You are on page 1of 343

Exercises in Quantum Mechanics

Kluwer Texts in the Mathematical Sciences


VOLUME 6

A Graduate-Level Book Series

The titles published in this series are listed at the end of this volume.
Exercises in
Quantum Mechanics
A Collection of Illustrative Problems
and Their Solutions
Second Revised Edition
by

Harry Mavromatis
Department of Physics,
King Fahd University of Petroleum and Minerals,
Dhahran, Saudi Arabia

Springer-Science+Business Media, B.V.


Library ofCongress Cataloging-in-Publication Data

"CIP-data available from the Publisher on request."

ISBN 978-94-010-5172-9 ISBN 978-94-011-2652-6 (eBook)


DOI 10.1007/978-94-011-2652-6

Printed on acid-free paper

AlI Rights Reserved


© 1992 Springer Science+Business Media Dordrecht
Originally published by Kluwer Academic Publishers in 1992
Softcover reprint ofthe hardcover lst edition 1992
No part ofthe material protected by this copyright notice may be reproduced or
uti1ized in any form or by any means, electronic or mechanical,
including photocopying, recording or by any information storage and
retrieval system, without written permission from the copyright owner.
To: Vasso

and

To: Anthony and Blanche

Tn f.p.a un f.uTiv
Table of Contents

Preface IX
Preface to Second Edition Xl
Schematic Illustration Xlll

Chapter 1 Wilson-Sommerfeld Quantization Condition 1


Chapter 2 The Delta Function, Completeness, and Closure 16
Chapter 3 Momentum Space 34
Chapter 4 Wavepackets and the Uncertainty Principle 47
Chapter 5 Uncertainty Principle and Ground-State Energies
of Quantum-Mechanical Systems 68
Chapter 6 Free Particles Incident on Potentials, Time Delay,
Phase Shifts, and the Born Approximation 74
Chapter 7 Heisenberg Representation 102
Chapter 8 Two-, Three-, and N-, Versus One-Dimensional
Problems 111
Chapter 9 'Kramers' Type Expressions, The Virial Theorem,
and Generalizations 124
Chapter 10 Upper Bounds and Parity Considerations 133
Chapter 11 Perturbation Theory 157
Chapter 12 Degeneracy 186
Chapter 13 The Inverse Problem 214
Chapter 14 The Dalgarno-Lewis Technique 227
Chapter 15 Angular Momentum and Coupled States 262
Chapter 16 Tensor Operators, and Evaluation of Matrix Elements 291
Chapter 17 Applications of Quantum Mechanics 312
Index 326
Preface

This monograph is written within the framework of the quantum mechanical


paradigm. It is modest in scope in that it is restricted to some observations and
solved illustrative problems not readily available in any of the many standard (and
several excellent) texts or books with solved problems that have been written on
this subject. Additionally a few more or less standard problems are included for
continuity and purposes of comparison.
The hope is that the points made and problems solved will give the student
some additional insights and a better grasp of this fascinating but mathematically
somewhat involved branch of physics.
The hundred and fourteen problems discussed have intentionally been chosen to
involve a minimum of technical complexity while still illustrating the consequences
of the quantum-mechanical formalism.
Concerning notation, useful expressions are displayed in rectangular boxes while
calculational details which one may wish to skip are included in square brackets.

Beirut HARRY A. MAVROMATIS


June, 1985

IX
Preface to Second Edition

More than five years have passed since I prepared the first edition of this mono-
graph.
The present revised edition is more attractive in layout than its predecessor, and
most, if not all of the errors in the original edition (many of which were kindly pointed
out by reviewers, colleagues, and students) have now been corrected.
Additionally the material in the original fourteen chapters has been extended with
significant additions to Chapters 8, 13, and 14.
Three new chapters, Chapters 15-17 were added in order to make this volume more
comprehensive and more complete in its coverage of Elementary Quantum Mechanics,
principally by including material on angular momentum coupling and tensor algebra.
As a result of these additions and the revisions of the original fourteen chapters, the
present monograph includes 228, as opposed to the 114 solved exercises in the first
edition.
The author would like to acknowledge the support of the King Fahd University
of Petroleum and Minerals in this project.

