and C β∀z in Eq.
(7.118).
(f) Show that the eigenvectors of each Pauli matrix correspond­
ing to the eigenvalues ±1 given by Eqs. (7.116), (7.117) and
(7.118) are orthonormal.

SQ7(8)(a) Pauli matrices are


0 1 0 −i 1 0
σx := , σ y := , σz := .
1 0 i 0 0 −1
We have
0 1 0 −i i 0
σx · σ y = · = ,
1 0 i 0 0 −i
0 −i 0 1 −i 0
σ y · σx = · = .
i 0 1 0 0 i
It follows that σx · σ y − σ y · σx is equal to
i 0 1 0
σx · σ y − σ y · σx = 2 = 2i = 2i σz .
0 −i 0 −1
0 0 0 0
σx · σ y + σ y · σx = 2 = .
0 0 0 0
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

30 Matrices and Their Relations with Vectors

Similar calculations prove the remaining commutation and anticom­


mutation relations. The fact that the square of a Pauli matrices is
equal to the 2 × 2 identity matrix is also similarly proved.
SQ7(8)(b) Expand the square we get
 2
ax σx + ay σ y + az σz = ax2 σx2 + a2y σz2 + az2 σz2 + cross terms.

The cross terms are all  zero since Pauli matrices anticommute, e.g.,
ax ay σx · σ y + σ y · σx = 02×2 . Next we can verify the fact that the
square of each Pauli matrix is the 2 × 2 identity matrix I 2×2 , e.g.,

0 1 0 1 1 0
σx2 := = = I 2×2 .
1 0 1 0 0 1

It follows that
 2  
ax σx + ay σ y + az σz = ax2 + ay2 + az2 I 2×2 .

SQ7(8)(c) The determinant of a Pauli matrix is equal to −1, i.e., we


have
     
0 1    0 
  = −1,  0 − i  = −1,  1
1 0 i 0  0 − 1  = −1.

The trace of every Pauli matrix is clearly 0.


SQ7(8)(d) The fact that σx · σx = I 2×2 implies that the inverse of
σx is equal to σx . The same result applies to σ y and σz , i.e., we have
σx−1 = σx , σ y−1 = σ y , σz−1 = σz .
SQ7(8)(e) We need to evaluate �C α∀ z | σx C α∀ z ∪, �σx C α∀ z |
C α∀ z ∪, �C α∀ z | σx C β∀z ∪, �σx C α∀ z | C β∀x ∪ and so on. First we have

0 1 1 0
σx C α∀ z = · = ,
1 0 0 1

and

0 1 0 1
σx C β∀z = · = .
1 0 1 0
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Matrices and Their Relations with Vectors 31

Then we have, using Eq. (7.79),


0
�C α∀ z | σx C α∀ z ∪ = (1 0) · = 0,
1
1
�σx C α∀ z | C α∀ z ∪ = (0 1) · = 0,
0
1
�C α∀ z | σx C β∀z ∪ = (1 0) · = 1,
0
0
�σx C α∀ z | C β∀z ∪ = (0 1) · = 1,
1
0
�C β∀z | σx C α∀ z ∪ = (0 1) · = 1,
1
1
�σx C β∀z | C α∀ z ∪ = (1 0) · = 1,
0
1
�C β∀z | σx C β∀z ∪ = (0 1) · = 0,
0
0
�σx C β∀z | C β∀z ∪ = (1 0) · = 0.
1
The required results follow immediately.
SQ7(8)(f) For normalisation we have
1 1
�C α∀ x | C α∀ x ∪ = (1 1) · = 1.
2 1
1 1
�C α∀ y | C α∀ y ∪ = (1 − i ) · = 1.
2 i
1
�C α∀ z | C α∀ z ∪ = (1 0) · = 1.
0
Similar calculations show that
�C β∀x | C β∀x ∪ = �C β∀y | C β∀y ∪ = �C β∀z | C β∀z ∪ = 1.
For orthogonality we have
1 1
�C α∀ x | C β∀x ∪ = (1 1) · = 0.
2 −1
1 1
�C α∀ y | C β∀y ∪ = (1 − i ) · = 0.
2 −i
0
�C α∀ z | C β∀z ∪ = (1 0) · = 0.
1
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

32 Matrices and Their Relations with Vectors

Q7(9) Show that

(a) If P is an n × n projection matrix then I n×n – P is also a


projection matrix.
(b) If P and Q are n × n projection matrices then P · Q is also a
projection matrix if P and Q commute.
(c) If P and Q are n × n projection matrices then P + Q is also
a projection matrix if P · Q = Q · P = 0, where 0 is the n × n
zero matrix.

SQ7(9)(a) First, the matrix I n×n − P is selfadjoint since P is


selfadjoint. Explicitly we have, by Eq. (7.60), that
 †

I n×n − P = I n×n − P † = I n×n − P.

Secondly, the matrix I n×n − P is idempotent, i.e.,


 2
I n×n − P = I 2n×n + P 2 − I n×n · P − P · I n×n
= I n×n − P,
since P 2 = P and P · I n×n = P · I n×n = P. The desired result follows
from Definition 7.7.5(1) on projection matrices.
SQ7(9)(b) First, we have
 †  2    
P · Q = Q † · P † = Q · P, P·Q = P·Q · P·Q .
If Q and P commute then
 †
P·Q =P·Q and

 2    
P·Q = P·Q · Q·P =P·Q·P=P·P·Q
= P · Q.
In other words P · Q is selfadjoint and idempotent. It follows that
P · Q is a projection matrix.
SQ7(9)(c) First, P+Q is clearly selfadjoint. For idempotence we
have
 2
P + Q = P2 + Q2 + P · Q + Q · P = P + Q,
since P · Q = Q · P = 0. It follows that P + Q is a projection matrix.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Matrices and Their Relations with Vectors 33

Q7(10) Show that the 2 × 2 matrices in Eqs. (7.178) to (7.183) are


projection matrices.1 Find their eigenvectors.
SQ7(10) The matrix P α∀ x in Eq. (7.178) is selfadjoint since it is
equal to its adjoint, i.e., it is equal to the complex conjugate of
its transpose. All we now need is to prove idempotence. This is
confirmed by explicitly calculations, i.e.,
1 1 1 1 1 1 2 2
P 2α∀ x = · = = P α∀ x .
4 1 1 1 1 4 2 2
Hence P α∀ x is a projection matrix. There are two orthonormal
eigenvectors C α∀ x and C β∀x corresponding to the eigenvalues 1 and
0 respectively, i.e., we have
 †   † 
P α∀ x C α∀ x = C α∀ x C α∀ x C α∀ x = C α∀ x C α∀ x C α∀ x = C α∀ x ,
 †   † 
P α∀ x C β∀x = C α∀ x C α∀ x C β∀x = C α∀ x C α∀ x C β∀x = 02×2 = 0 C β∀x .
We can also verify these results by explicit calculations in terms of
the matrix P α∀ x and the column vectors C α∀ x and C β∀x .
Similar calculations show that all the matrices in Eqs. (7.179)
to (7.183) are selfadjoint and idempotent, i.e., they are projection
matrices. They all have eigenvalues 0 and 1 in accordance with
Theorem 7.7.5(1). Explicitly we have
 † 
P β∀x C α∀ x = C β∀x C β∀ C α∀ x = 0 C α∀ x ,
x
 † 
P β∀x C β∀x = C β∀x C β∀ C β∀x = C β∀x .
x

 † 
P α∀ y C α∀ y = C α∀ y C α∀ y C α∀ y = C α∀ y ,
 † 
P α∀ y C β∀y = C α∀ y C α∀ y C β∀y = 0 C β∀y .

 † 
P β∀y C α∀ y = C β∀y C β∀ C α∀ y = 0 C α∀ y ,
y
 † 
P β∀y C β∀y = C β∀y C β∀ C β∀y = C β∀y .
y

 † 
P α∀ z C α∀ z = C α∀ z C α∀ z C α∀ z = C α∀ z ,
 † 
P α∀ z C β∀z = C α∀ z C α∀ z C β∀z = 0 C β∀z .

1 Isham p. 91.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

34 Matrices and Their Relations with Vectors

 † 
P β∀z C α∀ z = C β∀z C β∀ C α∀ z = 0 C α∀ z ,
z
 † 
P β∀z C β∀x = C β∀z C β∀ C β∀z = C β∀z .
x

We have used the orthonormality property of the eigenvectors


C α∀ x , C β∀x , C α∀ y , C β∀y , C α∀ z , C β∀z of Pauli matrices proved earlier in
SQ7(8)(f).
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 8

∀3
Operations on Vectors in IE

Q8(1) Show that an invertible operator has a unique inverse.


SQ8(1) Let A−1 be an inverse of A and let B  be another inverse of
A, i.e., we have A A = A A = II , B
−1 −1  A = A B
 = II . Then
   
A−1 = A−1 II = A−1 A B
 = A−1 A B =B .

Q8(2) Show that in IE∀ 3 the condition B  A = II is sufficient to imply


 
the invertibility of A with B as its inverse.
SQ8(2) Let u∀ be a vector such that Au∀ = 0. ∀ Then we have
  ∀   
B A u∀ = 0. But B A = II . It follows that we must have u∀ = 0. ∀
  −1
Hence A is invertible by Theorem 8.2.2(1). Let A be its inverse,
then multiplying the equation B  A = II from the left by A−1 we
 = A−1 .
immediately get the result that B
Q8(3) Let A be a matrix representation of an invertible operator
A. Show that the matrix representation of the inverse A−1 of A is the
inverse matrix A−1 to the matrix A.
SQ8(3) The matrix representations M Â , M Â−1 of A and A−1 in a
given basis {e∀ℓ } are given by their matrix elements
   
MÂ kℓ = �e∀k | Ae∀ℓ ∪, MÂ −1 ℓm = �e∀ℓ | A−1 e∀m ∪.

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

3

36 Operations on Vectors in IE

 
The product M Â · M Â −1 has elements M Â · M Â −1 km given by
   
MÂ kℓ
MÂ −1 ℓm


= �e∀k | Ae∀ℓ ∪�e∀ℓ | A−1 e∀m ∪


= � A† e∀k | e∀ℓ ∪�e∀ℓ | A−1 e∀m ∪ = �A† e∀k | A−1 e∀m ∪

= �e∀k | A A−1 e∀m ∪ = �e∀k | II e∀m ∪ = δkm .


It follows that M Â · M Â −1 = II � M Â −1 = M −Â 1 . We have used
the Pythagoras theorem in the form of Eq. (6.29).
Q8(4) Let {e∀ℓ , ℓ = 1, 2, 3} be a basis for IE∀ 3 and let A be an
invertible operator. Show that {e∀ �ℓ = Ae∀ℓ , ℓ = 1, 2, 3} is also a basis
for IE∀ 3 .1
SQ8(4) All we need is to show that e∀ �ℓ are linearly independent. We

can start by considering an arbitrary linear combination ℓ cℓ e∀ �ℓ ,
i.e.,
3
 3
 3

cℓ e∀ �ℓ = cℓ Ae∀ℓ = A cℓ e∀ℓ .
ℓ=1 ℓ=1 ℓ=1
Then
3
 3

∀ � A
cℓ e∀ �ℓ = 0 ∀
cℓ e∀ℓ = 0.
ℓ=1 ℓ=1

On account of A being invertible we have, by Theorem 8.2.2(1),


3
 3

A ∀ �
cℓ e∀ℓ = 0 ∀
cℓ e∀ℓ = 0.
ℓ=1 ℓ=1

Since e∀ℓ are linearly independent we must have cℓ = 0 ∞ℓ. In other


words we have
3
∀ � cℓ = 0 ∞ℓ.
cℓ e∀ �ℓ = 0
ℓ=1

This immediately implies the linear independence of e∀ �ℓ . It follows


that {e∀ �ℓ , ℓ = 1, 2, 3} is a basis for IE 3 .
1 Halmos p. 63.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

3

Operations on Vectors in IE 37

Q8(5) Show that an operator A is invertible if and only if every


vector v∀ ⇒ IE∀ 3 can be expressed as v∀ = Au∀ for some u∀ ⇒ IE∀ 3 .2

SQ8(5) First, suppose A is invertible. Then, by definition, A


generates a one-to-one mapping of IE∀ 3 onto IE∀ 3 . This means that
any vector v∀ ⇒ IE∀ 3 can be regarded as a vector in the range of the
operator. In other words there is a vector u∀ in the domain of the
operator such that Au∀ = v∀.
Secondly, suppose A is such that every vector v∀ in IE∀ 3 can be
expressed as v∀ = Au∀ for some vector u∀. Let {e∀ �ℓ } be a basis in IE∀ 3 .
Then there exists a set of vectors e∀ℓ such that e∀ �ℓ = Ae∀ℓ . Moreover
these vectors e∀ℓ are linearly independent, i.e.,

∀ � cℓ = 0 ∞ℓ.
cℓ e∀ℓ = 0

We can prove this as follows:


 
∀ � A
cℓ e∀ℓ = 0 ∀
cℓ e∀ℓ = 0
ℓ ℓ
 
� cℓ Ae∀ℓ = ∀ � cℓ = 0 ∞ℓ.
cℓ e∀ �ℓ = 0
ℓ ℓ

It follows {e∀ℓ } is also a basis for IE∀ 3 and any vector v∀ ⇒ IE∀ 3 can be

written as v∀ = ℓ vℓ e∀ℓ . Then
 
∀ �
Av∀ = 0 vℓ Ae∀ℓ = ∀
vℓ e∀ �ℓ = 0.
ℓ ℓ

Since e∀ �ℓ
are linearly independent we must have vℓ = 0 ∞ℓ, i.e.,
∀ We can conclude that A is invertible by Theorem 8.2.2(1).
v∀ = 0.

Q8(6) Show that the matrix representation of the adjoint of A is



the adjoint matrix M A� to the matrix M Â .

SQ8(6) The matrix element of M† are given by


 
MÂ † lm = �e∀l | A† e∀m ∪ = �Ae∀l | e∀m ∪ = �e∀m | Ae∀l ∪.
   
Since MÂ lm = �e∀l | Ae∀m ∪ and MÂ ml = �e∀m | Ae∀l ∪ we get
   
M† lm = M ml .

2 Halmos pp. 62–63.


June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

3

38 Operations on Vectors in IE

This means that the transpose of M is equal to M† . We arrive at the
desired result since the transpose of a real square matrix is equal the
adjoint of the matrix. All quantities are real in IE∀ 3 , e.g., �Ae∀l | e∀m ∪ =
�e∀m | Ae∀l ∪.
∀ ) = 0.
Q8(7) Let F be a linear functional on IE∀ 3 . Show that F (0
3
Prove a similar result for linear operators.
SQ8(7) For a linear functional F we have F (au∀ ) = aF (u∀ ). Now,
∀ it follows that
let a = 0 we get F (0u∀ ) = 0F (u∀ ) = 0. Since 0u∀ = 0

F (0 ) = 0.  
For a linear operator A we also have A au∀ = a Au∀. Now, let

a = 0 we get A(0u∀ = 0Au∀ = 0. ∀ Since 0u∀ = 0 ∀ it follows that
∀ = 0.
A 0 ∀

Q8(8) Prove Eqs. (8.38) to (8.42) on adjoint operations.


SQ8(8)
(1) For Eq. (8.17) and SP6.3.1(2) on scalar product we get, ∞ u, ∀ v∀,
 
�u∀ | a A + bB  v∀ ∪
= �u∀ | a Av∀ ∪ + �u∀ | bB  v∀ ∪ = �a A† u∀ | v∀ ∪ + �bB
 † u∀ | v∀ ∪
 †    †
= � a A + bB  † u∀ | v∀ ∪ � a A + bB  = a A† + bB †.

Equation (8.39) is a special case of Eq. (8.38).


(2) To prove Eq. (8.40) we have, using Eq. (8.34),
�u∀ | A†† v∀ ∪ = �u∀ | (A† )† v∀ ∪ = �A† u∀ | v∀ ∪ = �u∀ | Av∀ ∪.
Since this is true for all u∀ and v∀ we can conclude that A†† = A.
(3) To prove Eq. (8.41) we have, ∞ u,
∀ v∀,
 †
�u∀ | A B  v∀ ∪ = �B
 v∀ ∪ = �A† u∀ | B  † A† u∀ | v∀ ∪ � A B
  † A† .
=B

(4) To prove Eq. (8.42) we make used of Eq. (8.41) to get


 −1 †  −1 † †
A A = A A .

Since A A−1 = II and II † = II we get


 †  †  −1
II = A−1 A† � A−1 = A† .

3 Halmos pp. 20, 55.


June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

3

Operations on Vectors in IE 39

In (2) and (3) above we have used Eq. (6.31) which tells us that
�u∀ | Av∀ ∪ = �u∀ | B ∀ v∀ ⇒ IE∀ 3 � A = B
 v∀ ∪ ∞ u, ∀.

Q8(9) Associated with the projection operation given by Eq. (6.41)


we can define an operator Pi∀ by
P ∀ v∀ = v x i∀ ∞ v∀ ⇒ IE∀ 3 .
i
Show that
(a) The operator Pi∀ is idempotent.
(b) The operator Pi∀ is selfadjoint.
(c) The operator possesses only two eigenvalues 0 and 1. Find
the corresponding eigenvectors.
(d) The matrix representation MP̂ i∀ of Pi∀ in basis {i∀, ∀j , k∀} is
⎛ ⎞
1 0 0
MP̂ i∀ = ⎝ 0 0 0 ⎠ .
0 0 0
Show also that MP̂ i∀ is a projection matrix. Find the
eigenvalues and their corresponding eigenvectors of MP̂ i∀ .
 2  2
SQ8(9)(a) Pi∀ v∀ = Pi∀ Pi∀ v∀ = v x Pi∀ i = v x i∀ ∞∀
v � Pi∀ = Pi∀.

SQ8(9)(b) The operator Pi∀ is selfadjoint since


�u∀ | Pi∀ v∀ ∪ = �u∀ | v x i∀ ∪ = v x �u∀ | i∀ ∪ = v x u x .
�Pi∀ u∀ | v∀ ∪ = �u x i∀ | v∀ ∪ = u x �i∀ | v∀ ∪ = u x v x .
� �u∀ | Pi∀ v∀ ∪ = �Pi∀ u∀ | v∀ ∪ ∞u, ∀ v∀.
The concept of selfadjointness is introduced for square matrices by
Definition 7.7.4(1). A square matrix is selfadjoint if it satisfies that
selfadjointness condition stated after Definition 7.7.4(1). We have
applied the condition to show that the projector Pi∀ is selfadjoint.
All these features are introduced for projectors in P9.3.1(6) and for
operators in Definition 9.4.1(1).

SQ8(9)(c) Let e∀ be an eigenvector of Pi∀ corresponding to


eigenvalue a, i.e., Pi∀ e∀ = ae∀, where a is real (see Eq. (8.49)). We have,
on account of Pi∀ being idempotent,
 2
Pi∀ e∀ = Pi∀ e∀ = ae∀ and
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

3

40 Operations on Vectors in IE

 2  
Pi∀ e∀ = Pi∀ Pi∀ e∀ = a Pi∀ e∀ = a2 e∀ � a2 = a

� a = 0 or 1.
Eigenvectors corresponding to the eigenvalue 1 is of the form u x i∀
while eigenvectors corresponding to the eigenvalue 0 is of the form
u y ∀j + u z k∀. The eigenvalue 1 is nondegenerate since we do not
count two eigenvectors differing only by a multiplicative constant
as distinct. The eigenvalue 0 is degenerate with degeneracy 2.
Any linear combination of ∀j and k∀ is also an eigenvector of P ∀ i
corresponding to the eigenvalue 0.
SQ8(9)(d) The given matrix M P̂ i∀ is clearly selfadjoint and idempo­
tent. It follows that M P̂ i∀ is a projection matrix by Definition 7.7.5(1).
By the same argument presented for the preceding question we can
see that the matrix has two eigenvalues 1 and 0. The eigenvector
corresponding to the eigenvalue 1 is
⎛ ⎞
1
⎝0⎠,
0
and the eigenvectors corresponding to the eigenvalue 0 are linear
combinations of
⎛ ⎞ ⎛ ⎞
0 0
⎝ 1 ⎠ and ⎝ 0 ⎠ .
0 1
The eigenvalue 1 is nondegenerate while the eigenvalue 0 is
degenerate with degeneracy 2.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 9

Special Operators on IE∀ 3

Q9(1) Show that an operator on IE∀ 3 is orthogonal if and only if it


is invertible and its inverse is equal to its adjoint.
SQ9(1) Let R be an orthogonal operator. By Definition 9.2.2(1) we
have ||Ru∀ || = || u∀ ||. It follows that
∀ � || u∀ || = 0 � u∀ = 0.
Ru∀ = 0 ∀
Theorem 8.2.2(1) then tells us that R is invertible. Next we have, by
Eq. (8.34),
�u∀ | u∀ ∪ = �Ru∀ | Ru∀ ∪ = �u∀ | R† Ru∀ ∪ ∞u.

As pointed out in P9.4.2(4) the product operator R† R is selfadjoint.
Hence Corollary 9.4.1(1) applies, i.e., we have
R† R = II � R† = R−1 .
To prove the converse we first assume that R is invertible with
its inverse R−1 equal its adjoint R† . Then we have
�Ru∀ | Ru∀ ∪ = �u∀ | R† Ru∀ ∪ = �u∀ | R−1 Ru∀ ∪ = �u∀ | u∀ ∪ ∞u.

The operator is therefore orthogonal.
Q9(2) Show that orthogonal operators on IE∀ 3 have at most two
eigenvalues, i.e., ±1. Find the eigenvalues and eigenvectors of Rr x in
Eq. (9.4) and Rr x yz in Eq. (9.5).

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

42 Special Operators on IE∀ 3

SQ9(2) Let r be an eigenvalue an orthogonal operator R corre­


sponding to normalised eigenvector e∀, i.e.,

Re∀ = re∀ where �e∀ | e∀ ∪ = 1.

Then

�Re∀ | Re∀∪ = r 2 and �Re∀ | Re∀∪ = �e∀ | e∀ ∪ = 1

� r 2 = 1 � r = ±1.

The operator Rr x in Eq. (9.4) has eigenvalue −1, corresponding


to eigenvectors of the form v x i∀ and eigenvalue 1 corresponding to
eigenvectors of the form v y ∀j + v z k∀.
The operator Rr x yz in Eq. (9.5) has only one eigenvalue, i.e., −1,
which is degenerate corresponding to eigenvectors of the form v x i∀+
v y i∀ + v z i∀, i.e., any vector in IE∀ 3 . The operator acts like −II .
Q9(3) Let {e∀ℓ } be an orthonormal basis for IE∀ 3 show that their
orthogonal transforms {e∀ �ℓ } generated by an orthogonal operator R
is also an orthonormal basis for IE∀ 3 .
SQ9(3) An orthogonal transformation preserves the scalar prod­
uct. It follows that the transformed vectors e∀ �ℓ of the basis vectors
e∀ℓ would remain orthonormal. Since orthogonal vectors are also

linearly independent the transformed vectors e∀ℓ would form a
complete orthonormal set, i.e., an orthonormal basis.
Q9(4) Prove Eqs. (9.6) to (9.8) on orthogonal transformations.
SQ9(4) We can prove Eqs. (9.6) to (9.8) using the property R† R =
II shown in SQ9(1).
For Eq. (9.6) we have

�Ru∀ | Rv∀ ∪ = �u∀ | R† Rv∀ ∪ = �u∀ | v∀ ∪.

For Eq. (9.7) we have

�u∀ � | A � v∀ � ∪ = �Ru∀ | R A R† Rv∀ ∪ = �Ru∀ | R Av∀ ∪


= �R† R u∀ | Av∀ ∪ = �u∀ | Av∀ ∪.

Eq. (9.8) follows from Eq. (9.7).


June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Special Operators on IE∀ 3 43

Q9(5) Prove Theorem 9.3.2(1).


SQ9(5) By Definition 9.3.2(1) we have, ∞∀ v ⇒ IE 3 ,
 
P S∀ + P S∀� v∀ = P S∀v∀ + P S∀� v∀ = v∀S∀ + v∀S∀� = v∀.

We have used Eq (6.59). It follows that P S∀ + P S∀� = II .


Q9(6) Find the eigenvectors of the two-dimensional projector in
Eq. (9.43) corresponding to the eigenvalues 1 and 0.
SQ9(6) The vector e∀3 is an eigenvector of P S∀12 corresponding
to the eigenvalue 0 while e∀1 and e∀2 are eigenvectors of P S∀12
corresponding to the eigenvalue 1. The eigenvalue 0 is nonde­
generate while the eigenvalue 1 is degenerate with degeneracy 2.
Any linear combination of e∀1 and e∀2 is also an eigenvector of P S∀12
corresponding to the eigenvalue 1.
Q9(7) Verify Eq. (9.44).
   
SQ9(7) When �e∀1 | e∀2 ∪ = 0 we have P e∀1 Pe∀2 u∀ = P e∀1 P e∀2 u∀ = 0∀
since the projection P e∀2 u∀ of u∀ onto e∀2 has a zero projection onto e∀1 .
Next, suppose Pe∀1 Pe∀2 = 0. Then for any u∀ and v∀ we have

�P e∀1 u∀ | Pe∀2 v∀ ∪ = �u∀ | P e∀1 Pe∀2 v∀ ∪ = 0.

Since

P e∀1 u∀ = u1 e∀1 , u1 = �e∀1 | u∀ ∪, Pe∀2 v∀ = v2 e∀2 , v2 = �e∀2 | v∀ ∪,

we get

�P e∀1 u∀ | P e∀2 v∀ ∪ = �u1 e∀1 | v2 e∀2 ∪ = u1 v2 �e∀1 | e∀2 ∪ = 0.

It follows that �e∀1 | e∀2 ∪ = 0.


Q9(8) Prove Theorems 9.4.1(1), 9.4.1(2) and Corollary 9.4.1(1).
SQ9(8)1 We start with Eq. (6.30) which implies

v | Aw
�∀ v � Aw
∀ ∪ = 0 ∞∀  ∞w
∀ =0 ∀ � A = 
0.

The proof of Theorem 9.4.1(1) is based on showing that

�u∀ | Au∀ ∪ = 0 ∞u∀ � �∀


v | Aw
∀ ∪ = 0 ∞∀ ∀.
v, w
1 Halmos p. 138.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

44 Special Operators on IE∀ 3

This result can be established by expressing the scalar product


v | Aw
�∀ ∀ ∪ as a sum of the values of the quadratic form generated by A
for different vectors. We can  achieve this by expanding the quadratic
form �(v∀ + w ∀ ) | A(v∀ + w
∀ ∪, i.e.,

∀ ) | A(v∀ + w
�(v∀ + w ∀ ∪ = �∀ v | Av∀ ∪ + �w
∀ | Aw v | Aw
∀ ∪ + �∀ ∀ | Av∀ ∪.
∀ ∪ + �w
Since A is selfadjoint we have �∀ v | Aw ∀ ∪ = �Av∀ | w
∀ ∪. We also have

�A v∀ | w ∀ ∪ = �w ∀ | A v∀ ∪ since scalar product in IE∀ 3 is commutative.

The above equation becomes

�(v∀ + w∀ ) | A(v∀ + w
∀ ∪ = �∀v | Av∀ ∪ + �w∀ | Aw
∀ ∪ + 2�∀v | Aw
∀ ∪.
We get

v | Aw
2�∀ ∀ ) | A(v∀ + w
∀ ∪ = �(v∀ + w v | Av∀ ∪ − �w
∀ ∪ − �∀ ∀ | Aw
∀ ∪.
The condition �u∀ | Au∀ ∪ = 0 ∞u∀ implies that the values of the
quadratic form for v∀ + w ∀ , v∀ and w
∀ on the right hand side are all
zero. The above equation implies �∀ v | Aw∀ ∪ = 0 ∞∀v, w∀ which in turn

implies A = 0. 
Theorem 9.4.1(2) follows from Theorem 9.4.1(1) since
 
�u∀ | Au∀ ∪ = �u∀ | B
 u∀ ∪ ∞u∀ � �u∀ | A − B  u∀ ∪ = 0 ∞u∀
� A − B = 0.
Corollary 9.4.1(1) follows immediately from Theorem 9.4.1(2) since
�u∀ | Au∀ ∪ = �u∀ | u∀ ∪ = �u∀ | II u∀ ∪ ∞u∀ � A = II .

Q9(9) Show that all projectors satisfy the sefladjointness condi­


tion in Eq. (9.35) and that they are also idempotent.
SQ9(9) From Definition 9.3.1(1) it is easily verified that any one­
dimensional projector Pe∀ satisfies the selfadjoint condition in Eq.
(9.35), i.e., �u∀ | Pe∀ v∀ ∪ = ue ve = �P e∀ u∀ | v∀ ∪, where ue = �e∀ | u∀ ∪
and ve = �e∀ | v∀ ∪. Since
   
Pe∀ Pe∀ u∀ = P e∀ ue e∀ = ue Pe∀ e∀ = ue e∀ = P e∀ u∀,
the operator is also idempotent. A two-dimensional projector P S∀
is expressible as the sum of two orthogonal one-dimensional
projectors as shown in Eq. (9.43), i.e.,
P S∀ = P e∀1 + P e∀2 , �e∀1 | e∀2 ∪ = 0.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Special Operators on IE∀ 3 45

The selfadjointness property can be verified as follows:


 
�u∀ | P S∀ v∀ ∪ = �u∀ | P e∀1 + P e∀2 v∀ ∪ = �u∀ | P e∀1 v∀ ∪ + �u∀ | P e∀2 v∀ ∪
 
= �P e∀1 u∀ | v∀ ∪ + �P e∀2 u∀ | v∀ ∪ = � Pe∀1 + P e∀2 u∀ | v∀ ∪
= �P S∀ u∀ | v∀ ∪.
Since P e∀1 and P e∀2 are idempotent and orthogonal we have
 2
Pe∀1 + P e∀2 = P 2 + P 2 = Pe∀1 + P e∀2 ,
e∀1 e∀2

showing that P e∀1 + P e∀2 is idempotent.


Q9(10) Show that selfadjoint operators are represented by
selfadjoint matrices.
SQ9(10) The matrix representation of A† is the adjoint matrix to
the matrix representation to A, as shown in SQ8(6). For a selfadjoint
operator we have A† = A which implies that matrix representation
of A† is the same as matrix representation of A, i.e., the matrix is
selfadjoint.
Q9(11) Demonstrate the inequality in Eq. (9.53) with u∀ = u y ∀j and
v∀ = v x i∀. Verify Eq. (9.56).
SQ9(11) For the orthogonal operator Rz (θz ) in Eq. (9.2) and for
any two vectors u∀ and v∀ lying in the x -y plane, i.e., u∀ = u x i∀ + u y ∀j and
v∀ = v x i∀ + v y ∀j , we have, in accordance with Eq. (9.52),
Rz (θz )u∀ = (u x cos θz − u y sin θz ) i∀ + (u x sin θz + u y cos θz ) ∀j ,
Rz (θz )v∀ = (v x cos θz − v y sin θz ) i∀ + (v x sin θz + v y cos θz ) ∀j .
Then
�u∀ | Rz (θz )v∀ ∪ = u x (v x cos θz − v y sin θz )
+u y (v x sin θz + v y cos θz ).

�R z (θz )u∀ | v∀ ∪ = (u x cos θz − u y sin θz )v x
+(u x sin θz + u y cos θz )v y .
When u∀ = u y ∀j with u x = 0 and v∀ = v x i∀ with v y = 0 we get
v ∪ = u y v x sin θz �= � Rz (θz )u∀ | v∀ ∪ = −u y v x sin θz .
�u∀ | Rz (θz )∀
Hence the operator Rz (θz ) is not selfadjoint. The corresponding
matrix R z (θz ) in Eq. (7.148) is also not selfadjoint.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

46 Special Operators on IE∀ 3

When acting on u∀ = u x i∀ + u y ∀j the adjoint operator is given by


Eqs. (9.55), i.e.,

R†z (θz )u∀ = u�x i∀ + u�y ∀j ,

where

u�x = u x cos θz + u y sin θz , u�y = −u x sin θz + u y cos θz .

This operator is seen to be different from Rz (θz ). We have

�R†z (θz )u∀ | v∀ ∪ = (u x cos θz + u y sin θz )v x + (−u x sin θz + u y cos θz )v y .

This agrees with the expression for �u∀ | Rz (θz )v∀ ∪ obtained earlier.
Q9(12) Prove Eq. (9.70).
SQ9(12) Let

e∀ �21 = α1 e∀21 + α2 e∀22 , e∀ �22 = β1 e∀21 + β2 e∀22 .

Since e∀ �21 and e∀ �22 are orthonormal the coefficients must satisfy

α12 + α22 = 1, β12 + β22 = 1, and α1 β1 + α2 β2 = 0.

We can also express e∀21 and e∀21 in terms of e∀ �21 and e∀ �21 , i.e.,
β2 e∀ �21 − α2 e∀ �22 β1 e∀ �21 − α1 e∀ �22
e∀21 = , e∀22 = .
β2 α1 − α2 β1 β1 α2 − α1 β2
Since e∀21 and e∀22 are

(1) orthogonal the coefficients must satisfy

�β2 e∀ �21 − α2 e∀ �22 | β1 e∀ �21 − α1 e∀ �22 ∪ = 0 � α1 α2 + β1 β2 = 0,

(2) normalised the coefficients must satisfy

β22 + α22 = (β2 α1 − α2 β1 )2 = α12 + β12 .

Adding β12 and α12 on both sides we get

β22 + β12 + α22 + α12 = 2(α12 + β12 ).

Since β22 + β12 = α22 + α12 = 1 we get α12 + β12 = 1 which in turn
implies α22 + β22 = 1.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Special Operators on IE∀ 3 47

For any given u∀ we have


 
Pe∀21 + P e∀22 u∀ = u1 e∀21 + u2 e∀22 ,
where u1 = �e∀21 | u∀ ∪ and u2 = �e∀22 | u∀ ∪ and
 
�u∀ | P e∀21 + P e∀22 u∀ ∪ = u21 + u22 .
For the projectors generated by e∀ �21 and e∀ �22 we have
  
P e∀ �21 u∀ = �e∀ �21 | u∀ ∪ e∀ �21 = α1 u1 + α2 u2 α1 e∀21 + α2 e∀22 ,
  
P e∀ �22 u∀ = �e∀ �22 | u∀ ∪ e∀ �22 = β1 u1 + β2 u2 β1 e∀21 + β2 e∀22 .
We get
   
�u∀ | P e∀ �21 u∀ ∪ = α1 u1 + α2 u2 �u∀ | α1 e∀21 + α2 e∀22 ∪
  
= α1 u1 + α2 u2 α1 u1 + α2 u2
= α12 u12 + 2α1 α2 u1 u2 + α22 u22 ,
   
�u∀ | P e∀ �22 u∀ ∪ = β1 u1 + β2 u2 �u∀ | β1 e∀21 + β2 e∀22 ∪
  
= β1 u1 + β2 u2 β1 u1 + β2 u2
= β12 u21 + 2β1 β2 u1 u2 + β22 u22 .
Adding the above equations we get, for all u∀,
 
�u∀ | P e∀ �21 + P e∀ �22 u∀ ∪
   
= α12 + β12 u21 + α22 + β22 u22 due to (α1 α2 + β1 β2 ) = 0
= u21 + u22 due to α12 + β12 = 1 and α22 + β22 = 1
 
= �u∀ | Pe∀21 + P e∀22 u∀ ∪
� P e∀ �21 + P e∀ �22 = P e∀21 + P e∀22 .
The present question is relevant when the eigenvalue a2 of a
selfadjoint operator A is degenerate with degeneracy 2 correspond­
ing to two orthonormal eigenvectors e∀21 , e∀22 . The corresponding
eigensubspace is two-dimensional spanned by e∀21 , e∀22 . The projector
onto this two-dimensional subspace is given by Eq. (9.43), i.e., equal
to the sum of two projectors P e∀21 and Pe∀22 , i.e.,

P Â (a2 ) = P e∀21 + P e∀22 = |e∀21 ∪�e∀21 | + |e∀22 ∪�e∀22 |


= P e∀ �21 + P e∀ �22 = |e∀ �21 ∪�e∀ �21 | + |e∀ �22 ∪�e∀ �22 |.
which is Eq. (9.70).
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

48 Special Operators on IE∀ 3

Q9(13) Show that the selfadjoint operator A in Eq. (9.71) admits


aℓ as its eigenvalues with the unit vectors e∀ℓ as the corresponding
eigenvectors.
SQ9(13) Given the operator A in Eq. (9.71) we have

3  3
 3

Ae∀ℓ = aℓ� Pe∀ℓ� e∀ℓ = aℓ� Pe∀ℓ� e∀ℓ = aℓ� δℓ� ℓ e∀ℓ = aℓ e∀ℓ .
ℓ� ℓ� ℓ�

This shows that e∀ℓ is an eigenvector of A corresponding to


eigenvalue aℓ .
Q9(14) Show that the products of an operator with its adjoint, i.e.,
A† A and A A† , are selfadjoint and positive.
SQ9(14) Using Eqs. (8.40) and (8.41) we get
 † †    † †  
A A = A† A†† = A† A and A A = A†† A† = A A† .

These results imply that A† A and A A† are selfadjoint. Moreover we
have
 
v | A A† v∀ ∪ = �A† v∀ | A† v∀ ∪ = || A† v∀ ||2 ⊕ 0,
�∀
 
v | A† A v∀ ∪ = �Av∀ | Av∀ ∪ = || Av∀ ||2 ⊕ 0.
�∀

These two equations hold for all v∀, including the eigenvectors of A† A
and A A† . It follows that the product operators do not have negative
eigenvalues, i.e., both product operators are positive.
Q9(15) Show that the expression of the inverse in Eq. (9.86)
satisfy the condition A A−1 = II .
−1
SQ9(15) Taking the expressions for A in Eq. (9.79) and A in Eq.
(9.86) we get
−1
   
A A = am� P Â (am� ) −1  Â
am P (am )
m� m

= −1
am� am P Â (am� )P Â (am )
m� , m

−1  Â
= am� am P (am� )δm� m
m� , m

= P Â (am ) = II .
m
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Special Operators on IE∀ 3 49

We have used the result P Â (am� ) P Â (am ) = P Â (am ) δm� m and the
spectral decomposition of the identity in Eq. (9.80).

Q9(16) The square root  A of a positive operator A is defined by
 2
Eq. (9.87). Show that 
A = A.
SQ9(16) From Eq. (9.87) we get
 2   �   � 
A = am P Â (am ) am� P Â (am� )
m m�
� �
= am am� P Â (am ) δmm�
m, m�

= am P Â (am ) = A.
m

We have used the result P Â (am ) P Â (am� ) = P Â (am ) δmm� and the
spectral decomposition of the A in Eq. (9.79).
Q9(17) Show that the quadratic form Q(A, u∀ ) generated by a
positive selfadjoint operator in any u∀ is non-negative.
SQ9(17) Since the quadratic form Q(A, u∀ ) is given by �u∀ | A u∀ ∪ we
get, for a positive selfadjoint operator A,

Q(A, u∀ ) = �u∀ | am P Â (am ) u∀ ∪, am ⊕ 0
m

= am �u∀ | P Â (am ) u∀ ∪
m
⊕ 0,
since the eigenvalues am of a positive selfadjoint operator A is non­
negative by definition 9.4.4(2) and
 2
�u∀ | P Â (am ) u∀ ∪ = �u∀ | P Â (am ) u∀ ∪
= �P Â (am )u∀ | P Â (am ) u∀ ∪
= || P Â (am )u∀ ||2 ⊕ 0.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 10

Probability, Selfadjoint Operators, Unit


Vectors and the Need for Complexness

Q10(1) In the model theory in §10.2.1 the observable A has


three possible values, i.e., a1 , a2 , and a3 . Find the probabilities for
the value a1 when the state vector u∀ is one of the following unit
vectors:

1 2
u∀1 = e∀1 , u∀2 = e∀2 , and u∀3 = � e∀1 + e∀3 .
3 3

What are the expectation values E(A, u∀ ) and the uncertainties


(A, u∀ ) in these cases?

SQ10(1)
(1) For the first state vector u∀1 = e∀1 the probability of getting the
value a1 on a measurement of A is 1 with expectation value
E(A, u∀1 ) = a1 and zero uncertainty.
(2) For the second state vector u∀2 = e∀2 the probability of getting
the value a1 on a measurement of A is 0 with expectation value
E(A, u∀2 ) = a2 and zero uncertainty.
 �  �
(3) For the third state vector u∀3 = e∀1 + 2 e∀3 / 3 the probability
of getting the value a1 on a measurement of A is 1/3 with

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

52 Probability, Selfadjoint Operators, Unit Vectors and the Need for Complexness

expectation value
1 
E(A, u∀3 ) = a1 + 2a3 .
3
The variant can be calculated in accordance with Eq. (3.33), i.e.,
we have
 2 1  2 2
Var(A, u∀3 ) = a1 − E(A, u∀3 ) + a3 − E(A, u∀3 )
3 3
4 2 2 2
= a1 − a3 + − a1 + a3
27 27
2 2
= a1 − a3 .
9
The uncertainty is then given by

 2
( A, u∀3 ) = Var(A, u∀3 ) = | a1 − a3 |.
3
The variant can also be calculated in accordance with Eq. (3.36),
i.e., we have
 1 2 2
Var( A, u∀3 ) = a12 + a32 − E(A, u∀
3 3
1 2  1 2
= a1 + 2a32 − a1 + 2a3
3 9
2 2
= a1 − a3 .
9

Q10(2) Show that A and A� in Eq. (10.16) commute.