Dhahran Harry A. Mavromatis


October, 1991

Xl
Schematic illustration of
various approaches to
calculating Energy Levels
of
Quantum Mechanical Systems

Energy Levels:
Approach for
High-lying States:
Generally useful
approaches: 1) Wilson-Sommerfeld
Quantization Condition
1) Schrodinger equation (Chapter 1)
in Momentum Space
(Chapter 3)

2) Poles of Scattering
Amplitude
(Chapter 6)

3) Schrodinger equation
in Coordinate Space
(Chapters 8, 9, 12, 16)

4) Perturbation Theory
(Chapter 11) Approaches for Ground State:

1) Uncertainty Principle
5) Dalgarno-Lewis (Chapter 5)
Technique
(Chapter 14)
2) Variational Technique
(Chapter 10)

Xlll
CHAPTER 1

Wilson-Sommerfeld Quantization Condition

The hydrogen atom, when treated using Bohr's admixture of classical and quan-
tum concepts involves an electron circulating about a proton (subject to the attractive
Coulomb force - (e 2 147rfor2) r) in orbits which satisfy the condition:

27rr = nADe Broglie, n = 1, 2, 3 ....

Since ADe Broglie = hlp this reduces to p27rr = nh, which may be generalized to
the Wilson-Sommerfeld quantization condition:

f pdq = nh n = 1, 2, 3... , (1.1)

where § implies a complete cycle, and p and q are conjugate variables.


'By construction' Eq. (1.1) gives the correct quantized energies for the hydrogen
atom. But it also gives the correct energy spectrum for a particle in a box with
infinite walls:

0, 0< x < a,
V(x) = {
00, x < 0, x> a.

EXAMPLE 1.1
Find the energy levels for a particle in a box with infinite walls:

0, 0< x < a,
V(x) = {
00, x < 0, x> a.

In the region 0 < x < a, E = p2/2m.


Hence Eq. (1.1) in this case becomes

f .J2mEdx = nh
where a cycle involves x varying from 0 --7 a and from a --7 O.
Integrating one obtains:

n 2 h2
2a.J2mE = nh or E = - 82 ' n = 1, 2, 3 .... (1.2)
ma
2 Chapter 1
By contrast the quantum-mechanical treatment of this problem involves solving
the Schrodinger equation:

"liZ dZ
[- 2m dxz + V(x) ] 'I/;(x) = E'I/;(x), (1.3)

V( xl

~---- ______ ~ ____________ x


o a

Figure 1.1: Potential in Example 1.1.

for 0 < x < a with boundary conditions '1/;(0), 'I/;(a) = 0, 'I/;(x) being zero for
x < 0, x> a.
The properly normalized eigenfunctions of Eq. (1.3) which satisfy these boundary
conditions are:

(:2 k
"li 2 2
'I/;(x)=V~ sinkx with ka=mr, where 2m =E.

If
Thus
'1/;(0) = sinO = 0,

and

'I/;(a) = If sinka = 0 if ka = mr, n = 1, 2, 3... ,

while
Wilson-Sommerfeld Quantization Condition 3
by construction.
Since

this implies

exactly the result Eq. (1.2).

One can gain a little more insight as to the range of applicability of the Wilson-
Sommerfeld quantization condition by studying slightly more complicated systems.

EXAMPLE 1.2
Find the energy levels of a particle in the potential:

0, 0< x < a,

V(x)= 1 Vo, a<x<a+b,


00, x < 0, x > a + b.
(Assuming E > Vo which corresponds to the interesting case.)

V( x)

Vo

~~----------~--------------~--------x
a o+b

Figure 1.2: Potential in Example 1.2.


4 Chapter 1
The Wilson-Sommerfeld quantization condition can be immediately applied to
this case:
fa a+b
2 10 PI dx + 2 a P2 dx = nh
l
where

PI = hkl = J2mE, P2 = hk2 = V2m(E - Vo)


I.e.