SQ10(2) Writing
3
 3

A = aℓ P e∀ℓ and A� = a�ℓ� Pe∀ℓ� .
ℓ=1 ℓ� =1
we get
3
 3
 3

A A� = aℓ a�ℓ� Paℓ Paℓ� = aℓ aℓ� � Paℓ δℓℓ� = aℓ aℓ� Paℓ .
ℓ, ℓ� =1 ℓ, ℓ� =1 ℓ=1
3
 3
 3

A� A = aℓ� � aℓ Paℓ� Paℓ = aℓ� � aℓ Paℓ δℓℓ� = aℓ� aℓ Paℓ .
ℓ� , ℓ=1 ℓ, ℓ� =1 ℓ=1

It follows that A A − A A = 
 � �
0, i.e., A and A commute. �
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Probability, Selfadjoint Operators, Unit Vectors and the Need for Complexness 53

Q10(3) Prove Eq. (10.31).


SQ10(3) From Eq. (10.30) we get
� � �
|z� | = z�� z� = zz� = z� z = |z |.
More directly we have

|z� | = a2 + b2 = |z |.

For z = a + i b and w = c + i d. We have


  
|z| |w| = (a2 + b2 ) (c 2 + d 2 ) = (a2 + b2 )(c 2 + d 2 ).

From Eq. (10.29) on product of two complex numbers we get


zw = (ac − bd) + i (ad + bc).

The norm of zw is given by Eq. (10.30), i.e. we get



|zw| = (ac − bd)2 + (ad + bc)2

= (ac)2 + (bd)2 + (ad)2 + (bc)2

= (a2 + b2 )(c 2 + d 2 )
= |z| |w|.

Q10(4) Let θ , a and  b be real numbers. For the complex numbers


z = ei θ and z = ei θ a + i b) show that
|ei θ | = 1 and |z|2 = a2 + b2 .

SQ10(4) Using the Euler’s formula given in Eq. (10.26) we get


 �
ei θ = cos θ + i sin θ � ei θ = cos θ − i sin θ
 �
� |ei θ |2 = ei θ ei θ = (cos θ − i sin θ)(cos θ + i sin θ) = 1
� |z|2 = |ei θ |2 |(a + i b)|2 = a2 + b2 .
We have used Eqs. (10.30) and (10.31) and the formula sin2 θ +
cos2 θ = 1.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 11

Complex Vectors

Q11(1) Prove Eqs. (11.17), (11.18) and (11.19) in IE∀ 3c .


SQ11(1) To prove Eq. (11.17) we take the scalar product of e∀ℓc and
ζ∀ to get
3
 3

�e∀ℓc | ζ∀ ∪c = �e∀ℓc | ζℓ� e∀ℓc� ∪c = ζℓ� �e∀ℓc | e∀ℓc� ∪c = ζℓ ,
ℓ� =1 ℓ� =1
since �e∀ℓc | e∀ℓc� ∪c = δℓℓ� . A comparison with Eq. (11.12) shows that
ζℓ = uℓ + i vℓ . To prove Eq. (11.19) we have
3 3

∀ ∀ c
�ζ 1 | ζ 2 ∪ = � c
ζ1ℓ e∀ℓ | ζ2ℓ� e∀ℓc� ∪c
ℓ=1 ℓ� =1
3
 3

� �
= ζ1ℓ ζ2ℓ� �e∀ℓc | e∀ℓc� ∪c = ζ1ℓ ζ2ℓ .
ℓ, ℓ� =1 ℓ=1

Equation (11.18) is just a special case of Eq. (11.19).

Q11(2) Let
1   1  
ǫ∀1 = � i∀c + i ∀j c , ǫ∀2 = � i∀c − i ∀j c , ǫ∀3 = k∀ c .
2 2
(a) Write down the matrix representations of the above vectors
in the orthonormal basis {i∀c , ∀j c , k∀ c }.

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

56 Complex Vectors

(b) Show that ǫ∀ℓ , ℓ = 1, 2, 3 above form an orthonormal basis


for IE∀ 3c .
(c) Show that an arbitrary vector ζ∀ can be expressed as
3

ζ∀ = ζℓ ǫ∀ℓ , ζℓ = �∀ǫℓ | ζ∀ ∪c .
ℓ=1

SQ11(2)(a) The required matrix representations are given re­


spectively by
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 1 0
1 1
� ⎝ i ⎠ , � ⎝ −i ⎠ , ⎝ 0 ⎠ .
2 0 2 0 1

SQ11(2)(b) Since orthogonality implies linear independence all


we need is to show that the three vectors are orthonormal.1
(1) The vector ǫ∀3 is clearly normalised and orthogonal to ǫ∀1 and ǫ∀2 .
(2) The vectors ǫ∀1 and ǫ∀2 are normalised since
1 ∀c
�i + i ∀j c | i∀c + i ∀j c ∪c
�∀ǫ1 | ǫ∀1 ∪c =
2
1 1
= �i∀c | i∀c ∪c + i � i � ∀j c | ∀j c ∪c = 1.
2 2
1
�∀ǫ2 | ǫ∀2 ∪c = �i∀c − i ∀j c | i∀c − i ∀j c ∪c
2
1 1
= �i∀c | i∀c ∪c + i � i � ∀j c | ∀j c ∪c = 1.
2 2
(3) The vectors ǫ∀1 and ǫ∀2 are orthogonal since
1 ∀c
�∀ǫ1 | ǫ∀2 ∪c =�i + i ∀j c | i∀c − i ∀j c ∪c
2
1 1
= �i∀c | i∀c ∪c − i � i � ∀j c | ∀j c ∪c = 0.
2 2
∀ 3c , since
It follows that ǫ∀1 , ǫ∀2 and ǫ∀3 form an orthonormal basis in IE
we know that IE ∀ 3c is three-dimensional.

1 See Q6(4) and SQ6(4) or Q12(8) and SQ12(8).


June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Complex Vectors 57

SQ11(2)(c) The proof is the same as that in SQ11(1). Since ǫ∀1 , ǫ∀2
and ǫ∀3 form an orthonormal basis we can express any vector ζ∀ as
3
a linear combination of ǫ∀1 , ǫ∀2 and ǫ∀3 , i.e., we have ζ∀ = ∀ℓ .
ℓ=1 ζℓ ǫ
Taking the scalar product of ǫ∀ℓ and ζ∀ we get
3
 3

�∀ǫℓ | ζ∀ ∪c = �∀ǫℓ | ζℓ� ǫ∀ℓ� ∪c = ζℓ� �∀ǫℓ | ǫ∀ℓ� ∪c = ζℓ .
ℓ� =1 ℓ� =1

Q11(3) Show that Eqs. (11.18) and (11.19) hold in a general


complex orthonormal basis.

SQ11(3) The proof follows that of SQ11(1).


June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 12

N-Dimensional Complex Vectors

Q12(1) Prove Eq. (12.6).


SQ12(1) Evaluating �ζ∀ + η∀ | ζ∀ + η∀ ∪ and �ζ∀ + i η∀ | ζ∀ + i η∀ ∪ we get

�ζ∀ + η∀ | ζ∀ + η∀ ∪ = �ζ∀ | ζ∀ ∪ + �η∀ | η∀ ∪ + �ζ∀ | η∀ ∪ + �η∀ | ζ∀ ∪.


�ζ∀ + i η∀ | ζ∀ + i η∀ ∪ = �ζ∀ | ζ∀ ∪ + �η∀ | η∀ ∪ + i �ζ∀ | η∀ ∪ − i �η∀ | ζ∀ ∪.
� �ζ∀ + η∀ | ζ∀ + η∀ ∪ −i �ζ∀ + i η∀ | ζ∀ + i η∀ ∪
 
= (1 − i ) �ζ∀ | ζ∀ ∪ + �η∀ | η∀ ∪ + 2�ζ∀ | η∀ ∪.
Equation (12.6) follows immediately.
Q12(2) Show that
1
�ζ∀ | η∀ ∪ = ||ζ∀ + η∀ ||2 − ||ζ∀ − η∀ ||2
4
 
−i ||ζ∀ + i η∀ ||2 − ||ζ∀ − i η∀ ||2 .

SQ12(2) To prove the above equation we begin by evaluating


|| ζ∀ + η∀ ||2 − || ζ∀ − η∀ ||2 and || ζ∀ + i η∀ ||2 − || ζ∀ − i η∀ ||2 .
The calculations run as follows
|| ζ∀ + η∀ ||2 − || ζ∀ − η∀ ||2 = �ζ∀ + η∀ | ζ∀ + η∀ ∪ − �ζ∀ − η∀ | ζ∀ − η∀ ∪
= 2 �ζ∀ | η∀ ∪ + 2 �η∀ | ζ∀ ∪.
|| ζ∀ + i η∀ || − || ζ∀ − i η∀ || = 2i �ζ∀ | η∀ ∪ − 2i �η∀ | ζ∀ ∪.
2 2

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

60 N-Dimensional Complex Vectors

The desired result follows, i.e., we have


 
|| ζ∀ + η∀ ||2 − || ζ∀ − η∀ ||2 − i || ζ∀ + i η∀ ||2 + i || ζ∀ − i η∀ ||2 = 4�ζ∀ | η∀ ∪.

Q12(3) Prove the Schwarz inequality in VV∀ N :1


|�ζ∀ | η∀ ∪| ≤ ||ζ∀ || ||η∀ ||.

SQ12(3) The Schwarz inequality clearly holds if either ζ∀ or η∀ are


zero, or if ζ∀ = η.
∀ Indeed we have an equality in these cases. Generally,
when neither are zero we can define a new vector ξ∀ by
�η∀ | ζ∀ ∪ �η∀ | ζ∀ ∪
ξ∀ = ζ∀ − c η∀, c = = .
�η∀ | η∀ ∪ ||η∀ ||2
�ξ∀ | ξ∀ ∪ = �ζ∀ | ζ∀ ∪ + c � c�η∀ | η∀ ∪ − c�ζ∀ | η∀ ∪ − c � �η∀ | ζ∀ ∪
|�η∀ | ζ∀ ∪|2 �η∀ | ζ∀ ∪ �η∀ | ζ∀ ∪�
= ||ζ∀ ||2 + || ∀
η || 2
− �ζ∀ | η∀ ∪ − �η∀ | ζ∀ ∪
||η∀ ||4 ||η∀ ||2 ||η∀ ||2
|�η∀ | ζ∀ ∪|2 |�η∀ | ζ∀ ∪|2 |�η∀ | ζ∀ ∪|2
= ||ζ∀ ||2 + − −
||η∀ ||2 ||η∀ ||2 ||η∀ ||2
|�η∀ | ζ∀ ∪| 2
= ||ζ∀ ||2 − .
||η∀ ||2
The Schwarz inequality follows immediately, i.e.,
�ξ∀ | ξ∀ ∪ ⊕ 0 � || ζ∀ ||2 − |�ζ∀ | η∀ ∪|2 /∩|η∀ ||2 ⊕ 0
� ||ζ∀ || ||η∀ || ⊕ |�ζ∀ | η∀ ∪|.

Q12(4) Prove the following triangle inequalities in VV∀ N :2


||ζ∀ + η∀ || ≤ ||ζ∀ || + ||η∀ ||.
 
 ∀ 
 ||ζ || − ||η∀ ||  ≤ ||ζ∀ − η∀ ||.

SQ12(4) For the first triangle inequality we have

|| ζ∀ + η∀ ||2 = �ζ∀ + η∀ | ζ∀ + η∀ ∪
= �ζ∀ | ζ∀ ∪ + �η∀ | η∀ ∪ + �ζ∀ | η∀ ∪ + �η∀ | ζ∀ ∪
= ||ζ∀ ||2 + ||η∀ ||2 + 2 × real part of �η∀ | ζ∀ ∪
≤ ||ζ∀ ||2 + ||η∀ ||2 + 2 × |�η∀ | ζ∀ ∪|.
1 Roman p. 419. Prugovečki pp. 19–20.
2 Work out ||ζ∀ + η∀ ||2 = �ζ∀ + η∀ | ζ∀ + η∀ ∪ and ||ζ∀ − η∀ ||2 = �ζ∀ − η∀ | ζ∀ − η∀ ∪. Note that
�ζ∀ | η∀ ∪ + �η∀ | ζ∀ ∪ = 2× the real part of �ζ∀ | η∀ ∪ which is less than or equal to 2|�ζ∀ | η∀ ∪|.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

N-Dimensional Complex Vectors 61

The Schwarz inequality tells us that |�η∀ | ζ∀ ∪| ≤ ||ζ∀ || ||η∀ ||. So


 2
||ζ∀ + η∀ ||2 ≤ ||ζ∀ ||2 + ||η∀ ||2 + 2 ||ζ∀ || ||η∀ || = ||ζ∀ || + ||η∀ || ,
� ||ζ∀ + η∀ || ≤ ||ζ∀ || + ||η∀ ||.
For the second triangle inequality we have
|| ζ∀ − η∀ ||2 = �ζ∀ − η∀ | ζ∀ − η∀ ∪
= �ζ∀ | ζ∀ ∪ + �η∀ | η∀ ∪ − �ζ∀ | η∀ ∪ − �η∀ | ζ∀ ∪
= �ζ∀ | ζ∀ ∪ + �η∀ | η∀ ∪ − 2 × real part of �η∀ | ζ∀ ∪
⊕ ||ζ∀ ||2 + ||η∀ ||2 − 2 × |�η∀ | ζ∀ ∪|.
Since |�η∀ | ζ∀ ∪| ≤ ||ζ∀ || ||η∀ || we get
 2
||ζ∀ − η∀ ||2 ⊕ ||ζ∀ ||2 + ||η∀ ||2 − 2||ζ∀ || ||η∀ || = ||ζ∀|| − ||η∀ ||
 
� ||ζ∀ − η∀ || ⊕  ||ζ∀|| − ||η∀ || .

Q12(5) Show that the Frobenius expression in Eq. (12.23) satisfies


the properties CSP11.2.2(1), CSP11.2.2(2) and CSP11.2.2(3) of a
scalar product.
SQ12(5) The Frobenius expression in Eq. (12.23) is
∀ | N∀ ∪ := M11
�M � �
N11 + M12 �
N12 + M21 �
N21 + M22 N22 .
Property CSP11.2.2(1) is clearly satisfied. CSP11.2.2(2) is also
satisfied since
   
�M∀ | a N∀ + b L∀ ∪ = M11

aN11 + bL11 + M12 �
aN12 + bL12

  �
 
+ M21 aN21 + bL21 + M22 aN22 + bL22
= a�M ∀ | N∀ ∪ + b�M
∀ | L∀ ∪.

For property CPS11.2.2(3) we first observe that


∀ |M
�M ∀ ∪ = M11
� �
M11 + M12 �
M12 + M21 �
M21 + M22 M22 ⊕ 0,
since each term is non-negative. Also we have
� � � �
M11 M11 + M12 M12 + M21 M21 + M22 M22 = 0,
∀.
∀ =0
which implies each term must vanish, i.e., M
Q12(6) Show that the integral expression in Eq. (12.25) satisfies
the properties CSP11.2.2(1), CSP11.2.2(2) and CSP11.2.2(3) of a
scalar product.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

62 N-Dimensional Complex Vectors

SQ12(6) The integral expression clearly satisfies CSP11.2.2(1).


The linear property of integrals means that CSP11.2.2(2) is also
satisfied, i.e., we have
 L
 
ϕ � (x) aφ(x) + bψ(x) dx
0
 L  L
=a ϕ � (x)φ(x) dx + b ϕ � (x)bψ(x) dx.
0 0

For continuous functions we have


 L
φ � (x)φ(x) dx = 0 � φ(x) = 0,
0
which is CSP11.2.2(3).
Q12(7) Verify that the vectors ϕ∀ℓ corresponding to the functions
ϕℓ (x) in Eq. (12.29) are orthonormal.
SQ12(7) For ℓ = ℓ� the integrand becomes 1/L. Integrating over 0
and L the integral gives the value 1. For ℓ = � ℓ� we have
 L  L
1 �
ϕℓ� (x)ϕℓ� (x) dx = e2πi (ℓ −ℓ )x/Ldx
0 L 0
L
1 L 
2πi (ℓ� −ℓ )x/L
= �
e 
L 2πi (ℓ − ℓ ) 0
1
= (1 − 1) = 0,
2πi (ℓ� − ℓ )
bearing in mind that exp(2π ni ) = 1 for any positive or negative
integer n. Hence their corresponding vectors ϕ∀ℓ are orthonormal.
Q12(8) Show that two orthogonal vectors are linearly indepen­
dent.
SQ12(8) The proof follows that of SQ(6.4). Let ζ∀1 and ζ∀2 be two
orthogonal vectors and c1 and c2 be two complex numbers such that
∀ Taking scalar product of this equation with ζ∀1 and
c1 ζ∀1 + c2 ζ∀2 = 0.
ζ 2 in turn we get, on account of �ζ∀1 | ζ∀2 ∪ = �ζ∀2 | ζ∀1 ∪ = 0,

c1 �ζ∀1 | ζ∀1 ∪ = 0 � c1 = 0 and c2 �ζ∀2 | ζ∀2 ∪ = 0 � c2 = 0.
The two vectors are therefore linearly independent.
Q12(9) Show that the three Pauli matrices σx , σ y and σz in Eq. (7.9)
together with the 2 × 2 identity matrix form a basis for the vector
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

N-Dimensional Complex Vectors 63

space of 2 × 2 complex matrices with Frobenious scalar product. Is


this an orthonormal basis?
SQ12(9) The set of 2 × 2 matrices constitutes a four-dimensional
scalar product space with Frobenius scalar product. One can verify
that the Pauli matrices and the 2 × 2 identity matrix I 2×2 form an
orthogonal set with respect to the Frobenius scalar product, e.g.,
�σx | σ y ∪ = 0 × 0 + 1 × (−i ) + 1 × i + 0 × 0 = 0.
�σ y | σz ∪ = 0 × 1 + (−i )� × 0 + i � × 0 + 0 × 1 = 0.
�σz | σx ∪ = 1 × 0 + 0 × 1 + 0 × 1 + (−1) × 0 = 0.
�σx | I 2×2 ∪ = 0 × 1 + 1 × 0 + 1 × 0 + 0 × 1 = 0.
�σ y | I 2×2 ∪ = 0 × 1 + (−i )� × 0 + i � × 0 + 0 × 1 = 0.
�σz | I 2×2 ∪ = 1 × 1 + 0 × 0 + 0 × 0 + (−1) × 1 = 0.
It follows that the Pauli matrices and the 2 × 2 identity matrix I 2×2
form a linearly independent set, i.e., they form a basis for the four­
dimensional space of the set of 2 × 2 matrices. The basis is not
orthonormal since Pauli matrices are not normalised in Frobenius
scalar product.

Q12(10) Show that the mapping in Eq. (12.32) is unitary.


N
SQ12(10) The space C∀ is spanned by e∀ℓc . It follows that f in Eq.
N
(12.32) is a mapping of VV∀ N onto C∀ . The mapping is also one-to­
one since the coefficients cℓ associate every vector φ∀ in VV∀ N a unique
N
vector f (φ∀ ) ⇒ C∀ . In other words f establishes an isomorphism
N
between VV∀ N and C∀ .
The mapping also preserves the scalar product since for any
 
given φ∀ = ℓ cℓ ϕ∀ℓ and φ∀ � = ℓ cℓ� ϕ∀ℓ in VV∀ N we have
N

�φ∀ | φ∀ � ∪ = cℓ� c � ℓ and
ℓ=1
  N

� f (φ∀ ) | f (φ∀ � ) ∪ = � cℓ e∀ℓc | cℓ� e∀ℓc� ∪ = cℓ� c � ℓ .
ℓ ℓ� ℓ=1

The mapping is therefore unitary.


June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 13

Operators on N-Dimensional Complex


Vectors

Q13(1) Prove Eqs. (13.3) and (13.4).


SQ13(1) The proofs are similar to that given in SQ12(8). We have,
for all ζ∀, η∀ ⇒ VV∀ N ,
 
�ζ∀ | c1 A1 + c2 A2 η∀ ∪ = �ζ∀ | c1 A1 η∀ ∪ + �ζ∀ | c2 A2 η∀ ∪
† †
= �c1� A1 ζ∀ | η∀ ∪ + �c2� A2 ζ∀ | η∀ ∪
 † †
= � c � A + c � A ζ∀ | η∀ ∪
1 1 2 2
 † † †
� c1 A1 + c2 A2 = c1� A1 + c2� A2 .
   
�ζ∀ | A B  η∀ ∪ = �A† ζ∀ | B
 η∀ ∪ = �ζ∀ | A B  † A† ζ∀ | η∀ ∪
 η∀ ∪ = �B
 † †
=� B  A ζ∀ | η∀ ∪
 †
� A B  =B  † A† .

Q13(2) Verify Eq. (13.7).


   
SQ13(2) Evaluating �ζ∀ + η∀ | A ζ∀ + η∀ ∪ and �ζ∀ + i η∀ | A ζ∀ + i η∀ ∪
we get
 
�ζ∀ + η∀ | A ζ∀ + η∀ ∪ = �ζ∀ | Aζ∀ ∪ + �η∀ | Aη∀ ∪ + �ζ∀ | Aη∀ ∪ + �η∀ | Aζ∀ ∪.
 
�ζ∀ + i η∀ | A ζ∀ + i η∀ ∪ = �ζ∀ | Aζ∀ ∪ + �η∀ | Aη∀ ∪ + i �ζ∀ | Aη∀ ∪ − i �η∀ | Aζ∀ ∪.

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

66 Operators on N-Dimensional Complex Vectors

From the above equations we get


   
�ζ∀ + η∀ | A ζ∀ + η∀ ∪ − i �ζ∀ + i η∀ | A ζ∀ + i η∀ ∪
 
= (1 − i ) �ζ∀ | Aζ∀ ∪ + �η∀ | Aη∀ ∪ + 2�ζ∀ | Aη∀ ∪.
Equation (13.7) follows immediately.
Q13(3) Consider the x y-plane as a two-dimensional real vector
space in its own right. In basis {i∀, ∀j } an arbitrary vector v∀ on the
x y-plane has the matrix representation
vx
C v = .
vy
In basis {i∀, ∀j } the operator R p (π/2) which rotates any vector u∀ on
the x y-plane about the origin by an angle of π/2 has the following
matrix representation in accordance with Eq. (7.136):
0 −1
R p (π/2) = .
1 0
Show that �∀v | R p (π/2)v∀ ∪ = 0 and explain why this result does not
satisfy Theorem 13.1(1).
SQ13(3) We have, by Eqs. (7.136), (7.137) and (7.138),

v | R p (π/2)v∀ ∪ = �C v | Rp (π/2)C v ∪ = C v † · Rp (π/2)C v


�∀
  0 −1 vx
= v x� v �y
1 0 vy
  −v y
= vx v y = 0 since v x , v y are real.
vx
v | Rp (π/2)v∀ ∪ =
This result does not satisfy Theorem 13.1(1), i.e., �∀

0 for all v∀ in the x-y plane does not imply R p (π/2) = 0. ˆ Theorem
13.1(1) is proved using Eq. (13.7) which applies to complex vector
spaces. Hence the theorem does not apply to the x-y plane which is
a real vector space unless the operator is selfadjoint (see Theorem
9.4.1(1)). As pointed out in C9.4.1(6) the operator Rp (π/2) is not
selfadjoint.
Q13(4) Prove that every vector ζ∀ in VV∀ N can be decomposed
uniquely as a sum of a vector lying in a given subspace S∀ and another
one lying in its orthogonal complement S∀� as shown in Eq. (13.15).
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Operators on N-Dimensional Complex Vectors 67

SQ13(4)

(1) For uniqueness let us suppose there are two such decomposi­
tions
ζ∀ = ζ∀ S∀ + ζ∀ S∀� and ζ∀ = ζ∀ �S∀ + ζ∀ �S∀� .
Then we have
   
∀ = ζ∀ ∀ − ζ∀ �∀ + ζ∀ ∀� − ζ∀ �∀� .
0 S S S S

The two vectors (ζ∀ S∀ − ζ∀ �S∀) ⇒ S∀ and (ζ∀ S∀� − ζ∀ �S∀� ) ⇒ S∀� are
orthogonal and hence linearly independent, i.e., their sum
cannot be zero. It follows that the two vectors must be zero
separately, i.e., we must have
ζ∀ S∀ = ζ∀ �S∀ and ζ∀ S∀� = ζ∀ �S∀� .
This proves the uniqueness of the decomposition.
(2) We need to show the existence of such a decomposition. Since
S∀ is a subspace there exists an orthonormal basis { ε∀ j , j =
1, 2, · · · , M ≤ N } for S∀. Let
M

ζ∀ S∀ = ε j | ζ∀ ∪ ε∀ j
�∀ and ζ∀ S∀� = ζ∀ − ζ∀ S∀.
j =1

Clearly ζ∀ S∀ lies in the subspace S∀. Also ζ∀ S∀� lies in S∀� since
M

ε j | ζ∀ S∀� ∪ = �∀ǫ j | ζ∀ ∪ − �∀
�∀ εj | ε j � | ζ∀ ∪ ε∀ j � ∪
�∀
j � =1
M

ε j | ζ∀ ∪ −
= �∀ ε j � | ζ∀ ∪ �∀
�∀ ε j | ε∀ j � ∪
j � =1
M

ε j | ζ∀ ∪ −
= �∀ ε j � | ζ∀ ∪ δ j j � = 0.
�∀
j � =1

We have a desired decomposition, i.e., we have


ζ∀ S∀ ⇒ S∀, ζ∀ S∀� ⇒ S∀� and ζ∀ S∀ + ζ∀ S∀� = ζ∀.

Q13(5) Show that Eq. (13.20) is equivalent to Eq. (13.25) or


Eq. (13.26) in defining the order relation of projectors.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

68 Operators on N-Dimensional Complex Vectors

SQ13(5)1

(1) Proof of Eq. (13.20) implying Eq. (13.25):


N
Let us start with the obvious result P S∀1 ζ∀ ⇒ S∀1 for all ξ∀ ⇒ VV∀ .
If S∀1 ≥ S∀2 we have P S∀1 ζ∀ ⇒ S∀2 also. It follows that the projector
P S∀2 will have no effect on P S∀1 ζ∀, i.e.,

    N
P S∀2 P S∀1 ζ∀ = P S∀2 P S∀1 ζ∀ = P S∀1 ζ∀ ∞ζ∀ ⇒ VV∀ .

This means P S∀2 PS∀1 = P S∀1 . Secondly, if S∀1 ≥ S∀2 , projecting onto
S∀2 and then projecting onto S∀1 is the same as projecting onto
S∀1 alone. We can prove this by taking the adjoint of the equality
PS∀2 PS∀1 = P S∀1 , i.e., we have

†  † † †
PS∀1 = P S∀ = P S∀2 PS∀1 = PS∀ P S∀ = P S∀1 P S∀2 .
1 1 2

The result shows that Eq. (13.20) implies Eq. (13.25).


(2) Proof of Eq. (13.25) implying Eq. (13.26):
N
Given P S∀1 = PS∀1 PS∀2 we have PS∀1 η∀ = PS∀1 PS∀2 η∀ for all η∀ ⇒ VV∀ .
Using Eq. (8.15) or Eq. (17.7) we immediately get2
 
||P S∀1 η∀ || = ||P S∀1 PS∀2 η∀ || ≤ ||PS∀1 || ||P S∀2 η∀ || = ||P S∀2 η∀ ||.

We have used the result that the norm of a projector is equal to


1, as stated in Theorem 13.2.2(1).
(3) Proof of Eq. (13.26) implying Eq. (13.20):
Let η∀ = η∀ ∀ + η∀ ∀� = P ∀ η∀ + P ∀� η∀. Then we have
S1 S1 S1 S1

�η∀ | η∀ ∪ = ||η∀ ||2 = ||P S∀1 η∀ ||2 + ||PS∀� η∀ ||2 . (�)


1

Equation (�) above implies:


(a) If η∀ ⇒ S∀1 , i.e., PS∀� η∀ = 0,
∀ then ||η∀ || = || P ∀ η∀ ||.
S1
1

(b) If ||η∀ || = ||P S∀1 η∀ ||, then ||P S∀� η∀ || = 0 which implies P S∀� η∀ =
1 1

0.
1 Prugovečki p. 202. Roman Vol. 2 p. 538, p. 569.
2 See P.13.1(1) on the definition of norm of an operator on VV∀ N .
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Operators on N-Dimensional Complex Vectors 69

We can conclude that


η∀ ⇒ S∀1 if and only if ||η∀ || = ||PS∀1 η∀ ||.
Now assume Eq. (13.26) holds. Then we have, for any η∀ ⇒ S∀1 ,
||η∀ || = ||P ∀ η∀ || ≤ ||P ∀ η∀ ||.
S1 S2

||P S∀2 η∀ || ≤ ||P S∀2 || ||η∀ || = ||η∀ ||.


The second inequality arises from Eq. (8.15) or Eq. (17.7). We
can also arrive at this result from the fact that the length of any
projection ||P S∀2 η∀ || must be less or equal to the length ||η∀ || of
the original vector η∀. It follows that
||η∀ || ≤ ||P ∀ η∀ || ≤ ||η∀ || � ||P ∀ η∀ || = ||η∀ ||.
S2 S2

This in turn implies η∀ ⇒ S∀2 , i.e., any η∀ ⇒ S∀1 must also be in S∀2 . In
other words we have S∀1 ≥ S∀2 .
We have therefore showed that
Eq. (13.20) � Eq. (13.25) � Eq. (13.26) � Eq. (13.20).

Q13(6) Prove Eqs. (13.22) and (13.23).


SQ13(6)
  †   
(1) Equation (13.22), i.e., ζ∀ η∀  = η∀ ζ∀ , is proved by the
following scalar product calculation for any ξ∀ , ǫ∀ ⇒ VV∀ N :
  
�ξ∀ | ζ∀ η∀  ǫ∀ ∪ = �ξ∀ | �η∀ | ǫ∀ ∪ ζ∀ ∪ = �η∀ | ǫ∀ ∪ �ξ∀ | ζ∀ ∪,
  
� η∀ ζ∀  ξ∀ | ǫ∀ ∪ = ��ζ∀ | ξ∀ ∪ η∀ | ǫ∀ ∪ = �ζ∀ | ξ∀ ∪� �η∀ | ǫ∀ ∪
= �ξ∀ | ζ∀ ∪�η∀ | ǫ∀ ∪
     
� � η∀ ζ∀  ξ∀ | ǫ∀ ∪ = �ξ∀ | ζ∀ η∀  ǫ∀ ∪.

(2) Equation (13.23), i.e.


       
ζ∀1 η∀1  ζ∀2 η∀2  = �η∀1 | ζ∀2 ∪ ζ∀1 η∀2 ,

is proved in two steps. First we have, for any ξ∀ ⇒ VV∀ N ,


       
ζ∀1 η∀1  ζ∀2 η∀2  ξ∀ = ζ∀1 η∀1  �η∀2 | ξ∀ ∪ ζ∀2
  
= �η∀2 | ξ∀ ∪ ζ∀1 η∀1  ζ∀2
= �η∀2 | ξ∀ ∪�η∀1 | ζ∀2 ∪ ζ∀1 .
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

70 Operators on N-Dimensional Complex Vectors

Secondly we have
    
�η∀1 | ζ∀2 ∪ ζ∀1 η∀2  ξ∀ = �η∀1 | ζ∀2 ∪ �η∀2 | ξ∀ ∪ ζ∀1 .

The required result follows immediately.

Q13(7) Prove statement (3) of Theorem 13.2.2(1).


SQ13(7) Let P S∀1 and P S∀2 be two projectors onto the subspaces S∀1
and S∀2 respectively. Let ζ∀ be an arbitrary vector, and let P S∀1 ζ∀ = ζ∀1
and P S∀2 ζ∀ = ζ∀2 . By Definition 9.3.2(2) two projectors are orthogonal
if the subspaces onto which they project are orthogonal.

(1) If P S∀1 and P S∀2 are orthogonal then ζ∀2 has no projection onto ζ∀1
since �ζ∀1 | ζ∀2 ∪ = 0.
   

P S∀1 P S∀2 ζ∀ = P S∀1 P S∀2 ζ∀ = P S∀1 ζ∀2 = 0.

(2) If P S∀1 P S∀2 = 


0 then

∀ � ζ∀2 has no projection onto S∀1 .


P S∀1 P S∀2 ζ∀ = P S∀1 ζ∀2 = 0

Since P S∀1 P S∀2 = 


0 � P S∀2 P S∀1 = 
0 we can similarly deduce that
ζ 1 has no projection onto S 2 . In other words S∀1 and S∀2 are
∀ ∀
orthogonal.

Q13(8) Prove the two expressions in Eq. (13.29).


SQ13(8) We have, for any vector η∀,
N
 N

η∀ = cℓ ε∀ℓ , cℓ = �∀
ε ℓ | η∀ ∪ � �η∀ | η∀ ∪ = εℓ | η∀ ∪|2 .
|�∀
ℓ=1 ℓ=1

We also have
N
 N
 N

�η∀ | P ε∀ℓ η∀ ∪ = �η∀ | �∀
ε ℓ | η∀ ∪ ε∀ℓ ∪ = ε ℓ | η∀ ∪|2 .
| �∀
ℓ=1 ℓ=1 ℓ=1

It follows that
 N N
 
�η∀ | P ε∀ℓ η∀ ∪ = �η∀ | P ε∀ℓ η∀ ∪ = �η∀ | η∀ ∪.
ℓ=1 ℓ=1
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Operators on N-Dimensional Complex Vectors 71

We can deduce two results:


(1) For any unit vector η∀ we have
 N

�η∀ | P ε∀ℓ η∀ ∪ = 1.
ℓ=1
(2) We also have
N

P ε∀ℓ = II on account of Eq. (13.10).
ℓ=1
This is a major result stated in Eq. (13.28).
Next, for any unit vector ζ∀ we have by Schwarz inequality3
|�ζ∀ | P ε∀ℓ ζ∀ ∪| ≤ || ζ∀ || || P ε∀ℓ ζ∀ || = || P ε∀ℓ ζ∀ || ≤ || P ε∀ℓ || = 1.
We also have
�ζ∀ | P ε∀ℓ ζ∀ ∪ = �ζ∀ | P 2ε∀ℓ ζ∀ ∪ = |�P ε∀ℓ ζ∀ | P ε∀ℓ ζ∀ ∪ = ||P ε∀ℓ ζ∀ ||2 ⊕ 0.
The required result follows, i.e.,
�ζ∀ | Pε∀ℓ ζ∀ ∪ ⇒ [0, 1].

Q13(9) Prove Theorem 13.3.1(1) on selfadjoint operators.


SQ13(9) If A is selfadjoint, then, using the properties of scalar
product, we get
�ζ∀ | Aζ∀ ∪� = �Aζ∀ | ζ∀ ∪ = �ζ∀ | Aζ∀ ∪ � �ζ∀ | Aζ∀ ∪ ⇒ IR .
For the converse we first apply Eq. (13.7) we get
1 ∀
�ζ∀ | A η∀ ∪ = �ζ + η∀ | A(ζ∀ + η∀ ) ∪ − i �ζ∀ + i η∀ | A(ζ∀ + i η∀ ) ∪
2
 
+(i − 1) �ζ∀ | A ζ∀ ∪ + �η∀ | A η∀ ∪ ,
1
�η∀ | A ζ∀ ∪ = �η∀ + ζ∀ | A(η∀ + ζ∀ ) ∪ − i �η∀ + i ζ∀ | A(η∀ + i ζ∀ ) ∪
2
 
+(i − 1) �η∀ | A η∀ ∪ + �ζ∀ | A ζ∀ ∪ .

Now suppose �ξ∀ | A ξ∀ ∪ ⇒ IR for any ξ∀ , e.g.,


�ζ∀ + i η∀ | A(ζ∀ + i η∀ ) ∪ ⇒ IR , �η∀ + i ζ∀ | A(η∀ + i ζ∀ ) ∪ ⇒ IR .
3 See Q12(3) and its solutions. We also use Eq. (8.15) or Eq. (17.7) to get || P�ε∀ℓ ζ∀ || ≤
|| P�ε∀ℓ || || ζ∀ || = || P�ǫ∀ℓ ||.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

72 Operators on N-Dimensional Complex Vectors

Then
�A ζ∀ | η∀ ∪ = �η∀ | A ζ∀ ∪�
1
= �η∀ + ζ∀ | A(η∀ + ζ∀ ) ∪ + i �η∀ + i ζ∀ | A(η∀ + i ζ∀ ) ∪
2
 
−(i + 1) �η∀ | A η∀ ∪ + �ζ∀ | A ζ∀ ∪ .

By expanding the scalar product expressions on the right-hand sides


of the equations for �ζ∀ | A η∀ ∪ and �A ζ∀ | η∀ ∪ we get, for all ζ∀, η∀,
�A ζ∀ | η∀ ∪ = �ζ∀ | A η∀ ∪ � A† = A.

Q13(10) Prove Spectral Theorems 13.3.2(1) and 13.3.2(2).


SQ13(10) We can prove Spectral Theorem 13.3.2(1) in the same
way Theorem 9.4.5(1) is proved in the book. We know that A
possesses a complete orthonormal set of eigenvectors ε∀ℓ and a
corresponding complete orthogonal family of eigenprojectors P ε∀ℓ
N
corresponding to a set of eigenvalues aℓ . Then for any vector ζ∀ ⇒ VV∀
3
we have ζ∀ = ℓ=1 ζℓ ε∀ℓ , ζℓ = �∀ε ℓ | ζ∀ ∪ and
3
 3

Aζ∀ = ζℓ Aε∀ℓ = aℓ ζℓ ε∀ℓ .
ℓ=1 ℓ=1

Construct the operator


3

A� := aℓ Pε∀ℓ .
ℓ=1

Then
3
 3
 3

A� ζ∀ = aℓ P ε∀ℓ ζ∀ = aℓ �∀
ε ℓ | v∀ ∪∀
εℓ = aℓ ζℓ ε∀ℓ .
ℓ=1 ℓ=1 ℓ=1

We see that Aζ∀ = A� ζ∀ for every vector ζ∀. It follows that A� = A, i.e.,
the operator A has a spectral decomposition.
The spectral decomposition of A in Theorem 13.3.2(2) is just
a restatement of that of Theorem 13.3.2(1) by grouping all the
projectors of degenerate eigenvalues.
Q13(11) Show that the eigenprojectors of a selfadjoint operator
commutes with each other and that a selfadjoint operator commutes
with its eigenprojectors.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Operators on N-Dimensional Complex Vectors 73

SQ13(11) The eigenprojectors P ε∀ℓ of a selfadjoint operator A


in Theorem 13.3.2(1) are mutually orthogonal, i.e., Pε∀ℓ P ε∀ℓ� =

0 = P ε∀ℓ� P ε∀ℓ for ℓ = � ℓ� , and hence they mutually commute, i.e.,
[Pε∀ℓ , P ε∀ℓ� ] = 
0. Substituting the spectral decomposition
3

A = aℓ P ε∀ℓ .
ℓ=1

into the commutator [A, P ε∀ℓ� ] we get


3

[A, Pε∀ℓ� ] = 
aℓ [ P ε∀ℓ , P ε∀ℓ� ] = 0.
ℓ=1

The same proof applies to the eigenprojectors P Â (am ) of a


selfadjoint operator A in Theorem 13.3.2(2).
Q13(12) For a selfadjoint operator A with its spectral decomposi­
tion given by Eq. (13.36) show that
A P Â (am ) = am P Â (am ).

SQ13(12) From the orthogonal nature of eigenprojectors we get


 
A P Â (am ) = am� P Â (am� ) P Â (am )
m�

= am� P Â (am� )P Â (am )
m�

= am� P Â (am� )δm� m = am P Â (am ),
m�

which is the desired result.


Q13(13) Prove Eq. (13.50).
SQ13(13) To prove Eq. (13.50) for the expression for the adjoint
we have, for any ζ∀ and η∀
 
�ξ∀ | ei  η∀ ∪ = �ξ∀ | ei am P  (am ) η∀ ∪ = �ξ∀ | ei am P  (am ) η∀ ∪
m m
 
= �e −i am
P Â (am ) ξ∀ | η∀ ∪ = � e−i am P Â (am ) ξ∀ | η∀ ∪
m m
 †
= �e −i Â
ξ∀ | η∀ ∪ � e i Â
=e −i Â
.
We have used Eqs. (13.48) and (13.49).
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

74 Operators on N-Dimensional Complex Vectors

For the inverse we have, again using Eqs. (13.48) and (13.49),
   
ei am P Â (am ) e−i am� P Â (am� )
m m�

= e i am −i am�
e P (am� )P Â (am )
 Â
m, m�

= ei am e−i am� P Â (am� )δm� m
m, m�

= P Â (am ) = II
m
 −1 
� ei am P Â (am ) = e−i am� P Â (am� )
m m�
 
i  −1
 †  −1
� e = e−i  � ei  = ei  = e−i  .

Q13(14) Prove Theorem 13.3.4(1).


SQ13(14) Theorem 13.3.4(1) can be proved in two steps.
(1) Let the spectral decompositions of two selfadjoint operators A
 
 be A =  Â   B̂
and B m am P (am ) and B = m� bm� P (bm� ). Then

A B
= am bm� P Â (am )P B̂ (bm� ),
m, m�

 A =
B am bm� P B̂ (bm� )P Â (am ),
m, m�

and
 
[A, B
] = am bm� P Â (am )P B̂ (bm� ) − am bm� P B̂ (bm� )P Â (am )
m, m� m, m�

=  Â
am bm� [P (am ), P (bm� )].  B̂
m, m�

It follows that
[P Â (am ), P B̂ (bm� )] =  ] = 
0 ∞ m, m� � [A, B 0.

(2) To prove the converse let us assume that [ A, B ]= 0. From the
  Â  Â
result A P (am ) = am P (am ) in Eq. (13.118) proved in SQ13(12)
we deduce that
     
A B P Â (am ) ζ∀ = B  A P Â (am ) ζ∀ = am B P Â (am ) ζ∀ .
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Operators on N-Dimensional Complex Vectors 75

This means that B  P Â (am )ζ∀ belongs to the eigensubspace of A


corresponding to the eigenvalue am . In other words we have
   
P Â (am ) B  P Â (am )ζ∀ = B  P Â (am )ζ∀ ∞ζ∀

   
�  P Â (am ) ζ∀ = B
P Â (am )B  P Â (am ) ζ∀ ∞ζ∀

� P Â (am )B
 P Â (am ) = B
 P Â (am ).

Taking the adjoint of each side of the equation we get


 †  †
 P Â (am ) = B
P Â (am )B  P Â (am ) .

� P Â (am )B
 P Â (am ) = P Â (am )B
.
 P Â (am ) = P Â (am )B
We get B  , i.e.,
 , P Â (am )] = 
[B 0 and similarly [A, P B̂ (bm )] = 
0.
What we have shown is that if a selfadjoint operator A
 then
commutes with another selfadjoint operator B
(1) A commutes with every eigenprojector P B̂ (bm ) of B
 , and
(2) B commutes with every eigenprojector P (am ) of A.
  Â

It follows that P Â (am ) commutes with P B̂ (bm� ) since P Â (am )


commutes with B  , i.e., P Â (am ) will commute with all the
eigenprojectors of B 
 . Similarly the eigenprojectors P B̂ (bm ) of B
will commute with all the eigenprojectors P Â (am� ) of A.

Q13(15) Prove Eqs. (13.53) and (13.54).


SQ13(15)
(1) To prove Eq. (13.53) we start with
 
P Â (am ) = II and P B̂ (bm� ) = II .
m m�

Multiplying the above equations we immediately get



� P Â (am )P B̂ (bm� ) = II .
m, m�
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

76 Operators on N-Dimensional Complex Vectors

(2) To prove Eq. (13.54) we can use Eq. (13.118) proved in


SQ13(12), i.e., A P Â (am ) = am P Â (am ), we get

Aη∀mm� = A P Â (am )P B̂ (bm� )η∀ = am P Â (am )P B̂ (bm� )η∀


= am η∀mm� .
A similar analysis shows that
 η∀mm� = bm� η∀mm� .
B

Q13(16) Prove properties P13.4.1(1) to P13.4.1(4) of a unitary


operator listed right after Theorem 13.4.1(1).

SQ13(16)

(1) P13.4.1(1) is true on account of Theorem 13.4.1(1). Explicitly


we have, using Eq. (13.64),
�U ζ∀ | U η∀ ∪ = �U † U ζ∀ | η∀ ∪ = �ζ∀ | η∀ ∪.

(2) To prove P13.4.1(2) we have, for every vector ζ∀,


�U † ζ∀ | U † ζ∀ ∪ = �ζ∀ | U U † ζ∀ ∪ = �ζ∀ | ζ∀ ∪.
Hence U † is unitary. We have used Eq. (13.64) in the above
calculation.
(3) To prove P13.4.1(3) we consider the eigenvalue equation of a
unitary operator U , i.e., U η∀ = λη∀, where η∀ is an eigenvector and
λ is the associated eigenvalue. Then
�U η∀ | U η∀ ∪ = �λη∀ | λη∀ ∪ = λ� λ�η∀ | η∀ ∪.
Since a unitary operator preserves the scalar product we have
�U η∀ | U η∀ ∪ = �η∀ | η∀ ∪.
It follows that λ� λ = |λ|2 = 1 which implies that λ is of the form
ei a , a ⇒ IR.
(4) To prove P13.4.1(4) we start with the matrix elements MU� mn of
the matrix representation M U� of a given unitary operator U in a
ε ℓ }, i.e.,
given orthonormal basis {∀
εm | U ε∀n ∪.
MU� mn = �∀
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Operators on N-Dimensional Complex Vectors 77


By definition the adjoint matrix M U� has elements

εℓ | U ε∀n ∪� = �U ε∀n | ε∀ℓ ∪.
MU� nℓ = MU�� ℓn = �∀


The product matrix of M U� and M U� has elements
 †

MU�mn MU� = ε m | U ε∀n ∪�U ε∀n | ε∀ℓ ∪
�∀
nℓ
n n

= εn | U † ε∀ℓ ∪
�U † ε∀m | ε∀n ∪�∀
n

= �U † ε∀m | U † ε∀ℓ ∪ = �∀


εm | U U † ε∀ℓ ∪ = δmℓ

� M U� M U� = I .

Hence matrix is unitary. We have used Eqs. (13.35) and (13.64).

Q13(17) Show that the operator in Eq. (13.69) is unitary.

SQ13(17) For the operator in Eq. (13.69) we have, using


Eqs. (13.48), (13.50) and the orthogonal property of the eigenpro­
jectors,
N

U † = e−i aℓ Pε∀ℓ
ℓ=1
N
 N

� U † U = e−i aℓ ei am Pε∀ℓ Pε∀m = P ε∀ℓ = II .
ℓ, m=1 ℓ

Hence U is invertible (see Theorem 8.2.2(2) and P13.1(3)) with its


inverse equal to its adjoint, i.e., U † = U −1 . It follows that U is unitary
by Theorem 13.4.1(1).

Q13(18) Show that the operator U in Eq. (13.70) is unitary.

SQ13(18) For the operator in Eq. (13.70) we can write down its
adjoint, using Eq. (13.22), i.e.,
N
   � 
U † =  ε∀ℓ ε∀ ℓ .
ℓ=1
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

78 Operators on N-Dimensional Complex Vectors

Then

N
    N
 �  
U † U =  ε∀ℓ ε∀ �ℓ   ε∀ k ε∀k 
ℓ=1 k=1
N 
      N
   
=  ε∀ℓ ε∀ �ℓ   ε∀ �k ε∀k  = ε �ℓ | ε∀ �k ∪  ε∀ℓ ε∀k 
�∀
ℓ, k=1 ℓ, k=1
N
   
=  ε∀k ε∀k  since �∀
ε �ℓ | ε∀ �k ∪ = δkℓ
k=1

= II .
Hence U is invertible (see SQ13(17)) with its adjoint equal to its
inverse, i.e., U † = U −1 . Hence U is unitary. We have used Eq. (13.23).
Q13(19) Prove Eq. (13.74) on the preservation of commutation
relations.
SQ13(19) To prove the preservation of commutation relations in
Eq. (13.75) we let A� = U AU † , B  � = U B
U † and C� = U CU † . Then
     
 � ] = A� B
[ A� , B � − B � A� = U AU † U B U † − U B U † U AU †
 
= U A BU † − U B  AU † = U A B−B  A U † = U CU †
= C� .
Q13(20) Prove Eq. (13.77) on the preservation of the quadratic
form.
SQ13(20) To prove the preservation of quadratic form in Eq.
(13.75) we let A� = U AU † and ζ∀� = U ζ∀. Then
 
�ζ∀� | A� ζ∀� ∪ = �U ζ∀ | U AU † U ζ∀ ∪ = �U ζ∀ | U Aζ∀ ∪ = �ζ∀ | Aζ∀ ∪.

Q13(21)4 Let B  a selfadjoint operator and let B � be the unitary


transform of B  generated by a unitary operator U . Show that B
 and
 � possess the same set of eigenvalues and that their corresponding
B
eigenvectors are unitary transforms of each other.5
4 This question is on the important property stated in P13.4.2(6).
5 First
show that B�� possesses the same eigenvalues as B � and that the eigenvectors
�� are the corresponding unitary transforms of the eigenvectors of B
B �. Then show
� possesses the same eigenvalues as B
that B �� and that the eigenvectors B� are the
�� .
corresponding unitary transforms of the eigenvectors of B
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Operators on N-Dimensional Complex Vectors 79

SQ13(21) Let η∀ be an eigenvector of B  corresponding to the


eigenvalue b, i.e., B  η∀ = bη∀. Let η∀ and B
�  be the unitary transforms

 generated by the unitary operator U . Then


of η∀ and B
 
 � η∀ � = U B
B U † U η∀ = U B η∀ = b U η∀ = b η∀ � ,
showing that an eigenvalue of B  is also an eigenvalue of B  � with the
corresponding unitary transform η∀ � as eigenvector.
Conversely the operator B  can be regarded as the unitary
transform of B generated by the unitary operator U � = U † . It
 �

follows that an eigenvalue b� of B  � corresponding to an eigenvector


η∀ � must also be an eigenvalue of B  with the corresponding unitary
 � �
transform U η∀ as eigenvector.
The conclusion is that B  and B  � possess the same set of
eigenvalues and that their corresponding eigenvectors are unitary
transforms of each other.

Q13(22) Let U (t) be a continuous one-parameter group of unitary


operators. Let A be the generator of U (t) in accordance with
Theorem 13.4.3(2) of Stone. Let ξ∀ (0) be an eigenvector of A
corresponding to a non-denegerate eigenvalue a. Show that
− i at
U (t)ξ∀ (0) = e ¯ ξ∀ (0),
and that
ξ∀ (t) = U (t)ξ∀ (0),
is a solution of Eq. (13.100).

SQ13(22) From Theorem 13.4.3(2) of Stone we know that



− i At i
U (t) = e ¯ , –i = ,

showing that U is an exponential function of A. This function can be
expressed in the form of Eq. (13.48), i.e.,
N
− ia t
U (t) = e ¯ ℓ P ε∀ℓ .
ℓ=1
Now choose a1 = a and ε∀1 = ξ∀ (0) in the above equation.6 Then we
∀ for all ℓ =
have Pε∀1 ξ∀ (0) = ξ∀ (0) and Pε∀ℓ ξ∀ (0) = 0 � 1. It follows that
N
 − ia t − i at
U (t)ξ∀ (0) = e ¯ ℓ P ε∀ℓ ξ∀ (0) = e ¯ ξ∀ (0).
ℓ=1

6 Assuming ξ∀ (0) is normalised. If not we can insert a normalisation constant to


normalise it.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

80 Operators on N-Dimensional Complex Vectors

Next we have
   − i at 
dξ∀ (t) d U (t)ξ∀ (0) d e ¯ ξ∀ (0)
i� = i� = i�
dt dt dt
 − i at   − i at 
=a e ¯ ∀
ξ (0) = e ¯  ∀
A ξ (0)
 − i at   − i at 
= e ¯ A ξ∀ (0) = A e ¯ ξ∀ (0)
= A ξ∀ (t).
This shows that ξ∀ (t) is a solution of Eq. (13.100).
The question assumes that A has a nondegenerate eigenvalue a.
A similar proof can be carried out for an eigenvector corresponding
to a degenerate eigenvalue. We will then use the expression of U in
Eq. (13.49).
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 14

Model Theories Based on Complex


Vector Spaces

Q14(1) Verify the matrix representation of MŜ x in Eq. (14.30) and


of MŜ y in Eq. (14.41).
SQ14(1) The matrix elements are defined by Eq. (13.107). Using
Eq. (14.25) we express α∀ z and β∀z in terms of α∀ x and β∀x , i.e.,
1   1  
α∀ z = � α∀ x + β∀x , β∀z = � α∀ x − β∀x .
2 2
Then we have
  1 � 
MŜ x 11 = �∀ αz | Sx α∀ z ∪ = �∀ αx + β∀x | α∀ x − β∀x ∪ = 0.
2 2
  1 �  �
MŜ x 12 = �∀ αz | Sx β∀z ∪ = �∀ αx + β∀x | α∀ x + β∀x ∪ = .
2 2 2
  1 �  �
MŜ x 21 = �β∀z | Sx α∀ z ∪ = �∀ αx − β∀x | α∀ x − β∀x ∪ = .
2 2 2
  1 � 
MŜ x 22 = �β∀z | S x β∀z ∪ = �∀
 αx − β∀x | α∀ x + β∀x ∪ = 0.
2 2
These are the matrix elements of MŜ x in Eq. (14.30).
Using Eq. (14.37) we express α∀ z and β∀z in terms of α∀ y and β∀y , i.e.,
1   i  
α∀ z = � α∀ y + β∀y , β∀z = − � α∀ y − β∀y .
2 2

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

82 Model Theories Based on Complex Vector Spaces

Then we have
  1 � 
αz | Sy α∀ z ∪ = �∀
MŜ y 11 = �∀ α y + β∀y | α∀ y − β∀y ∪ = 0.
2 2
  i �  �
αz | Sy β∀z ∪ = − �∀
MŜ y 12 = �∀ α y + β∀y | α∀ y + β∀y ∪ = −i .
2 2 2
  i �  �
MŜ y 21 = �β∀z | Sy α∀ z ∪ = �∀ α y − β∀y | α∀ y − β∀y ∪ = i .
2 2 2
  1 � 
MŜ y 22 = �β∀z | Sy β∀z ∪ = − �∀ α y − β∀y | α∀ y + β∀y ∪ = 0.
2 2

Q14(2) Find the matrix representation of projectors Pα∀ y and P β∀y


 
in Eq. (14.38) in basis α∀ z , β∀z .
SQ14(2) Using Eq. (14.37) we express α∀ y and β∀y in terms of α∀ z and
β∀z , i.e.,
1   i  
α∀ z = � α∀ y + β∀y , β∀z = − � α∀ y − β∀y .
2 2
Then we have
1
P α∀ y α∀ z = �∀ α y = � α∀ y ,
α y | α∀ z ∪∀
2
i
P α∀ y β∀z = �∀ α y | β∀z ∪∀α y = − � α∀ y .
2
1
P β∀y α∀ z = �β∀y | α∀ z ∪β∀y = � β∀y ,
2
i
P β∀y β∀z = �β∀y | β∀z ∪β∀y = � β∀y .
2
The matrix M P̂ α∀ y is defined by the following matrix elements:
  1 1
M P̂ α∀ y 11
αz | P α∀ y α∀ z ∪ = � �∀
= �∀ αz | α∀ y ∪ = .
2 2
  i i
M Pˆ α∀ y 12
αz | Pα∀ y β∀z ∪ = − � �∀
= �∀ αz | α∀ y ∪ = − .
2 2
  1 i
M P̂ α∀ y 21
= �β∀z | P α∀ y α∀ z ∪ = � �β∀z | α∀ y ∪ = .
2 2
  i 1
M P̂ α∀ y 22
= �β∀z | P α∀ y β∀z ∪ = − � �β∀z | α∀ y ∪ = .
2 2
This is the projection matrix in Eq. (7.180).
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Model Theories Based on Complex Vector Spaces 83

The matrix M P̂ β∀ is defined by the following matrix elements:


y

  1 1
M P̂ β∀ 11
αz | P β∀y α∀ z ∪ = � �∀
= �∀ αz | β∀ y ∪ = .
y
2 2
  i i
M P̂ β∀ 12
αz | P β∀y β∀z ∪ = � �∀
= �∀ αz | β∀y ∪ = .
y
2 2
  1 i
M P̂ β∀ 21
= �β∀z | P β∀y α∀ z ∪ = � �β∀z | β∀y ∪ = − .
y
2 2
  i 1
M P̂ β∀ 22
= �β∀z | P β∀y β∀z ∪ = � �β∀z | β∀y ∪ = .
y
2 2
This is the projection matrix in Eq. (7.181).
Q14(3) Find the probability distribution function for the x ­
component spin values in the state given by the vector α∀ z . What is
the corresponding expectation value?
SQ14(3) Equations (14.56) and (14.58) apply here. The state
vector α∀ z is related to the eigenvectors of Sx by Eq. (14.25), i.e., we
have
1 1
α∀ x = � (α∀ z + β∀z ), β∀x = � (α∀ z − β∀z ).
2 2

(1) The value of the distribution function F Ŝ x (α∀ z , τ ) for τ < − 12 � is


zero.
(2) The value rises to 1/2 when τ = − 12 � since
1 1
αz | β∀x ∪ |2 = .
αz | Pβ∀x α∀ z ∪ = |�∀
℘ Ŝ x (α∀ z , − �) = �∀
2 2
1
(3) The value rises to 1 when τ = 2
� since
1 1
℘ Ŝ x (α∀ z , − �) + ℘ Ŝ x (α∀ z , �)
2 2
 
= �∀ 
αz | P β∀x α∀ z ∪ + �∀  αz | P β∀x + P α∀ x α∀ z ∪
αz | P α∀ x α∀ z ∪ = �∀
αz | II α∀ z ∪ = 1.
= �∀
1
We can check this result by calculating ℘ Ŝ x (α∀ z , 2
�) explicitly, i.e.,
1 1
℘ Ŝ x (α∀ z , αz | P α∀ x α∀ z ∪ = |�∀
�) = �∀ αz | α∀ x ∪ |2 = .
2 2
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

84 Model Theories Based on Complex Vector Spaces

The probability distribution function


⎧ is
⎪ �

⎪ 0 τ <−

⎨ 2
1 � �
F Ŝ x (α∀ z , τ ) = − ≤τ < .

⎪ 2 2 2

⎪ �
⎩1 τ⊕
2
The expectation value is equal to
 1  1  1  1
− � ℘ Ŝ x (α∀ z , − �) + � ℘ Ŝ x (α∀ z , �) = 0.
2 2 2 2
Q14(4) In the two-dimensional vector space VV∀ 2 we have the
following matrix representations in basis α∀ z , β∀z :

(1) The matrix representations of the projectors P α∀ x and Pβ∀x are


given by M P̂ α∀ x and M P̂ β∀ in Eq. (14.29).
x

(2) The matrix representations of projectors P α∀ y and P β∀y are given


by the matrices P α∀ y and P β∀y in Eqs. (7.180) and (7.181). In the
notation of this chapter these matrices are relabelled as M P̂ α∀ y
and M P̂ β∀ .
y

(3) The vector η∀ is represented by the column vector C η∀ in


Eq. (14.3).

Using the expression for M P̂ α∀ x = C α∀ x C α∀ x in Eq. (7.178) evaluate
M P̂ α∀ xC η∀. Show that the same result is obtained using explicit matrix
representations of C α∀ x and M P̂ α∀ x in Eqs. (14.3), (14.26) and (14.29).
Explain how the result confirms the projection nature of the matrix
M P̂ α∀ x.
Carry out a similar evaluation of M P̂ β∀ C η∀, M P̂ α∀ yC η∀ and M P̂ β∀ C η∀.
x y

SQ14(4) The evaluation is straightforward. We can evaluate


M P̂ α∀ xC η∀ in two different ways:
(1) Using Eq. (7.178) for the expression of M Pˆ α∀ x we get
 †   † 
M P̂ α∀ x C η∀ = C α∀ x C α∀ x C η∀ = C α∀ x C α∀ x C η∀
= �C α∀ x | C η∀∪C α∀ x ,
1
where
1 1 1
C α∀ x = � and �C α∀ x | C η∀∪ = � (c+ + c− ).
2 1 2
1 See Eq. (7.116) for the expression of C
α∀ x .
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Model Theories Based on Complex Vector Spaces 85

(2) Using explicit matrix representations in Eqs. (14.3) and (14.29)


and we obtain the same result, i.e.,

1 1 1 c+ c+ + c− 1
M P̂ α∀ x C η∀ = =
2 1 1 c− 2 1
1
= � (c+ + c− )C α∀ x .
2
The result shows that the matrix M P̂ α∀ x projects the column vector C η∀
onto the unit column vector C α∀ x .
Next we need to evaluate M P̂ β∀ C η∀, M P̂ α∀ yC η∀ and M P̂ β∀ C η∀.
x y

To evaluate M P̂ β∀ C η∀ we have:
x

(1) Using Eq. (7.179) for the expression of M P̂ β∀ we get


x
 †   † 
M P̂ β∀ C η∀ = C β∀x C β∀ C η∀ = C β∀x C β∀ C η∀
x x x

= �C β∀x | C η∀∪C β∀x ,


where2
1 1 1
C β∀x = � and �C β∀x | C η∀∪ = � (c+ − c− ).
2 −1 2
(2) Using explicit matrix representations in Eqs. (14.3) and (14.29)
and we have
1 1 −1 c+ c+ − c− 1
M P̂ β∀ C η∀ = = .
x 2 −1 1 c− 2 −1
The matrix M P̂ β∀ projects C η∀ onto the unit column vector C β∀x .
x

To evaluate M P̂ α∀ yC η∀ we have:
(1) Using Eq. (7.180) for the expression of M P̂ α∀ y we get
 †   † 
M P̂ α∀ y C η∀ = C α∀ y C α∀ y C η∀ = C α∀ y C α∀ y C η∀
= �C α∀ y | C η∀∪C α∀ y ,
where3
1 1 1
C α∀ y = � and �C α∀ y | C η∀∪ = � (c+ − i c− ).
2 i 2
2 See Eq. (7.116) for the expression of C
β∀x .
3 See Eq. (7.117) for the expression of C
α∀ y .
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

86 Model Theories Based on Complex Vector Spaces

(2) Using explicit matrix representations in Eqs. (14.3), (14.40) and


(7.180) get
1 1 −i c+ c+ − i c− 1
M Pˆ α∀ x C η∀ = = .
2 i 1 c− 2 i
The matrix M P̂ α∀ x projects C η∀ onto the unit column vector C α∀y .
Finally to evaluate M P̂ β∀ C η∀ we have:
y

(1) Using Eq. (7.181) for the expression of M P̂ β∀ we get4


y

 †   † 
M P̂ β∀ C η∀ = C β∀y C β∀ C η∀ = C β∀y C β∀ C η∀ = �C β∀y | C η∀∪C β∀y .
y y y

where
1 1 1
C β∀y = � and �C β∀y | C η∀∪ = � (c+ + i c− ).
2 −i 2

(2) Using Eq. (14.3) and (7.181) we have


1 1 i c+ c+ + i c− 1
M Pˆ β∀ C η∀ = = .
y 2 −i 1 c− 2 −i

The matrix M P̂ β∀ projects C η∀ onto the unit column vector C β∀y .


y

Q14(5) Write down a formal expression of the probability


density function w  (η∀, τ ) for the piecewise-constant probability
distribution function F Â (η∀, τ ) in Eq. (14.58) in terms of Dirac
delta functions.5 Write down the Stieltjes integral in Eq. (14.60) for
the expectation value in terms of the probability density function
w  (η∀, τ ) obtained above and evaluate the integral.

SQ14(5) In terms of Dirac delta functions the probability density


function for the distribution function in Eq. (14.58) is expressible
formally as

dF Â (η∀, τ )
ω Â (η∀, τ ) = = ℘ Â (η∀, a1 ) δ(τ − a1 )

+ ℘ Â (η∀, a2 ) δ(τ − a2 ) + · · · + ℘ Â (η∀, aN ) δ(τ − aN ).

4 See Eq. (7.117) for the expression of C which agrees with Eq. (14.40).
β∀ y
5 See Eq. (4.74).
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Model Theories Based on Complex Vector Spaces 87

We have used Eq. (4.74) here to express the derivative of a piecewise


constant function in terms of delta functions. The expectation value
is
 →  →
E(A, η∀ ) = τ dF Â (η∀, τ ) = τ ω Â (η∀, τ ) dτ
−→ −→
 → 
= τ ℘ Â (η∀, a1 ) δ(τ − a1 ) + ℘ Â (η∀, a2 ) δ(τ − a2 )
−→

+ · · · + ℘ Â (η∀, aN ) δ(τ − aN ) dτ

= a1 ℘ Â (η∀, a1 ) + a2 ℘ Â (η∀, a2 ) + · · · + aN ℘ Â (η∀, aN ),


as expected.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 15

Spectral Theory in Terms of Stieltjes


Integrals

Q15(1) Given a spectral function F (τ ) show that F (τ2 ) − F (τ1 )


for t2 > t1 is a projector. Is F (τ1 ) − F (τ2 ) also a projector?

SQ15(1) For selfadjointness we have


 †
F (τ2 ) − F (τ1 ) = F (τ2 )† − F (τ1 )† = F (τ2 ) − F (τ1 ).

For idempotence we have, using Eq. (15.11),


 2
F (τ2 ) − F (τ1 ) = F (τ2 )2 + F (τ1 )2 − F (τ1 )F (τ2 ) − F (τ2 )F (τ1 )

= F (τ2 ) + F (τ1 ) − 2F (τ1 )

= F (τ2 ) − F (τ1 ).

Hence F (τ2 ) − F (τ1 ) is a projector by Theorem 13.2.2(1). If P is a


projector then −P is not a projector
 since −P is not idempotent.
Similarly F (τ1 ) − F (τ2 ) = − F (τ2 ) − F (τ1 ) is not a projector.
  

Q15(2) Using Eqs. (15.15), (15.19) and (15.20) prove Eqs. (15.21),
(15.22) and (15.23).

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

90 Spectral Theory in Terms of Stieltjes Integrals

SQ15(2)
(1) To prove Eq. (15.21) we have
     
M [τ1 , τ2 ] = M  {τ1 } ∅ (τ1 , τ2 ] = M ({τ1 }) + M  (τ1 , τ2 ]
= F (τ1 ) − F (τ1 − 0) + F (τ2 ) − F (τ1 )
= F (τ2 ) − F (τ1 − 0).
(2) To prove Eq. (15.22) we have
   
 [τ1 , τ2 ] = M
M  [τ1 , τ2 ) ∅ {τ2 }
 
=M  [τ1 , τ2 ) + M  ({τ2 }).
     
 [τ1 , τ2 ) = M
M  [τ1 , τ2 ] − M  ({τ2 }) = F (τ2 ) − F (τ1 − 0)
 
− F (τ2 ) − F (τ2 − 0)
= F (τ2 − 0) − F (τ1 − 0).
(3) To prove Eq. (15.23) we have
   
 [τ1 , τ2 ) = M
M  {τ1 } ∅ (τ1 , τ2 )
 
=M  ({τ1 }) + M (τ1 , τ2 ) .
    
M (τ1 , τ2 ) = M  [τ1 , τ2 ) − M  ({τ1 })
   
= F (τ2 − 0) − F (τ1 − 0) − F (τ1 ) − F (τ1 − 0)
= F (τ2 − 0) − F (τ1 ).

Q15(3) Show that F(η∀, τ ) and M(η∀, ) in Eq. (15.50) define


a probability distribution function and a probability measure,
respectively.
SQ15(3)
(1) We need to show that F(η∀, τ ) in Eq. (15.50) satisfies proper­
ties MP3.6(1), MP3.6(2) and MP3.6(3) required in Definition
3.6(2).1
(a) MP3.6(1) is satisfied since F (τ2 ) − F (τ1 ) is selfadjoint and
idempotent, i.e., we have
 
�η∀ | F (τ2 )η∀ ∪ − �η∀ | F (τ1 )η∀ ∪ = �η∀ | F (τ2 ) − F (τ1 ) η∀ ∪
 
= || F (τ2 ) − F (τ1 ) η∀ ||2 ⊕ 0.

1 See also Definition 4.3.1(1).


June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Spectral Theory in Terms of Stieltjes Integrals 91

(b) MP3.6(2) is obviously satisfied on account of property


SF15.1(3) of spectral functions given in Definition 15.1(2),
i.e., F (−→) = 
0 and F (→) = II
(c) MP3.6(3) is also satisfied due to SF15.1(3) in Definition
15.1(2).

(2) Measures are defined by Definition 4.1.2(1) and a probability


measure is a measure M such that M(IR ) = 1. Properties
SM15.1(1) and SM15.1(2) of spectral measures are given in
Definition 15.1(3). We can show that �η∀ | M ()η∀ ∪ satisfies the
properties of a measure in the following manner:
(a) The non-negativity property is obvious, i.e., we have �η∀ |
 ()η∀ ∪ ⊕ 0 since M
M  () is selfadjoint and idempotent.
(b) For the empty set we have �η∀ | M  (∈) = 
 (∈)η∀ ∪ = 0 since M 0
from Eq. (15.12).
(c) For additivity we have, for a set of mutually disjoint Borel
sets ℓ ,
    
 ∅ℓ ℓ η∀ ∪ = �η∀ |
�η∀ | M  ℓ η∀ ∪
M

  
=  ℓ η∀ ∪.
�η∀ | M

(d) We have shown that �η∀ | M ()η∀ ∪ is a measure. It is also a


probability measure since
 (IR )η∀ ∪ = �η∀ | II η∀ ∪ = 1.
�η∀ | M
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 16

Infinite-Dimensional Complex Vectors


and Hilbert Spaces

Q16(1) Prove the inequalities in Eq. (16.4).


SQ16(1) Let z = a + i b and w = c + i d be two complex numbers,
where i is the imaginary unit and a, b, c, d are real numbers. To
prove the second inequality we start with, using Eq. (10.31),
 2   
|z� w |2 = |z� | |w| = |z |2 |w |2 = a2 + b2 c 2 + d 2 ,
   
|z |2 + |w |2 = a2 + b2 + c 2 + d 2 ,
 2   2
|z | − |w | = a2 + b2 − c 2 + d 2
     
= a2 + b2 + c 2 + d 2 − 2 (a2 + b2 c 2 + d 2
= |z |2 + |w |2 − 2 |z� w |.
 2
Since |z | − |w | ⊕ 0 we get
|z |2 + |w |2 ⊕ 2 |z� w |.
For the first inequality we start with
 �  
|z + w |2 = z + w z + w = |z |2 + |w |2 + z� w + zw �
= |z |2 + |w |2 + 2 × real part of z� w
≤ |z |2 + |w |2 + 2 |z� w | ≤ |z |2 + |w |2 + |z |2 + |w |2
 
= 2 |z |2 + |w |2 ,
Quantum Mechanics: Problems and Solutions
K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

94 Infinite-Dimensional Complex Vectors and Hilbert Spaces

since |z� w | ⊕ real part of z� w, and |z|2 + |w|2 ⊕ 2 |z� w |2 .


Q16(2) Verify Eqs. (16.21), (16.22) and (16.23).
SQ16(2)
� First we express ϕHn (x) in Eq. (16.18) as a function of y =
λ x, i.e.,
 � 1/2
� λ 2
ϕHn (y/ λ) = � n e−y /2 H n (y). (�)
π 2 n!

We can directly verify Eq. (16.21), i.e., A ϕH0 (y/ λ ) = 0, using the
fact that H 0 (y) = 1 in Eq. (16.14).
Next, using Eqs. (�) and (16.16) we get

A ϕHn (x) = A ϕHn (y/ λ )
 � 1/2
1 d λ 2
= � y+ � n e−y /2 H n (y)
2 dy π 2 n!
 � 1/2
1 λ 2
= � � (n−1) (n − 1)!
e−y /2 2nH n−1
2 π 2n 2
� � �
= n ϕ H (n−1) (y/ λ) = n ϕ H (n−1) (x).
This is Eq. (16.22).
Finally, for Eq. (16.23) we can carry out similar calculations using
Eqs. (�) and (16.17), i.e.,
 � 1/2
 � 1 d λ 2
A ϕHn (x) = � y− � n e−y /2 H n (y)
2 dy π 2 n!
 � 1/2
1 λ 2  
= � � n e−y /2 2y H n (y) − 2nH n−1 (y)
2 π 2 n!
 � 1/2
1 λ 2
= � � n e−y /2 H n+1 (y)
2 π 2 n!
 � 1/2
1 λ 2(n + 1) 2
= � � (n+1) e−y /2 H n+1 (y)
2 π 2 (n + 1)!
� � �
= n + 1 ϕ H (n+1) (y/ λ ) = n + 1 ϕ H (n+1) (x).

Q16(3) Prove the equality


||ϕ∀ + φ∀ ||2 + ||ϕ∀ − φ∀ ||2 = 2||ϕ∀ ||2 + 2||φ∀ ||2 .
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Infinite-Dimensional Complex Vectors and Hilbert Spaces 95

SQ16(3) Simply work out the norm explicitly:


||ϕ∀ + φ∀ ||2 = �ϕ∀ + φ∀ | ϕ∀ + φ∀ ∪ = ||ϕ∀ ||2 + ||φ∀ ||2 + �ϕ∀ | φ∀ ∪ + �φ | ϕ∪.
||ϕ∀ − φ∀ ||2 = �ϕ∀ − φ∀ | ϕ∀ − φ∀ ∪ = ||ϕ∀ ||2 + ||φ ||2 − �ϕ∀ | φ∀ ∪ − �φ∀ | ϕ∀ ∪.
Adding the above two equations produces the result.

Q16(4) Prove the Schwarz inequality in Eq. (16.75).


SQ16(4) The Schwarz inequality can be proved in the same way
as in a finite-dimensional complex vector space shown in SQ12(3)
to question Q12(3) in Exercises and Problems Chapter 12.
Q16(5) Prove the triangle inequalities in Eqs. (16.76) and (16.77).
SQ16(5) The triangle inequalities can be proved in the same way
as in a finite-dimensional complex vector space shown in SQ12(4) to
question Q12(4) in Exercises and Problems for Chapter 12.
Q16(6) Show that φ(x ), ψ(x ) ⇒ L2 (IR ) � φ(x ) + ψ(x ) ⇒ L2 (IR )
and that the integral on the right hand side of Eq. (16.28) is finite for
functions in L2 (IR ).

SQ16(6)1 Since
 �  
|φ(x ) + ψ(x )|2 = φ(x ) + ψ(x ) φ(x ) + ψ(x )
= |φ(x )|2 + |ψ(x )|2 + φ � (x )ψ(x ) + φ(x )ψ(x )� ,
 �  
|φ(x ) − ψ(x )|2 = φ(x ) − ψ(x ) φ(x ) − ψ(x )
= |φ(x )|2 + |ψ(x )|2 − φ � (x )ψ(x ) − φ(x )ψ(x )� ,
we get, by adding the above two equations,
 
|φ(x ) + ψ(x )|2 = 2 |φ(x )|2 + |ψ(x )|2 − |φ(x ) − ψ(x )|2
 
≤ 2 |φ(x )|2 + |ψ(x )|2 .

Since |φ(x )|2 and |ψ(x )|2 are separately integrable over IR their sum
is integrable over IR . It follows that |φ(x ) + ψ(x )|2 is integrable over
IR and that φ(x ) + ψ(x ) is a member of L2 (IR ).
Similarly |φ(x ) ± i ψ(x )| are square-integrable and hence are
members of L2 (IR ).
1 Fano pp. 240–241.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

96 Infinite-Dimensional Complex Vectors and Hilbert Spaces

To prove Eq. (16.28) we start with similar calculations, i.e.,


 �  
|φ(x ) + i ψ(x )|2 = φ(x ) + i ψ(x ) φ(x ) + i ψ(x )
= |φ(x )|2 + |ψ(x )|2 + i φ � (x )ψ(x ) − i φ(x )ψ(x )� .
 �  
|φ(x ) − i ψ(x )|2 = φ(x ) − i ψ(x ) φ(x ) − i ψ(x )
= |φ(x )|2 + |ψ(x )|2 − i φ � (x )ψ(x ) + i φ(x )ψ(x )� .
Then
i |φ(x ) − i ψ(x )|2 − i |φ(x ) + i ψ(x )|2
= 2φ � (x )ψ(x ) − 2φ(x )ψ(x )� .
From earlier calculations we get
|φ(x ) + ψ(x )|2 − |φ(x ) − ψ(x )|2
= 2φ � (x )ψ(x ) + 2φ(x )ψ(x )� .

Finally we get,2 by adding the above equations,


1
φ � (x )ψ(x ) = |φ(x ) + ψ(x )|2 − |φ(x ) − ψ(x )|2
4 
+ i |φ(x ) − i ψ(x )|2 − i |φ(x ) + i ψ(x )|2 .

Since all the terms on the right of the equality sign are integrable
over IR we conclude that φ(x )� ψ(x ) is integrable over IR .
Q16(7) Show that the sequence of vectors f∀n defined by functions
fn (x ) in Eq. (16.60) is a Cauchy sequence in the space C∀().3
SQ16(7)4 The function fn (x ) has the value 0 for x ≤ 1 − 1/n so
that in the limit as n � → the function has the value zero for all x <
1. It follows that in the limit as n, m � → the function fn (x ) − fm (x )
also has the value zero for all x < 1. Given that fn (x ) and fm (x ) have
the value 1 for all x ⊕ 1 and hence their difference
 fn (x ) − fm (x ) has
the value 0 for all x ⊕ 1 we can deduce that fn (x ) − fm (x ) � 0 for
all x as n, m � → and that
 2
| fn (x ) − fm (x )|2 dx � 0 as n, m � →.
0

2 Fano p. 241.
3 Fano p. 251.
4 Fano pp. 250–251.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Infinite-Dimensional Complex Vectors and Hilbert Spaces 97

Hence the set of functions fn (x) defines a Cauchy sequence of


vectors f∀n in C∀(). Note that
 2
|| f∀n − f∀m ||2 = | fn (x ) − fm (x )|2 dx .
0
Q16(8) Express the spherical harmonics in Eqs. (16.64) to (16.67)
in the Cartesian coordinates x, y, z which are related to the spherical
coordinates r, θ, ϕ by Eq. (16.41).

SQ16(8) In terms of Cartesian coordinates related to the spherical


coordinates by Eq. (16.41) the expression for Y0, 0 is unchanged
while the others become

3  
Y1, −1 = cos ϕ − i sin ϕ sin θ


3  
= cos ϕ sin θ − i sin ϕ sin θ


3 x − iy
= ,
8π r

3 z
Y1, 0 = ,
4π r

3 x + iy
Y1, 1 = − ,
8π r

where r = x 2 + y 2 + z2 .
Q16(9) Show that Eq. (16.81) defines a continuous linear func­
∀.
tional on H
SQ16(9) In Eq. (16.81) we have a functional F ϕ∀ on H ∀ defined by
F ϕ∀ (φ∀ ) = �ϕ∀ | φ∀ ∪. This functional is linear since the scalar product is
linear in φ∀.
Let {φ∀n , n = 1, 2, . . .} be a sequence of vectors converging to the
vector φ∀. Using the Schwarz inequality we get
|�ϕ∀ | φ∀ ∪ − �ϕ∀ | φ∀n ∪| = |�ϕ∀ | φ∀ − φ∀n ∪|
≤ ||ϕ∀ || || φ∀ − φ∀n || � 0
as φ∀n � φ∀. The functional F ϕ∀ is therefore continuous.
Q16(10) In the Dirac notation a scalar product is denoted by
�ϕ∀ | φ∀ ∪. Dirac formally consider the notation as the product of
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

98 Infinite-Dimensional Complex Vectors and Hilbert Spaces

two quantities: (1) �ϕ∀ | called a bra and (2) |φ∀ ∪ called a ket. Their
product forms a bracket �ϕ∀ | φ∀ ∪ which is the scalar product. Explain
how we can interpret bras and kets in terms of vectors and linear
functionals.5
SQ16(10) A ket is just a vector by definition, i.e., |φ∪ := φ∀. Since
a bra �ϕ∀ | operates on a ket φ∀ produces the scalar product �ϕ∀ |
φ∀ ∪ which is equal to the linear functional F ϕ∀ generated by ϕ∀, i.e.,
F ϕ∀ (φ∀ ) = �ϕ∀ | φ∀ ∪ we can see that the bra �ϕ∀ | is identifiable with
the linear functional F ϕ∀ .
Q16(11) Explain the concept of separability in a Hilbert space.
SQ16(11) An infinite-dimensional vector space does not nec­
essarily possess a countable orthonormal basis. The concept of
separability of a space is defined by the existence of a countable
orthonormal basis, i.e., a Hilbert space is separable if there is a
countable orthonormal basis for the space.

5 Jauch p. 32.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 17


Operators in a Hilbert Space H

Q17(1) Show that every linear operator on a finite-dimensional


Hilbert space is bounded.1

SQ17(1) Let A be a linear operator in a finite-dimensional space


N N
VV∀ . Let {∀
εℓ , ℓ = 1, 2, · · · , N} be an orthonormal basis for VV∀ . By
definition A is defined on all the basis vectors ε∀ℓ , i.e., Aε∀ℓ are vectors
N
in VV∀ with a finite norm ||Aε∀ℓ ||. Let M be the maximum of the set
N
of norms ||Aε∀ℓ ||. Let u∀ be an arbitrary unit vector in VV∀ . We have

u∀ = ℓ cℓ ε∀ℓ . For the norm of Au∀ we have, by the triangle inequality
in Eq. (16.75) and Schwarz inequality in Eq. (16.74),

N
 N
 N

||Au∀ || = || cℓ Aǫ∀ℓ || ≤ |cℓ | ||Aε∀ℓ || ≤ |cℓ | M
ℓ=1 ℓ=1 ℓ=1
N

=M |cℓ | ≤ MN,
ℓ=1

since |cℓ | = |�∀ εℓ || ||u∀ || = 1. It follows that A is bounded


εℓ | u∀ ∪| ≤ ||∀
by Definition 17.1(2).