( 1.4)

On the other hand, solving the Schrodinger equation in the regions 0 < x < a
and a < x < a + b ('IjJ(x) being zero for x:::; 0, x ~ a + b) yields

'IjJ(x) = AsinkIx, 0 < x < a,

'IjJ(x)=Bsink2(x-a-b), a<x<a+b,
since 'IjJ(0) and 'IjJ(a + b) must be zero.
The continuity of 'IjJ(x) , d'IjJ(x)jdx at x = a then implies

kl
tan kl a = - k2 tan k2b. (1.5)

One notes that the conditions specified in Eq. (1.4) and Eq. (1.5) are differ-
ent. Only if k2 rv kl i.e. E > > Vo that is if the total energy is large compared to
the potential energy does Eq. (1.5) reduce to Eq. (1.4) since then tan kla rv - tan k2b
which is satisfied if
kla + k2b = mr, n = 1, 2, 3 ....
A second example which illustrates the range of applicability of the Wilson-
Sommerfeld quantization condition is the finite square-well problem.

EXAMPLE 1.3
Find the energy levels for a particle in a square well

0, Ixl > a,
v- {
-IVoI, Ixl < a,
where E < o.
Wilson-Sommerfeld Quantization Condition 5
The kinetic energy
p2
T = -
2m
= IVoI - lEI for Ixl < a.
The Wilson-Sommerfeld quantization condition therefore yields in this case

l.e.
(1.6)

v (x)

-0 o
----------------.---------r--------.---------------- x
- - - - - - - - - - - - - - - - 1 - - - - - -lEI
T
L---_ _ t -IVol
+--_~

Figure 1.3: Potential in Example 1.3.

On the other hand, solving the Schrodinger equation in the regions Ixl < a,
Ix I > a yields respectively

1/Je(x) = Acoskx Ixl < a, 1/Je(x) = Bexp{-Klxl} Ixl > a,


where
k= K-
_J2m EI l
~'
or

1/Jo(x) = Asinkx Ixl < a, 1/Jo(x) = ±Bexp {-Klxl} x~ ± a,


(where the subscripts e, 0 stand for even and odd solutions under the interchange
x ---+ -x.)
6 Chapter 1
The continuity of 'l/J'f,(x) and d'l/J'f,(x)/dx at x = ± a (independently for the odd
and even solutions) implies:
K
tan ka = (even solution),
k
and ( 1.7)
cot ka = (odd solution).
k
As in the previous example, Eq. (1.6) and Eq. (1.7) are not identical. Only if the
total energy is large compared to the potential energy, i.e. IVoI » lEI
(since these quantities are negative) i.e. K/k -+ 0 do they agree, since then

tanka = 0 7r,27r ...


and i.e. ka = {
cot ka = 0 7r /2, 37r /2 ... ,
or
n7r
ka = 2' n = 1, 2, 3... ,
yielding

I.e.

which is identical to Eq. (1.6).

Though the Wilson-Sommerfeld quantization condition was superseded by Quan-


tum Mechanics (with the Schrodinger and Heisenberg formulations in the early twen-
ties), as a calculational aid it has the advantage over the Schrodinger equation that it
is easier to work with since it involves an integral rather than a differential equation.
However, as indicated in Examples 1.2-3, it generally gives results which are reason-
ably accurate (i.e. in agreement with Quantum Mechanics) only when the energy is
large compared to the potential under consideration.

If V(x) = Alxl P one can obtain the form of the energy sequence according to the
Wilson-Sommerfeld quantization condition Eq. (1.1) as follows:

can be written (for En > 0) as


Wilson-Sommerfeld Quantization Condition 7
Hence

or generally:

{ I
AII/Ph }2P/(P+2)
lEn I = n 2p /(p+2) n = 1, 2, ... (p > -2), (1.8)
V2m I(p)

f -Iul du
where
I(p) = )1 p if En > 0,

I(p) = f Vlul
and
du P - 1 if En < o.
As p - t 00, En in Eq. (1.8) becomes ex n 2 , the result Eq. (1.2) for a particle in a
box with infinite walls. [In detail 1(00) =4, E(X> = n 2Aoh 2/32m = n 2h 2 /32m. This
corresponds to a = 2 in Eq. (1.2), i.e. V = 0, Ixl < 1, V = 00, Ixl > 1.]
In several cases I(p) can be easily evaluated directly.

EXAMPLE 1.4
Find the energy levels for a particle in the well

V(x) = Ax2, -00 < X < 00,

using the Wilson-Sommerfeld quantization condition.


If p = 2 Eq. (1.8) reduces to

(
A )1/2 h
En = n 2m 1(2) ,

where

l.e.
En = n'hlff., n= 1, 2, 3 .... (1.9)

This result can be compared to the familiar Schrodinger equation result:

(1.10)
8 Chapter 1
EXAMPLE 1.5
Find the energy levels for a particle in the well:

VeX) = { AX2, x> 0,

00, x < 0,
using the Wilson-Sommerfeld quantization condition.
This problem goes through like Example 1.4, except

1(2) = 2 1
o
1
~ du = -
1r

2
and hence

En = 2nhV{2A
-;;;:, n = 1, 2, 3 ... , (1.11)

as opposed to the Schrodinger equation result:

En = (2n - Dhfg, n= 1, 2, 3.... (1.12)

EXAMPLE 1.6
Find the energy levels for a particle in the well

Vex) = Alxl, all x (A> 0).


Here p = 1 and Eq. (1.8) reduces to:
2/3
E - n2/3 ( Ah )
n - y'2r;i"1(1) ,

1 vr=x
where
1 8
1(1) = 4 dx = -,
o 3
l.e.
n = 1, 2, 3.... (1.13)

as compared to the solution of the Schrodinger equation (see Eq. (3.31) ) for this
problem in the limit of large En namely:

En = {n - ~r/3 (A~2) 1/3 c~r/3, n= 1, 2, 3....


Wilson-Sommerfeld Quantization Condition 9
EXAMPLE 1.7
Find the energy levels for a particle in the well

Alxl, x > 0 (A> 0),


V(x) = {
00, x < o.
This problem goes through as in Example 1.6 except

1(1) = 210 vT=X dx =


1
4/3,

and hence

_ 2/3 (A21i,2)1/3 (~)2/3


En - n /0 ' n = 1, 2, 3... , (1.14)
m 2v2
as compared to the solution of the Schrodinger equation (see Eq. (3.18)) for this
problem in the limit of large En namely:

EXAMPLE 1.8
Find the energy levels for a particle in the well:

IAI
V(x) = -~' all x.

Here p = -1 and E < o. Thus

where
1( -1) = 410 J~ - 1 du = 27r,
1

I.e.

E __ 2mlAI2 1
n - 1i,2 n2 '
(1.15)
which is identical to the energy levels obtained for this problem using the Schrodinger
equation (see Eq. (3.32) ).
10 Chapter 1
EXAMPLE 1.9
Find the energy levels for a particle in the well:

-IAI/x, x> 0,
V(x) = {
00, x < O.

One proceeds as in Example 1.8 except

1(-1) = 210 J~ -1 du =
1
7[",

and hence
n = 1, 2, 3 ... , (1.16)

which is identical to the energy levels obtained for this problem using the Schrodinger
equation (see Eq. (3.23) ).

EXAMPLE 1.10
Find the energy levels for a particle in the well

V(x) = AlxI 1/ 2 , all x.

Here p = 1/2, and E > O.


Thus
E - n 2/ 5 {
A2h }2/5
n - V2rri 1(1/2) ,
where

1.e.

En = n 2/ 5 (
A2h 15
V2rri
)2/5 ' n = 1, 2, 3.... (1.17)
2m 32

EXAMPLE 1.11
Find the energy levels for a particle in the well:
Ax1/2, x> 0,
V(x) = {
00, x < O.

One proceeds as in Example 1.10 except


Wilson-Sommerfeld Quantization Condition 11
I.e.
n = 1, 2, 3 .... (1.18)

EXAMPLE 1.12
Find the energy levels for a particle in the well:
IAI all x.
V(x) = -lxI 1/ 2 '

Here p = -1/2 and E < O.


Thus
E - n- 2 / 3 { IAI-2h }-2/3
n = 1, 2, 3, ... ,
I nl- .,fiiTi 1(-1/2) ,
where

1.e.
IAI- 2h) -2/3
( .,fiiTi
En = _n- 2 / 3 7r ,n = 1, 2, 3 .... (1.19)

EXAMPLE 1.13
Find the energy levels for a particle in the well:
-IAlx-l/2, x> 0,
V(x) = { (E < 0)
00, x < O.
One proceeds as in Example 1.12 except

1( -~) = 2101 VU- 1/ 2 - 1 du = i'


1.e.
_
En - -n
-2/3 (IAI-22h)
.,fiiTi
-2/3 , n = 1, 2, 3 .... (1.20)
2m 7r

Generally

I(p) = f VI -Iul p du = 4101 viI - uP du = ~ 101(1 - W)1/2 w {1!P)-1 dw

( 1.21)
= ±B (~ ~) _ 2v1iT ( {I / p} + 1 )
p 2' p - r( {1/p} + {3/2} ) .
12 Chapter 1
EXAMPLE 1.14
Find the energy levels for a particle in each of the following wells:

Vi(x) = Ax 1 / 3 , V2(x) = Ax 1 / \ ltJ(x) = AX 1/ 5 , all x,

using Eq. (1.21).