1 Halmos p. 177.

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)


100 Operators in a Hilbert Space H

Q17(2) Show that projectors in a Hilbert space are bounded with


norm 1.
SQ17(2) Let P S∀ be the projector onto the subspace S∀, and let φ∀ be
a unit vector. Since φ∀ = φ∀S∀ + φ∀S∀� and φ∀S∀� and φ∀S∀� are orthogonal
we get
�φ∀ | φ∀ ∪ = �φ∀S∀ | φ∀S∀ ∪ + �φ∀S∀� | φ∀S∀� ∪ ⊕ �φ∀S∀ | φ∀S∀ ∪

� ||φ∀S∀ || ≤ ||φ∀ || = 1.
Since P S∀ φ∀ = φ∀S∀ we deduce that
||P S∀ φ∀ || ≤ ||φ∀ || = 1 ∞ φ∀.
The upper bounded of ||PS∀ φ∀ || is 1 for any φ∀ ⇒ S∀. It follows that the
norm of PS∀ is 1.
Q17(3) Are projectors invertible? Are they reducible?
SQ17(3) Projectors are not invertible since they violate Theorem
17.5(1). The projector P S∀ onto the subspace S∀ would have zero
projection of any vector in the orthogonal complement S∀� of S∀, i.e.,
φ∀ ⇒ S∀� � PS∀ φ∀ = 0. ∀

The subspace S is clearly invariant under PS∀ , i.e., φ∀ ⇒ S∀ �
P S∀ φ∀ ⇒ S∀. Since PS∀ is selfadjoint and bounded Theorem 17.9(1)


applies, i.e., P ∀ is reducible by S.
S

Q17(4) Verify that the spherical harmonics given by Eqs. (16.64)


to (16.67) are eigenfunctions of the operator L z (Su ) in Eq. (17.42),
i.e.,
∂Yℓ, mℓ (θ, ϕ)
−i � = mℓ � Yℓ, mℓ (θ, ϕ),
∂ϕ
or
z (Su )Y∀ℓ, mℓ = mℓ � Y∀ℓ, mℓ .
L

SQ17(4) Using the differential expression for  Lz in Eq. (17.42), i.e.,



Lz = −i � ∂/∂ϕ we obtain the following results:
(1) For Y0, 0 in Eq. (16.63) we have −i � ∂Y0, 0 /∂ϕ = 0, since Y0, 0 is a
constant.
(2) For Y1, −1 in Eq. (16.66) we have −i � ∂Y1, −1 /∂ϕ = −�Y1, −1 . The
value −� comes from differentiating exp(−i ϕ) in Y1, −1 .
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)


Operators in a Hilbert Space H 101

(3) For Y1, 0 in Eq. (16.65) we have −i � ∂Y1, 0 /∂ϕ = 0, since Y1, 0 is
independent of ϕ.
(4) For Y1, 1 in Eq. (16.66) we have −i � ∂Y1, 1 /∂ϕ = �Y1, 1 . The value
� comes from differentiating exp(i ϕ) in Y1, 1 .
We can conclude that the spherical harmonics in Eqs. (16.64) to
(16.67) are eigenfunctions of Lz corresponding to eigenvalues 0, −�,
0 and � respectively.2
Q17(5) Prove properties P17.10(1), P17.10(2) and P17.10(3) for
a pair of operators a and a† in Definitions 17.10(1) and 17.10(2).
SQ17(5) For property P17.10(1) we have
�ϕ∀1 | aϕ∀0 ∪ = 0 and �aϕ∀1 | ϕ∀0 ∪ = �ϕ∀0 | ϕ∀0 ∪ = 1.
In other words we have �ϕ∀1 | aϕ∀0 ∪ �= �aϕ∀1 | ϕ∀0 ∪. This implies
that a is not selfadjoint since Eq. (17.102) which is necessary for
selfadjointness is violated. Similarly we have
�ϕ∀1 | a† ϕ∀0 ∪ = �ϕ∀1 | ϕ∀1 ∪ = 1 and

�a† ϕ∀1 | ϕ∀0 ∪ = 2�ϕ∀2 | ϕ∀0 ∪ = 0.
Since �ϕ∀1 | a† ϕ∀0 ∪ �= �a† ϕ∀1 | ϕ∀0 ∪ the operator a† is not selfadjoint.
For property P17.10(2) we have
 ϕ∀0 = a† a ϕ∀0 = 0 ϕ∀0 .
N
� � 
 ϕ∀n = a† a ϕ∀n = a† n ϕ∀n−1 = n (n − 1) + 1 ϕ∀(n−1)+1
N
= nϕ∀n ∞n = 1, 2, 3, · · · .
 corresponding
These results show that ϕ∀n are the eigenvectors of N
to eigenvalues n = 0, 1, 2, · · · .
For property P17.10(3) we have, by the definition of the
operators, for n ⊕ 1,

aa† ϕ∀n = a n + 1 ϕn+1 = (n + 1)ϕ∀n ,

a† a ϕ∀n = a n ϕ∀n−1 = nϕ∀n
 
� aa† − a† a ϕ∀n = ϕ∀n
� [ a, a† ] = II
� [ a† , a] = −II .
2 See
SQ36(1) in Exercises and Problems for Chapter 36 for calculations in Cartesian
coordinates.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)


102 Operators in a Hilbert Space H

The above results also hold n = 0 since


 
aa† − a† a ϕ∀0 = aa† ϕ∀0 = ϕ∀0 .

Using the formula [A, B  C ] = [A, B


 ]C + B
 [A, C ] and the commuta­
† 
tion relation [a, a ] = II we get
[ a, N ] = [ a, a† a ] = [ a, a† ] a = a,
 ] = [ a† , a† a] = a† [ a† , a] = −a† .
[ a† , N

Q17(6)  := a† a.


Find the domain of the operator N
SQ17(6) Let φ∀ be a vector in the domain D ∀ (N
 ) of N
 . Expressing

∀ ∀
φ as a linear combination of ϕ∀n , i.e., writing φ = n cn ϕ∀n we have,
using property P17.10(2),

 →

 φ∀ =
N  ϕ∀n =
cn N cn nϕ∀n ,
n=o n=0

 →

 φ∀ ||2 =
||N cn� cm nm �ϕ∀n | ϕ∀m ∪ = |cn |2 n2 .
n, m=0 n=0

 φ∀ to have a finite norm, i.e., we must have


We must require N


 φ∀ ||2 < → or equivalently
||N |�ϕ∀n | φ∀ ∪|2 n2 ≤ →,
0

since cn = �ϕ∀n | φ∀ ∪. It follows that the domain of N


 is


∀ (N
D ∀:
 ) = {φ∀ ⇒ H |�ϕ∀n | φ∀ ∪|2 n2 ≤ →}.
n=0

This is consistent with Eq. (17.58), i.e.,


∀ (N
D ∀ (a) : aφ∀ ⇒ D
 ) = {φ∀ ⇒ D ∀ (a† ) = D
∀ (a)}.

Q17(7) Let {ϕ∀n , n = 0, 1, 2, . . .} be an orthonormal basis for a


given Hilbert space. Show that the vector
→
1 2 zn
∀ z = exp −
 |z| � ϕ∀n , z ⇒ C,
2 n=0
n!
is the eigenvector of the annihilation operator a associated with
the orthonormal basis defined by Eqs. (17.114) and (17.115)
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)


Operators in a Hilbert Space H 103

∀ z is
corresponding to the eigenvalue z. Show also that the vector 
normalised.
Finally show that
∀z | N
� ∀ z ∪ = | z |2 ,
 = a† a.
where N
SQ17(7) First we have
→
1 2 zn
∀ z = exp
a  − |z| � aϕ∀n
2 n=0
n!
→
1 2 zn �
= exp − |z| � n ϕ∀n−1
2 n=1
n!


1 2 zn−1
= exp − |z| z� ϕ∀n−1
2 n=1
(n − 1)!
! →
"
1  zn−1
= z exp − | z |2 � ϕ∀n−1
2 n=1
(n − 1)!
! →
"
1 2  zn
= z exp − |z| � ϕ∀n = z ∀ z,
2 n=0
n!
∀ z is an eigenvector of the annihilation operator a
showing that 
corresponding to the eigenvalue z.
For normalisation we have
 →
(zn )� zm
� ∀ z ∪ = exp(−| z |2 )
∀z |  � � �ϕ∀n | ϕ∀m ∪
n, m=0
n! m!
 →
(zn )� zm
= exp(−| z |2 ) � � δnm
n, m=0
n! m!

 (| z |2 )n
= exp(−| z |2 )
n=0
n!
= exp(−| z | ) exp(|z|2 ) = 1.
2

∀z = z
Using the result a  ∀ z we get
∀z | N
� ∀ z ∪ = �
∀ z | a a 
† ∀ z∪
= �a ∀ z ∪ = �z
∀ z | a ∀ z | z
∀ z ∪ = z� z �
∀z | 
∀z∪
= | z |2 .
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)


104 Operators in a Hilbert Space H

We can explicitly construct of the eigenvectors of  a. Let ∀ z be an


a corresponding to the eigenvalue z, i.e., 
eigenvector of  ∀ z = z
a ∀ z.
Now expand  ∀ z as a linear combination of ϕ∀ n , i.e.,


∀z =
 �∀ ∀ z ∪∀
ϕn |  ϕn . (�)
n=0

Then we have

�∀
ϕn | 
a∀ z ∪ = � ∀ z∪ =
a† ϕ∀ n |  n + 1 �∀ ∀ z∪
ϕn+1 | 


� z �∀ ∀z ∪ =
ϕn |  n + 1 �∀ ∀ z ∪,
ϕn+1 | 

∀ z∪ = � z ∀ z ∪.
ϕn+1 | 
� �∀ �∀
ϕn | 
n+1
This is valid for any n. Explicitly we have:

(1) For n = 0 we get


z
∀ z ∪ = � �∀
ϕ1 | 
�∀ ∀ z ∪.
ϕ0 | 
1
(2) For n = 1 we get
z z2
�∀ ∀ z ∪ = � �∀
ϕ2 |  ∀z ∪ = �
ϕ1 |  �∀ ∀ z ∪.
ϕ0 | 
2 2×1
(3) For n = 2 we get
3
�∀ ∀ z ∪ = �z �∀
ϕ3 |  ∀z ∪ = � z
ϕ2 |  �∀ ∀z ∪
ϕ0 | 
3 3×2×1
z3
= � �∀ ∀ z ∪.
ϕ0 | 
3!
(4) By repeating the process we get, for ⊕ 1,
z
∀ z ∪ = � �∀
ϕn | 
�∀ ϕn−1 | ∀z ∪
n
z z z z
= � � � · · · � �∀ ∀z ∪
ϕ0 | 
n n−1 n−2 1
zn
= � �∀ ∀ z ∪.
ϕ0 | 
n!
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)


Operators in a Hilbert Space H 105

It follows that Eq. (�) becomes


→ →
zn zn

z = � �∀ ∀
ϕ0 | z ∪ ϕ∀ n = �∀ ∀
ϕ0 | z ∪ � ϕ∀ n .
n=0
n! n=0
n!
By identifying �∀ ∀ 2
ϕ0 | z ∪ with the normalisation constant e−|z| /2 this
eigenvector agrees with  ∀ z.
To make the proof of the statement
zn
∀ z ∪ = � �∀
ϕn | 
�∀ ∀z ∪
ϕ0 | 
n!
more rigorous for all n we can use the method of (mathematical)
induction. This is a standard method to prove a sequence of
statement Sn to be true. The method goes as follows3 :
(1) Show that the first statement S1 is true.
(2) Show that if the nth statement is true then the (n + 1)th
statement must also be true, i.e.,
Sn is true � Sn+1 is true.
(3) It follows that S2 is true, since S1 is true, and then S3 is true, since
S2 is true, and so on. We can then conclude that the statement Sn
is true for all n.
In the present case statement S1 is the expression for �ϕ∀1 | ∀ z ∪ given
in item (1) and statement Sn is the expression for �ϕ∀n | ∀ z ∪ given in
item (4) earlier. We have already shown that S1 is true, i.e.,
z
�ϕ∀1 | ∀ z ∪ = z�ϕ∀0 | ∀ z ∪ = � �∀ ∀ z ∪.
ϕ0 | 
1!
Next assume that statement Sn is true, i.e., the following equation
zn
�ϕ∀n | ∀ z ∪ = � �ϕ∀0 | ∀ z ∪
n!
is true for any given n > 1. Then we have
1 1
�ϕ∀n+1 | ∀ z ∪ = � �a† ϕ∀n | ∀ z ∪ = � �∀
ϕn | 
a∀z ∪
n+1 n+1
z
= � �ϕ∀n | ∀ z ∪
n+1
z  zn 
= � � �ϕ∀0 | ∀ z ∪
(n + 1) n!
zn+1
= � �ϕ∀0 | ∀ z ∪.
(n + 1)!
3 See SQ27(8) and SQ27(13) for further examples.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)


106 Operators in a Hilbert Space H

In other words the statement is then true for n + 1. We can now


conclude that the expression is true for all n.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 18


Bounded Operators on H

Q18(1) The characteristic function χ  (x ) of the interval  in IR


defines a multiplication operator χ   on L∀2 (IR ) by Eq. (17.11). Show

that χ  is a projector. Find the eigenvalues and their corresponding
eigenvectors of χ .
SQ18(1) The multiplication operator χ   defined by the character­
istic function χ  (x ) is clearly selfadjoint since χ  (x ) is real-valued
so that we have, for all φ∀, ψ∀ ⇒ L∀2 (IR ),
 →
  ψ∀ ∪ =
�φ∀ | χ φ(x )� χ  (x )ψ(x ) dx
−→
 →
 
= χ  (x )φ(x ) � ψ(x ) dx = � χ  φ∀ | ψ∀ ∪.
−→

It is also idempotent since


     
χ  2 ψ∀ = χ  χ  ψ∀ := χ  (x ) 2 ψ(x ) = χ  (x )ψ(x ).
The operator χ   is therefore projector.
As is the case for all projectors it has two eigenvalues, 1 and
0. The eigenvectors corresponding to the eigenvalue 1 are vectors
defined by functions in L2 (IR ) which vanish outside  while the
eigenvectors corresponding to the eigenvalue 0 are vectors defined
by functions in L2 (IR ) which vanish inside .

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)


108 Bounded Operators on H

x () on L∀2 () in


Q18(2) Show that the multiplication operator 
Eq. (17.22) is bounded and selfadjoint.
SQ18(2) Let M be the upper bound of x 2 for x in the interval .
Then we have, for every unit vector φ∀ ⇒ L∀2 (),
 
x ()φ∀ ||2 =
|| x 2 |φ(x)|2 dx ≤ M2 |φ(x)|2 dx = M.
 
Hence the operator is bounded. This is not true if  is an unbounded
interval.
As a multiplication operator  x is selfadjoint since the operator
satisfies the selfadjointness condition in Eq. (18.3), i.e., for any ψ∀ and
φ∀ in L∀2 () we have


�ψ |  ∀
x ()φ ∪ = ψ � (x)x φ(x) dx = �x ()ψ∀ | φ∀ ∪.


Q18(3) Show that plane waves f p (x) in Eq. (18.12) is not square­
integrable over IR.

SQ18(3) The plane wave�f p (x) has an absolutely value of 1/ 2π �
for all x, i.e., | f p (x)| = 1/ 2π � and | f p (x)|2 = 1/2π �. Hence the
function is not integrable over IR since
 →  →
2 1
| f p (x)| dx = dx = →.
−→ 2π � −→

Q18(4) Show that


(1) The trace of a projector is equal to the dimension of the
subspace onto which the projector projects.
(2) If a density operator is a projector then it is a one-dimensional
projector.

SQ18(4)
(1) Let PS∀ be a projector onto the subspace S∀ of a Hilbert space H ∀.

Let ϕ∀m be an orthonormal basis for S and let ψ∀ n be an orthonormal
basis for S∀� . Then the set of vectors {ϕ∀m , ψ∀ n } form an orthonormal
basis for H∀ . In accordance of Eq. (18.14) the trace of P ∀ is given by
S
 
tr(P S∀ ) = �ϕ∀m | P S∀ ϕ∀m ∪ + �ψ∀ n | PS∀ ψ∀ n ∪
m n

= �ϕ∀m | P S∀ ϕ∀m ∪,
m
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)


Bounded Operators on H 109

∀ Since �ϕ∀m | P ∀ ϕ∀m ∪ = �ϕ∀m | ϕ∀m ∪ = 1


using the fact that P S∀ ψ∀ n = 0. S
for all ϕ∀m the final sum above is equal to the dimension of the
subspace S∀.
(2) A density operator D  has trace 1. It follows that if it is also
a projector then it must be a projector onto a one-dimensional
subspace, i.e., it must be a one-dimensional projector.
Q18(5) Show that a convex combination of density operators as
shown in Eq. (18.24) defines a density operator.
SQ18(5) First the operator is bounded since for every unit vector
ϕ∀ we have, on account of the triangle inequality in Eq. (16.76) and
the result on the norm of an operator in Eq. (17.7),
n
 n
 n


||Dϕ∀ || = || 
wℓ Dℓ ϕ∀ || ≤ 
|| wℓ Dℓ ϕ∀ || =  ℓ ϕ∀ ||
wℓ || D
ℓ=1 ℓ=1 ℓ=1
n
 n

≤  ℓ || ||ϕ∀ || =
wℓ || D wℓ M ≤ M,
ℓ=1 ℓ=1
 ℓ ||. Since all the operators involved are
where M is the biggest of || D
bounded we can use Eq. (17.95) to show that D  is selfadjoint, i.e., we
have
 
† =
D wℓ D † = wℓ Dℓ = D.

ℓ ℓ
The operator is also positive since
n

�φ∀ | D
 φ∀ ∪ = wℓ �φ∀ | D
 ℓ φ∀ ∪
ℓ=1
is a sum of terms which are all positive.1 Let {ϕ∀m , m = 1, 2, . . .} be
an orthonormal basis for the Hilbert space. The trace of D  is
  n

tr D) =  ϕ∀m ∪ =
�ϕ∀m | D �ϕ∀m | wℓ D ℓ ϕ∀m ∪
m m ℓ=1
n
 
= wℓ  ℓ ϕ∀m ∪
�ϕ∀m | D
ℓ=1 m
n
 n


= ℓ) =
wℓ tr D wℓ = 1.
ℓ=1 ℓ=1

1 See Definition 13.3.1 and P18.1(1) for definition of positive operators.


June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)


110 Bounded Operators on H

 is a density operator.
Hence D
Q18(6) Show that the density operator D z in Eq. (18.29) can be
decomposed in terms of the eigenprojectors P α∀ y and Pβ∀y of Sy
shown in Eq. (14.38) as
z = 1 P α∀ y + 1 P ∀ .
D
2 2 βy

SQ18(6) Since Pα∀ z and P β∀z constitute a complete orthogonal


family of projectors on VV∀ 2 we have P α∀ z + P β∀z = II . It follows that
 
 z = P α∀ z + P ∀ /2 = II /2.
D βz

Since Pα∀ y and P β∀y also constitute a complete orthogonal family of


projectors on VV∀ 2 we have P α∀ y + P β∀y = II . It follows that
 
 z = II /2 = P α∀ y + P ∀ /2.
D βy

 Find
Q18(7)  the matrix representations (density matrices) in
basis α∀ z , β∀z of the density operators D
z and D
x in Eqs. (18.29)
and (18.30).
SQ18(7) Using the formula in Eq. (13.107) the matrix elements of
the density matrix D z for D  z are easily worked out to be
  1   1
D z 11 = �∀
αz | D z α∀ z ∪ = �∀
αz | P α∀ z + P β∀z α∀ z ∪ = ,
2 2
  1  
D z 12 = �∀
αz | D z β∀z ∪ = �∀αz | P α∀ z + P β∀z β∀z ∪ = 0,
2
   
 z α∀ z ∪ = �β∀z | 1 P α∀ z + P ∀ α∀ z ∪ = 0,
D z 21 = �β∀z | D βz
2
  1   1
D z 22 = �β∀z | Dz β∀z ∪ = �β∀z |
 P α∀ z + P β∀z β∀z ∪ = .
2 2
The desired density matrix is
1/2 0 1 10
Dz = = .
0 1/2 2 01
The density matrix D x for D  x can be similarly worked out. Since
x = D
D  z by Eq. (18.32) the same density matrix is obtained, i.e., we
have
1 10
Dx = .
2 01
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)


Bounded Operators on H 111

Q18(8) Show that unitary operators on a Hilbert space are


invertible.
SQ18(8) For a unitary operator U we have �U φ∀ | U φ∀ ∪ = �φ∀ | φ∀ ∪.
∀ implies �φ∀ | φ∀ ∪ = 0. This in turn implies
It follows that U φ∀ = 0
∀ ∀
φ = 0. Theorem 17.5(1) then tells us that U is invertible.
Q18(9) Show that a unitary transformation preserves the trace of
the product operator in Eq. (18.15), i.e.,
   � �
D
tr B  = tr B D  ,

where B � are the unitary transforms of B


 � and D  and D
 respectively

generated by a unitary operator U in accordance with Definition
13.4.2(1).
SQ18(9) Let {ϕ∀ℓ } be an orthonormal basis. Then
    
tr B D
 = D
�ϕ∀ℓ | B  ϕ∀ℓ ∪.

We also have
 � �   � �   
tr BD = D
�ϕ∀ℓ | B  ϕ∀ℓ ∪ = U † U D
�ϕ∀ℓ | U B  U † ϕ∀ℓ ∪
ℓ ℓ
 
= † D
�U ϕ∀ℓ | B  U ϕ∀ℓ ∪ =
† D
�ϕ∀ �ℓ | B  ϕ∀ �ℓ ∪,
ℓ ℓ

where ϕ∀ �ℓ †
= U ϕ∀ℓ . These new vectors, being unitary transforms of
the basis vectors of an orthonormal basis, form a new orthonormal
basis. Since the trace is independent of the choice of basis, we get
    � �  
D
�ϕ∀ �ℓ | B  ϕ∀ �ℓ ∪ = tr B
D � tr BD = tr B D .

Q18(10) Show that the unitary transform of a one-dimensional


projector is again a one-dimensional projector.
SQ18(10) Let P be a one-dimensional projector, i.e., P = |φ∀ ∪�φ∀ |
for some unit vector φ∀. Let φ∀� and P� be the unitary transforms of φ∀
and P generated by a unitary operator U , i.e., φ∀� = U φ∀ and P� =
U PU † . We have, using Eq. (18.40),
|φ∀� ∪�φ∀� | = |U φ∀ ∪�U φ∀ | = U |φ∀ ∪�φ∀ | U † = P � .
In other words P� is the projector generated by the unit vector φ∀� . It
follows that P� is one-dimensional.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)


112 Bounded Operators on H

Q18(11) Verify Eq. (18.53).


SQ18(11) Using the properties of Dirac delta functions shown in
Eqs. (18.44) and (18.47) we get
 →
ϕ( p)�⊗ϕ( p) dp

−→
 →  → �  →

= f p (x) φ(x) dx f p� (x � ) φ(x � ) dx � dp
−→ −→ −→
 → →  →
� �
= φ (x)φ(x ) f p (x) f p� (x � ) dp dxdx �
−→ −→ −→
 → →
= φ � (x)φ(x � ) δ(x − x � ) dxdx �
−→ −→
 →
= φ � (x)φ(x) dx.
−→

Q18(12) Using the expression i �d/dp and p for x (IR)



(IR)
and p ⊗
verify the commutation relation in Eq. (18.69).

SQ18(12)
 d d   d  d 
[ (IR ] = −i � x
x (IR), p − x = −i � x − 1+x = i �.
dx dx dx dx

 d d   d  d 
[
x⊗ (IR),

(IR)]
p ⊗
= i� p− p = i� 1+ p −p = i �.
⊗ dp dp dp dp

Q18(13) The Fourier transform ⊗ϕ∀ of ϕ∀ ⇒ L∀2 (IR) is given by the


following characteristic function on the momentum space:

⎨ 0, � p ≤ − p0 ,
ϕ( p) = 1/ 2 p0 , p ⇒ (− p0 , p0 ],
⊗ ⎩
0, p ⊕ p0 ,

where p0 is a real and positive constant. By performing the


inverse Fourier transform show that the function ϕ(x) ⇒ L2 (IR)
corresponding to the vector ϕ∀ is given by
#
� sin( p0 x/�)
ϕ(x) = .
π p0 x
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)


Bounded Operators on H 113

SQ18(13) The inverse Fourier transform of ⊗ϕ( p) is


 →
1
ϕ(x) = � ei x p/� ⊗ϕ( p) dp
2π � −→
 p0
1 1
= � ei x p/� � dp
2π � − p0 2 p0
1 1 �  i x p0 /� 
= � � e − e−i x p0 /�
2π � 2 p0 i x
#
� sin( p0 x/�)
= .
π p0 x
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 19

Symmetric and Selfadjoint Operators



in H

Q19(1) Show that symmetric operators generate real quadratic


forms, as shown in Eq. (19.2).
SQ19(1) For a symmetric operator A we have, for any φ∀ in the
domain of A,
   
Q A,  ∀ | φ∀ ∪ = �φ∀ | Aφ∀ ∪� = Q A,
 φ∀ = �φ∀ | Aφ∀ ∪ = � Aφ  φ∀ � .
 
Hence Q A, φ∀ is real.
Q19(2) By evaluating �φ∀ | N  ψ∀ ∪ and �N
 φ∀ | ψ∀ ∪ separately verify

that the number operator N in Eq. (19.5) satisfies
�φ∀ | N
 ψ∀ ∪ = �N φ∀ | ψ∀ ∪ ∞φ∀, ψ∀ ⇒ D ∀ (N
 ).
 
SQ19(2) We have, for every φ∀, ψ∀ ⇒ D ∀ N
 ,
 ψ∀ ∪ = �φ∀ | a† a ψ∀ ∪ = � aφ∀ | aψ∀ ∪.
�φ∀ | N
�N φ∀ | ψ∀ ∪ = � a† a φ∀ | ψ∀ ∪ = � aφ∀ | aψ∀ ∪.
 
Hence �φ∀ | N ψ∀ ∪ = �N  φ∀ | ψ∀ ∪. Note that the domain D ∀ N
 which is
   †
smaller than D ∀ a = D ∀ a due Eq. (17.58).1 It follows that a† and
 
a can act on φ∀ and ψ∀ which are in D ∀ N  .


1D �) is worked out in SQ17(6).
N

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)


116 Symmetric and Selfadjoint Operators in H

Q19(3) Show that the number operator N  in Eq. (19.5) is


bounded below. Show also that the square of a selfadjoint operator
is also bounded below.
SQ19(3) We  ∀
 need to show that the quadratic form Q(N , φ ) for
all φ∀ ⇒ D ∀ N ) is bounded below. This is seen in the following
calculation:
�φ∀ | N
 φ∀ ∪ = �φ∀ | a† a φ∀ ∪ = �a φ∀ | a φ∀ ∪ = ||a φ∀ ||2 ⊕ 0.
Hence N  is bounded below by 0 in accordance with Definition
19.1(5).
Similarly the square of a selfadjoint operator is bounded below
by 0, i.e., we have �φ∀ | A 2 ψ∀ ∪ = �Aφ∀ | Aφ∀ ∪ ⊕ 0.
Q19(4) Explain why H  → () cannot act on ϕ∀λ=0, n () in Eq.
D
 → ()ϕ∀λ=0, n ()∪ is undefined.2
(19.34) and that �ϕ∀λ=0, n () | H D

SQ19(4) The vector ϕ∀λ=0, n () is not in the domain of the


Hamiltonian H  → () since ϕ∀λ=0, n () violates the Dirichlet boundary
D
condition in Eq. (17.31). It follows that
 → ()ϕ∀λ=0, n ()∪
�ϕ∀λ=0, n () | H D

is undefined. Formal calculations without reference to this result


is likely to lead to erroneous conclusions. This is demonstrated by
SQ19(5) below.
Q19(5) The eigenvectors ϕ∀ D, → →
ℓ () of H D () in Eq. (19.43) form
an orthonormal basis for L∀ (). We can expression ϕ∀λ=0, n () in
2

Eq. (19.34) as a linear combination of ϕ∀ D→, ℓ (). Consider the vector


ϕ∀λ=0, n=0 () which corresponds to a constant function. We have


ϕ∀λ=0, n=0 () = cℓ ϕ∀ D→, ℓ ().
ℓ=1

Evaluate the coefficients cℓ .


Investigate whether any one of the following two proce­
dures
 → would yield a meaningful
 value for the quadratic form
Q H (), ϕ∀λ=0, n=0 () given formally by
D
 → ()ϕ∀λ=0, n=0 () ∪.
�ϕ∀λ=0, n=0 () | H (�)
D

2 ϕ∀λ=0, n () are the eigenvectors of operator �


pλ=0 () in E19.3(2).
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)


Symmetric and Selfadjoint Operators in H 117

(1) Assume that H  → ()ϕ∀λ=0, n=0 () can be calculated by


D
→

H → ()ϕ∀λ=0, n=0 () = H
 → () cℓ ϕ∀ → ()
D D D, ℓ
ℓ=1


=  → ()ϕ∀ → ()
cℓ H D D, ℓ
ℓ=1


= cℓ E → →
∀ D,
D, ℓ ()ϕ ℓ (), (��)
ℓ=1

where E → →
D, ℓ () are the corresponding eigenvalues of H D () in
Eq. (19.44). Then assume that the expression in Eq. (�) can be
calculated by


�ϕ∀λ=0, n=0 () | cℓ E → →
∀ D,
D, ℓ ()ϕ ℓ ()∪
ℓ=1


= cℓ E → →
∀λ=0, n=0 () | ϕ∀ D,
D, ℓ ()�ϕ ℓ ()∪
ℓ=1


= | c ℓ |2 E →
D, ℓ (). (���)
ℓ=1

Determine whether the above sum converges.


(2) Assume that H  → ()ϕ∀λ=0, n=0 () can be calculated by formal dif­
D
ferentiation using Eq. (19.42). What vector would be obtained?
Is the resulting value of the expression
 → ()ϕ∀λ=0, n=0 ()∪
�ϕ∀λ=0, n=0 () | H D

meaningful?

SQ19(5) The coefficients are given by3


�  L
2 ℓπ x
cℓ = �ϕ∀ D→, ℓ () | ϕ∀λ=0, n=0 ()∪ = sin dx
L 0 L
� $
2  0� if ℓ is even
= 1 − cos ℓπ = .
ℓπ 2 2/ℓπ if ℓ is odd

3 Equation
�(19.34) shows that the function which defines ϕ∀λ=0, n=0 () has a constant
value 1/ L for all x ⇒ , i.e., the function violates the Dirichlet boundary condition
in Eq. (17.31).
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)


118 Symmetric and Selfadjoint Operators in H

(1) The sum in Eq. (���) over odd integers values of ℓ, i.e., ℓ =
1, 3, 5 · · · , diverges, i.e.,
→ →
 � 2
  2 2 π 2 �2 ℓ2
2 →
| cℓ | E D, ℓ () =
ℓ=odd ℓ=odd
ℓπ 2mL2
→
4�2
= = →.
ℓ=odd
mL2

It follows that the above procedure fails to produce a finite value


for the expression in Eq. (�).

(2) Since ϕ∀λ=0, n=0 corresponds to a constant function of value 1/ L
for all x ⇒  a formal differentiation of this constant function
lead to the zero function which would corresponds to the zero
vector in L∀2 (), i.e., we will get

H D

 → ()ϕ∀λ=0, n=0 () = 0(). (����)

This would imply


 → ()ϕ∀λ=0, n=0 ()∪ = 0.
�ϕ∀λ=0, n=0 () | H D

This cannot be correct since Eq. (����) above implies that


 → () corresponding to an
ϕ∀λ=0, n=0 () is an eigenvector of H D
 → () does not have
eigenvalue 0. But Eq. (19.44) tells us that H D
a zero eigenvalue. Another way of looking at this is to realise
that
 → ()ϕ∀λ=0, n=0 ()∪
�ϕ∀λ=0, n=0 () | H D

cannot be smaller than the smallest eigenvalue E →


D, 1 () of
 → () given by Eq. (19.44). This is because
H D

 →

| c ℓ |2 E →
D, ℓ () ⊕ | c ℓ |2 E → →
D, 1 () = E D, 1 ().
ℓ=odd ℓ=odd

The conclusion is that �ϕ∀λ=0, n () | H  → ()ϕ∀λ=0, n ()∪ does


D
not have a meaningful value is due to the fact that an unbounded
operator A cannot act on a vector ψ∀ lying outside its domain. The
sum in Eq. (��) does not define a vector in L∀2 () since the sum does
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)


Symmetric and Selfadjoint Operators in H 119

not give a finite value for the norm of the formal vector expression,
i.e., carrying out formal calculation we get
→ →

� cℓ E →
D, ℓ ∀
()ϕ →
ℓ () | cm E → ∀m→ ()∪
D, m ()ϕ
ℓ=1 m=1

 →
  2
= cℓ� cm E → →
D, ℓ ()E m ()δℓm = |cℓ |2 E →
D, ℓ ()
ℓ, m=1 ℓ
→ →
 � 2 2
   2
 2 2 π 2 �2 ℓ2
= |cℓ |2 E →
D, ℓ () = = →.
ℓ=odd ℓ=odd
ℓπ 2mL2
So the expression in Eq. (���) would not produce a meaningful result
due to the fact that Eq. (��) fails to define a vector. Any attempt to get
round this would fail to produce any meaning result.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 20

Spectral Theory of Selfadjoint Operators



in H

Q20(1) Show that the spectral functions of a projector P , of the


zero operator 
0 and of the identity operator II are given, respectively
by1
!

0 τ <0
F (τ ) = 

.
II τ ⊕ 0
!

0 τ <1
 IÎ
F (τ ) =  .
II τ ⊕ 1


⎨0 τ <0
 P̂  
F (τ ) = II − P 0 ≤ τ < 1 .

⎩ II τ ⊕1
SQ20(1) We are dealing with operators with a discrete spectrum.
The spectral functions of these operators are therefore of the form
of Eq. (20.19), i.e., it is piecewise constant, being 
0 at −→ and II at
→ and with discontinuities at the eigenvalues.
(1) The zero operator  0 has one eigenvalue, i.e., 0. This means that
F 0̂ (τ ) = 
0 for τ < 0 and at τ = 0 the spectral function would
1 Weidmann p. 195. Wan (2006) p. 152.

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)


122 Spectral Theory of Selfadjoint Operators in H

jump to a value F 0̂ (0). The function remains constant for all τ ⊕


0 as there are no eigenvalues above 0. It follows that F 0̂ (0) is
equal to its value at → which is II . In other words we have
!

0 τ <0
F (τ ) = 

.
II τ ⊕ 0

(2) The identity operator II has one eigenvalue, i.e., 1. This means
that F IÎ (τ ) = 
0 for τ < 1 and at τ = 1 the spectral function
would jump to a value F IÎ (1). The function remains constant for
all τ ⊕ 1 as there are no eigenvalues above 1. It follows that
F 0̂ (1) is equal to its value at → which is II . In other words we
have !

0 τ <1
 IÎ
F (τ ) =  .
II τ ⊕ 1

(3) The projector operator P has two eigenvalue, i.e., 0 and 1. This
means that F P̂ (τ ) =  0 for τ < 0 and at τ = 0 the spectral
function would jump to a value F P̂ (0). The function F P̂ (τ )
remains constant for all τ ⊕ 1 as there are no eigenvalues above
1, i.e., F P̂ (1) is equal to its value at → which is II . So we have
F P̂ (τ ) = 0 for τ < 0 and F P̂ (τ ) = II for τ ⊕ 1.
The function F P̂ (τ ) for τ ⇒ [0, 1) can be determined as follows:
(a) Being a selfadjoint operator the projector P has a spectral
decomposition in the form of Eq. (20.20). Now let the
eigenvalues of P be denoted by a1 and a2 , i.e., a1 = 0 and
a2 = 1, and let the eigenprojectors associated with these
two eigenvalues be denoted by P P̂ (a1 ) and P P̂ (a2 ). These
eigenprojectors are related to the spectral function F P̂ (τ ),
i.e., we have, by Eq. (15.10),

P P̂ (a1 ) = F P̂ (a1 ) − F P̂ (a1 − 0),


P P̂ (a2 ) = F P̂ (a2 ) − F P̂ (a2 − 0).
The spectral decomposition of P becomes2
P = a1 P P̂ (a1 ) + a2 P P̂ (a2 ) = 0P P̂ (0) + 1 P P̂ (1)
= F P̂ (1) − F P̂ (1 − 0).
2 Note that F� P̂ (1 − 0) is meant to be F� P̂ (a2 − 0) where a2 = 1.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)


Spectral Theory of Selfadjoint Operators in H 123

Since F P̂ (1) = II we get


F P̂ (1 − 0) = II − P.
Since F P̂ (τ ) is constant for τ ⇒ [0, 1) we conclude that
F P̂ (τ ) = II − P ∞τ ⇒ [0, 1).
(b) Another way to look at the problem is to note that at
τ = 1 the increment of the spectral function is equal to
the eigenprojector associated with the eigenvalue 1. From
Eqs. (20.16) and (20.19) we can see that this eigenprojector
is P itself. It follows that

F P̂ (1) − F P̂ (1 − 0) = P � F P̂ (1 − 0) = II − P.


To sum up we have


⎨0 τ <0

F (τ ) = II − P

0≤τ <1 .

⎩ II τ ⊕1
Q20(2) Show that on an interval  of IR the spectral measure
 P̂ () of a projector P is related to P by3
M

⎪ II − P if  = {0}


M () = P

if  = {1} .

⎩0 if  does not contain 0 or 1

SQ20(2) For the singleton set {0} and {1} we get, by Eq. (15.20),4
 P̂ ({0}) =
M F P̂ (0) − F P̂ (0 − 0) = II − P ,
 P̂ ({1}) =
M F P̂ (1) − F P̂ (1 − 0) = P .
The same results can be obtained directly from Eq. (20.19).
Finally, since the spectral function is constant over any interval 
not containing 0 or 1, we get M  P̂ () = 
0 for such intervals.

Q20(3) Prove that Theorem 9.4.4(1) remains valid for a selfad­


joint operator with a discrete spectrum in an infinite-dimensional
Hilbert space.
3 Wan (2006) p. 152. See Q28(2) for an application.
4 As in SQ20(1) the notation F� P̂ (0−0) is the same as F� P̂ (a1 −0), a1 = 0, and F� P̂ (1−0)
is the same as F� P̂ (a2 − 0), a2 = 1.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)


124 Spectral Theory of Selfadjoint Operators in H

SQ20(3)
(1) Let ϕ∀ be an eigenvector associated with an eigenvalue a of a
selfadjoint operator A. We have
�ϕ∀ | Aϕ∀ ∪ = �Aϕ∀ | ϕ∀ ∪ � �ϕ∀ | aϕ∀ ∪ = �aϕ∀ | ϕ∀ ∪
� a�ϕ∀ | ϕ∀ ∪ = a� �ϕ∀ | ϕ∀ ∪
� a = a� � a ⇒ IR .

(2) Let ϕ∀1 and ϕ∀2 be two eigenvectors associated with eigenvalue a1
and a2 respectively. We have
�Aϕ∀1 | ϕ∀2 ∪ = �ϕ∀1 | Aϕ∀2 ∪
� �a1 ϕ∀1 | ϕ∀2 ∪ = �ϕ∀ | a2 ϕ∀2 ∪
� ( a1 − a2 )�ϕ∀1 | ϕ∀2 ∪ = 0
� �ϕ∀1 | ϕ∀2 ∪ = 0 if a1 =
� a2 .

(3) When the eigenvalues are all nondegenerate the spectral


decomposition of the identity in Eq. (20.20) becomes II =


ℓ P ϕ∀ℓ as shown in Eq. (20.21) which, as pointed out in
P20.3(4), shows the completeness of the eigenvectors, i.e., the
eigenvectors form an orthonormal basis.
(4) Suppose the eigenvalue a1 is degenerate and all other eigen­
values are nondegenerate. Let S∀(a1 ) be the eigensubspace
associated with this degenerate eigenvalue a1 with degeneracy
d1 which can be infinite. We can choose an orthonormal
basis ϕ∀1 j , j = 1, 2, · · · , d1 , for the subspace S∀(a1 ). These are
eigenvectors of the operator corresponding to the eigenvalue a1
and they are orthogonal to the eigenvectors ϕ∀ℓ= � 1 corresponding
to eigenvalues to aℓ�=1 . Then the eigenprojector onto the
eigensubspace S∀(a1 ) is
d1

P Â (a1 ) = P ϕ∀1 j ,
j =1

where all the projectors P ϕ∀1 j are one-dimensional. The spectral


decomposition of the identity in Eq. (20.20) can be written as

II = P Â (a1 ) + P ϕ∀ℓ .
ℓ�=1
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)


Spectral Theory of Selfadjoint Operators in H 125

It follows that the corresponding eigenvectors ϕ∀1 j and ϕ∀ℓ=� 1


form a complete set, i.e., they form an orthonormal basis. This
set is not unique since ϕ∀1 j are not unique. The analysis can
be extended to apply to an operator with many degenerate
eigenvalues.
(5) The projectors generated by a complete orthonormal set of
eigenvectors form a complete orthogonal family of projectors.