From Eq. (1.21): I(1/3) = 64/35, I(1/4) = 512/315, I(1/5) = 1024/693.
Hence
A3h 35 }2/7
En(P = 1/3) = n2/7...;2m
{ , (1.22)
2m 64

E ( = 1/4) = n2/9 { A4h 315 }2 / 9 (1.23)


n P ...;2m 512 '

E (p = 1/5) = n2/11 { A5h 693 }2/11 (1.24)


n ...;2m 1024
The results of Example 1.4-14 are summarized in Table 1.1.

For two-dimensional systems (or three-dimensional systems where a particle moves


in a plane chosen for convenience to be the x-y plane), where the potential only
depends on p,
P~ P~
E = -2
m
+ -mp
22 + V(p). (1.25)

f
Hence
P¢ de/> = n¢h, n¢ = 1, 2, 3 ... , (1.26)
which implies
P¢ = n¢n,
and
fpp dp= nph, np = 1, 2, 3 ... , (1.27)
I.e.

f 2mE - (n¢h)2 - 2mV(p) dp = nph,


p2
np = 1, 2, 3... . (1.28)

The integral in Eq. (1.28) can be evaluated analytically for certain problems.

EXAMPLE 1.15
Find the energy levels for a particle of mass m in the potential
Wilson-Sommerfeld Quantization Condition 13
TABLE 1.1
Energy levels for various potentials Vex) = AlxlP

Wilson-Sommerfeld
p Range quantization condition Schrodinger result

1 x ;:: 0 E
n
= n 2/3 ( A2/;2
m
)1/3 ( 3". )2/3
2T2 En
= (n _ 1)2/3
4
( A2/;2t3( 311: y/3
m 2T2
(for large E)

1 -00 < x < 00 E n = n 2/3 ( A2/;2


m
r/ 472 y/3
3 ( 3". E n = (n 2 ( A2/;2t3(
_ 1)2/3 m 311: 3
472 t
(for large E)

2 x;::o En = 2n C~) 1/2 ~ A~) 1/2 -1h:


E n = (~2n- -2\) (--2m

2 (:Ay/2
= (n - t)
2 -00 <x< 00 En = n (A~f/2
2m
!l '7r
En 2m !l
".

-1 x;::o En = _n- 2(!n;:t_) En = _n- 2(m;~2_)

-1 -00 <X < 00 En = _n- 2(?m~tI2_) E n -- -n -2 (?mltI2_)


Ii.

1/2 x;:: 0 E n-n


- 2/5 (A
,~
h15;r/5
2

1/2 -00 <X < 00 E n -- n 2/5 (A 2 h15 r / 5


,Fm32 i

-1/2 x;::o E n = _n- 2/3(lAI- 2 2hJ- 2/ 3


,Fm"";

-1/2 -00 < X < 00 En -__ n -2/3 (~)-2/3


2m ".

1/3 -00
- 2/7 (A h35i )2/7
< X < 00 E n-n 3
,~,

1/4 -00 < X < 00 E n = n 2/9 CA4h 315 )2/9


2m 512i

1/5 -00 < X < 00 E n = n 2/11 ASh 1024


(V2m 693 ) 2/7
14 Chapter 1
From Eq. (1.28) one has:

f V2mEp2 - (n ",n)2 - m 2w 2p4


p dp = nph. (1.29)

One can either evaluate the integral in Eq. (1.29) using complex integration l , or
by elementary methods.
Thus defining

A == - (nq,n)2, B == 2mE, C == _m 2w 2, u == p2,

2nph = f VA + ~u + Cu 2 du

= 2 A (l umax du
Umin UV A + Bu + Cu 2
B l
+-
2 umin
umax du
VA + Bu + Cu 2
+ ~ lumax d(A + Bu + CU 2 ))
2 umin VA + Bu + Cu 2