Q20(4) In L∀2 (IR) show that for every φ(x) ⇒ C c→ (IR) we have
d f (x)
[ f ( (IR) ]φ∀ := i �
x ), p φ(x).
dx
SQ20(4)
 d d 
[ f ( (IR)]φ∀ := −i � f (x)
x ), p − f (x) φ(x)
dx dx
 d  d f (x) d 
= −i � f (x) − + f (x) φ(x)
dx dx dx
d f (x)
= i� φ(x).
dx
Q20(5) Show that selfadjoint operators having a discrete spec­
trum are reducible by their eigensubspaces.
SQ20(5) A selfadjoint operator with a discrete spectrum com­
mutes with its eigenprojectors. Let P Â (am ) be the projector onto
the eigensubspace S∀ Â (am ) associated with the eigenvalue am of a
selfadjoint operator A. Then P Â (am ) commutes with A. Theorem
17.9(1) then tells us that A is reducible by S∀ Â (am ).
Q20(6) Show that in L∀2 (IR) the position operator  x (IR) is
(IR) is also reducible.5
reducible and the momentum operator p
SQ20(6) Generally a selfadjoint operator commutes with its
spectral projectors. For the position operator x in L∀2 (IR) its spectral
projector associated with an interval  is given by Eq. (20.28), i.e.,
χ  . The subspace S∀x̂ () associated with the projector is defined
by all the functions in L2 (IR) vanishing outside the interval . Since
χ  commute with x Theorem 17.9(1) tells us x is reducible by the
5 This
is in contrast to the fact discussed in E20.7(1) that together the position and
momentum operators form an irreducible set.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)


126 Spectral Theory of Selfadjoint Operators in H

subspace S∀x̂ (). The same analysis in the momentum space shows
that the momentum operator is similarly reducible.
x in L∀2 (IR ) constitutes a
Q20(7) Show that the position operator 
complete set of selfadjoint operators.
SQ20(7)6 The proof presented below is based on an application of
Theorem 20.6(1), i.e., we want to show that any bounded operator
on L∀2 (IR ) commuting with  x is a function of x.
Let φ(x ) be a function in L2 (IR ) and let ψ(x ) be a real positive
function in L2 (IR ), i.e., ψ(x ) > 0 for all x . Let φ∀ and ψ∀ be the vectors
in L∀2 (IR ) defined by φ(x ) and ψ(x ). We can introduce a function
G(x ) = φ(x )/ψ(x ) since ψ(x ) > 0 for all x . This function G(x )
defines a multiplication operator G in L∀2 (IR ) by
φ(x )
Gϕ∀ := ϕ(x ),
ψ(x )

where the vector ϕ∀ := ϕ(x ) ⇒ L2 (IR ). We have


φ(x )
φ∀ = Gψ∀ since Gψ∀ := ψ(x ) = φ(x ).
ψ(x )
The operator G is seen to be a multiplication operator defined by the
function G(x ) := φ(x )/ψ(x ). As such a multiplication operator G is
a function of the position operator  x , i.e., G = G(
x ). Note that once
2
a real positive function ψ(x ) ⇒ L (IR ) is chosen we can define an
operator G for every function φ(x ) in L2 (IR ).
 be a bounded operator on L∀2 (IR ) which commutes with the
Let B
position operator   commutes with G since G is a function
x . Then B
∀  ∀
x . Let  = B ψ and let (x ) be the function in L2 (IR ) which
of 
corresponds to the vector  ∀ , i.e., 
∀ =B  ψ∀ := (x ). Then

  φ(x ) (x )
 φ∀ = B
B  Gψ∀ = G B ∀ :=
 ψ∀ = G  (x ) = φ(x ).
ψ(x ) ψ(x )
 is a multiplication operator corresponding to the
It follows that B
function f (x ) = (x )/ψ(x ). By Definition 20.6(3) the position
operator constitutes a complete set of selfadjoint operators in
L∀2 (IR ).

6 Jordan p. 58.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 21

Spectral Theory of Unitary Operators



on H

Q21(1) Let U be a unitary operator in the Hilbert space H ∀ with


a spectral decomposition given by Eq. (21.4). Let φ∀ be an arbitrary
vector in H∀ . Show that

U φ∀ = cℓ ei aℓ ϕ∀ℓ , cℓ = �ϕ∀ℓ | φ∀∪.

SQ21(1) Using Pϕ∀ℓ φ∀ = cℓ ϕ∀ℓ , cℓ = �ϕ∀ℓ | φ∀ ∪ we get



→  →
   →

U φ∀ = e i aℓ
P ϕ∀ℓ φ∀ = e i aℓ
P ϕ∀ℓ φ∀ = cℓ ei aℓ ϕ∀ℓ .
ℓ=1 ℓ=1 ℓ=1

¨
Q21(2) Equation (10.27), which is a time dependent Schrodinger
equation, can be written in the form of Eq. (21.18) which follows
from Stone’s theorem as a vector equation in the Hilbert space
L∀2 (IR), i.e.,
dφ∀(t)  φ∀(t),
i� =H
dt

where φ∀(t) is in the domain of H  . Explain why the norm of vector


φ∀(t) is preserved in time, i.e., ||φ∀(t1 )|| = ||φ∀(t2 )|| for any t1 and t2 .

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)


128 Spectral Theory of Unitary Operators on H

SQ21(2) The solution of Eq. (21.20) for φ∀(0) := φ∀(t = 0) is easily


verified to be
�t
−iH
φ∀(t) = U (t)φ∀(0), U (t) = e ¯ ,
since
dφ∀(t) dU (t) ∀  φ∀(t).
i� = i� φ (0) = H
dt dt
In other words φ∀(t) is a unitary transform of φ∀(0). A unitary
transform preserves the norm of vectors, i.e., ||φ∀(t) || = ||φ∀(0) ||.
Q21(3) Consider the following unitary transformations of an
operator A:
A(t) = U † (t)AU (t),

− i tH �
where U (t) = e ¯ is a one-parameter family of unitary operators
in Eq. (21.13).1 Show that2
d A(t)
i� = [A(t), H
 ].
dt
SQ21(3)
d A(t)  dU †(t) dU (t) 
i� = i� A U (t) + U †(t) A
dt dt dt
  †    † 
= −H U (t) AU (t) + U (t) A H U (t)  
= −H  A(t) + A(t)H  = [A(t), H  ].

1 The operators � and H


A � are independent of t and A
�(t = 0) = A
�.
2 See Eq. (29.19) for the physical relevance of this result.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 22

Selfadjoint Operators, Unit Vectors and


Probability Distributions

Q22(1) The commutation relation between the selfadjoint posi­


tion and momentum operators  x = x and p
 = −i � d /dx in the
Hilbert space L∀2 (IR ) is often written as

[  ] = i �.
x, p

As pointed out in the discussion in §17.7 this relation should be


expressed as an inequality, i.e.,

[  ] ≥ i � II ,
x, p

or

[  ] φ∀ = i � φ∀
x, p (�)

for an appropriate set of vectors φ∀ in L∀2 (IR ). What are the conditions
φ∀ must satisfy in order for the equality to hold?
 
SQ22(1) What we want is the domain D ∀ [  ] of the com-
x, p
mutator [  ]. The domains D
x, p ∀ (
x ) and D ∀(p ) of the position and
momentum operators are given by Eqs. (17.13) and (17.49). The
commutator is  − p
xp x . So we must specify the domains of  
xp

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

130 Selfadjoint Operators, Unit Vectors and Probability Distributions


and p x first in terms that of D ∀ (
x ) and D ∀(p ), and then specify the
domain of the commutator. We have, by Eq. (17.58),
 
D∀ (
xp ∀(p
 ) := φ∀ ⇒ L∀(IR) : φ∀ ⇒ D  ), p ∀ (
 φ∀ ⇒ D x) .
 
∀(p
D  ∀ (
x ) := φ∀ ⇒ L∀(IR) : φ∀ ⇒ D x φ∀ ⇒ D
x ),  ∀(p) .
Finally we have, by Eq. (17.56), the domain
 
D∀ [ ] = D
x, p ∀ (
xp) ⊥ D∀(p
 x ).
 
∀ ∀
Vectors φ in Eq. (�) must belong to D [ ] .
x, p
Q22(2) Consider a particle confined in an infinite potential well
of width [0, L]. All its wave functions φ(x) must vanish outside the
well, i.e., φ(x) = 0 ∞x ⇒ / [0, L]. Hence the state space of the trapped
particle is taken to be L∀2 (),  = [0, L] rather than L∀2 (IR).

(a) Taking the operator 


x () in Eq. (17.22) as the position operator
show that the uncertainty in position cannot be bigger than the
width of the well.
λ=0 () in Eq. (17.36) as the momentum
(b) Taking the operator p
operator show that the momentum uncertainty
 
 p λ=0 (), ϕ∀λ=0, n ()
is zero. Here ϕ∀λ=0, n () are eigenvectors of p
λ=0 () in Eq.
(19.34).
(c) Bearing in mind the above results investigate whether or not
an uncertainty relation similar to that shown in Eq. (22.28)
λ=0 ().1
remains valid for eigenvectors of p

SQ22(2)(a) Let φ∀ be a unit vector in L∀2 () defined by a


normalised function φ(x) ⇒ L2 (). For the infinite the potential well
of width L from x = 0 to x = L we have x ≤ L. We have
 L  L

�φ |  2∀
x ( ) φ ∪ = 2 2
x |φ(x)| dx ≤ L 2
|φ(x)|2 dx = L2 .
0 0
It follows from Eq. (22.4) that
x ( ), φ∀ )2 = �φ∀ | 
( x ( )2 φ∀ ∪ − �φ∀ | 
x ( )φ∀ ∪2
x ( )2 φ∀ ∪ ≤ L2
≤ �φ∀ | 
� ( x ( ), φ∀ ) ≤ L.
1 Fanopp. 407–408 for a similar problem with the uncertainty relation in the Hilbert
space L∀2 (Ca ). See also §28.3.3 on a particle in circular motion.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Selfadjoint Operators, Unit Vectors and Probability Distributions 131

SQ22(2)(b) For an eigenvector ϕ∀λ=0, n () of p λ=0 () correspond­


ing to the eigenvalue pλ=0, n () in Eq. (19.35) we have
λ=0 ()ϕ∀λ=0, n ()∪ = pλ=0, n (),
�ϕ∀λ=0, n () | p
 2
λ=0 ()2 ϕ∀λ=0, n ()∪ = pλ=0, n () .
�ϕ∀λ=0, n () | p
It follows that
 
 p λ=0 ( )2 ϕ∀λ=0, n ()∪
λ=0 ( ), ϕ∀λ=0, n () = �ϕ∀λ=0, n () | p
λ=0 ( )ϕ∀λ=0, n ()∪2
−�ϕ∀λ=0, n () | p
= 0,
a result which is intuitively obvious.
SQ22(2)(c) The results in SQ22(2)(a) and SQ22(2)(b) imply
   
  x ( ), ϕ∀λ=0, n ()  pλ=0 ( ), ϕ∀λ=0, n () = 0.

This result violates the inequality in Eq. (22.28), i.e., the inequality
cannot be applied to eigenvectors of p λ=0 (). The reason for
this violation lies in the domain of the commutator discussed in
SQ22(1), i.e., the commutator [  λ=0 () ] is not defined on the
x ( ), p
eigenvector ϕ∀λ=0, n () in Eq. (19.34). To see this let the commutator
acts on a vector φ∀ ⇒ L∀2 (), i.e., we have
[ λ=0 () ]φ∀ = 
x ( ), p λ=0 () φ∀ − p
x () p x () φ∀.
λ=0 () 
For above operation to be meaningful we require that:
(1) The vector φ∀ must in the domain of p λ=0 (), i.e., apart from
differentiability the corresponding function φ(x) ⇒ L2 () must
satisfy the periodic boundary condition φ(0) = φ(L).
(2) The function xφ(x) corresponding to the vector  x ()φ∀ must
also satisfy the periodic boundary condition.
Clearly the function xϕλ=0, n ()(x) corresponding to the vector
x ( ) ϕ∀λ=0, n (), i.e.,

% &
x 2nπ
� exp i x ,
L L
does not satisfy the periodic boundary condition. This means
x ( ) ϕ∀λ=0, n () is not in the domain of p
that the vector  λ=0 ().
Consequently pλ=0 () cannot act on  x () ϕ∀λ=0, n (). It follows that
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

132 Selfadjoint Operators, Unit Vectors and Probability Distributions

the commutator cannot act on ϕ∀λ=0, n (), which in turn implies that
the inequality in Eq. (22.28) cannot be applied to ϕ∀λ=0, n ().
A similar problem arises with the uncertainty relation in the
Hilbert space L∀2 (Ca ) for a particle in circular motion. This is
discussed in §28.3.2.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 23

Physics of Unitary Transformations

Q23(1) Explain why a coordinate representation space and its


corresponding momentum representation space introduced in
§18.4.2 give rise to two mathematically different descriptions of the
position and momentum which are physically equivalent.
SQ23(1) The coordinate and momentum representations are
related by a Fourier transformation which is a unitary transforma­
tion. The two representations are therefore physically equivalent.
Take the example of a particle in one-dimensional motion. The
representation space is L∀2 (IR ) in the coordinate representation. The
position and momentum operators  x (IR ) and p (IR ) are defined
by Eqs. (17.12), (17.14), (17.49) and (17.50). The representation
space is L∀2 (IR)

in the momentum representation. The position
and momentum operators  x⊗ (IR)

and p (IR)

in the momentum

representation are defined by Eqs. (18.58) and (18.61). The
quantities in two representation spaces are related by a Fourier
transformation, i.e., from Eqs. (18.54), (18.57) and(18.60),
ϕ∀ = U F ϕ∀, ϕ∀ ⇒ L∀2 (IR ), ϕ∀ ⇒ L∀2 (IR),
⊗ ⊗ ⊗


x⊗ (IR)

x (IR ) U F−1 , ⊗p
= U F  (IR)

(IR ) U F−1 ,
= U F p
where U F is the Fourier transform operator defined by Eq. (18.54).
Since U F is unitary all the measured values of physical quantities,

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

134 Physics of Unitary Transformations

e.g., eigenvalues, expectation values, probabilities, are the same in


the two representations on account of Eq. (23.2), e.g.,
x (IR ), ϕ∀ ) = Q(
Q( x⊗ (IR),

ϕ∀), Q( p

(IR ), ϕ∀ ) = Q( p

(IR),

ϕ∀).

The two representations are therefore physically equivalent.


June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 24

Direct Sums and Tensor Products of


Hilbert Spaces and Operators

Q24(1) Consider a three-dimensional Hilbert space H ∀ with a


preferred decomposition as the direct sum of a complete orthogonal
family of one-dimensional subspaces S∀ (−) , S∀ (0) and S∀ (+) , i.e.,
∀ =H
H ∀ � = S∀ (−) � S∀ (0) � S∀ (+) .

Let η∀ (−) , η∀ (0) and η∀ (+) be unit vectors in S∀(−) , S∀(0) and S∀(+) ,
respectively.1 A vector in H ∀ � is of the form

η∀ � = c− η∀ (−) � c0 η∀ (0) � c+ η∀ (+)


= c− η∀ (−)� + c0 η∀ (0)� + c+ η∀ (+)� ,
where c− , c0 , c+ ⇒ C .
∀ � are
(a) Show that selfadjoint decomposable operators A� on H
diagonalisable and of the form
A� = a− II (−) � a0 II (0) � a+ II (+) , a− , a0 , a+ ⇒ IR .
What are the eigenvalues and eigenvectors of A� ?

1 The notation such as η


∀ (−)� ∀ (0) �0
follows that of Eq. (24.23), i.e., η∀ (−)� = η∀ (−) �0 ∀ (+)
∀ �.
which is a vector in H

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

136 Direct Sums and Tensor Products of Hilbert Spaces and Operators

(b) Define three operators 


L− , 
L+ and  ∀ � by2
L on H


L− η∀ (−)� = η∀ (0)� , 
L− η∀ (0)� = η∀ (−)� ,  ∀ �.
L− η∀ (+)� = 0

L+ η∀ (+)� = η∀ (0)� , 
L+ η∀ (0)� = η∀ (+)� ,  ∀�.
L+ η∀ (−)� = 0

L= 
L− + 
L+ .

Show that these operators are selfadjoint but not decomposable.3


SQ24(1)(a) By Definition 24.1.2(2) a selfadjoint decomposable
∀ � is of the form
operator A� on H

A� = A(−) � A(0) � A(+) ,

where A(−) , A(0) , A(+) are sefladjoint operators defined on the


spaces S∀(−) , S∀(0) , S∀(+) respectively. Since these spaces are one-
dimensional A(−) , A(0) , A(+) must be proportional to the identity
operators in the spaces, i.e.,
A(−) = a− II (−) , A(0) = a0 II (0) , A(+) = a+ II (+) ,
where a1 , a2 , a3 real numbers. The eigenvalues of A� are a1 , a2 , a3
corresponding respectively to eigenvectors
η∀ (−)� = η∀ (−) � 0 ∀ (+) , η∀ (0)� = 0
∀ (0) � 0 ∀ (−) � η∀ (0) � 0
∀ (+) ,

∀ (−) � 0
η∀ (+)� = 0 ∀ (0) � η∀ (+) .

SQ24(1)(b) We can appreciate the action of the operators 


L− and

L+ as follows:
(1)  L− interchanges η∀ (−)� and η∀ (0)� while annihilating η∀ (+)� , i.e.,
L− η∀ (−)� = η∀ (0)� , 
 L− η∀ (0)� = η∀ (−)� ,  ∀ �.
L− η∀ (+)� = 0
(2)  L+ interchanges η∀ (+)� and η∀ (0)� while annihilating η∀ (−)� , i.e.,
L+ η∀ (+)� = η∀ (0)� , 
 L+ η∀ (0)� = η∀ (+)� ,  ∀ �.
L+ η∀ (−)� = 0
2 Wan (2006) p. 356. These operators are not the direct sums of operators on
S∀(−) , S∀(0) and S∀(+) . Hence they are not denoted with a superscript �. Here 0∀�
is the zero operator on H ∀ �.
3 See §32.3 and §34.7 for physical applications and Q32(3) for a similar operator in a

two-dimensional space.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Direct Sums and Tensor Products of Hilbert Spaces and Operators 137

We need to verify the selfadjointness condition on the basis


vectors η∀ (−)� , η∀ (0)� , η∀ (+)� in H ∀ � , e.g., for  L− we need to verify
 (ℓ� )�  (ℓ� )�
that �η∀ (ℓ)�
| L− η∀ ∪ = � L− η∀ (ℓ)�
| η∀ ∪ ∞ℓ, ℓ� , where ℓ, ℓ�
stand for −, 0, +. These scalar products can be evaluated using the
orthonormal properties of the basis vectors, i.e., we get

�η∀ (−)� | 
L− η∀ (−)� ∪ = 0, �
L− η∀ (−)� | η∀ (−)� ∪ = 0.
�η∀ (−)� | 
L− η∀ (0)� ∪ = 1, �
L− η∀ (−)� | η∀ (0)� ∪ = 1.

�η∀ (−)� | 
L− η∀ (+)� ∪ = 0, �
L− η∀ (−)� | η∀ (+)� ∪ = 0.
�η∀ (0)� | 
L− η∀ (−)� ∪ = 1, �
L− η∀ (0)� | η∀ (−)� ∪ = 1.

�η∀ (0)� | 
L− η∀ (0)� ∪ = 0, �
L− η∀ (0)� | η∀ (0)� ∪ = 0.
�η∀ (0)� | 
L− η∀ (+)� ∪ = 0, �
L− η∀ (0)� | η∀ (+)� ∪ = 0.
�η∀ (+)� | 
L− η∀ (−)� ∪ = 0, �
L− η∀ (+)� | η∀ (−)� ∪ = 0.

L− η∀ (0)� ∪ = 0, �
�η∀ (+)� |  L− η∀ (+)� | η∀ (0)� ∪ = 0.
�η∀ (+)� | 
L− η∀ (+)� ∪ = 0, �
L− η∀ (+)� | η∀ (+)� ∪ = 0.

Similarly we have

�η∀ (−)� | 
L+ η∀ (−)� ∪ = 0, �
L+ η∀ (−)� | η∀ (−)� ∪ = 0.

�η∀ (−)� | 
L+ η∀ (0)� ∪ = 0, �
L+ η∀ (−)� | η∀ (0)� ∪ = 0.
�η∀ (−)� | 
L+ η∀ (+)� ∪ = 0, �
L+ η∀ (−)� | η∀ (+)� ∪ = 0.

�η∀ (0)� | 
L+ η∀ (−)� ∪ = 0, �
L+ η∀ (0)� | η∀ (−)� ∪ = 0.
�η∀ (0)� | 
L+ η∀ (0)� ∪ = 0, �
L+ η∀ (0)� | η∀ (0)� ∪ = 0.

�η∀ (0)� | 
L+ η∀ (+)� ∪ = 1, �
L+ η∀ (0)� | η∀ (+)� ∪ = 1.
�η∀ (+)� | 
L+ η∀ (−)� ∪ = 0, �
L+ η∀ (+)� | η∀ (−)� ∪ = 0.

�η∀ (+)� | 
L+ η∀ (0)� ∪ = 1, �
L+ η∀ (+)� | η∀ (0)� ∪ = 1.

�η∀ (+)� | 
L+ η∀ (+)� ∪ = 0, �
L+ η∀ (+)� | η∀ (+)� ∪ = 0.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

138 Direct Sums and Tensor Products of Hilbert Spaces and Operators

These results shows that  L− and  L+ are selfadjoint. It follows that



L= L− +  L+ is selfadjoint.
These operator is not decomposable. The subspaces S∀ (−)� , S∀ (0)�
and S∀ (+0)� are not invariant under any of these operators, e.g.,4

L− η∀ (−)� = η∀ (0)� ⇒/ S∀(−)� ,

L+ η∀ (+)� = η∀ (0)� ⇒/ S∀(+)� ,
L− + 
( / S∀(0)� .
L+ )η∀ (0)� = η∀ (−)� + η∀ (+)� ⇒

Q24(2) Show that the tensor products of two projectors, P (1) in


Hilbert space H ∀ (2) , is a projector in
∀ (1) and P(2) in Hilbert space H
∀ ∗ ∀
the tensor product space H = H ∗ H .(1) ∀ (2)

SQ24(2) On account of Eqs. (24.52) and (24.53) the tensor


product P 1 ∗ P2 is idempotent and selfadjoint in the tensor product
∀ ∗ . It is therefore a projector on H
space H ∀ ∗.

Q24(3) Consider the tensor product H ∀∗ = H ∀ ∗H ∀ . Let {ϕ∀m } be an



orthonormal basis for H. Then {ϕ∀m ∗ ϕ∀n } is an orthonormal basis for
H∀ ∗ . Define the permutation operator U p on H ∀ ∗ by5
  
U p cmn ϕ∀m ∗ ϕ∀n := cmn ϕ∀n ∗ ϕ∀m .
m, n m, n

(a) Show that this is a bounded operator with H ∀ ∗ as its domain.


(b) Show that the square of U p is equal to the identity, and U p is
unitary and selfadjoint, i.e.,
U p2 = II , U p† = U p−1 = U p .
(c) For two bounded operators A and B ∀ ∗ show that6
 on H
 
U p A ∗ B
 U † = B
p
 ∗ A.
(d) Show that
1   1  
P(s) := II + U p , P (a) := II − U p ,
2 2
are projectors which are orthogonal to each other. Find
examples of vectors in H ∀ ∗ which are unchanged by each of
these two projectors.

4 See the notation in Eq. (24.21).


5 See §33.3 for physical applications of these operators.
6 See Eq. (33.15) in §33.3.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Direct Sums and Tensor Products of Hilbert Spaces and Operators 139

(e) Find the eigenvalues and eigenvectors of Up .

∀ ∗ can be written as
∀ ∗ in H
SQ24(3)(a) An arbitrary vector 

∀∗ =
 cmn ϕ∀m ∗ ϕ∀n .
m, n

Then
 
� cmn ϕ∀m ∗ ϕ∀n | cm� n� ϕ∀m� ∗ ϕ∀n� ∪
m, n m� , n�

= ∀ ∗ ||2 < →.
|cmn |2 = || 
m, n

Since
   
Up cmn ϕ∀m ∗ ϕ∀n = cmn ϕ∀n ∗ ϕ∀m ,
m, n m, n

we have
 
||Up cmn ϕ∀m ∗ ϕ∀n ||2
m, n
 
=� cmn ϕ∀n ∗ ϕ∀m | cm� n� ϕ∀n� ∗ ϕ∀m� ∪
m, n m� , n�

= |cmn | = ||  ||2 .
2 ∀∗
m, n

Hence the operator Up preserves the norm and it is defined on every
∀ ∗ . Hence it is bounded operator of the norm ||Up || = 1.
vector in H

SQ24(3)(b) Given an arbitrary vector  ∀ � = m, n cmn ϕ∀m ∗ ϕ∀n ⇒
H∀ ∗ we can establish the following results:

(1) The square of Up is equal to itself, since


 
U p2 cmn ϕ∀m ∗ ϕ∀n
m, n

 
= Up Up cmn ϕ∀m ∗ ϕ∀n
m, n
 
= Up cmn ϕ∀n ∗ ϕ∀m
m, n

= cmn ϕ∀m ∗ ϕ∀n � U p2 = II
.
m, n
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

140 Direct Sums and Tensor Products of Hilbert Spaces and Operators

(2) From SQ24(a) we know that Up preserves the norm of vectors
∀ ∗ . It is invertible by Theorem 17.5(1), since
in H

Up  ∀∗
∀∗ = 0 � ∀ ∗ || = ||
||Up  ∀ ∗ || = 0
∀∗.
∀∗ = 0
�

Its inverse is equal to itself, i.e., U p−1 = Up , since U p2 = II


.
(3) The adjoint of the operator is equal to its inverse since U p† Up =
Up U p† = II
 . To prove this result we have, for all  ∀ ∗,

� ∀ ∗ ∪ = �
∀∗ |  ∀ ∗ | Up Up 
∀ ∗ ∪ = �Up† 
∀ ∗ | Up 
∀∗∪

∀∗ | 
= �Up Up†  ∀∗∪

 by Eq. (18.2).
� Up U p† = II
∀∗ | 
� ∀ ∗ | Up 
∀ ∗ ∪ = �Up  ∀ ∗ ∪ = �U p† Up 
∀∗ | 
∀∗∪

.
� U p† Up = II

(4) Since the operator has been shown to be bounded, preserve the
norm of all vectors, invertible with its inverse equal its adjoint it
is therefore unitary by Theorem 18.3(1).

∀ � ∗ = m� , n� cm
(5) For selfadjointness let  �
∀m� ∗ ϕ∀n� . Then:
� n� ϕ

 
� ∀ �∗∪ = �
∀ ∗ | Up  cmn ϕ∀m ∗ ϕ∀n | �
cm ∀n� ∗ ϕ∀m� ∪
� n� ϕ

m, n m� , n�

� �
= cmn cm � n� δmn� δnm� .

m, n, m� , n�
 
�Up  ∀ �∗∪ = �
∀∗ |  cmn ϕ∀n ∗ ϕ∀m | �
cm ∀m� ∗ ϕ∀n� ∪
� n� ϕ

m, n m� , n�

� �
= cmn cm � n� δnm� δmn� .

m, n, m� , n�

∀ � ∗ ∪ = �Up 
∀ ∗ | Up 
It follows that � ∀ � ∗ ∪ which implies the
∀∗ | 
selfadjointness of Up .

From Up = U p−1 we get Up = Up† = U p−1 . All this confirms the
previous conclusion that Up is unitary by Theorem 18.3(1).
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Direct Sums and Tensor Products of Hilbert Spaces and Operators 141


∀∗ =
SQ24(3)(c) Given an arbitrary vector  m, n cmn ϕ∀m ∗ ϕ∀n in
H∀ we have

    ∗    
Up A ∗ B p

 U †  = p A ∗ B
U  c mn ∀
ϕ n ∗ ∀
ϕ m
m, n
  
= Up cmn Aϕ∀n ∗ B
 ϕ∀m
m, n
  
= cmn B ϕ∀m ∗ Aϕ∀n
m, n
   
 ∗ A
= B cmn ϕ∀m ∗ ϕ∀n .
m, n

SQ24(3)(d) The operators P (s) and P (a) are clearly selfadjoint


on account of Eq. (17.96) and the selfadjointness of Up . Using the
property U p2 = II we have

 1  2
2 
P (s) II + 2IIUp + U p2 = P (s) .
=
4
 (a) 2 1  2 
P = II − 2IIUp + U p2 = P (a) .
4
Hence P and P are projector. They are orthogonal since
(s) (a)

1  2  
P(s) P (a) = II − U p + Up − Up2 =  0.
4
The followings are examples of vectors unchanged by these
projectors:
(1) Vectors of the form
  
∀ ∗s =
 cmn ϕ∀m ∗ ϕ∀n + ϕ∀n ∗ ϕ∀m
m, n

are unchanged by P(s) since


1   
∀ ∗s =
P (s)  cmn ϕ∀m ∗ ϕ∀n + ϕ∀n ∗ ϕ∀m
2 m, n
  
+ cmn ϕ∀n ∗ ϕ∀m + ϕ∀m ∗ ϕ∀n
m, n
  
= ∀ ∗s .
cmn ϕ∀m ∗ ϕ∀n + ϕ∀n ∗ ϕ∀m = 
m, n
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

142 Direct Sums and Tensor Products of Hilbert Spaces and Operators

These vectors are shown in Eq. (33.19) and they are said to be
symmetrical in Definition 33.3.1(1).
(2) Vectors of the form
  
∀ ∗a =
 cmn ϕ∀m ∗ ϕ∀n − ϕ∀n ∗ ϕ∀m
m, n

are similarly unchanged by P (a) . These vectors are shown in


Eq. (33.20) and they are said to be antsymmetrical in Definition
33.3.1(1).
SQ24(3)(e) Consider the eigenvalue equation of Up , i.e., Up  ∀ λ∗ =
λ∀ λ , where λ ⇒ IR . Then U p 
∗  2 ∀λ = λ 
∗ 2 ∀ λ . Since U p = II and
∗  2 
 
the eigenvalues of U p are real (since U p is selfadjoint) we can
immediately conclude that λ2 = 1 which implies that λ is equal
∀ ∗ and
to either 1 or -1. Their corresponding eigenvectors are P(s ) 
∀ ∗ , i.e.,
P (a) 
   
∀ ∗ = + P (s ) 
Up P(s )  ∀∗ ,
   
∀ ∗ = − P (a) 
Up P (a)  ∀∗ .
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 25

Pure States

Q25(1) Explain the concept of pure states.


SQ25(1) Quantum systems satisfy property QMP5.3(1), i.e., not
all observables are simultaneously measurable. Simultaneously
measurable observables are said to be compatible. A maximum
amount of information about the system corresponds to a set of
simultaneously measured values of a maximum set of compatible
discrete observables of the system, known as a complete set of
discrete observables. Such an amount of information characterises
a state of the system. States characterised by such a maximum
amount of information about the system are called pure states. If
we do not have a maximum amount of information about the system
we cannot determine a pure state. Such a situation does occur in
many practical cases. It is still desirable to have a characterisation
of the system based on the information practically available. Such
a characterisation which is based on less than a maximum set of
data about the system leads to the notation of mixed states which
is discussed in Chapter 31.
Note that a measurement of a continuous observables like
position and linear momentum cannot produce a precise value (see
P4.3.2(5) and C28.2(3)). Hence they are not used in the above
discussion of the concept of pure states.

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

144 Pure States

Q25(2) Explain why only unit vectors are used for state descrip­
tion in Postulate 25.1(PS).1
SQ25(2) Theorem 22.1(1) tells us that a unit vector in a Hilbert
space together with the spectral function of a selfadjoint operator in
the Hilbert space can generate a probability distribution function. It
follows that if we describe a state by a unit vector and an observable
by a selfadjoint operator we can generate a probability distribution
function which can be taken to describe the probability distribution
of the values of the observable. Such a description of states and
observables is demonstrated in the model theory for electron spin
in §14.1.1. A formal statement that observables correspond to the
selfadjoint operators is given by Postulate 26.1(OV).
A vector which is not normalised cannot be used in Theorem
22.1(1) to produce a probability distribution function. It needs to
be normalised first.

Q25(3) Explain why pure states do not correspond one-to-one to


unit vectors in the state space.

SQ25(3) According to Theorem 22.1(1) two unit vectors which


differ only by a phase factor, i.e., a multiplicative constant of
magnitude 1, would generate the same probability distribution
function. It is therefore not possible to distinguish two such unit
vectors physically as far as state description is concerned. In other
words all the unit vectors different by a phase factor would describe
the same state. It follows that pure states do not correspond one-to­
one to unit vectors.
All the unit vectors which differ only by a phase factor lie
in a one-dimensional subspace. Hence pure states correspond
to one-dimensional subspaces. Since one-dimensional subspaces
correspond to one-dimensional projectors we have a correspon­
dence between pure states and one-dimensional projectors. The
correspondence is one-to-one for orthodox quantum system defined
by Definition 26.1(1).

1 See also the discussion in §14.1.1 and §22.1.


June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 26

Observables and Their Values

Q26(1) Discuss the fundamental differences between classical


observables and quantum observables.
SQ26(1)
Classical observables possess the following properties:

(1) They are all compatible, i.e., they are all simultaneously
measurable.
(2) Classical observables are related to the state in that they are
described by numerical functions defined on the classical state
space, i.e., they are numerical functions of the state. It follows
that a given state would determine the values of all observables.
In other words a classical system in a given state would possess
a value of every observables. These values can be revealed by
measurement.
(3) As the the state evolves in time observables would evolve in
time accordingly. The values of observables at a later time are
determined by their initial values.

Quantum observables possess the following contrasting properties:

(1) They are not all compatible, i.e., they are not all simultaneously
measurable. Only a limited number of quantum observables are
compatible.
Quantum Mechanics: Problems and Solutions
K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

146 Observables and Their Values

(2) Observables are not mathematically related to the state in such


a way that their values are determined by the state, e.g., they are
not functions of the state. A state does not determine the value of
an arbitrary observable. A state can only generate a probability
distribution of the values of an observable. The prescription for
generating such a probability distribution needs to be explicitly
stated. Details are discussed in Chapter 28.
(3) Time evolution of the state does not automatically determine
the time evolution of observables. The future values of the
observables are not generally determined by their initial values.
Details are discussed in Chapter 29.

The different relationship between states and observables in


classical and quantum mechanics entails a fundamental difference
in the meaning of conservation laws for classical and quantum
observables. This will become clear after the discussion of quantum
time evolution in Chapter 29.1

Q26(2) What are the measurable values of a function f (A) of a


discrete observable A described by a selfadjoint operator A which
has a discrete spectrum spd = {a1 , a2 , . . .}?

SQ26(2) Given an observable A described by a selfadjoint


operator A the observable corresponding to the function f (A) of A
is described by the operator f (A ). This is stated in C26.1(6). When
A has a purely discrete spectrum {am } the operator f (A) has the
following spectral decomposition2 :

f (A) = f (am ) P Â (am ).
m

It follows that f (A) has a discrete set of eigenvalues f (am ) which


are then the measurable values of the observable f (A).

Q26(3) Give a brief account of the concept of propositions in


quantum mechanics.

1 See Definition 29.1.2(1) and Q29(6) in Exercises and Problems for Chapter 29.
2 See §20.5 for discussion on functions of selfadjoint operators. Definition 13.3.3(1),
i.e., Eqs. (13.37) and (13.38), applies to discrete observables.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Observables and Their Values 147

SQ26(3) Physically a proposition is a statement about the system


which is either true or false.3 An experiment can be performed to
test whether a proposition is true or false. Such an experiment is
called a yes-no experiment. In quantum mechanics a proposition is
a discrete observable described by a projector. A proposition has
only two values, i.e., 1 and 0 which correspond to the eigenvalues
of its associated projector. The yes-no experiment is arranged such
that the value 1 corresponds to the yes answer and the value 0
corresponds to the no answer to the proposition.
A general observable has a set of propositions associated with
it. These propositions correspond to the spectral projectors of the
selfadjoint operator representing the observable.
Q26(4) The spectral projector M  x̂ () of the position operator 
x
for an interval  is defined by a characteristic function in Eq. (20.28).
What is the physical meaning of M  x̂ () as a proposition? What
physical devices are capable of measuring M  x̂ ()?

SQ26(4) The spectral projector M  x̂ () describes the proposition


that a measurement of the position of the particle will result in a
value τ in the interval . In other words M  x̂ () represents the
proposition that a measurement of the position of the particle will
result in a value in the interval . These propositions are called local
position observables. Detectors such as Geiger counters can serve as
a measuring device to carry out the yes-no experiment to measure
the proposition. A detailed discussion of position measurement is
given in §30.2.2.
Q26(5) A state φ s corresponds to the one-dimensional projector
P φ∀ generated by the state vector φ∀. What is the meaning of the
proposition corresponding to the projector Pφ∀ ?

SQ26(5) The projector P φ∀ describes the proposition that the


system is in the state φ s described by the state vector φ∀. An ideal
measurement of the proposition resulting in the value 1 would tell
us that the state is described by the unit vector φ∀ immediately after
the measurement.

3 Isham pp. 168–178 for a discussion of various interpretations. This kind of


observables also exist in classical mechanics (see Isham pp. 61–65).
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 27

Canonical Quantisation

Q27(1) Show that Hamilton’s equations of motion in Eq. (27.8) for


a classical harmonic oscillator whose Hamiltonian is given by Eq.
(27.11) are equivalent to Newton’s equation of motion.
SQ27(1) For a classical oscillator we have q j = x, pcj = p and
H ho = p2 /2m + mω2 x 2 /2. The Hamilton’s equations become
dx ∂ H ho (x, p) 1 dp ∂ H ho (x, p)
= = p, =− = −mω2 x
dt ∂p m dt ∂x
d2 x 1 dp
� = = −ω2 x.
dt2 m dt
Newton’s equation of motion is
dx 2 dV (x)
m =− .
dt2 dx
For the oscillator the potential energy is V (x) = mω2 x 2 /2. The
above Newton’s equation becomes
dx 2
= −ω2 x,
dt2
which agrees with Hamilton’s equations of motion.
Q27(2) Verify the properties of Poisson brackets shown in Eqs.
(27.54) to (27.59).

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

150 Canonical Quantisation

SQ27(2) It is obvious that


{A, A} = 0, {A, c} = 0,
{A, B} = −{B, A}, {A, c B} = c{A, B}.
Next we have
3
 ∂ A ∂(B + C ) ∂ A ∂(B + C )
{A, B + C } = −
j =1
∂ x j ∂ pcj ∂ pcj ∂xj
3
 ∂A ∂B ∂A ∂B
= −
j =1
∂ x j ∂ pcj ∂ pcj ∂ x j
3
 ∂ A ∂C ∂ A ∂C
+ −
j =1
∂ x j ∂ pcj ∂ pcj ∂ x j
= {A, B} + {A, C }.

3
 ∂ A ∂(BC ) ∂ A ∂(BC )
{A, BC } = −
j =1
∂ x j ∂ pcj ∂ pcj ∂ x j
3
 ∂A ∂B ∂A ∂B
= C− C
j =1
∂ x j ∂ pcj ∂ pcj ∂ x j
3
 ∂ A ∂C ∂ A ∂C
+ B −B
j =1
∂ x j ∂ pcj ∂ pcj ∂ x j
= {A, B}C + B{A, C }.
The equations for {A + B, C } and {A B, C } are similarly proved.
Q27(3) Verify the Poisson bracket relations in Eq. (27.61) be­
tween the components of the canonical angular momentum L∀ c .
SQ27(3) Using properties in Eqs. (27.54) to (27.59) and the
Poisson brackets of the canonical variables in Eq. (27.60) we get
{Lcx , Lcy } = {ypcz − zpcy , zpcx − x pcz }
= {ypcz , zpcx − x pcz } − {zpcy , zpcx − x pcz }
= {ypcz , zpcx } + {zpcy , x pcz } = −ypcx + x pcy
= Lcz ,
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Canonical Quantisation 151

{Lcz , Lcx } = {x pcy − ypcx , ypcz − zpcy }


= {x pcy , ypcz − zpcy } − {ypcx , ypcz − zpcy }
= {x pcy , ypcz } + {ypcx , zpcy } = −x pcz + zpcx
= Lcy ,
{Lcy , Lcz } = {zpcx − x pcz , x pcy − ypcx }
= {zpcx , x pcy − ypcx } − {x pcz , x pcy − ypcx }
= {zpcx , x pcy } + {x pcz , ypcx } = −zpcy + ypcz
= Lcx .

Q27(4) Show that the equation of motion (27.53) in terms of


Poisson bracket reduces to the Hamilton’s equations when we
replace A by xi and pi .
SQ27(4) Bearing in mind that xi and pcj are independent, i.e.,
∂ xi ∂ pci ∂ pi ∂ xi
= δi j , = δi j , = 0, = 0,
∂xj ∂ pcj ∂xj ∂ pcj
we get the corresponding Hamilton’s equations of motion, i.e.,
 3
dxi ∂ xi ∂ H ∂ xi ∂ H ∂H
= − = ,
dt j =1
∂ x j ∂ pcj ∂ pcj ∂ x j ∂ pci
 3
dpci ∂ pci ∂ H ∂ pci ∂ H ∂H
= − =− .
dt j =1
∂ x j ∂ pcj ∂ pcj ∂ pcj ∂ pci

Q27(5) Show that a classical observable is a constant of motion,


i.e., it is time-independent, if it has a zero Poisson bracket with the
Hamiltonian.1
SQ27(5) If A has a zero Poisson bracket with the Hamiltonian,
then d A/dt = 0 by Eq. (27.53). The observable is time-independent,
i.e., it is a constant of motion.
Q27(6) Verify Eq. (27.70). Show that Postulate 27.2(CQ) as
expressed in Eq. (27.69) cannot be valid without the imaginary
number i .
1 Recall
that we confine ourselves to observables which are not explicitly time
dependent unless otherwise is stated.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

152 Canonical Quantisation

SQ27(6) First let us have a quick look at how Eq. (27.70) can
 i and 
come about. Using the selfadjointness of Q P j and treating Eqs.
(17.100) and (17.101) as equalities we get
 †  †
i, 
[Q P j ]† = Q Pj − 
 i i = 
PjQ PjQi − Q i i, 
P j = −[ Q P j ].

i, 
This shows that [Q P j ] is not selfadjoint. Let us suppose

P j ] = �II .
 j, 
[Q

Then the right-hand side of the above equation is selfadjoint while


the left-hand side is not. Explicitly we have

[Q P j ] = �II � [ Q
 j,  P j ]† = �II † = �II = [ Q
 j,  i, 
Pj ]
   
� −[ Q j , P j , ] = [ Q j , P j ],

which is a contradiction since [ Q  j, 


Pj ] =  It follows that the above
� 0.
equation cannot be true. An additional factor i on the right-hand
side of the commutation relation as stated in Postulate 27.2(CQ), i.e.,
P j ] = i �II , would produce the desired minus sign on both sides
 j, 
[Q
of the commutation relation to avoid this contradiction.
Now let us examine Eq. (27.70) more carefully again. This
equation is meant to hold in a restricted domain and not on the
entire Hilbert space since the operators involved are unbounded.
The same applies to all the equations presented above. However,
the calculation above can be justified without a full specification
of the domain of operations of those equations if we can find a
subset of vectors on which all those equations would hold. This
point is discussed in P27.2(4). To be definite let us consider the
situation in L∀2 (IR ) where we have Q  = x and P= p . The Schwartz
space S∀s (IR ) is invariant under  x and p .2 Restricted to such an
invariant subspace Eqs. (17.100) and (17.101) become equalities
which justify subsequent calculation.
� �
Q27(7) Show that U in Eq. (27.79) is unitary and that Q  ,
P in Eq.
 
(27.78) are the unitary transforms of Q , P generated by this unitary
operator.

2 See §16.1.2.4 for the definition of Schwartz functions. See also E.17.3.2.5(2),
Eq. (17.62) and P27.2(4).
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Canonical Quantisation 153

SQ27(7) Since f (x) is a real-valued function of x ⇒ IR the


following integral

1 x
− f (x)dx

define a selfadjoint multiplication operator. The operator U in Eq.
(27.79) is an exponential function of this selfadjoint multiplication
operator and is hence unitary by Eq. (21.3). We can also verify that
U in Eq. (27.79) satisfies Theorem 18.3(1), i.e., U † = U −1 .
For the unitary transforms we first have Q  � = U Q U † = Q since
Q := x commutes with U . Next we have, for φ∀ ⇒ D( p
 ),
 $  x  
d i
 U † φ∀ := −i �
p exp f (x)dx φ(x)
dx �
$  x 
i i
= −i � f (x) exp f (x)dx φ(x)
� �
$  x 
i dφ(x)
+ exp f (x)dx .
� dx
It follows that
 U † φ∀ = f (
p x )U † φ∀ + U † p
φ∀
 
� p  � = U p  U † φ∀ = U f ( x )U † φ∀ + U U † p
φ∀ = f ( x) + p φ∀.
We have used the fact that U commutes with f ( x ) and U U † = II .
Q27(8) Prove Eq. (27.104) by the method of induction.
SQ27(8) The method of induction is a method used to prove a
sequence of statements to be true.3 We can apply this method here.
So, to prove
[
x, p  n−1
 n] = i � n p x n ] = −i � n
, 
and [ p x n−1
we proceed in three steps:
Step 1: The above equations are true for n = 1, on account of the
canonical quantisation rule, i.e.,
[  ] = i �,
x, p x ] = −i �.
, 
[p
Step 2: Suppose the equations are true for any n. We desire to
prove that they are also true for n + 1, i.e.,
[
x, p  n−1 � [ 
n ] = i� n p  (n+1) ] = i � (n + 1) p
x, p  n,
x n ] = −i � n
, 
[p x n−1 � [ p x (n+1) ] = −i � (n + 1)
,  x n.
3 Such a method is set out in SQ17(7) and used also in SQ27(13).
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

154 Canonical Quantisation

All we need is to reduce [   (n+1) ], [ p


x, p x (n+1) ] in terms of [ 
,   n]
x, p
, n
and [ p x ] using the formula
[A, B
 C ] = [A, B
 ] C + B
 [A, C ].

We have
[  (n+1) ] = [
x, p n p
x, p  ] = [ n ] p
x, p  n [
+ p ]
x, p
 (n−1)
 n
= i� np   + (p
p  ) i�
 n,
= i � (n + 1) p

x (n+1) ] = [ p
, 
[p x n] 
,  x +xn[ p , 
x]
 (n−1)

= −i � n x x +
 x n (−i �)
x n.
= −i � (n + 1)
Step 3: We can now conclude that the desired equations are true
for all n.
Q27(9) Verify the commutation relations in Eqs. (27.106) and
(27.107).
SQ27(9) Consider an observable A which is a polynomial in 
x and
, i.e.,
p
N 
 M
A = xnp
cmn   m.
n=1 m=1

Using Eq. (27.104) we get


M
 M

x , A ] =
[ cmn [ 
x, 
x pn
 ]= m
x n [
cmn  m ]
x, p
m=1 m=1
M
   ∂ A
= cmn   (m−1) = i �
xn i� m p .
∂p 
m=1
N 
 M N 
 M
, A ] =
[p , 
cmn [ p m ] =
xnp , 
cmn [ p m
xn] p
n=1 m=1 n=1 m=1
M
   m ∂ A
= x (n−1) p
cmn i � n   = −i � .
m=1
∂x

Note that the differentiations are just formal expressions.


June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Canonical Quantisation 155

Q27(10) Verify the commutation relations in Eqs. (27.108),


(27.109) and (27.110).
SQ27(10) Treating H  as a polynomial functions of   and
x and p
using Eqs. (27.106) and (27.107) we get

 
[  ] = i� ∂ H .
x, H  ] = −i � ∂ H .
, H
and [ p
∂p ∂x

For the harmonic oscillator Hamiltonian we get


[  ho ] = i � ∂ H ho = i � p
x, H ,
∂p m

[p  ho ] = −i � ∂ H ho = −i �mω2 
, H x.
x
∂

Q27(11) Verify the commutation relations in Eqs. (27.111) to


(27.114) for angular momentum operators.
SQ27(11) From Eqs. (27.64) to (27.68) and the canonical commu­
tation relations for the canonical variables we get4

[
Lx ,  z − z p
Ly ] = [ y p x − 
y , z p xpz ]
= [ y p x − 
z , z p xpz ] − [ z p x − 
y , z p xpz ]
= [ y p x ] + [ z p
z , z p y , 
xpz ]
= −i �y p x + i � xp 
y = i � Lz .
Lz , 
[ Lx ] = [ 
xpy − y p z − z p
x , y p y ]
= [
xpy , y p y ] − [ y p
z − z p z − z p
x , y p y ]
= [
xp z ] + [ y p
y , y p x , z p
y ]
x pz + i �z px = i � Ly .
= −i �   
Ly , 
[ x − 
Lz ] = [ z p xpz , 
xpy − y p
x ]
= [ z p
x , 
xp x ] − [ 
y − y p xpz , 
xpy − y p
x ]
= [ z p
x , 
xpy ] + [ xpz , y p
x ]
= −i � z p z = i �
y + i � y p Lx .

4 See SQ27(3).
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

156 Canonical Quantisation

[
Lz , 
L2 ] = [
Lz , 
L2x ] + [
Lz , 
L2y ] + [
Lz , 
L2z ]
= [
Lz , 
L2x ] + [
Lz , 
L2y ]
=Lx [Lz , 
Lx ] + [Lz , 
Lx ] 
Lx
     
Ly [ Lz , Ly ] + [ Lz , Ly ] Ly
   
= i�  Lx  Ly 
Ly +  Lx − i �  Ly  Lx 
Lx +  Ly
=0.

Q27(12) Verify the commutation relations in Eq. (27.122).


SQ27(12) Using Eqs. (27.64) to (27.68) and (27.121) we get

[ a, N ] = [ a, a† a ] = [ a, a† ] a = a.


 ] = [ a† , a† a ] = a† [ a† , a ] = −a† .
[ a† , N

Q27(13) Prove, by the method of induction, that the eigenvectors


in Eq. (27.126) are normalised.
SQ27(13) The eigenvector ϕ∀0 is taken as normalised. Then:
(1) ϕ∀1 is normalised since

�ϕ∀1 | ϕ∀1 ∪ = �a† ϕ∀0 | a† ϕ∀0 ∪ = �ϕ∀0 | aa† ϕ∀0 ∪


 
= �ϕ∀0 | a† a + II ϕ∀0 ∪ = �ϕ∀0 | II ϕ∀0 ∪ = 1.
 n �
(2) Assume that ϕ∀n = a† ϕ∀0 / n! is normalised, i.e., �ϕ∀n | ϕ∀n ∪ = 1.
Then we have
1  † n+1
ϕ∀n+1 = � a ϕ∀0
(n + 1)!
1 1 n 1
= � a† � a† ϕ∀0 = � a† ϕ∀n
n+1 n! n + 1
1 1
� �ϕ∀n+1 | ϕ∀n+1 ∪ = �a ϕ∀n | a† ϕ∀n ∪ =

�ϕ∀n | aa† ϕ∀n ∪
n+1 n+1
1  
= �ϕ∀n | a† a + II ϕ∀n ∪
n+1
1  
= �ϕ∀n | n + 1 ϕ∀n ∪ = 1.
n+1
We can conclude that all ϕ∀n are normalised by the method of
induction.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Canonical Quantisation 157

Q27(14) Let  ∀ be a vector in L∀2 (IR 3 ) defined by the product


of a function of the radial variable r and a spherical harmonics
Yℓ, mℓ (θ, ϕ), i.e.,
∀ := (r, θ, ϕ) = φ(r)Yℓ, mℓ (θ, ϕ).

∀ be another vector defined in the same way, i.e.,
Let 
∀ := (r, θ, ϕ) = ψ(r)Yℓ, mℓ (θ, ϕ).


Furthermore the functions φ(r) and ψ(r) satisfy the boundary


condition

lim r|φ(r)| = 0 and lim r|ψ(r)| = 0. (�)


r�0 r�0

Working in spherical coordinates show that5


∀ |p
� ∀ ∪ = �p
r  ∀ |
r  ∀ ∪,

where pr is the radial momentum operator introduced by Eq.


(27.148).6 Explain why p
r can be symmetric but not selfadjoint.
SQ27(14) For functions (r, θ, ϕ) and (r, θ, ϕ) in L2 (IR 3 ) the
scalar product integral in spherical coordinates r, θ, ϕ is with respect
to the volume element dx 3 = r 2 sin θdrdθdϕ, i.e., the scalar product
∀ |p
� ∀ ∪ is
r 
 →  π  2π ∂ 1
−i �  � (r, θ, ϕ) + (r, θ, ϕ) dx 3 .
r=0 θ=0 ϕ=0 ∂r r

Spherical harmonics are functions in L2 (Su ) and they are normalised


by integrating with respect to the volume element sin θ dθdϕ on Su ,
as shown in Eqs. (16.43) and (16.44). It follows that
 → $ ∂ 
∀ 1

� | pr  ∪ =
 �
ψ(r) −i � + φ(r) r 2 dr.
0 ∂r r

5 See Eq. (16.44). In spherical coordinates the scalar product is given by

� → � 2π � π � �
∂ 1
∀ |�
� ∀ ∪ = −i �
pr  (r, θ, ϕ)� + (r, θ, ϕ)r 2 sin θ drdθ dϕ.
0 0 0 ∂r r

6 We assume that φ(r) and ψ(r) are differentiable with respect to r, i.e., they are
absolutely continuous in r. See Wan pp. 174–175 for more details.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

158 Canonical Quantisation

Using integration in parts we get


 → → 
∀ |p ∀ ∪ = −i � � ∂φ(r) 2 1
� r  ψ(r) r dr − i � ψ(r)� φ(r)r 2 dr
0 ∂r 0 r
 → � 2
' (→ ∂(ψ(r) r )
= −i � ψ � (r)φ(r) r 2 0 + i � φ(r)dr
0 ∂r
 →
1
−i � ψ(r)� φ(r)r 2 dr.
0 r
The first term vanishes due to the boundary condition
limr�0 rφ(r) = 0 and limr�0 rψ(r) = 0 at the origin and at r = →.7
As a result we get
 →  →
∂(ψ(r)�r 2 ) 1
∀ |p
� ∀ ∪ = i�
r  φ(r)dr − i � ψ(r)� φ(r)r 2 dr
∂r r
0 → 0
∂ψ(r)� 2
= i� r + ψ(r)� 2rφ(r) dr
0 ∂r
 →
−i � ψ(r)�rφ(r) dr
0
 →
∂ψ(r)� 2
= i� r + ψ(r)�rφ(r) dr
0 ∂r
 → ∂ 1 �
= −i � + ψ(r) φ(r) r 2 dr
0 ∂r r
= �p ∀ |
r  ∀ ∪.

This shows that for p r to be symmetric it must be defined on


a domain of vectors φ∀ which correspond to functions in L2 (IR 3 )
which are differentiable with respect to r and satisfy the boundary
condition given by Eq. (�) in Q27(14) at the origin r = 0. Since the
differential expression for the operator can also meaningfully act on
functions ψ(r) not satisfying the boundary condition at the origin
we can see that the adjoint of the operator would have a bigger
domain than that of pr . So, the operator can be symmetric but not
selfadjoint.
Q27(15) Explain why the product of operators  L(Ca ) and 
θ(Ca )
in L∀ (Ca ), i.e., the operator 
2
L(Ca ) 
θ (Ca ), cannot operate on the
7 We have lim = 0 and limr�→ rψ(r) = 0 at infinity since φ(r) and ψ(r) are
r�→ rφ(r)
square-integrable with respect to the volume element r 2 dr over the range (0, →).
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Canonical Quantisation 159

eigenvectors ϕ∀n (Ca ) of 


L(Ca ) in Eq. (19.36), and that8
θ (Ca ), 
[ L(Ca ) ] ϕ∀n (Ca )
is not defined.9
SQ27(15) The situation here is similar to the discussion in
§27.10.3 and in Q22(2) in Exercises and Problems for Chapter 22.
Here we have
θ (Ca ), 
[ θ(Ca )
L(Ca ) ] ϕ∀n (Ca ) =  L(Ca )ϕ∀n (Ca ) − 
L(Ca )
θ(Ca )ϕ∀n (Ca ).

The vector 
θ (Ca )ϕ∀n := θ ϕn (θ) does not satisfy the periodic boundary
condition in Eq. (17.37).10 This means that  θ (Ca )ϕ∀n (Ca ) is not in the
domain of L(Ca ). Hence the commutator is not defined on ϕ∀n (Ca ).

8 See Eqs. (17.23) and (27.111) for the definitions of �


θ(Ca ) and �
L(Ca ).
9 Fano pp. 407–408. See also Q22(2).
10 ϕ
n (θ ) is the exponential function on the right-hand side of Eq. (19.36).
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 28

States, Observables and Probability


Distributions

Q28(1) An electron spin is in state αxs .1 Find the probability of a


measurement of the z-component spin resulting in the value �/2.
SQ28(1) The probability is given by Postulate 28.1(PDDO), i.e.,

αx | P α∀z α∀ x ∪ = 1/2.
℘ Ŝ z (αxs , �/2) = �∀

We have used Eq. (14.25) to express α∀ x in terms of α∀ z and β∀z .


Q28(2) Using Eqs. (20.68) and (20.69), show that the probability
distribution function and the probability measure of a proposition
(as a discrete observable) represented by projector P in state
vector φ∀ are given by

⎨0 τ <0
F P̂ (φ∀, τ ) = 1 − �φ∀ | P φ∀ ∪ 0 ≤ τ < 1 .

1 τ ⊕1

⎨ 1 − �φ∀ | P φ∀ ∪ if  = {0}

P̂ ∀
M (φ , ) = �φ∀ | P φ∀ ∪ if  = {1}

⎩0 if  does not contain 0 or 1.

1 See §14.1.1 and §36.3 for the theory for electron spin.

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

162 States, Observables and Probability Distributions

SQ28(2) By Postulate 28.1(PDDO) or Postulate 28.2(PD) in


C28.2(5) the probability distribution function F P̂ (φ∀, τ ) for a
proposition in state φ s is given by Eq. (28.9) in terms of the spectral
function F P̂ (τ ) of the projector P . The spectral function F P̂ (τ ) is
given by Eq. (20.68). We get
F P̂ (φ∀, τ ) = �φ∀ | F P̂ (τ )φ∀ ∪

⎨0 τ <0
= 1 − �φ∀ | P φ∀ ∪ 0≤τ <1 .

1 τ ⊕1
The spectral measure M  P̂ () of P is given by Eq. (20.69). The
corresponding probability measure is

 Pˆ ()φ∀ ∪
M P̂ (φ∀,  ) = �φ∀ | M
⎧  
⎪ �φ∀ | II − P φ∀ ∪
⎨ if  = {0},
= �φ∀ | P φ∀ ∪ if  = {1},

⎩ �φ∀ | 0φ
 ∀∪ if  does not contain 0 or 1.

⎨ 1 − �φ∀ | Pφ∀ ∪ if
⎪  = {0},
= �φ∀ | P φ∀ ∪ if  = {1},

⎩0 if  does not contain 0 or 1.

Q28(3) For a particle in circular motion, the Hamiltonian H  (Ca ) is


given by Eq. (27.120). Show that the eigenvalues of the Hamiltonian
is degenerate with eigenvectors ϕ∀n given by Eq. (19.36). Write down
the spectral decomposition of the Hamiltonian in the form of Eq.
(20.20).
 
SQ28(3) The Hamiltonian 
 H Ca shares the same eigenvectors as
the momentum operator p  Ca but with different eigenvalues, i.e.,
       
 Ca ϕ∀n Ca = E n ϕ∀n Ca , E n = 1 p2 Ca = 1 (n�)2 .
H n
  2m 2ma2
The ground state, i.e., ϕ∀n=0 Ca is nondegenerate while all other
eigenvalues are degenerate with degeneracy  2 since E n = E −n .
The spectral decomposition of H Ca in terms of the projectors
 
generated by ϕ∀n Ca is given by Eq. (20.20), i.e.,
  
H Ca = E n P Ĥ (Ca ) (E n ), n = 0, 1, 2, · · · ,
n
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

States, Observables and Probability Distributions 163

where
   
P Ĥ (Ca ) (E 0 ) = |ϕ∀0 Ca ∪�ϕ∀0 Ca |,
       
P Ĥ (Ca ) (E n ) = |ϕ∀n Ca ∪�ϕ∀n Ca | + |ϕ∀−n Ca ∪�ϕ∀−n Ca |, n ⊕ 1.

Q28(4) A pair of annihilation and creation operators a, a† are


defined in terms of an orthonormal basis {ϕ∀n , n = 0, 1, 2, . . .} in
∀ of a quantum system.2 The corresponding number
the state space H

operator is N = a† a.3 The energy of the system is represented by
the Hamiltonian operator H  = E0 N.

(a) What are the possible energy values of the system?


(b) Let ∀ z be the unit vector in Eq. (17.129). Find the probability
mass function for the probability distribution of energy values
∀ z.
of the system in state sz described by the state vector 
(c) Find the energy expectation values in state sz .

SQ28(4)(a) Possible energy values are the eigenvalues of the


operator H , i.e., E n = nE 0 , n = 0, 1, 2, · · · , corresponding to the
eigenstates defined by the eigenvectors ϕ∀n of H .

SQ28(4)(b) The probability ℘ H (sz , E n ) of a measurement ob­


taining the value E n is given by
℘ H (sz , E n ) = � ∀ z ∪,
∀ z | P ϕ∀n 
where P ϕ∀n = |Pϕ∀n ∪�P ϕ∀n | is the projector generated by the
eigenvector ϕ∀n . Since
 1  zn
∀ z = exp − |z|2 � ϕ∀n ,
P ϕ∀n 
2 n!
we have
 1  zn
℘ H (sz , E n ) = �∀ z | exp − |z|2 � ϕ∀n ∪
2 n!
 1  zn
= exp − |z|2 � � ∀ z | ϕ∀n ∪
2 n!
 1  zn  1  z�n
= exp − |z|2 � exp − |z|2 �
2 n! 2 n!
2n
 2 n
  |z|   |z|
= exp − |z|2 = exp − |z|2 .
n! n!
2 See Definitions 17.10(1) and 17.10(2) in §17.10.
3 See Definition 19.1(4).
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

164 States, Observables and Probability Distributions

In accordance with Theorem 22.2(1) these values define a probabil­


ity mass function ℘ Ĥ ( ∀ z , E n ) on the spectrum spd (H ) = {E n } of H
,
H s Ĥ ∀
i.e., ℘ (z , E n ) = ℘ (z , E n ).
As a consistency check we have
→ →  n
H
 2
 |z|2
℘ (E n ) = exp − |z|
n=0 n=0
n!
   
= exp − | z|2 exp |z|2 = 1.

SQ28(4)(c) The energy expectation value is



 →    |z|2 n
H 2
℘ (E n )E n = exp − |z| nE 0
n=0 n=0
n!
  →  n
2 |z|2
= E 0 exp − |z|
n=1
(n − 1)!
  →  2 (n−1) 2
|z| |z|
= E 0 exp − |z|2 exp
n=1
(n − 1)!
  →  n−1
2 2 |z|2
= |z| E 0 exp − |z|
n=1
(n − 1)!
   |z|2 m
→ 
2 2
= |z| E 0 exp − |z| , m=n−1
m=0
m!
   
= |z|2 E 0 exp − |z|2 exp |z|2 = | z |2 E 0 .

We have used the result


→ →
x m  x (n−1)
exp x = = .
n=0
m! n=1
(n − 1)!

Q28(5) The Fourier transform ⊗ϕ( p) of a normalised wave function


ϕ(x) of a particle is

⎨0 � p ≤ − p0 ,
ϕ( p) = 1/ 2 p0 p ⇒ (− p0 , p0 ],
⊗ ⎩
0 p ⊕ p0 ,
where p0 ⇒ IR. What is the probability of a momentum measurement
resulting in a value in the range (− p0 , p0 ]? What is the momentum
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

States, Observables and Probability Distributions 165

expectation value and uncertainty? Write down the position proba­


bility density function in terms of the inverse Fourier transform of
ϕ( p) in Eq. (18.73).

SQ28(5)4 In accordance with Eq. (28.19) the probability of a


momentum measurement the state ϕ s a value in the range (τ1 =
− p0 , τ2 = p0 ] is given by
 τ2  p0
 
℘ p ϕ s , (τ1 , τ2 ] = ω p (ϕ s , τ ) dτ = |ϕ(

p)|2 dp
τ1 − p0
  2
p0  1 
= �  dp = 1.
 2p 
− p0 0

The momentum expectation value is calculated in the momentum


representation to be
 →  p0
s 2 1
E( p, ϕ ) = p |ϕ ( p)| dτ = p dp = 0.
−→
⊗ 2 p0 − p0
This is as expected since we are as likely to get a negative value as a
positive value of the momentum.
The corresponding uncertainty is given by Eq. (28.21), i.e.,
 
( p, ϕ s ) =  2 ϕ∀ ∪ − �ϕ∀ | p
�ϕ∀ | p  2 ϕ∀ ∪
 ϕ∀ ∪2 = �ϕ∀ | p
 p0 1/2
1 1
= p2 dp = � p0 .
2 p0 − p0 3
The inverse Fourier transform ϕ(x) of ϕ )( p) is given by Eq.
(18.73), i.e., we have
#
� sin( p0 x/�)
ϕ(x) = .
π p0 x
It follows from Eq. (28.13) that the position probability density
function wφx (x) is given by
2
� sin( p0 x/�)
ω x (ϕ s , x) = |ϕ(x)|2 = .
π p0 x

Q28(6) Working in the momentum representation and using Eqs.


(20.54) and (20.57) find the probability distribution function for the
4 See SQ30(2)(b).
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

166 States, Observables and Probability Distributions

kinetic energy of a particle of half the unit mass in one-dimensional


motion along the x -axis in a given state ϕ s .
SQ28(6) The state space of the particle is the Hilbert space L∀2 (IR ).
The state vector for the state ϕ s is denoted by ϕ∀ which is a unit
vector in L∀2 (IR ) corresponding to a normalised function ϕ(x ) ⇒
L2 (IR ). In the momentum representation the kinetic energy operator
of a particle of half the unit mass is the multiplication operator
(IR) = p
K (IR) 2
acting on the momentum space wave functions
⊗ ⊗ ⊗ ⊗

in L∀ (IR), e.g., ϕ( p) which is the Fourier transforms of ϕ(x ). The


2
⊗ ⊗
spectral function of p (IR)

as a multiplication operator in L2 (IR) ⊗
is

the characteristic function χ ⊗ (−→, τ ]
( p) given by Eq. (20.29). From
Eqs. (20.54), (20.56) and (20.57) we obtain the spectral function of
(IR) 2
p
⊗ ⊗
, i.e.,
!
ˆ

0 τ <0
F⊗K⊗ (IR⊗ ) (τ ) := .
χ⊗ [−�τ , �τ ] ( p) τ ⊕0

The probability distribution function for state ϕ s is, for τ > 0,


 �τ
ˆ (IR )
K  ˆ (IR )
K
F

⊗ ⊗ (τ ) = �⊗ϕ∀ | F⊗ ⊗ ⊗ (τ )ϕ

∀∪ = � |ϕ(

p)|2 dp.
− τ

This is as expected. The kinetic energy does not take any negative
values. The probability of the kinetic energy having a value in the
range (−→, τ ] for τ > 0 is the same as the probability of the kinetic
energy having a value in the range [0, τ ]. This is the same as the
� �
momentum having a value in the range [− τ , τ ].
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 29

Time Evolution

Q29(1) Using the expression for the spectral decomposition of


unitary operators in Eq. (21.6) and the result in Eq. (21.19),
show that Eq. (29.5) can be obtained from the unitary evolution
Eq. (29.10).
SQ29(1) For a Hamiltonian H  with a purely discrete set of
eigenvalues E ℓ corresponding a complete orthonormal set of
eigenvectors ϕ∀ℓ and eigenprojectors P ϕ∀ℓ we have, by Eq. (21.6),
 , t) = e− ¯i H� t
U (H
 −iE t
= e ¯ ℓ Pϕ∀ℓ .

φ∀(t) = U (H
 , t)φ∀(0)
 −iE t  
= e ¯ ℓ Pϕ∀ℓ φ∀(0)

 −iE t
= e ¯ ℓ cℓ ϕ∀ℓ , cℓ = �ϕ∀ℓ | φ∀(0)∪.

¨
Q29(2) Verify Eq. (29.12) in the Schrodinger picture and Eq.
(29.33) in the Heisenberg picture.
SQ29(2)
 ¨
In the Schrodinger picture the expectation value
E A, φ s (t) at time t is given by �φ∀(t) | Aφ∀(t) ∪, where A is time-

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

168 Time Evolution

independent. Differentiating this with respect to t and using the


¨
Schrodinger equation we get
 
dE A, φ s (t) dφ∀(t)  dφ∀(t)
=� | A φ∀(t) ∪ + �φ∀(t) | A ∪
dt dt dt
1  1 
= � H φ∀(t) | Aφ∀(t) ∪ + �φ∀(t) | A H φ∀(t) ∪
i� i�
1   
= −�H φ∀(t) | Aφ∀(t) ∪ + �φ∀(t) | A H  φ∀(t) ∪
i�
1  ∀  Aφ∀(t) ∪ + �φ∀(t) | A H φ∀(t) ∪

= −�φ (t) | H
i�
1
= �φ∀(t) | [ A, H
 ] φ∀(t) ∪.
i�
In the Heisenberg picture the expectation value at time t is given by
�φ∀ | A(t)φ∀ ∪, where φ∀ is time-independent. Differentiating this with
respect to t and using the Heisenberg equation we get

d A(t) 1
�φ∀ | φ∀ ∪ = �φ∀ | [ A(t), H
 ] φ∀ ∪.
dt i�
For brevity we have omitted the subscripts Sch for Schrodinger¨
picture quantities and Hei for Heisenberg quantities. We have also
assumed that the initial and the evolved vectors are in the domain of
relevant operators, e.g., φ∀(t) is in the domain of A.
Q29(3) The Hamiltonian of a system at time t = 0 is given in terms
of a pair of annihilation and creation operators a and a† by

 = 1
H a† a + �ω.
2
(a) Using the method of induction and the commutation relation of
a and a† , prove

H  − �ω)n ,
 n a = a (H n = 0, 1, 2, 3, · · · . (�)

(b) Assuming that the time dependence of the annihilation and


creation operators in the Heisenberg picture are given by the
Heisenberg equation of motion in Eq. (29.20), show that

aHei (t) = aHei (0) e−i ωt , †


aHei (t) = a†Hei (0) ei ωt .
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Time Evolution 169

¨
(c) The annihilation operator in the Schrodinger picture, denoted
 
by a Sch (t), is related to aHei (t) by
� �
aSch t) = e H t/i � aHei (t) e− H t/i � .
By expanding the exponential in term of a series, i.e.,1
→ n
� 1 t  n,
e H t/i � = H
n=0
n! i�

and using Eq. (�) show that


� �
e H t/i � aHei (t) = aHei (t) ei ωt e H t/i � .
Hence verify explicitly that aSch (t) is time-independent.

SQ29(3)(a) The equation H  − �ω)n can be proved by


 n a = a(H
induction:

(1) The equation is trivially true for n = 0 since H  − �ω)0


 0 and (H
2
are equal to the identity operator.
(2) Next we can show that it is true for n = 1, i.e.,
 ] = [ a, a† a ] �ω = [ a, a† ] a �ω = a �ω
[ a, H
� aH −H  a = a �ω
�  a = a (H
H  − �ω).

(3) Finally assume the equation to be true for an n > 1. Then we


have
 n+1 a = H
H H n a = H
 a(H
 − �ω)n
 − �ω)(H
= a(H  − �ω)n
 − �ω)n+1 .
= a(H
It follows that the equation is true for all n.

Note that this equality is valid in the Schrödinger picture and in


the Heisenberg picture for all times since the unitary transformation
which relates Schrödinger picture and Heisenberg picture also
preserves commutation relations.
1 Assuming an appropriate domain on which the expansion is valid.
2 See Definition 19.5(2) for �.
A�0 = II
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

170 Time Evolution

SQ29(3)(b) The annihilation and creation operators at time t in



the Heisenberg and the Schrodinger ¨ pictures aHei (t), aHei (t) and

aSch (t), aSch (t) are unitarily related with
aHei (0) = aSch (0) = aSch (t) = a,
† † †
aHei (0) = aSch (0) = aSch (t) = a† .
The operators in the Heisenberg picture are obtainable from the
¨
corresponding quantities in the Schrodinger picture by a unitary
transformation given by Eq. (29.13), i.e.,
aHei (t) = U (t)† aHei (0)U (t),
† †
aHei (t) = U (t)† aHei (0)U (t),
 t). Since unitary transformations preserve
where U (t) = exp(−–i H
commutation relations we have
† † †
aHei (t), aHei (t)] = [
[ a Sch (t), aSch (t)] = [aSch (0), aSch (0)] = 1.
In accordance with Eq. (29.15) the Hamiltonian in the Heisenberg

picture at time t is expressible in terms of aHei (t), and aHei (t) as3

 (t) = U (t)† H
 (0)U (t) = �ω † 1
H aHei (t)aHei (t) + .
2
It follows that
% &
 † 1
[ aHei (t), H (t)] = aHei (t), �ω aHei (t)aHei (t) +
2

= �ω [aHei (t), aHei (t)
aHei (t)]
= �ω aHei (t).
The Heisenberg equation for aHei (t) is
d  (t)] = �ω aHei (t)
aHei (t) = [aHei (t), H
i�
dt
d
� aHei (t) = −i ω aHei (t).
dt
The solution is clearly
aHei (t) = aHei (0)e−i ωt ,
3 The Hamiltonian is time-independent, � (t), i.e.,we have
even though it is written as H
� (t) = H
H � (0) = H
� , as explicitly shown in Q29(3)(c).
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Time Evolution 171

where aHei (0) can be identified with the corresponding Schrodinger


¨

picture operator aSch (0). The time dependence of aHei (t) is
†  † †
aHei (t) = aHei (0)e−i ωt = aHei (0)ei ωt .

¨
SQ29(3)(c) The annihilation operator in the Schrodinger picture,
denoted by aSch (t), is related to aHei (t) by
� �
aSch (t) = e H t/i � aHei (t) e− H t/i � .
Our problem is to work out the right-hand side and show explictly
that aSch (t) is time-independent. To work out the right-hand side we
first consider4

e H t/i � aHei (t)

 n
 1 H t
= aHei (t)
n=0
n! i �

 n
 1 t
H
−i ωt
=e aHei (0)
n=0
n! i�
→ n
1 t  n a since aHei (0) = a
= e−i ωt H
n=0
n! i�
→
1 t n  − �ω)n
= e−i ωt a(H by Eq. (�) in Q29(3)(a)
n=0
n!
i�

 n
 1 (H  − �ω) t
−i ωt
=e a
n=0
n! i�
� �
= e−i ωt a e(H −�ω)t/i � = a e H t/i � .
It follows that
� �
e H t/i � aHei (t) e− H t/i � = a = aSch (t),
which is time-independent as expected.
Q29(4) Let A be the selfadjoint operator representing an observ­
¨
able A in the Schrodinger picture and let F Â (τ ) be its spectral
function. Let the corresponding operator at time t in the Heisenberg
picture be denoted by AHei (t) and let F Â Hei (t) (τ ) be its spectral
4 Making used of Eq. (29.19) which relate a
�Hei (t) to a�Hei (0).
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

172 Time Evolution

function. How are these two spectral functions related and how are
the probability distribution functions generated by these spectral
functions in a given state related?
SQ29(4) At time t = 0 let the state vector be φ∀ and the operator
for the observable in question be A. At time t > 0 let the evolved
state vector be denoted by φ∀(t) in the Schrodinger
¨ picture. In the
Heisenberg picture the evolved vector φ∀Hei (t) at time t remains φ∀,
i.e., we have φ∀Hei (t) = φ∀. So, we can simply rewrite φ∀Hei (t) as φ∀
without the variable t. The two vectors φ∀(t) and φ∀Hei are related by
Eq. (29.36), i.e., we have by φ∀(t) = U (t)φ∀Hei . The evolved operator
remains A in the Schrodinger ¨ picture, while in the Heisenberg
picture the evolved operator becomes AHei (t) = U (t)† AU (t) as
shown in Eq. (29.39). For brevity of notation we have omitted the
subscripts Sch for quantities in the the Schrödinger picture.
In accordance of Theorem 15.3(2) the spectral functions F Â (τ )
and F Â Hei (t) (τ ) of A and AHei (t) are related by the unitary
transformation generated by the unitary operator U (t) in the same
way the operators of the two pictures are related by Eq. (29.43), i.e.,

F Â (τ ) = U (t)F Â Hei (t) (τ )U (t)† .

¨
The probability distribution function in the Schrodinger picture in a
given state at time t is given by

F Â (φ∀(t), τ ) = �φ∀(t) | F Â (τ )φ∀(t)∪.


The probability distribution function in the Heisenberg picture in
the same given state at time t is given by

F ÂHei (t) (φ∀Hei , τ ) = �φ∀Hei | F Â Hei (t) (τ )φ∀Hei ∪.


Then we have
 
F Â (φ∀(t), τ ) = �U (t)φ∀Hei | U (t)F Â Hei (t) (τ )U (t)† U (t)φ∀Hei ∪

= �U (t)φ∀Hei | U (t)F Â Hei (t) (τ )φ∀Hei ∪


= �φ∀Hei | F ÂHei (t) (τ )φ∀Hei ∪ = F ÂHei (t) (φ∀, τ ).
This is as expected since the probability distribution function is
measurable and it should not depend on any particular mathemati­
cal description.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Time Evolution 173

Q29(5) The Hamiltonian of a quantum particle of mass m in one­


dimensional motion along the x-axis with potential energy V ( x ) is
given by

H = 1 p  2 + V (
x ).
2m
x ) to be a polynomial function of 
Assuming V ( x show that in the
Heisenberg picture, we have5
d
m �φ∀ | x φ∀ ∪ = �φ∀ | p
 φ∀ ∪,
dt
d dV (
x)
�φ∀ | p
φ∀ ∪ = −�φ∀ | φ∀ ∪.
dt dx
Establish the following Ehrenfest’s theorem:
d2 ∀ dV (
x)
m x φ∀ ∪ = �φ∀ | F φ∀ ∪, where F = −
�φ |  .
dt2 dx
Discuss the physical significance of this result,
SQ29(5) From Eq. (29.32) we get6
d  (t)
∂H 1
x (t) =
 = p (t),
dt (t)
∂p m
d ∂H (t) ∂ V (x (t))
(t) = −
p =− .
dt x (t)
∂ x
∂

It follows that
d2 1 d d2 ∂ V (
x (t))
2

x (t) = 
p (t) � m 2
x (t) = −
 .
dt m dt dt x
∂
Taking the expectation values on both sides we get
d2 ∀ d
m x (t)φ∀ ∪ = �φ∀ | F φ∀ ∪,
�φ | 
dt2 dt
where
dV (
x (t))
F = − ,
dx (t)
which is the Ehrenfest’s theorem. This is the equivalence of Newton’s
second law in classical mechanics. The significance here is that it
5 The subscript Hei for Heisenberg picture quantities is omitted for brevity. We also
assume that φ∀ is in an appropriate domain of all the operators involved.
6 The subscript H ei for Heisenberg picture quantities is omitted for brevity.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

174 Time Evolution

is the second order time derivative of the expectation value of the


position which is related to the expectation value of the force.
Q29(6) Discuss the fundamental differences in the concept of
constants of motion in classical and quantum mechanics.
SQ29(6) In classical mechanics a constant of motion is an
observable whose value mains constant and independent of time. A
quantum observable is a constant of motion if its expectation value
in a given state is a constant and independent of time. Bearing in
mind that in a given state the observable may not possess a value it
is not meaningful to say that the observable would have a definite
initial value and as time evolves the observable would still possess
¨
this value. In the Schrodinger ∀
picture let the initial state be φ(0)

and the final state be φ (T ). Then an initial measured value in state
described φ∀(0) may well not be the same as an individual measured
value of the observable in state described φ∀(T ) at a later time T . But
the expectation value of the observable at time T is the same as that
at time 0.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 30

On States after Measurement

Q30(1) An electron spin is in state αxs .1 What is the state and the
state vector immediately after a measurement of the z-component
spin resulting in the value �/2?
SQ30(1) Immediately after a measurement of the z-component
spin resulting in the value �/2 state is αzs by Postulate 30.1.1(PPDO).
The corresponding state vector is α∀ z .
Q30(2) Find the subspace associated with the projector F x̂ (x2 ) −
F x̂ (x1 ) on L∀2 (IR ).
SQ30(2) The subspace consists of all the vectors ϕ∀ in L∀2 (IR )
satisfying the equation
 
F x̂ (x2 ) − F x̂ (x1 ) ϕ∀ = ϕ.

In terms of functions in L2 (IR ) this equation becomes

χ (x1 , x2 ] (x )ϕ(x ) = ϕ(x ).


The solutions to this equation are all the functions in L2 (IR ) which
vanish outside the interval  = (x1 , x2 ]. This subspace is identifiable
with L∀2 ().
1 See §14.1.1 for the theory for electron spin.

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

176 On States after Measurement

Q30(3) Consider a particle in one-dimensional motion along


the x-axis. According to Postulate 30.2.1(PPCO), if a momentum
measurement yields a value in the interval ( p1 , p2 ] the state vector ⊗ϕ∀
in the momentum representation right after the measurement must
satisfy  p̂ 

F ⊗ ( p2 ) − F ⊗ ( p1 ) ϕ∀ = ϕ,
⊗ ⊗
∀ ⊗ ⊗
(�)

where ⊗F ⊗ ( p) is the spectral function of the momentum operator

given in Theorem 20.4.2(1), i.e., ⊗F ⊗ ( p) is defined by the
characteristic function χ ⊗
(−→, τ ] ( p) of the interval (−→, τ ] on the
momentum space IR. ⊗
Express Eq. (�) in terms of χ ⊗
(−→, τ ] ( p) and functions ⊗ϕ( p) in
L2 (IR).