__ (1 .[
A~arcsm
2
Bu + 2A
UVB2_4AC
] B 1 .
+"2 V_Carcsm
[-2CU - B ]
VB2_4AC

+VA + Bu + CU2rm~)
mIn U

where Umax and umin are obtained by determining where the integrand

VA+ Bu + Cu 2
u
vanishes
I.e.
-B-VB2_4AC
Umax = + (since B > 0; A, C < 0)
min 2C
with the constraint
B2 :2': 4AC I.e. E:2': nq,nw.
This yields:

or
E = (2np + nq,) nw, where np} = 1, 2, 3....
nq,
(1.30)
Wilson-Sommerfeld Quantization Condition 15
By comparison, the standard Schrodinger equation result for this system (see
Chapters 8 and 12) is:

E = (2n +m - 2)hw : } = 1, 2, 3.... (1.31 )

Besides its shortcoming, that it gives quantized energies which are as a rule only
approximately correct for energies large compared to the potentials involved, the
Wilson-Sommerfeld procedure says nothing about the evaluation of probability dis-
tributions, transition rates etc. for which there are standard techniques in Quantum
Mechanics.
On a more positive note the quantization condition Eq. (1.1) is also a conse-
quence of applying the W.K.B. approximation to the Schrodinger equation2 with the
modification that nh must be replaced by (n + {1/2}) h in order to get the W. K. B.
approximation result.

References

1. H. Goldstein, Classical Mechanics, Addison-Wesley (1950), p. 300.

2. E. Merzbacher, Quantum Mechanics, Wiley (1970), p. 123.


CHAPTER 2

The Delta Function, Completeness, and Closure

The delta function is defined to have the following properties (in one dimension):

6(x - x') = 0, x f= x',

i: 6(x - x') dx = 1.
( 2.1)

i:
These have as a consequence that

f(x)6(x - x') dx = f(x' ). (2.2)

One way to get some insight into this useful function and expressions for it in
terms of standard functions is to use 'the principle of completeness'. The principle
of completeness allows one to expand an arbitrary function in terms of any complete
orthonormal set. Thus if 'IjJ( x) is an arbitrary function,

(2.3)
n

if the complete set 4>n (x) chosen is a discrete set, or

(2.4)

i: i:
if the complete set 4>k(X) involves continuous functions, where

an = 4>~(x)'IjJ(x) dx, a(k) = 4>;;(x)'IjJ(x) dx. (2.5)

Expanding the delta function 6(x - x') in terms of a complete set of discrete
functions implies:

6(x - x') = I>n4>n(x)


n
where

i.e. 00

6(x - x') = I: 4>~(x')4>n(x), (2.6)


n=O
The Delta Function, Completeness, and Closure 17
and similarly expanding the delta function in terms of a complete set of continuous
functions implies
(2.7)

EXAMPLE 2.1
Suppose one uses as a complete discrete set the eigenfunctions of a particle in an
infinite square-well potential (a box with infinite walls):
0, -a/2 .:::; x.:::; a/2,
V(x) ={
00, x < -a/2, x > a/2,
for the expansion in Eq. (2.6), with the choice x' = 0. The normalized even subset of
the above eigenfunctions (the rest, i.e. the odd subset is zero at x =:= 0, and does not
contribute to the integral in Eq. (2.5) for an, and hence to the sum in Eq. (2.6) ) is

<Pn(X) = ~a2 cos (2n a+ 1) ",.x,


" n = 0, 1, 2, ... ,
a a
-2'':::; x .:::; 2' '
( 2.8)
a a
x < -2" x> 2"
Hence a possible representation of the delta function is

h'(x) = ~
a n=O
f cos (2n + 1)n:x
a
--2 <
-
a
x <-
- 2 '
a

( 2.9)
a
= 0, x <--
2'

EXAMPLE 2.2
Show Eq. (2.9) is consistent with Eq. (2.1). One notes, interchanging integration
and summation7 that

1 -2 L cos ---7rX
00

-00 a
2n + 1
00

n=O
dx
a
= -
a
2
Ll
00

n=O
a/ 2

-a/2
2n + 1
cos ---7rX dx
a

= :;
4
E
00 (_l)n
2n + 1 = 1,
consistent with the integral in Eq. (2.1).
Considering the first two terms in the expansion of Eq. (2.9) as a crude approxi-
mation one gets the approximate representation:
2 (7rX 37rX) 4 7rX 27rx
h'(x) rv - cos-+cos-- = -cos-cos--, (2.10)
a a a a a a
18 Chapter 2
which is plotted in Figure 2.1.
a!)lx)=6"lc})

x
a

Figure 2.1: Plot of c5(x) '" (4/ a) cos (7rX/ a) cos (27rx / a).