Give an example of such a function in the momentum space
and its corresponding function in the coordinate space.
SQ30(3) In terms of χ (−→, τ ] ( p) and ⊗ϕ( p) Eq. (�) becomes
 ⊗

χ⊗ (−→, p2 ] ( p) − χ⊗ (−→, p1 ] ( p) ⊗ϕ( p) = ⊗ϕ( p).
Since
χ⊗ (−→, p2 ] ( p) − χ⊗ (−→, p1 ] ( p) = χ⊗ ( p1 , p2 ] ( p),
the equation becomes
χ⊗ ( p1 , p2 ] ( p)⊗ϕ( p) = ⊗ϕ( p).
This is similar to the situation in position measurement discussed in
SQ30(2). Clearly the following normalised function
$ �
1/ p2 − p1 p ⇒ ( p1 , p2 ]
ϕ( p) = ,
⊗ 0 p⇒/ ( p1 , p2 ]
in the momentum space satisfies the above equation. The corre­
sponding function ϕ(x) in the coordinate representation space is the
inverse Fourier transform of ⊗ϕ( p), i.e.,
 p2
1 i px
ϕ(x) = � e ¯ dp
2π �( p2 − p1 ) p1
1 ei p2 x/� − ei p1 x/�
= �
2π �( p2 − p1 ) i x/�
#
i p2 x/�
� e − ei p1 x/�
= .
2π ( p2 − p1 ) ix
This function reduces to that in Eq. (18.73) when p2 = p0 and p1 =
− p0 .
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 31

Pure and Mixed States

Q31(1) Explain why the transition shown in Eq. (31.1) due to


measurement cannot be generated by a unitary transformation.1
¨
SQ31(1) In the Schrodinger picture on quantum time evolution as
stated in Postulate 29.1.2 (TESP) we have:

(1) If an initial state is described by a state vector the evolved state


would be described by the evolved state vector in Eq. (29.10),
since a unitary transformation will yield a unique transformed
vector at any given time t from a given vector. A pure state would
evolves into another pure state.
(2) If the initial state is represented by a one-dimensional pro­
jector then the evolved state is again represented by a one­
dimensional projector since the unitary transform of a one­
dimensional projector is again a one-dimensional projector.2 A
pure state would evolves into another pure state.

1 This means that the transition as a time evolution does not satisfy Postulate
29.1.2(TESP) which applies to quantum evolution in the Schrödinger picture.
2 See Q18(9).

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

178 Pure and Mixed States

The transition in Eq. (31.1) shows a transition from a pure state to a


classical mixture of pure states. This contradicts what has been said
above. Hence such a transition cannot be a unitary transformation
of the initial pure state. We can also discuss the situation in terms of
state vectors or one-dimensional projectors:

(1) The transition shown in Eq. (31.1) means that an initial state
vector is transformed into many different final state vectors. It
follows that the transition cannot be a unitary transformation
from the initial state vector.
(2) A one-dimensional projector cannot evolve unitarily to a density
operator which is not a projector.

Q31(2) Prove Eqs. (31.6) and (31.7).


SQ31(2) To prove Eq. (31.6) we can choose an orthonormal basis
{ϕ∀ℓ } such that ϕ∀1 = φ∀. Then φ∀ will be orthogonal to all ϕ∀ℓ for ℓ ⊕ 2
∀ except for ℓ = 1. Then
so that P φ∀ ϕ∀ℓ = 0
  
tr B P ∀ =  P ∀ ϕ∀ℓ ∪
�ϕ∀ℓ | B
φ φ

= �ϕ∀1 | B  P ∀ φ∀ ∪ = �φ∀ | B
 P ∀ ϕ∀1 ∪ = �φ∀ | B  φ∀ ∪.
φ φ

Equation (31.7) can be proved in the same way.


Q31(3) Prove Eqs. (31.17) and (31.18).
SQ31(3) A general density operator D  is decomposable as a
convex combination of projectors in accordance with Eq. (18.22) or

 =
Eq. (18.24), i.e., D 
ℓ ωℓ P ϕ∀ℓ . Since the trace operation is linear
the right-hand side of Eq. (31.15) becomes
  
tr (F B̂ (τ )D
 ) = tr F B̂ (τ ) ωℓ P ϕ∀ℓ

  
= ωℓ tr F B̂ (τ )Pϕ∀ℓ .

By the result obtained in SQ31(2) we have


 
tr F B̂ (τ )Pϕ∀ℓ = �ϕ∀ℓ | F B̂ (τ )ϕ∀ℓ ∪

� tr (F B̂ (τ )D ) = ωℓ �ϕ∀ℓ | F B̂ (τ )ϕ∀ℓ ∪.

June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Pure and Mixed States 179

By Postulate 28.2(PD), i.e., Eq. (28.9), we have


F B̂ (ϕ∀ℓ , τ ) = �ϕ∀ℓ | F B̂ (τ )ϕ∀ℓ ∪

� tr (F B̂ (τ )D) = ωℓ F B̂ (ϕ∀ℓ , τ ),

which is the desired Eq. (31.17).


Equation (31.18) can be proved in the same way.
Q31(4) Prove Eq. (31.20) for the expectation value.
SQ31(4) On account of Eqs. (31.15) and (31.17) the integral in
Eq. (31.19) can be evaluated as follows:
 →  →  
  
τ dτ tr F B̂ (τ )D
) = τ dτ ωℓ F B̂ (ϕ∀ℓ , τ )
−→ −→ ℓ
  →  
= ωℓ τ dτ F B̂ (ϕ∀ℓ , τ ) =  ϕ∀ℓ ∪.
ωℓ �ϕ∀ℓ | B
ℓ −→ ℓ

We have used Eq. (22.3), i.e.,


 →
 ϕ∀ℓ ∪.
τ dτ F B̂ (ϕ∀ℓ , τ ) = �ϕ∀ℓ | B
−→

On the other hand we have


  
D
tr (B  ) = tr B
 ωℓ P ϕ∀ℓ

 
=  P ϕ∀ℓ ) =
ωℓ tr (B  ϕ∀ℓ ∪.
ωℓ �ϕ∀ℓ | B
ℓ ℓ

Equation (31.20) then follows, i.e.,



, D
E(B  ) = tr (B
D) =  ϕ∀ℓ ∪.
ωℓ �ϕ∀ℓ | B

June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 32

Superselection Rules

Q32(1) Prove Eq. (32.4).


SQ32(1) Since S∀(n) is a complete orthogonal family of subspaces
their corresponding projectors P (n) would constitute a complete
orthogonal family of projectors. It follows that their sum is equal to

the identity operator, i.e., n P (n) = II . Then
   
re =
B re
P (n) B P(m)
n m
 
= re P (n) +
P(n) B P (n) B
re P(m) .
n � m
n=

Being reducible by S∀(n) the operator B re would commute with P (n)
by Theorem 17.9(1) so that the terms in the second sum above can
re P (n) P (m) . Since the projectors are orthogonal these
be written as B
terms vanish. We get
 
re =
B P (n) B
re P(n) =  (n) ,
B
n n

where B  (n) =P (n) 


B re P  (n) .
Q32(2) For the example in §32.3 discuss how the initial state η (0)�
given by Eq. (32.15) would evolve in the Schrödinger picture under a
selfadjoint and decomposable Hamiltonian H � = B  (−) � B
 (0) � B
 (+) .

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

182 Superselection Rules

SQ32(2) The given initial state is η∀ (0)� = 0 ∀ (−) � η∀ (0) � 0


∀ (+) . Since
the subspace S∀ is one-dimensional the vector η∀ must be an
(0) (0)

 (0) in Eq. (32.17), i.e., we have


eigenvector B
 (0) η∀ (0) = b0 η∀ (0)
B for some b0 ⇒ IR .

For the given Hamiltonian H  (−) � B


� = B  (0) � B
 (+) we have

 � η∀ (0)� = b0 η∀ (0)� .
H

It follows from Eq. (29.4) that in the Schrödinger picture the evolved
state vector for the initial state vector η∀0� is given by
− ib t
η∀ (0)� (t) = e ¯ 0 η∀ (0)� .

Here both the initial and the final states are pure. The state vector
η∀ (0)� (t) remains in the subspace S∀(0) for all times.

Q32(3) A system has a two-dimensional state space H ∀. A



superselection rule operates with the state space H having the
following preferred direct sum decomposition:
∀ =H
H ∀� = H ∀ (2) ,
∀ (1) � H

where H ∀ and H
∀ (1) is spanned by a unit vector η∀ (1) ⇒ H ∀ (2) is spanned
∀ which is orthogonal to η∀ (1) . Let 
by a unit vector η∀ (2) ⇒ H L be an

operator on H defined by

Lη∀ (1) = η∀ (2) , 


L η∀ (2) = η∀ (1) .

(a) Show that 


L is selfadjoint and explain why 
L cannot represent
an observable.
(b) Suppose the Hamiltonian of the system is λ  L, where λ ⇒ IR .
¨
Show that in the Schrodinger picture the initial state vector
(2)
η∀(0) = i η∀ at time t = 0 evolves in time to the following state
vector at time t:

η∀(t) = sin(λt/�)η∀ (1) + i cos(λt/�)η∀ (2) .

(c) Discuss how the evolution may cause a transition from a pure
state to a mixture.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Superselection Rules 183

SQ32(3)(a) The operator  L satisfies the selfadjointness condition


�η∀ (m) | 
Lη∀ (n) ∪ = �
Lη∀ (m) | η∀ (n) ∪ ∞n, m = 1, 2, since
�η∀ (1) | 
Lη∀ (1) ∪ = �
Lη∀ (1) | η∀ (1) ∪ = 0,
�η∀ (1) | 
Lη∀ (2) ∪ = �
Lη∀ (1) | η∀ (2) ∪ = 1,
�η∀ (2) | 
Lη∀ (1) ∪ = �
Lη∀ (2) | η∀ (1) ∪ = 1,
�η∀ (2) | 
Lη∀ (2) ∪ = �
Lη∀ (2) | η∀ (2) ∪ = 0.
It follows that 
L is selfadjoint. By Definition 32.1(5) the operator 
L
cannot represent an observable since it is not decomposable.

SQ32(3)(b) First the state vector η∀(t) satisfies the initial condi­
tion, i.e., η∀(t = 0) = η(0).
∀ ¨
It also satisfies the Schrodinger equation,
i.e.,
dη∀(t)
i� = i λ cos(λt/�) η∀ (1) + λ sin(λt/�) η∀ (2) .
dt
H η∀(t) = λ sin(λt/�) η∀ (2) + i λ cos b(λt/�) η∀ (1) .
dη∀(t)  η∀(t).
� i� =H
dt

SQ32(3)(c) The initial state vector η∀(0) = i η∀ (2) describes a pure


state while the evolved state vector η∀(t) at time t represents a
classical mixture of η∀ (1) and η∀ (2) , except for certain specific times
tn when either sin(λtn /�) or cos(λtn /�) is zero. Here we have an
example of an evolution from a pure state to a mixed state. This is
because that the Hamiltonian is not decomposable, and it does not
represent an observable of the system. The physical significance of
this is discussed in §34.7.2 on quantum measurement.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 33

Many-Particle Systems

Q33(1) Calculate the expectation value of the two-particle observ­


able represented by the operator in Eq. (33.9) in the state described
∀ (c) in Eq. (33.2).
by the vector 
SQ33(1) We can make use of Eqs. (33.5) and (33.8). For the two­
particle observable in Eq. (33.9) the expression for the expectation
value in becomes
 
∀ (c) | A (1 ∗ II (2) + II (1) ∗ A (2) 
� ∀ (c) ∪∗
   
= �∀ (c) | A (1 ∗ II (2)  ∀ (c) ∪∗ + � ∀ (c) | II (1) ∗ A (2)  ∀ (c) ∪∗
 (1) (2)
= c �j , k cm, n �φ∀ j | A(1) φ∀m(1) ∪(1) �φ∀k | II (2) φ∀n(2) ∪(2)
j , k, m, n
 (1) (2)
+ c �j , k cm, n �φ∀ j | II (1) φ∀m(1) ∪(1) �φ∀k | A(2) φ∀n(2) ∪(2)
j , k, m, n
 (1)
= c �j , k cm, k �φ∀ j | A(1) φ∀m(1) ∪(1)
j , k, m
 (2)
+ c�j , k c j , n �φ∀k | A(2) φ∀n(2) ∪(2) .
j , k, n

Q33(2) Show that symmetrical vectors in Eq. (33.19) form a


∀ (c) .
subspace of H

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

186 Many-Particle Systems

SQ33(2) The symmetrical vectors clearly form a linear subset


since any finite linear combination combinations of symmetrical
vectors are again symmetrical. To see that the set of symmetrical
vectors form a subspace we need to check that any Cauchy sequence
of symmetrical vectors would converge to a symmetrical vector.
Generally let φ∀ℓ be a Cauchy sequence of vectors in H ∀ (c) . Such
∀ ∀ (c)
a sequence would converge to a vector φ in H . The permutation
operator Up in Eq. (33.14) is a bounded operator. By Theorem
17.1(1) it is also a continuous operator. This means that the
sequence Up φ∀ℓ would converge to Up φ∀.
(s)
A Cauchy sequence of symmetrical vectors ψ∀ ℓ would converge
to a vector ψ∀ in H ∀ (c) . By definition the permutation operator does
(s) (s)
not affect symmetrical vectors, i.e., Up ψ∀ ℓ = ψ∀ ℓ as shown in Eq.
(s)
(33.16). It follows that the sequence U P ψ∀ ℓ would also converge to
(s)
ψ∀ . On the other hand the sequence Up ψ∀ ℓ must also converge to
Up ψ∀ since Up is continuous. This means that the Up ψ∀ = ψ∀ which
implies that the limit vector ψ∀ is symmetrical. We can conclude that
symmetrical vectors form a subspace of H ∀ (c) .
We can also use the projector P(cs)  in Eq. (33.22) to check the
result. Since Up P(cs) = P (cs) we get Up P (cs) φ∀ = P (cs) φ∀, i.e., Up has
no effect on the vector P (cs) φ∀. This implies that the projector P (cs)
projects every vector φ∀ in H ∀ (c) onto symmetrical vector. This means
that the set of symmetrical vectors form the subspace onto which
P (cs) projects.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 34

Conceptual Issues

Q34(1) Verify that  ∀ (q, m) (t) in Eq. (34.22) satisfies the initial
condition in Eq. (34.16) and the Schrödinger equation (34.21).
SQ34(1) By setting t = 0 in Eq. (34.22) we can clearly see that
∀ (q, m) (t) satisfies the initial condition in Eq. (34.16). To prove

∀ (q, m) (t) satisfies the Schrodinger
 ¨ equation we first have
d (q, m)
i�  ∀ (t)
dt
 
= iρc− ϕ∀− ∗ − sin(ρt/�) η∀ (0)� − i cos(ρt/�) η∀ (−)�
 
+ iρc2 ϕ∀+ ∗ − sin(ρt/�) η (0)� − i cos(ρt/�) η∀ (+)�
 
= ρc− ϕ∀− ∗ − i sin(ρt/�) η∀ (0)� + cos(ρt/�) η∀ (−)�
 
+ ρc2 ϕ∀+ ∗ − i sin(ρt/�) η (0)� + cos(ρt/�) η∀ (+)� .

Next we have, using the Hamiltonian in Eq. (34.17)


∀ (q, m) (t)
 (q, m) 
H
 
(q) (m) (q) (m) ∀ (q, m)
= ρ P − ∗  L− + P + ∗ 
L+  (t)
 
= ρc− ϕ∀− ∗ cos(ρt/�)η∀ (−)� − i sin(ρt/�)η∀ (0)�
 
+ ρc+ ϕ∀+ ∗ cos(ρt/�)η (+)� − i sin(ρt/�)η (0)� .

The results above show that  ∀ (q, m) (t) satisfies the Schrodinger
¨
equation, i.e., it satisfies Eq. (34.21).
Quantum Mechanics: Problems and Solutions
K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 35

Harmonic and Isotropic Oscillators

Q35(1) Verify the results in Eqs. (35.5) to (35.8).


SQ35(1) The probabilities are given by Postulate 28.1(PPDO), i.e.,
by Eq. (28.1). In other words we have
℘ Hho (φ s , E 1 ) = �φ∀ | P Ĥho (E 1 )φ∀ ∪ = �φ∀ | |ϕ∀1 ∪�ϕ∀1 |φ∀ ∪ = 1/3,
℘ Hho (φ s , E 2 ) = �φ∀ | P Ĥho (E 2 )φ∀ ∪ = �φ∀ | |ϕ∀2 ∪�ϕ∀2 |φ∀ ∪ = 2/3,

� 1, 2 since |ϕ∀n ∪�ϕ∀n | φ∀ = 0.
℘ Hho (φ s , E n ) = 0 if n =
Given that E 1 = 3�ω/2 and E 2 = 5�ω/2 we can calculate the
expectation value E(H ho , φ s ), i.e.,
→ 1 3 2 5 13
E(H ho , φ s ) = ℘ Hho (φ s , E n ) E n = + �ω = �ω.
n=0
3 2 3 2 6
This is also is equal to �φ∀ | H  ho φ∀ ∪ since for the state vector φ∀ in Eq.
(35.4) we have  
1 2  1 2 
∀  ∀
�φ | H ho φ ∪ = � � ϕ∀1 + 
ϕ∀2 | H ho � ϕ∀1 + ϕ∀2 ∪
3 3 3 3
  1  
1 2 2
= � � ϕ∀1 + ϕ∀2 | � E 1 ϕ∀1 + E 2 ϕ∀2 ∪
3 3 3 3
  
1 1 2 2
= � � ϕ∀1 | E 1 � ϕ∀1 ∪ + � ϕ∀2 | E 2 ϕ∀2 ∪
3 3 3 3
1 2 1 3 2 5 13
= E1 + E2 = �ω + �ω = �ω.
3 3 3 2 3 2 6
Quantum Mechanics: Problems and Solutions
K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

190 Harmonic and Isotropic Oscillators

Q35(2) Verify Eq. (35.10).


SQ35(2) Using Eq. (35.9) we get
λ i i
[ a, a† ] = [ 
x+ x−
, 
p ]
p
2 mω mω
λ i i
= − [ ] +
x, p x ] = II ,
, 
[p
2 mω mω
using [   ] = i �II and λ = mω/�. This is the result in Eq. (35.10),
x, p
subject to domain restriction.
Q35(3) Verify that ϕ0 (x) in Eq. (35.13) satisfies Eq. (35.12).
SQ35(3) First differentiate ϕ0 (x) given in Eq. (35.13) to get
1
d λ 4 1 2
ϕ0 (x) = (−λx)e− 2 λ x = −λ x ϕ0 (x).
dx π
With λ = mω/� Eq. (35.12) follows immediately.
Q35(4) Obtain the first and the second excited state eigenvectors
from the expression in Eq. (35.19) in terms of ϕH0 (x).
SQ35(4) The general expression of Hermite polynomials are given
by Eqs. (16.14) and (16.15). Explicitly we have
H 0 (y) = 1, H 1 (y) = 2y, H 2 (y) = 4y 2 − 2,

where y = λ x. The corresponding Hermite functions defined by
Eq. (16.18) are1
ϕH 0 (x) = ϕ0 (x) in Eq. (35.13),
� �
ϕH 1 (x) = 2λ xϕ0 (x) = 2λ xϕH 0 ,
1 1
ϕH 2 (x) = � (2λ x 2 − 1) ϕ0 (x) = � (2λ x 2 − 1) ϕH 0 (x).
2 2
Our problem is to check that Eq. (35.19) leads to the same results.
Letting n = 1 and 2 in Eq. (35.17) we get2

† λ � d
ϕ∀H 1 (x) = a ϕ∀H 0 := x− ϕH 0 (x)
2 mω dx

λ 1 d
= xϕH 0 (x) − ϕ (x)
2 λ dx H 0

= 2λ xϕH 0 (x) = ϕH 1 .
1 Note that the factorial of zero is equal to 1, i.e., 0! = 1.
2 The results here can be compared with Eqs. (16.19) to (16.22).
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Harmonic and Isotropic Oscillators 191

1
ϕ∀H 2 (x) = � a† ϕ∀H 1
2

λ � d
:= x− ϕH 1 (x)
2 mω dx

λ 1 d
= xϕH 1 (x) − ϕ (x)
2 λ dx H 1

λ� 1 d 
= 2λ x 2 ϕH 0 (x) − ϕH 0 (x) + x ϕH 0 (x)
2 λ dx
1  2 1 
= � λ x ϕH 0 (x) − ϕH 0 (x) + x 2 ϕH 0 (x)
2 λ
1
= � (2λ x 2 − 1) ϕH 0 (x) = ϕH 2 .
2
Q35(5) Verify Eq. (35.14) and (35.15).
SQ35(5) To verify Eq. (35.14) we have
λ i i
a† a = x−
 
p x+
 
p
2 mω mω
λ 1 i i
= x2 +
 2
2 +
p 
xp− 
p x
2 (mω) mω mω
λ 1 i
= x2 +
 2
2 +
p [ ]
x, p
2 (mω) mω
mω 1 � 
= x2 +
 2
2 −
p II .
2� (mω) mω
To verify Eq. (35.15) we have
1 2 1 1
�ωa† a =  + mω2 
p x 2 − �ω
2m 2 2
 
� H ho = �ω N + 1 .
2
Q35(6) Verify that φ(x, t) in Eq. (35.39) is normalised and that
¨
it also satisfies the Schrodinger equation for time evolution of the
harmonic oscillator.
SQ35(6) Using Eqs. (35.43) for |φ(x, t) |2 and following formula of
integration
 → 
−a(x−b)2 π
e dx =
−→ a
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

192 Harmonic and Isotropic Oscillators

we can verify the normalisation of φ(x, t) immediately, i.e.,


 →
�φ∀(t) | φ∀(t)∪ = |φ(x, t) |2
−→
 →
1  
λ 2 −λ x−xc (t) 2
= e dx = 1.
−→ π
¨
To verify the Schrodinger equation we can perform the following
calculations. First let
i mω  2
f (x, t) = exp pc (t)x − x − xc (t) .
� 2�
Then we have φ(x, t) = C (t) f (x, t). We can carry out the following
calculations:
(1) For the time derivative we have
dxc (t) 1
= −ω xc (0) sin ωt = pc (t),
dt m
dpc (t)
= −mω2 xc (0) cos ωt = −mω2 xc (t),
dt
dC (t) 1 2 1 1
i� = p (t) − mω2 xc2 (t) + �ω C (t),
dt 2m c 2 2
∂ f (x, t)    
i� = mω2 xc (t)x + i ω x − xc (t) pc (t) f (x, t),
∂t
∂φ(x, t) dC (t)  ∂ f (x, t) 
i� = i� f (x, t) + C (t) i �
∂t dt ∂t
1 2 1 1
= p (t) − mω2 xc2 (t) + �ω φ(x, t)
2m c 2 2
   
2
+ mω xc (t) x + i ω x − xc (t) pc (t) φ(x, t).
(2) For the spatial derivatives derivative we have
  
 φ(x, t) = pc (t) + i mω x − xc (t) φ(x, t).
p
  2
 2 φ(x, t) = m�ω + pc (t) + i mω x − xc (t)
p φ(x, t).

2
p 1 1 2  
φ(x, t) = �ω + pc (t) + i ω pc (t) x − xc (t)
2m 2 2m
1 1 
− mω2 x 2 + mω2 x xc (t) − mω2 xc2 (t) φ(x, t).
2 2
(3) Comparing the above results we get
∂φ(x, t)  ho φ(x, t).
i� =H
∂t
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Harmonic and Isotropic Oscillators 193

¨
In other words the wave function φ(x, t) satisfies the Schrodinger
equation.
Q35(7) Verify Eqs. (35.44), (35.45).
SQ35(7) Using Eqs. (35.23) and (35.27) for φ(x, t) and |φ(x, t) |2
we get
 →
�φ∀(t) | 
x φ∀(t)∪ = x |φ(x, t) |2
−→
 →
1  2
λ 2
= x e−λ x−xc (t)
dx.
−→ π
We can rewrite the integral as the sum of two integrals, i.e.,

 −λ x−xc (t)2
1
→ 
λ 2
x − xc (t) e dx
−→ π
 →
1  2
λ 2 1
+ xc (t) e− 2 λ x−xc (t)
dx.
−→ π
The first integral can be written as
 →
1
λ 2
2
ye−λ y dy, y = x − xc (t).
−→ π
This integral vanishes on account of the integrand being an odd
function of y. Since the function φ(x, t) is normalised the second
integral is equal to xc (t), i.e., we get

�φ∀(t) | 
x φ∀(t)∪ = xc (t).

To calculate �φ∀(t) | p
 φ∀(t)∪ we first observe that
  
 φ(x, t) = pc (t) + i mω x − xc (t) φ(x, t).
p

The integral for �φ∀(t) | p φ∀(t)∪ becomes


 →  

pc (t) + i mω x − xc (t) | φ(x, t)|2 dx
−→
 →  →
  
= pc (t) | φ(x, t)|2 dx + − i mω x − xc (t) | φ(x, t)|2 dx
−→ −→
= pc (t),

since the second integral vanishes due to an odd integrand.


June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

194 Harmonic and Isotropic Oscillators

Q35(8) Verify Eq. (35.46).


SQ35(8) First we evaluate �φ∀(t) | x 2 φ∀(t)∪ which is given by
 → 1  
λ 2 2 −λ x−xc (t) 2

�φ (t) |  2 ∀
x φ (t)∪ = x e dx. (�)
−→ π
In order to employ the formula
 → 
2 1 π
x 2 e−ax dx = , (��)
−→ 2a a
we replace x 2 in the integrand of the integral in Eq. (�) above by
 2  2
x − xc (t) + xc (t) = x − xc (t) + 2(x − xc (t))xc (t) + xc (t)2 .
The integral in Eq. (�) becomes a sum of three integrals, i.e.,
 →  2 
1
λ 2 2 
x − xc (t) exp − λ x − xc (t) dx
−→ π
 →  2 
1
λ 2   
+ 2 x − xc (t) xc (t) exp − λ x − xc (t) dx
−→ π
 →  2 
1
λ 2 
+ xc (t)2 exp − λ x − xc (t) dx.
−→ π
The values of these integrals are as follows:
(1) On account of the formula in Eq. (��) above for the first integral
can be evaluated to give the value 1/2λ = �/2mω.
(2) The second integral vanishes due to an odd integrand.
(3) The third integral has a value of xc (t)2 due to φ(x, t) being
normalised.
The final result is

x 2 φ∀t ∪ = xc (t)2 +
�φ∀t |  .
2mω
Next we evaluate �φ∀(t) | p
 2 φ∀(t)∪ which is given by
 →

�φ (t) | p 2 ∀
 φ (t)∪ = φ � (x, t) p
 2 φ(x, t)dx.
−→
Using the result obtained for SQ35(6) for p  2 φ(x, t) we can see that
 2 φ∀(t)∪ is given by the following integral:
�φ∀(t) | p
 
→  2
�mω + pc (t) + i mω x − xc (t) |φ(x, t)|2 dx.
−→
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Harmonic and Isotropic Oscillators 195

Making used of previous results this integral can be evaluated as


follows:
(1) The integral over �mω has the value �mω.
(2) The integral over pc (t)2 has the value pc (t)2 .
(3) The integral over 2i pc (t)mω x − xc (t) has the value 0 due to an
odd integrand.  2
(4) The integral over −m2 ω2 x − xc (t) has the value −�mω/2.
The final result is
1
�φ∀(t) | p
 2 φ∀(t)∪ = pc (t)2 + m�ω.
2
It follows that the expectation value of the Hamiltonian is given by
 φ∀(t)∪ = 1 �φ∀(t) | p
�φ∀(t) | H
1
 2 φ∀(t)∪ + mω2 �φ∀(t) | 
x 2 φ∀(t)∪
2m 2
1 1 1
= pc (t)2 + mω2 xc (t)2 + �ω.
2m 2 2
This differs from the corresponding classical value by �ω/2.
Q35(9) Using the expressions in Eqs. (35.40) and (35.41) for
xc (t) and pc (t) show that energy expectation value in Eq. (35.46)
is conserved, i.e., the value is time-independent. Explain how
the expectation value can approximate the energy of a classical
harmonic oscillator.
SQ35(9) The explicit time-dependence of the energy expectation
value in Eq. (35.46) is given as follows:
�φ∀(t) | H  φ∀(t)∪
1 2 1 1
= p (t) + mω2 xc2 (t) + �ω
2m c 2 2
1  2 1 1
= − mωxc (0) sin ωt + mω2 xc2 (0) cos2 ωt + �ω
2m 2 2
mω2 2   1
= xc (0) cos2 ωt + sin2 ωt + �ω
2 2
mω2 2 1
= xc (0) + �ω,
2 2
which
 is  time-independent, i.e., the energy expectation value
 ∀
E H ho , φ is conserved.
The first term in the above expression for the expectation value is
equal to the classical value for a classical oscillator initially placed at
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

196 Harmonic and Isotropic Oscillators

rest at position xc (0). For sufficient large oscillations due to a large


initial displacement xc (0) the second term �ω/2 becomes negligible.
The above expectation value then approximates the energy of a
corresponding classical harmonic oscillator.
Q35(10) Verify Eqs. (35.58) and (35.59).
SQ35(10) Using Eqs. (35.56) and (35.57) we have:
(1) For �φ∀(0) | 
xHei (t)φ∀(0)∪ we have
�φ∀(0) | 
xHei (t)φ∀(0)∪
sin ωt
= (cos ωt) �φ∀(0) | 
x φ∀(0)∪ + �φ∀(0) | p
 φ∀(0)∪

= (cos ωt) xc (0) = xc (t).
We have used Eqs. (35.44) and (35.55) at t = 0 proved in
x φ∀(0)∪ = xc (0) and �φ∀(0) | p
SQ35(7), i.e., �φ∀(0) |   φ∀(0)∪ =
pc (0) = 0.
(2) For �φ∀(0) | p
Hei (t)φ∀(0)∪ we have
�φ∀(0) | p
Hei (t)φ∀(0)∪
x φ∀(0)∪ + (cos ωt) �φ∀(0) | p
= (−mω sin ωt) �φ∀(0) |   φ∀(0)∪
= (−mω sin ωt) xc (0) = pc (t).
Q35(11) Verify Eq. (35.60) by explicit calculation of H  hoHei (t)
using  Hei (t) in Eqs. (35.56) and (35.57).
xHei (t) and p
SQ35(11) From Eq. (35.47) we have
 hoH ei (t) = 1 p
H
1
 2 (t) + mω2  2
xHei (t)
2m Hei 2
1  2
= x + (cos ωt) p
(−mω sin ωt)  
2m
1  sin ωt 2
+ mω2 (cos ωt)  x+ 
p
2 mω
1 2 1  hoH ei (0).
=  + mω2 
p x2 = H
2m 2
We have used the following results:
 2
(−mω sin ωt) x + (cos ωt) p  = (−mω sin ωt)2  2
x 2 + (cos ωt)2 p
− mω(sin ωt)(cos ωt)(
xp+ p
x ),
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Harmonic and Isotropic Oscillators 197

2
sin ωt sin ωt 2 2
x+
(cos ωt)  
p = (cos ωt)2 
x2 + 
p
mω mω
(sin ωt)(cos ωt)
+ (
xp+ p
x ),

(sin ωt)2 + (cos ωt)2 = 1.

Q35(12) The Hamiltonian of a forced harmonic oscillator in the


Schrödinger picture is given in the usual notation by3
 
 Sch = 1 p
H 2
Sch
1
+ mω2  2
x Sch −g x Sch +
1
Sch ,
p
2m 2 mω
where g is a real number. Let aSch and a†Sch be a pair of operators
related to  Sch by Eq. (35.9).
x Sch and p

(a) Show that


 
 Sch = �ω a† aSch + 1 + γ aSch + γ � a† ,
H Sch
2 Sch

where


γ =− (1 − i )g.
2mω
(b) In the Heisenberg picture the annihilation and creation opera­

tors become aHei (t) and aHei (t) and the Hamiltonian becomes
 
H Hei (t) = �ω a† (t)aHei (t) + 1 + γ aHei (t) + γ � a† (t).
Hei
2 Hei

Show that aHei (t) satisfies the Heisenberg equation of motion


and that the Heisenberg equation can be wrtten as
d i
aHei (t) = −i ωaHei (t) − γ � .
dt �

Show further that this equation can be rewritten in the form


d   i
aHei (t)ei ωt = − γ � ei ωt .
dt �
Integrate this equation to obtain an explicit expression for the
time dependence of aHei (t).
3 Merzbacher pp. 335–336.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

198 Harmonic and Isotropic Oscillators

SQ35(12)(a) ¨
In the Schrodinger picture we have
  
λ i
aSch := x Sch +
 Sch ,
p
2 mω
  
λ i
a†Sch := x Sch −
 Sch .
p
2 mω

We can calculate the product aSch aSch in terms of  Sch as in
x Sch and p
SQ35(5), i.e.,
mω i i
a†Sch aSch = x Sch −
 Sch
p x Sch +
 
p
2� mω mω Sch
mω 2 1 2 i
= 
x Sch + Sch
p + (x p  −p Sch 
x Sch )
2� (mω) 2 mω Sch Sch
mω 2 1 2 1
= x Sch +
 Sch
p − .
2� 2m�ω 2
It follows that
1 2 1 1
�ω a†Sch aSch =  + mω2 
p 2
x Sch − �ω.
2m Sch 2 2
We can express  Sch in terms of the annihilation and creation
x Sch and p
operators, i.e.,

�  
x Sch =
 aSch + a†Sch ,
 2mω
�mω  
Sch = −i
p aSch − a†Sch .
2
It follows that
  
1 �  †
 1 �mω  

x Sch +  =
p aSch + aSch + −i aSch − a†Sch
mω Sch 2mω mω 2
 
�   �  
= 1 − i aSch + 1 + i a†Sch
2mω 2mω
 1 
� −g  x Sch + 
p
 mω Sch 
�   �  
= −g 1 − i aSch − g 1 + i a†Sch
2mω 2mω
= γ aSch + γ � a†Sch
 
 Sch = �ω a† aSch + 1 + γ aSch + γ � a† .
�H Sch
2 Sch
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Harmonic and Isotropic Oscillators 199

SQ35(12)(b) Not being selfadjoint the annihilation and creation


operators do not represent any observables. Postulate 29.3.1(HP)
does not directly apply to them. However, they are related to position
and momentum at time t = 0 by Eqs. (35.9), i.e.,
 
λ i † λ i
a = x+
  , a =
p x−
  .
p
2 mω 2 mω
These annihilation and creation operators are identifiable with aSch
and a†Sch in SQ35(12)(a). In the Heisenberg picture the position and
momentum operators will evolve into  Hei (t). It follows
xHei (t) and p
that the annihilation and creation operators will evolve into

λ i
aHei (t) = xHei (t) +
  (t) ,
p
2 mω Hei

† λ i
aHei (t) = xHei (t) −
  (t) ,
p
2 mω Hei
and that the Hamiltonian H  Sch will evolve into

H Hei (t) = �ωa† (t)aHei (t) + 1 �ω + γ aHei (t) + γ � a† (t).


Hei
2 Hei

These time-dependent operators possess the following properties:

(1) Using the commutation relation between  Hei (t), i.e.,


xHei (t) and p
[ Hei (t)] = i �II , we can verify that the commutation
xHei (t), p
relation between the annihilation and creation operators are
preserved, i.e., we have, from Eq. (35.10),

[ aHei (t), aHei (t)] = II .
(2) aHei (t) satisfies the Heisenberg equation of motion since

d λ d i d
i � aHei (t) =
 i�  x (t) + i� p (t)
dt 2 dt Hei mω dt Hei

=
λ
[  Hei (t)] + i [ p
xHei (t), H  Hei (t)]
 (t), H
2 mω Hei
 Hei (t)].
= [ aHei (t), H

The above commutator can be calculated, i.e.,


 Hei ] = [ aHei (t), �ω a† (t)aHei (t) + 1
[ aHei (t), H Hei
2
+ γ aHei (t) + γ � aHei

(t) ]
= �ω aHei (t) + γ � .
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

200 Harmonic and Isotropic Oscillators

The Heisenberg equation becomes

d
i� a (t) = �ωaHei (t) + γ � , or
dt Hei
d i
a (t) = −i ωaHei (t) − γ � .
dt Hei �
It follows that:

d   d
aHei (t)ei ωt = a (t) ei ωt + i ωaHei (t)ei ωt
dt dt Hei
i �
= −i ωaHei (t) − γ ei ωt + i ωaHei (t)ei ωt

i
= − γ � ei ωt .

Integrating the above equation, i.e.,
d   i
a (t)ei ωt = − γ � ei ωt ,
dt Hei �
from 0 to t we get
1 � i ωt 1 �
aHei (t)ei ωt − aHei (0) = − γ e + γ
�ω �ω
1 � 
� aHei (t) = aHei (0)e−i ωt − γ 1 − e−i ωt .
�ω

Q35(13) What is the degeneracy of the eigenvalues E 1, 1 , E 2, 2


of the two-dimensional isotropic oscillator? Are all eigenvalues
degenerate?
SQ35(13) From Eq. (35.86) we can see that the eigenvalue E 1, 1
of a two-dimensional isotropic oscillator is degenerate with a
degeneracy of 3 corresponding to the eigenvectors ϕ∀1, 1 , ϕ∀2, 0 and
ϕ∀0, 2 . The eigenvalue E 2, 2 is degenerate with a degeneracy of 5
corresponding to the eigenvectors ϕ∀2, 2 , ϕ∀4, 0 , ϕ∀1, 3 , ϕ∀3, 1 and ϕ∀0, 4 . Not
all eigenvalues are degenerate. The ground state energy eigenvalue
E 0, 0 is nondegenerate with corresponding eigenvector ϕ∀0, 0 .
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 36

Angular Momenta

Q36(1) The spherical harmonics Y1, 1 (θ, ϕ), Y1, 0 (θ, ϕ), Y1, −1 (θ, ϕ)
are given by Eqs. (16.65) to (16.67). In Cartesian coordinates, these
functions are given by1

3 x − iy
Y1, −1 (x, y, z) = ,
8π r

3 z
Y1, 0 (x, y, z) = ,
4π r

3 x + iy
Y1, 1 (x, y, z) = − .
8π r
(a) Using the expression in Cartesian coordinates for the quan­
tised orbital angular momentum operator  L cz in Eq. (27.86),
verify by explicit calculations that Y1, 1 (x, y, z), Y1, 0 (x, y, z) and
Y1, −1 (x, y, z) are the eigenfunctions of  L cz corresponding to
eigenvalues �, 0 and −�.
(b) Consider the following coordinate transformations:
x � z� , y � x �, z � y� .
1 Zettili
pp. 293–934. For convenience we have used the same symbols, e.g., Y1, 0 for
the functions in Cartesian coordinates.

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

202 Angular Momenta

(1) Show that the component of the orbital angular momentum


operator along the z� -direction is the same as that along the
L cz� = 
x-direction, i.e., show that  L cz .
(2) Show that the simultaneous eigenfunctions of  L 2c and 
L cx
 2 2
corresponding to eigenvalues of L equal to 2� are given
c
in Cartesian coordinates by

3 y − iz
X 1, −1 (x, y, z) = ,
8π r

3 x
X 1, 0 (x, y, z) = ,
4π r

3 y + iz
X 1, 1 (x, y, z) = − .
8π r

SQ36(1)(a) In Cartesian coordinates we have2 :



3 ∂ x − iy
x Y∀1, −1 := −i �
p
8π ∂x r

3 1 x − iy ∂
= −i � − r
8π r r2 ∂x

3 1 x − iy x
= −i � −
8π r r2 r

3 1  2 
= −i � 3
r − x2 + i xy .
8π r

3 ∂ x − iy
y Y∀1, −1
p := −i �
8π ∂y r

3 −i x − iy ∂
= −i � − r
8π r r2 ∂y

3 −i x − iy y
= −i � −
8π r r2 r

3 1  2 2

= −i � −ir − x y + i y .
8π r3

2 We have omitted the subscript c in the operators, e.g., writing �


pcx as �
px in
accordance with Eq. (27.90).
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Angular Momenta 203

 
Lz Y∀1, −1 := x p
 y − y p
x Y1, −1

3 1
= −i � (−i xr 2 − x 2 y + i x y 2 )
8π r 3

−(yr 2 − yx 2 + i x y 2 )

3 1 
= −i � 3
− i xr 2 − yr 2
8π r

3 1
= −� {x − i y} = −� Y1, −1 .
8π r
Next we have
  
3 ∂ z 3 z ∂ 3 zx
x Y∀1, 0 := −i �
p = −i � − 2 r = i� .
4π ∂ x r 4π r ∂x 4π r 3
 
3 ∂ z 3 z ∂
y Y∀1, 0 := −i �
p = −i � − r
4π ∂ y r 4π r2 ∂y

3 zy
= i� .
4π r 3
 
Lz Y∀1, 0
 := x py − y p
x Y1, 0
   
3 zy 3 zx
= x i� 3
− y i� = 0.
4π r 4π r 3

Finally we have
 
3 ∂ x + iy 3 1 x + iy ∂
x Y∀1, 1 := i �
p = −i � − r
8π ∂x r 8π r r2 ∂x

3 1 x + iy x
= i� −
8π r r2 r

3 1  2 
= i� 3
r − x2 − i xy ,
8π r
 
3 ∂ x + iy 3 i x + iy ∂
y Y∀1, 1 := i �
p = i� − r
8π ∂y r 8π r r2 ∂y

3 i x + iy y
= i� −
8π r r2 r

3 1  2 
= i� 3
ir − x y − i y 2 ,
8π r
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

204 Angular Momenta

 
Lz Y∀1, 1 := x p
 y − y p
x Y1, 1

3 1  2 
= i� i xr − x 2 y − i x y 2
8π r 3
 
= − yr 2 − yx 2 − i x y 2

3 1  2 
= i� 3
i xr − yr 2
8π r

3 1 
= −� x + i y = � Y1, 1 .
8π r

SQ36(1)(b)
(1) Consider a transformation from the original coordinate system
(x, y, z) to a new coordinate system (x � , y � , z� ):
x � z� , y � x � , z � y � .
We can see that the new z� -axis coincides with the original x-axis,
and that
 
r = x 2 + y 2 + z2 = x �2 + y �2 + z�2 = r � .
Physically we expect the angular momentum about the x-axis to
be the same as the angular momentum about the z� -axis. We can
confirm this mathematically by showing that
 ∂ ∂
Lx = y p z − z py = −i � y − z
∂z ∂y
is equal to
 ∂ ∂
Lz� = x � p x � = −i � x � � − y � � .
y � − y � p
∂y ∂x
This is obvious since x � = y, y � = z. It follows that the eigenfunctions
and eigenvalues of  Lx are the same in terms of the dashed
coordinates as that of  Lz� .
(2) We know the eigenfunctions Yℓ�, mℓ (x � , y � , z� ) and eigenvalues
of 
Lz� in the transformed coordinates (x � , y � , z� ) as they are of the
same form as those of the eigenfunctions and eigenvalues of  Lz in
the original coordinates, i.e., we have
 
� � � � 3 x � − i y� � � � � 3 z�
Y1, −1 (x , y , z ) = �
, Y1, 0 (x , y , z ) = ,
8π r 4π r �

3 x � + i y�
Y1,� 1 (x � , y � , z� ) = − .
8π r�
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Angular Momenta 205

Since Lz� = 
Lx these are also eigenfunctions of  Lx . Written in terms
of the original coordinates these functions become
 
� 3 y − iz � 3 x
Y1, −1 = , Y1, 0 = ,
8π r 4π r

3 y + iz
Y1,� 1 = − .
8π r
We can conclude that the required eigenfunctions  Lx , denoted more
conveniently by X 1, 1 , X 1, 0 , X 1, −1 are

3 y − iz
X 1, −1 = ,
8π r

3 x
X 1, 0 = ,
4π r

3 y + iz
X 1, 1 = − .
8π r
Q36(2) Suppose  L 2c and  L cz are measured giving the eigenvalues
2� and −�, respectively. A measurement of 
2
L cx is then made.
What are the possible results of the measurement of  L cx ? Find the
probability of each of these possible results.
SQ36(2) The first measurement serves as a state preparation
process, assuming the projection postulate. From the measured
values of  L2 and 
Lz we deduce that ℓ = 1, mℓ = −1. The projection
postulate tells us that the state after the first measurement is
described by the eigenfunction

3 x − iy
Y1, −1 = .
8π r
It also follows that the possible values of  Lx must be −�, 0, �, since
the orbital angular momentum quantum number ℓ is 1. According
to Postulate 28.1(PDDO) the probability of obtaining any of these
results are:
℘−1 = �Y∀1, −1 | P X∀ 1, −1 Y∀1, −1 ∪
 
= �Y∀1, −1 | |X∀ 1, −1 ∪�X∀ 1, −1 | Y∀1, −1 ∪

= �Y∀1, −1 | X∀ 1, −1 ∪ �X∀ 1, −1 | Y∀1, −1 ∪


 2
 
= �X∀ 1, −1 | Y∀1, −1 ∪ .
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

206 Angular Momenta

℘0 = �Y∀1, −1 | P X∀ 1, 0 Y∀1, −1 ∪
 
= �Y∀1, −1 | |X∀ 1, 0 ∪�X∀ 1, 0 | Y∀1, −1 ∪

= �Y∀1, −1 | X∀ 1, 0 ∪ � X∀ 1, 0 | Y∀1, −1 ∪
 2
 
= �X∀ 1, 0 | Y∀1, −1 ∪ .