Considering the first four terms of the expansion given in Eq. (2.9) one gets a
somewhat better approximate representation of the delta function:

2 (7rX 37rx 57rx 77rX)


c5(x) '" -
a
cos -
a
+ cos -a
- + cos - - + cos - - ,
a a
( 2.11)
8 7rX 27rX 47rX
= -cos-cos--cos--.
a a a a
This is plotted in Figure 2.2.
One notes that the central maximum (about x = 0) gets progressively sharper
and the secondary maxima less important as the number of terms increases. Also the
normalization of these two expressions (J~oo c5(x) dx) , 8/37r '" 0.85 and 304/105 7r '"
0.92 gets closer to 1 as the number of terms one takes increases.
Eq. (2.9) resembles a Fourier expansion with the constraint the expansion is zero
at x = ±a/2 and only valid for -a/2 < x < a/2.

EXAMPLE 2.3
Consider instead for the complete, discrete set in the expansion of the delta func-
tion, the eigenfunctions of a particle in a harmonic-oscillator potential:
The Delta Function, Completeness, and Closure 19
The eigenfunctions of this potential which are non zero at x = 0 are:

<Pn{x)
1 (1)1/2(
= 2n/2n!
1)1/4 exp (X2)
7rb2
(X)
- 2b2 Hn b ' (n even). (2.12)

as I x I=SI-a-I
8

-----------4-+4-4-~-+~+----------.-L
a

Figure 2.2: Plot of 8( x) rv (8/ a) cos (7rX / a) cos (27rx / a) cos (47rx/ a).

Hence an alternative representation of the delta function is:

(2.13)

where one need not specify the sum is only over even n since Hn(O) is zero for odd n.

The mathematical identity (known as Mehler's formula l )

with the substitution y = 0, X -t x/b becomes:


20 Chapter 2
Substituting this expression in Eq. (2.13) yields:

8() 1 r [1 {X2 (1 + t 2) }]
X = ..jit~ b~exp -2b2(1-t2 ) ,

or in other words, provided one makes the substitution E = l? (1 - t 2 ) ,

(2.15)

Figure 2.3: Plot of 8(x/b) = b 8(x) '" (3/ [2..ji ])exp (-x 2 /2b 2 ) (1- 2x 2 /3b2 ).

The Eq. (2.15) is a standard representation of the delta function in terms of a


limit.
In Figures 2.3 and 2.4 are plotted two approximate expressions for the delta
function using Eq. (2.13). The first involves including the first two non-zero terms
and the second the first three non-zero terms of this expression,

( 2.16)
The Delta Function, Completeness, and Closure 21

8( x) ~ 8~ b exp ( - ;;2) (1 - ~~: + :;; ) ,


respectively, with normalizations (J~008(x) dX) ,../2/2 ~ 0.707, and (7v'2)/8 ~
1.24, respectively. ----/
The general features and trends are similar to those of Figures 2.1, and 2.2.

S(t1:bS( xl

-5 4 5

igure 2.4: Plot of 8(x/b) = b 8(x) ~ (15/ [8 Ji]) exp (_x 2 /2b2) (1 - 4x 2/3b2 + 4x 4 /15b4 ) •
EXAMPLE 2.4
Show that Eq. (2.13) is consistent with Eq. (2.2) for f(x) = exp (_x 2 /2b 2 ).
One notes, interchanging integration and summation, that since Ho(O = 1, and
the H's are orthonormal9 ,

1-0000 (X2)
exp - 2b2 8(x) dx = Ji1 b 1-0000 (X2) E
exp - b2
00 1 (X)
2nn!Hn b Hn(O) dx =

~ b ~ 2:n!Hn(O) f: exp (- ~:)Hn(~) Ho(~) dx = ~ ~ 2:n!Hn(O)8no Ji = 1,


consistent with Eq. (2.2).