℘1 = �Y∀1, −1 | P X∀ 1, 1 Y∀1, −1 ∪
 
= �Y∀1, −1 | |X∀ 1, 1 ∪�X∀ 1, 1 | Y∀1, −1 ∪

= �Y∀1, −1 | X∀ 1, 1 ∪ � X∀ 1, 1 | Y∀1, −1 ∪
 2
 
= �X∀ 1, 1 | Y∀1, −1 ∪ .

It is tedious to evaluate all these scalar products. A simpler way to


proceed is to recognise that if we expand the initial state Y∀1, −1 in
terms of the eigenvectors of  Lx , i.e.,
Y∀1, −1 = c−1 X∀ 1, −1 + c0 X∀ 1, 0 + c1 X∀ 1, 1 ,
then according to Eq. (28.4) the desired probabilities are equal to
the absolute value squares of the coefficients of expansion, i.e.,
℘−1 = |c−1 |2 , ℘0 = |c0 |2 , ℘1 = |c1 |2 .
Explicitly we have

3 x − iy
8π r
   
3 y − iz 3 x 3 y + iz
= c−1 + c0 + c1 − .
8π r 4π r 8π r

� x − i y = c−1 (y − i z) + c0 2 x − c1 (y + i z)

= 2 c0 x + (c−1 − c1 )y − i (c−1 + c1 )z.
Now, equate the coefficients of x, y, and z on both sides in turn we
can obtain the values of c1 , c0 , and c−1 :

1 = 2c0 , −i = c−1 − c1 , 0 = c−1 + c1
It follows that

c0 = 1/ 2, c−1 = −i /2, c1 = i /2,
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Angular Momenta 207

and
1 1 1
℘−1 = |c−1 |2 =, ℘0 = |c0 |2 = , ℘1 = |c1 |2 = .
4 2 4
The total probability is 1 as it must be.

Q36(3) The Hamiltonian of a classical particle of mass m


constrained to move freely on the surface of a sphere of radius a is
1 2
H = L ,
2I
where I = ma2 is the moment of inertia of the particle and L2
is the total orbital angular momentum square of the particle, both
with respect to the origin. Quantise the system and find the energy
eigenvalues and the corresponding eigenfunctions of the quantised
system. What are the degeneracy of the energy eigenvalues?3

SQ36(3) The state space of the system (a rigid rotator) is the


Hilbert space L∀2 (Su ) introduced in §16.1.2.8 and in E16.2.2(3),
i.e., it is defined by the set L2 (Su ) of square-integrable functions
defined on the unit sphere Su centered at the coordinate origin.
This set is spanned by the spherical harmonics Yℓ, mℓ (θ, ϕ) first
introduced in Eq. (16.63).4 Following the discussion in §36.1.1 the
square of the orbital angular momentum of the rigid rotator is
quantised as the operator  L2 (Su ) in Eq. (19.50).5 It follows that the
Hamiltonian is quantised as the operator H  = L2 (Su )/2I . The eigen­
functions are the spherical harmonics Yℓ, mℓ (θ, ϕ) corresponding to
eigenvalue
E ℓ, mℓ = ℓ(ℓ + 1)�2 /2I, ℓ = 0, 1, 2, · · · .
Apart from the eigenvalue E 0, 0 all the eigenvalues are degenerate
with degeneracy equal to the number of different mℓ values for a
given ℓ, i.e., 2ℓ + 1 since mℓ ranges from −ℓ to ℓ.

3 Zettili
pp. 296–297. Such a system is known as a rigid rotator which can be used to
model a diatomic molecule. For the state space of a quantum rigid rotator, see the
comment on a footnote in §27.8.
4 A general and explicit expression of Y
ℓ, mℓ (θ, ϕ) is given by Eq. (36.42).
5 See Eq. (36.8) and the discussion there.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

208 Angular Momenta

Q36(4) Verify the commutation relation in Eq. (36.51).


SQ36(4) Using the linearity and product rules for commutators

[A, B
 + C ] = [A, B
 ] + [A, C ],

[ A, B
 C ] = [A, B
 ] C + B
 [A, C ],

and the commutation relations of the annihilation and creation


operators, e.g.,
 
† † † † † †  1 (N
 2 + 1),
a1 a2 a2 a1 = a1 a1 a2 a2 = a1 a1 a2 a2 + 1 = N

J x, 
we can calculate the commutation relation [  J y ] as follows:

�2 † † † †
[
J x, 
Jy ] = [ (a1 a2 + a2 a1 ), (
a1 a2 − a2 a1 ) ]
4i
�2  † † †
= a1 a2 − a2 a1 ) ]
[ (a1 a2 , (
4i 
† † †
+ [ a2 a1 , (
a1 a2 − a2 a1 ) ]
�2  † † † †

= − [ a1 a2 , a2 a1 ] + [ a2 a1 , a1 a2 ]
4i
�2  † † † †

= − a1 [ a2 , a2 a1 ] − [a1 , a2 a1 ] a2
4i 
† † † †
+ a2 [ a1 , a1 a2 ] + [ a2 , a1 a2 ] a1
�2  † †
 
† †

= − a1 a1 + a2 a2 + a2 a2 − a1 a1
4i
�2  † †

= a2 a2 − a1 a1 = i �  J z.
2i

Q36(5) Verify Eqs. (36.54), (36.55) and (36.60).


SQ36(5) To verify Eq. (36.54) we only need to observe that N 1 =
†  2 = a† a2 . To verify Eq. (36.55) we need to express  2 2
a1 a1 and N 2 J x , Jy
and  J z2 in terms of the annihilation and creation operators, i.e.,

�2  
 † † † † † †
J x2 = (a1 a2 )2 + (a2 a1 )2 + a1 a2 a2 a1 + a2 a1 a1 a2
4
�2  
† †  1 (N
 2 + 1) + N
 2 (N
 1 + 1) ,
= (a1 a2 )2 + (a2 a1 )2 + N
4
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Angular Momenta 209

 �2  † 2 †  1 (N
 2 + 1) − N 2 (N

 1 + 1) ,
J y2 = − (a1 a2 ) + (a2 a1 )2 − N
4
2  
 � 2 + N 2 − N1 N2 − N 2 N
1 ,
J z2 = N 1 2
4
2  

� 
J2= 2 + N
N 1
2 + N
2
1 N2 + N 2 N
 1 + 2N  1 + 2N
2
4
�2   
2 2 + 2 N

1 + N2

= N1 + N
4
 N
N 
= + 1 �2 .
2 2
To verify Eq. (36.60) we observe that

 � † † † †

J+ =  Jx + iJy = (a1 a2 + a2 a1 ) + (a1 a2 − a2 a1 )
2

= � a1 a2 .
 � † † † †

J− =  Jx − iJy = (a1 a2 + a2 a1 ) − (a1 a2 − a2 a1 )
2

= � a2 a1 .

Q36(6) Verify properties P36.2.2(1), P36.2.2(2) and P36.2.2(3).


including Eqs. (36.73) and (36.74).

SQ36(6) The starting point is Eq. (36.62).


(1) To prove P36.2.2(1) we first note that n1 and n2 takes 0 and
positive integers values. Consequently m j = (n1 − n2 )/2 would
take integer and half integer values. It is positive when n1 > n2
and negative when n1 < n2 . It takes the value 0 when n1 = n2 .
Since j = (n1 + n2 )/2 it can only take the value 0, positive half
integer and integer values.
(2) To prove P36.2.2(2) we have
1 2  1 2 
m2 = n1 − 2n1 n2 + n22 , j 2 = n1 + 2n1 n2 + n22
4 4
� m2 ≤ j 2 � − j ≤ m ≤ j,
since n1 and n2 are both non-negative. If either n1 or n2 is zero
then J 2 = m2 . Since m = (n1 − n2 )/2 is negative when n1 = 0
we have m = − j and when n1 = 0 and m = j if n2 = 0.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

210 Angular Momenta

(3) To prove P36.2.2(3) we use the expressions


 †  †
| j, m∪ = ϕ∀n1 , n2 , J + = � a1 a2 , J − = � a2 a1 .

We get



J + | j, m∪ = � a1 a2 ϕ∀n1 , n2 = � (n1 + 1) n2 ϕ∀n1 +1, n2 −1 . (�)

Next we desire to express ϕ∀n1 +1, n2 −1 in the notation | j � , m� ∪. This


can be done by using Eq. (36.62) and rewriting ϕ∀n1 +1, n2 −1 as
ϕ∀n�1 , n�2 , where n�1 = n1 + 1 and n�2 = n2 − 1. Then we have:
1 � 1 
j� =(n1 + n�2 ) = (n1 + 1) + (n2 − 1) = j,
2 2
� 1 � � 1 
m = (n1 − n2 ) = (n1 + 1) − (n2 − 1) = m + 1,
2 2
� ϕ∀n1 +1, n2 −1 = | j, m + 1∪.

Equation (�) becomes




J + | j, m∪ = � (n1 + 1) n2 | j, m + 1∪. (��)

Since j = (n1 + n2 )/2 and m = (n1 − n2 )/2 we get n2 = j − m


and n1 = j + m, i.e., we have
 
n2 (n1 + 1) = ( j − m)( j + m + 1).

Equation (��) can be rewritten in the form of Eq. (36.73), i.e.,




J + | j, m∪ = � ( j − m)( j + m + 1) | j, m + 1∪.

Equation (36.74) is proved similarly, i.e., we have





J − | j, m∪ = � a2 a1 ϕ∀n1 , n2 = � (n2 + 1) n1 ϕ∀n1 −1, n2 +1 .

Writing ϕ∀n1 −1, n2 +1 in the notation | j �� , m�� ∪ and using Eq. (36.62)
we get
1 �� 1 
j �� =(n + n��2 ) = (n1 − 1) + (n2 + 1) = j,
2 1 2
1 1 
m�� = (n��1 − n��2 ) = (n1 − 1) − (n2 + 1) = m − 1,
2 2

n1 (n2 + 1) = ( j + m)( j − m + 1),

which immediately leads to Eq. (36.74).


June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Angular Momenta 211

Q36(7) ∀ in Eq. (36.108) normalised?


How is the vector 
SQ36(7) First we note that the functions φ+ (x∀ ) and φ− (x∀ ) are
normalised separately in accordance of Eq. (36.89), and that the spin
vectors α∀ and β∀ are orthonormal by definition. The vector 
∀ is then
normalised by requiring
 →
�∀ |∀ ∪ = |c+ |2 |φ+ (x∀)|2 dxdydz
−→
 →
2
|c− | |φ− (x∀)|2 dxdydz
−→
= |c+ | + |c− |2 = 1.
2

Q36(8) Prove Eqs. (36.124) and (36.125) directly using the


defining properties of spin operators given in P36.3(1) to P36.3(4)
without using Eqs. (36.82), (36.83) and (36.84).
SQ36(8) The answers are already given in item 8 of §36.3.5. First
we can construct the operators S± = Sx ± i Sy . These are operators
on VV∀ 2 . Then Eqs. (36.13) and (36.17) apply to spin operators since
these equations are derived using only the commutation relations of
the operators concerned. For examples we have
S+ S− = S2 − S2 + �Sz and S− S+ = S2 − S2 − �Sz
z z
Sz S+ = S+ Sz + �S+ and Sz S− = S− Sz − �S− .
Equation (36.124) is proved as follows:
(1) Applying Sz S+ = S+ Sz + �S+ to β∀z we get
  1
Sz S+ β∀z = S+ Sz + �S+ β∀z = − � + � S+ β∀z
2
1
= � S+ β∀z .
2
This means that S+ β∀z is an eigenvector of Sz corresponding to
eigenvalue �/2 which is nondegenerate. It follows that we must
have S+ β∀z = c+ α∀ z . We then have �S+ β∀z | S+ β∀z ∪ = |c+ |2 . The
constant c+ can be calculated as follows:
|c+ |2 = �S+ β∀z | S+ β∀z ∪ = �S− S+ β∀z | β∀z ∪
 
= � S2 − S2z − �Sz ) β∀z | β∀z ∪
3 1 1
= � − + �2 β∀z | β∀z ∪ = �2 .
4 4 2
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

212 Angular Momenta

Choosing c+ = � gives the answer.6


(2) Applying Sz S− = S− Sz − �S− to α∀ z we get
  1 1
Sz S− α∀ z = S− Sz − �S− α∀ z = � − � S− α∀ z = − � S− α∀ z .
2 2
This means that S− α∀ z is an eigenvector of Sz corresponding
to eigenvalue −�/2 which is nondegenerate. It follows that we
must have S− α∀ z = c− β∀z . The constant c− can be calculated as
follows:
|c− |2 = �S− α∀ z | S− α∀ z ∪ = �S+ S− α∀ z | α∀ z ∪
 
= � S2 − S2 + �Sz ) α∀ z | α∀ z ∪
z
3 1 1
= � − + �2 α∀ z | α∀ z ∪ = �2 .
4 4 2
Choosing c− = � gives the answer.

Equation (36.125) is proved as follows:


(1) Applying Sz S+ = S+ Sz + �S+ to α∀ z we get
  1
Sz S+ α∀ z = S+ Sz + �S+ α∀ z = � + � S+ α∀ z
2
3 
= � S + α∀ z .
2
This means that S+ α∀ z is an eigenvector of Sz corresponding
to the eigenvalue 3�/2. This is a contradiction since by
definition Sz does not have 3�/2 as an eigenvalue. To avoid the
contradiction we must have S+ α∀ z = 0. ∀ Another way of looking
∀ then it would be linear independent of
at this is that if S+ α∀ z �= 0
α∀ z and β∀z . This would make the state space three-dimensional,
contradicting defining property P36.3(2). So, we can conclude

that S+ α∀ z = 0.
(2) Applying the result Sz S− = S− Sz − �S− to β∀z we get
  1
Sz S− β∀z = S− Sz − �S+ β∀z = − � − � S− β∀ z
2
3
= − � S− β∀z .
2
6 See Merzbacher p. 240 for comments on this choice.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Angular Momenta 213

This means that S− β∀z is an eigenvector of Sz corresponding


to the eigenvalue −3�/2. This is a contradiction since by
definition Sz does not have −3�/2 as an eigenvalue. To avoid

the contradiction we must have S− β∀z = 0.
Q36(9) Using the matrix representation of S(1)x , S(1)y , S(1)z for a
spin-1 particle in Eqs. (14.48), (14.49) and (14.50) show that the
matrices for
S(1)+ = S(1)x + i S(1)y , S(1)− = S(1)x − i S(1)y , S 2 ,
(1)
are
⎛ ⎞
� 0 1 0
MŜ (1)+ = 2� ⎝ 0 0 1⎠,
0 0 0
⎛ ⎞
� 0 0 0
MŜ (1)− = 2� ⎝ 1 0 0 ⎠,
0 1 0
⎛ ⎞
1 0 0
MŜ (1)
2 = 2�2 ⎝ 0 1 0⎠.
0 0 1
SQ36(9) Using the expressions for MŜ (1)x and MŜ (1)y in Eqs. (14.48)
and (14.49) we get
⎛ ⎞
� 0 1 0
MŜ (1)+ = MŜ (1)x + i MŜ (1)y = 2� ⎝ 0 0 1 ⎠ ,
0 0 0
⎛ ⎞
� 0 0 0
MŜ (1)− = MŜ (1)x − i MŜ (1)y = 2� ⎝ 1 0 0 ⎠ .
0 1 0
For MŜ 2(1) we have to evaluate MŜ 2(1)x , MŜ 2(1)y and MŜ 2(1)z first. We have
⎛ ⎞2 ⎛ ⎞
2 0 1 0 2 1 0 1
� ⎝1 0 1⎠ = � ⎝0 2 0⎠,
MS2ˆ(1)x =
2 2
0 1 0 1 0 1
⎛ ⎞2 ⎛ ⎞
2 0 −i 0 2 1 0 −1
� ⎝ � ⎝
MS2ˆ(1)y = i 0 −i ⎠ = 0 2 0 ⎠,
2 2
0 i 0 −1 0 1
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

214 Angular Momenta

⎛ ⎞2 ⎛ ⎞
1 0 0 1 0 0
MŜ2(1)z = �2 ⎝ 0 0 0 ⎠ = �2 ⎝ 0 0 0 ⎠
0 0 −1 0 0 1
⎛ ⎞
2 0 0
�2 ⎝
= 0 0 0 ⎠.
2
0 0 2

It follows that
⎛ ⎞
1 0 0
MŜ 2(1) = MŜ2(1)x + MŜ2(1)y + MŜ2(1)z = 2�2 ⎝ 0 1 0 ⎠ .
0 0 1

Q36(10) Using Eqs. (36.126) and (36.127) verify that α∀ x , β∀x


defined by Eq. (36.137) are eigenvectors of Sx and α∀ y and β∀ y defined
by Eq. (36.138) are eigenvectors of Sy .
SQ36(10) Using the results of Sx and Sy acting on α∀ z and β∀z in Eqs.
(36.126) and (36.127) we get
1   1 1
Sx α∀ x = � Sx α∀ z + Sx β∀z = � � (β∀z + α∀ z )
2 2 2
1
= � α∀ x .
2
1   1 1
Sx β∀x = � Sx α∀ z − Sx β∀z = � � (β∀z − α∀ z )
2 2 2
1 ∀
= − � βx.
2
1   1 1
Sy α∀ y = � Sy α∀ z + i Sy β∀z = � � (i β∀z + α∀ z )
2 2 2
1
= � α∀ y .
2
1   1 1  
Sy β∀y = � Sy α∀ z − i Sy β∀z = � � i β∀z − α∀ z
2 2 2
1
= − � β∀y .
2
These results show that α∀ x and β∀x are eigenvectors of Sx , and α∀ y and
β∀y are eigenvectors of Sy . The corresponding eigenvalues are ±�/2.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Angular Momenta 215

Q36(11) The z-component spin is measured giving a value − 12 �.


What are the possible outcomes of a measurement of the x­
component spin? Find the probabilities of these possible measured
outcomes.

SQ36(11) This question is similar to Q36(2). The first measure­


ment of the z-component spin serves as a state preparation process,
i.e., the state after is given by the state vector β∀z . Then the question
turns to the outcomes of a measurement of the x-component spin
given state vector β∀z . There are always two possible outcomes
for a spin measurement, i.e., possible outcomes of a measurement
of the x-component spin are ±�/2. To find the probabilities let
P α∀ x = |∀
αx ∪�∀
αx | be the projector generated by α∀ x . Then Postulate
28.1(PDDO) tells us that the probability of a measurement of Sx
given state β∀z yield the value +�/2 is
 
℘x, + = �β∀z | P x β∀z ∪ = �β∀z | |∀ αx | β∀z ∪
αx ∪�∀

αx | β∀z ∪ = |�β∀z | α∀ x ∪ |2 .
= �β∀z | α∀ x ∪ �∀

Using the expression for α∀ x in Eq. (36.139) and the orthonormality


of α∀ z and β∀z the scalar product �β∀z | α∀ x ∪ is easily calculated, i.e., we
have
1  
�β∀z | α∀ x ∪ = �β∀z | � α∀ z + β∀z ∪
2
1 1
= � � ℘x, + = .
2 2
The probability for the value −�/2 is obtained in the same way, i.e.,
we have
1  
�β∀z | β∀x ∪ = �β∀z | � α∀ z − β∀z ∪
2
1 1
= −� � ℘x, − = .
2 2

Another approach would be to express the initial state vector β∀z in


terms of the eigenvectors of Sx , i.e., we write β∀z = c+ α∀ x + c− β∀x .
Using the expression for α∀ x and β∀x in Eqs. (36.137) and (36.138) we
can work out c+ and c− immediately, by equating the coefficients of
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

216 Angular Momenta

α∀ z and β∀z on both sides of the above equation, i.e., we have

1 1
β∀z = c+ � (α∀ z + β∀z ) + c− � (α∀ z − β∀z )
2 2
1 1
= � (c+ + c− )α∀ z + � (c+ − c− )β∀z
2 2
1
� (c+ + c− ) = 0, � (c+ − c− ) = 1
2
� �
� c+ = 1/ 2, c− = −1/ 2.

The required probabilities are |c+ |2 = 1/2 and |c− |2 = 1/2 which
agree with previous results.

Q36(12) Verify that Eqs. (36.146), (36.148) and (37.147) are


satisfied by the matrices in Eqs. (36.143), (36.145) and (36.144).

SQ36(12) First we have, from Eqs. (36.139) to (36.141),

1 0
C α∀ z = , C β∀z = ;
0 1

1 1 1 1
C α∀ x = � , C β∀x = � ;
2 1 2 −1

1 1 1 1
C α∀ y = � , C β∀y = � .
2 i 2 −i

Next wehave the matrix representations of Sz , Sx and Sy in the basis

α∀ z , β∀z given by Eqs. (36.143) to (36.145), i.e.,

1 1 0
MŜ z = � ,
2 0 −1

1 0 1
MŜ x = � ,
2 1 0

1 0 −i
MŜ y = � .
2 i 0
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Angular Momenta 217

It is then a matter of matrix multiplication, i.e., we have:


� 1 0 1 � 1
MŜ z C α∀ z = =
2 0 −1 0 2 0

= C α∀ ,
2 z
� 1 0 0 � 0
MŜ z C β∀z = =
2 0 −1 1 2 −1

= − C β∀z ,
2
� 0 1 1 � 1
MŜ x C α∀ x = � = �
2 2 1 0 1 2 2 1

= C α∀ ,
2 x

� 0 1 1 � −1
MŜ x C β∀x = � = �
2 2 1 0 −1 2 2 1

= − C β∀x ,
2
� 0 −i 1 � 1
MŜ y C α∀ y = � = �
2 2 i 0 i 2 2 i

= C α∀ y ,
2
� 0 −i 1 � −1
MŜ y C β∀y = � = �
2 2 i 0 −i 2 2 i

= − C β∀y .
2
Q36(13)7 A unit vector n∀ in the 3-dimensional IE∀ 3 aligned at an
angle θ to the z-axis on the x z plane is given by n∀ = sin θ i∀ + cos θ k∀.
The spin operator Sn∀ in the direction of the unit vector n∀ is given by
Sn∀ = n∀ · S, where
n∀ · S = nx Sx + n y Sy + nz Sz = sin θ Sx + cos θ Sz .

7 See Zettili pp. 298–316 for more examples. IE∀ 3 corresponds to the physical space we
live in.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

218 Angular Momenta

(a) Show that the matrix representation of Sn∀ in basis { α∀ z , β∀z } is

1 cos θ sin θ
MŜ n∀ = � ,
2 sin θ − cos θ

and that MŜ n∀ admits


⎛ ⎞ ⎛ ⎞
cos(θ/2) − sin(θ/2)
C η∀n∀+ = ⎝ ⎠, C η∀n∀− = ⎝ ⎠ (�)
sin(θ/2) cos(θ/2)

as eigenvectors corresponding eigenvalues ±�/2.


(b) Find matrix representation of the state vector of a spin aligned
in the direction n∀.8
(c) A spin is aligned in the positive direction n∀. Find the probabil­
ities of a measurement of spin along the z-axis resulting in the
values ± 12 �.
(d) A beam of spin- 21 particles with its spin aligned in the positive
direction n∀ is fed into a Stern–Gerlach apparatus oriented
to measure the component of the spin along the z-axis. The
incoming beam will split into two with the upper beam
corresponding to spin-up along the z-axis and the lower beam
corresponding to spin-down along the z-axis. Find the ratio of
the intensities of the emerging beams.

SQ36(13)(a) Using the matrix representation MŜ z of Sz in Eq.


(36.143) and the matrix representation MŜ x of Sx in Eq. (36.144)
we immediately obtain the matrix representation MŜ n∀ of Sn∀ in basis
α , β∀ }, i.e., we have
{∀
MŜ n∀ = (sin θ )MŜ x + (cos θ )MŜ z
� 0 1 � 1 0
= (sin θ ) + (cos θ )
2 1 0 2 0 −1

1 cos θ sin θ
= � .
2 sin θ − cos θ

8 Thenotation C η∀n∀ in Eq. (�) shows that the spin aligned in the direction n∀ should be
denoted by η∀n∀+ .
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Angular Momenta 219

Using the formulae


cos(α ± β) = cos α cos β √ sin α sin β,

sin(α ± β) = sin α cos β ± cos α sin β,


we get
1 cos θ sin θ cos(θ/2)
MŜ n∀ C η∀n∀+ = �
2 sin θ − cos θ sin(θ/2)

1 cos θ cos(θ/2) + sin θ sin(θ/2)


= �
2 sin θ cos(θ/2) − cos θ sin(θ/2)

1 cos(θ/2) �
= � = C η∀ ,
2 sin(θ/2) 2 n∀+
and
1 cos θ sin θ − sin(θ/2)
MŜ n∀ C η∀n−

= �
2 sin θ − cos θ cos(θ/2)
1 − cos θ sin(θ/2) + sin θ cos(θ/2)
= �
2 − sin θ sin(θ/2) − cos θ cos(θ/2)
1 sin(θ/2) �
= � = − C η∀n∀− .
2 − cos(θ/2) 2

These results shows that C η∀n∀+ and C η∀n∀− are eigenvectors of MŜ n∀
corresponding to eigenvalues �/2 and −�/2 respectively.
SQ36(13)(b) The required state vector must be the eigenvector
of the spin operator Sn∀ corresponding to the eigenvalue �/2. In
matrix representation this means that the matrix representation of
the state vector must be the eigenvector of MŜ n∀ corresponding to
the eigenvalue �/2. This eigenvector has been shown in SQ36(13)(a)
to be C η∀n∀+ in Eq. (�). This explains the notation, i.e., η∀n∀ + is the state
vector for the spin aligned along the n∀ direction and C η∀n∀+ is its matrix
representation in basis α∀ z , β∀z .
SQ36(13)(c) We can express the state vector η∀n∀ + in terms of a
linear combination of the spin-up state vector α∀ z and spin-down
state vector β∀z . In matrix terms this means that
cos(θ/2) 1 0
= cos(θ/2) + sin(θ/2) .
sin(θ/2) 0 1
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

220 Angular Momenta

The required probabilities for spin up and spin down are given
by the squares of the coefficients in the above expression, i.e., the
probabilities are given respectively by
 2  2
cos(θ/2) and sin(θ/2) .
We can check that total probability is 1, i.e.,
 2  2
cos(θ/2) + sin(θ/2) = 1.

SQ36(13)(d) For the Stern-Gerlach experiment the intensities


of the spin-up and spin-down beams are proportional to the
probabilities of a particle emerging to be spin-up or spin-down. It
follows that the required intensities ratio is
2 2
cos(θ/2) θ
= cot .
sin(θ/2) 2

Q36(14) The Pauli matrix σ y is given by


0 −i
σy = .
i 0

(a) Show that the Pauli matrix σ y possesses the following proper­
ties9 :
σ y0 = σ y2 = σ y4 = · · · = σ y2k = I 2×2 ,

σ y = σ y3 = σ y5 = · · · = σ y2k+1 ,


  →
1 1
ecσ y = c 2k I 2×2 + c2k+1 σ y , (�)
k=0
(2k)! k=0
(2k + 1)!

where I 2×2 is the 2 × 2 identity matrix, k = 0, 1, 2, 3, · · · , and


c ⇒ C.

9 For the exponential function, we can use the expansion


� 1 � �n
ecσ y = cσ y .
n!
n=0
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Angular Momenta 221

(b) Show that10


⎛ ⎞
cos 21 θ − sin 12 θ
1
e−i 2 θσ y =⎝ ⎠.
sin 12 θ cos 21 θ

(c) Let
1
U (θ) = e− 2 i θσ y .

Show that U (θ) is unitary and evaluate the unitary transform of


C α∀z in Eq. (36.139) generated by U (θ). Give an account on the
physical meaning of the transformation.

SQ36(14)(a) First we have σ y0 is equal to the 2 × 2 identity matrix


I 2×2 by definition, i.e., by definition the zero power of a square
matrix of order n is equal to the identity matrix of the same order.
Next we know from Eqs. (7.47) to (7.49) that the square of any of the
Pauli matrices are equal to the 2 × 2 identity matrix, e.g., σ y2 = I 2×2 ,
results which can be easily verified. It follows that any even power
of σ y is equal to I 2×2 . For an odd power of σ y we have

σ y2k+1 = σ y2k · σ y = I 2×2 · σ y = σ y .

Carrying out the expansion of the exponential in Eq. (�) we get

→
1  n
ecσ y = cσ y
n=0
n!

 →
1  2k  1  (2k+1)
= cσ y + cσ y
k=0
(2k)! k=0
(2k + 1)!

  →
1 1
= c 2k I 2×2 + c(2k+1) σ y .
k=0
(2k)! k=0
(2k + 1)!

10 Use the following expansions of cos x and sin x:



� →

(−1)k (−1)k
cos x = (x)2k , sin x = (x)2k+1 .
(2k)! (2k + 1)!
k=0 k=0
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

222 Angular Momenta

SQ36(14)(b) Replacing c by −i 12 θ in the above expansion of the


exponential we obtain an expansion of exp −i 12 θ σ y , i.e.,
 1   →
1
exp − i θ σ y = (−i θ/2)2k I 2×2
2 k=0
(2k)!

 1
=+ (−i (θ/2))(2k+1) σ y .
k=0
(2k + 1)!

Using the formular x mn = (x n )m we get


 k
(−i )2k = (−i )2 = (−1)k ,

(−i )(2k+1) = (−i )2k (−i ) = (−1)k (−i ).

The exponential function exp −i 12 θ σ y becomes


 1   →
(−1)k
exp − i θ σ y = ( θ/2)2k I 2×2
2 k=0
(2k)!

 (−1)k  (2k+1)
−i θ/2 σy
k=0
(2k + 1)!
= cos(θ/2) I 2×2 − i sin(θ/2) σ y
1 0 0 −i
= cos(θ/2) − i sin(θ/2)
0 1 i 0
1 0 0 −1
= cos(θ/2) + sin θ/2 ,
0 1 1 0
cos(θ/2) − sin(θ/2)
= .
sin(θ/2) cos(θ/2)

SQ36(14)(c) A square matrix is unitary if it satisfies Eq. (7.158),


i.e., its adjoint is equal to its inverse. The adjoint of U (θ) is

cos(θ/2) sin(θ/2)
.
− sin(θ/2) cos(θ/2)

This is equal to the inverse of U (θ ) since

cos(θ/2) sin(θ/2) cos(θ/2) − sin(θ/2) 1 0


= .
− sin(θ/2) cos(θ/2) sin(θ/2) cos(θ/2) 0 1
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Angular Momenta 223

The unitary transform of C α∀ z generated by U (θ) is given by


cos(θ/2) − sin(θ/2) 1
U (θ)C α∀ z =
sin(θ/2) cos(θ/2) 0
cos(θ/2)
= = C η∀n∀+ .
sin(θ/2)
Since C α∀ z in Eq. (36.139) and C η∀n∀+ in Eq. (�) correspond to the
spin orientation along the positive directions of z-axis and the vector
n∀ in Q36(13) respectively the transformation from C α∀ z to U (θ)C α∀ z
corresponds to a change of spin orientation from z-axis to n∀ .
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Chapter 37

Particles in Static Magnetic Fields

Q37(1) Verify Eq. (37.2).


SQ37(1) The vector potential is

∀ = − 1 y B i∀ + 1 x B ∀j .
A
2 2
Using the expression for the curl in Eq. (27.33) we get, due to A z = 0,

∀ = ∂ Ay ∂ Ax ∂ Ay ∂ Ax
�×A − i∀ + ∀j + − k∀
∂z ∂z ∂x ∂y
= B k∀.
Q37(2) Using Eqs. (37.8) to (37.10), show that the magnetic field
in Eq. (37.7) is derivable from the vector potential in Eq. (37.14) and
that the magnetic field in Eq. (37.16) is derivable from the vector
potential in Eq. (37.17).
SQ37(2) In cylindrical coordinates the vector potential A ∀ in Eq.
(37.14) has a non-zero component only in the θ direction, i.e.,

1
Ar = 0, A z = 0 and A θ (r) = Br.
2
Then Eqs. (37.8), (37.9) and (37.10) tell us that:
(1) Since r and A θ are independent of z and A z = 0 we get Br = 0.

Quantum Mechanics: Problems and Solutions


K. Kong Wan
c 2021 Jenny Stanford Publishing Pte. Ltd.
Copyright 
ISBN 978-981-4800-72-3 (Paperback), 978-0-429-29647-5 (eBook)
www.jennystanford.com
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

226 Particles in Static Magnetic Fields

(2) Since A z = 0 and Ar = 0 have Bθ = 0.


(3) So, the curl, and hence the resulting magnetic field, has only an
z component Bz (r) given by

1 ∂(r A θ ) ∂ Ar 1 ∂(Br 2 /2)


Bz = − = = B.
r ∂r ∂θ r ∂r

For the vector potential in Eq. (37.17) we have Ar = 0, A z = 0


and
$
Br/2, r < b (proportional to r)
A θ (r) = ,
b /2πr, r > b (inversely proportional to r)
where b = π b2 B is the magnetic flux enclosed by the circle of
radius b in the x-y plane centered at the origin. To show that this
is the correct potential for the given field we can carry out the
following calculations:
(1) Since A z = 0, r and A θ are independent of z we have Br = 0.
(2) Since A z = 0 and Ar = 0 have Bθ = 0.
(3) So, the curl, and hence the resulting magnetic field, has only a z
component Bz (r) given by
1 ∂(r A θ ) ∂ Ar
Bz = −
r ∂r ∂θ
⎧ 2

⎪ 1 ∂(Br /2)
⎨ = B for r < R
= r ∂r .
⎪ 1 ∂(b /2π ) = 0 for r > R


r ∂r

Q37(3) When spatial motion is neglected, the Hamiltonian of an


electron of charge −e and mass m in a uniform and static magnetic
field of magnitude B pointing along the z-direction is given by
 (s) = e B Sz .
H
m
¨
Write down the Pauli–Schrodinger equation for the evolution of a
spin state in the Schrödinger picture. Show that the following initial
spin state
1  
η∀(0) = � α∀ z + β∀z
2
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Particles in Static Magnetic Fields 227

will evolve to a new state η∀(t) at time t given

1  − 1 i ωt 1

η∀(t) = � e 2 α∀ z + e 2 i ωt β∀z ,
2
where ω = eB/m. What are the spin orientations initially at t = 0
and later at t = π/2ω?
SQ37(3) ¨
The Pauli-Schrodinger equation is
d  (s) η∀(t) = eB Sz η∀(t).
i� η∀(t) = H
dt m
The given vector η∀(t) satisfies the initial condition, i.e., we have
1
∀ = 0) = � (α∀ z + β∀z ).
η∀(0) = η(t
2
The given vector η(t)∀ also satisfies the Pauli-Schrodinger ¨ equation,
i.e., we have
dη∀(t) 1 d  − 1 i ωt 1

i� = i�� e 2 α∀ z + e 2 i ωt β∀z
dt 2 dt
$ 
1 1 − 12 i ωt 1 1
i ωt ∀
= i�� − iω e α∀ z + i ω e 2 βz
2 2 2
1 1  1 1

= �ω � e− 2 i ωt α∀ z − e 2 i ωt β∀z .
2 2

 
 η∀(t) = eB �1 e− 21 i ωt Sz α∀ z + e 21 i ωt Sz β∀z
H
m 2
$ 
eB 1 1 � 1 −�
= � e− 2 i ωt α∀ z + e 2 i ωt β∀z
m 2 2 2
� eB 1  1 1

= � e− 2 i ωt α∀ z − e 2 i ωt β∀z
2 m 2
dη∀(t) eB
= i� since = ω.
dt m

Hence, η∀(t) is a solution of the Pauli-Schrodinger


¨ equation. We can
conclude that η∀(t) is the state vector evolved from the initial state
vector η∀(0).
According to Eq. (36.137) the initial state vector η∀(0) is an
eigenvector of Sx corresponding to eigenvalue �/2. This implies
that initially the spin is oriented in the positive x-direction. At time
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

228 Particles in Static Magnetic Fields

t = π/2ω we have
1  − 1 iω π 1 π

η∀(π/2ω) = � e 2 2ω α∀ z + e 2 i ω 2ω β∀z
2
1  −i π π

= � e 4 α∀ z + ei 4 β∀z
2
1 −i π  π π

= � e 4 α∀ z + ei 4 ei 4 β∀z
2
1 −i π  π

= � e 4 α∀ z + ei 2 β∀z
2
1 −i π  
= � e 4 α∀ z + i β∀z .
2
According Eq. (36.138) this represents a spin orientation along
positive y-direction. The complex phase factor does not affect this
conclusion. The spin evolves from x-direction to y-direction.
June 23, 2020 10:50 JSP Book - 9in x 6in SolutionsManual(C2019)

Bibliography

Fano, G. (1992). Mathematical Methods of Quantum Mechanics, McGraw-Hill,


New York.
Halmos, P. R. (1958). Finite-Dimensional Vector Spaces, Van Nostrand,
Princeton.
Isham, C. J. (1995). Lectures on Quantum Theory, Imperial College Press,
London.
Jauch, J. M. (1968). Foundations of Quantum Mechanics, Addison-Wesley,
Reading, Mass.
Jordan, T. F. (1969). Linear Operators for Quantum Mechanics, Thomas F.
Jordan, Duluth, Minnesota, USA.
Merzbacher, E. (1998). Quantum Mechanics, 3rd edition, Wiley, New York.
Papoulis, A. (1965). Probability, Random Variables, and Stochastic Processes,
MaGraw-Hill, New York.
Prugoveˇcki, E. (1981). Quantum Mechanics in Hilbert Space, 2nd edition,
Academic Press, New York.
Roman, P. (1975). Some Modern Mathematics for Physicists and Other
Outsiders, Vol. 2, Pergamon, New York.
Wan, K. K. (2006). From Micro to Macro Quantum Systems, Imperial College
Press, London.
Wan, K. K. (2019). Quantum Mechanics: A Fundamental Approach, Jenny
Stanford Publishing, Singapore.
Weidmann, J. (1980). Linear Operators in Hilbert Spaces, Springer-Verlag,
New York.
Zettili, N. (2001). Quantum Mechanics, Wiley, Chichester.

You might also like