EXAMPLE 2.5
Show that Eq. (2.13) is consistent with Eq. (2.1).
22 Chapter 2
One must verify that

-oo,jii b exp - 2b2 E2nn!Hn b Hn(O) dx = 1.


100 8(x) dx = 100
-00
1 (X2) 00 1 (X)
Consider

Interchanging the integration and the summation and defining ~ = x/b in this
expression,

Hn(O) = (-It/2n!/(n/2)! (n even),


while the integral6

With the substitution y2 = e/2, X = y'2, this reduces to

Substitution then yields 8 the desired result:

EXAMPLE 2.6
Use as a continuous representation of the delta function the set of free-particle
wave functions (V(x) = 0, all x),

1/Jk(X) = Aexp (ikx).

This implies, using Eqs. (2.4), (2.5) that:

8(x) = IAI2jOO exp (ikx) dk = IAI2 lim exp (ikx) -: exp (-ikx) = 21AI2 lim sin kx .
-00 k-+oo zx k-+oo X
The Delta. Function, Completeness, a.nd Closure 23
One can obtain, using the property of Eq. (2.1) J~ 6(x) dx = 1,
that:

1 = 2\A\2 hm .1
k--+oo
00

-00
sinkx
- - dx = 2\A\211",
X
(a result independent of k).

Hence
C() = -
uX
1 1.I msin-kx
-, (2.17)
11" k--+oo X
is a second expression for the delta function in terms of a limit. Using the above
value of A, to within an overall phase the set of free-particle wave functions with
'delta function' normalization becomes:

4>k(X) = ~exp(ikx), (2.18)


V 211"
while 6( x) can also be written

8(x) = 1
-2
11"
1 exp(ikx) dk
00

-00
= -1
211"
1 coskx dk 11" 1 coskx dk,
00

-00
= -1
0
00
( 2.19)

since the integral J::"oo sin kx dk = 0 , the integrand here being an odd function, while
the integrand cos kx being even,

1 cos kx dk
00
-00
= 2
Jo
roo cos kx dk .

The three-dimensional delta function 8(r' - fJ) in spherical coordinates can be


written as follows:
6(r - r')
6(r' - iJ) = 2 6( cos 0 - cos 0') 8( 4> - 4>')
r
( 2.20)
L R:1(r)Rn1(r')y;.l(o, 4»Y~(O', 4>'),
n,l,m

J8(r' - iJ)
where
dr= 1,

11
I.e.
roo 8(r -
Jo
r') dr
-1
8( cos 0 - cos 0') d cos 0 r
Jo
21f
6(4)- 4>') d4> = l.

If one multiplies both sides of Eq. (2.20) by ~q(O, 4»~*q(O', 4>'), and integrates
over dO. dO.' one obtains
8(r - r')
~--:-2~
r
= ,,*
L.J Rnq ()
n
r Rnq (')
r ,
24 Chapter 2
and for three-dimensional systems quite generally one can thus write for the delta
function (for any choice of orbital quantum number 1):
00

6(r - r') = E u:l(r')Unl(r), (2.21)


n=O

where unl(r) is the solution, subject to the condition Unl(O) 0, of the radial
Schrodinger equation:

The choice of central potential V (r) determines the detailed form of Unl (r). The
corresponding one-dimensional expression is Eq. (2.6) where x extends over all space,
and ¢n (x) is the solution of the one-dimensional Schrodinger equation:

1i2 cP }
{ -2mdx 2 + V(x) - E ¢n(x) = 0, (-00 < x < 00),

with the choice of V (x) similarly determining the detailed form of ¢n (x).
For continuum states Eq. (2.21) is replaced by

(2.22)

which is analogous to the one dimensional Eq. (2.7).


Eqs. (2.6), (2.7), (2.21), (2.22) are illustrations of the 'closure' property of quantum-
mechanical wave functions.

EXAMPLE 2.7
Obtain an expression for 6(r - r') if V(r) is the three-dimensional harmonic-
oscillator potential:
1i 2r 2/2mb\ r 2:: 0,
V(r) = {
undefined, r < 0.
Inserting the detailed solutions Unl( r) for this potential (d. Eq. (8.13) ) in Eq.
(2.21) one obtains:

( 2.23)
The Delta Function, Completeness, and Closure 25
for each permissible value of 1 (1 = 0, 1, 2... ), and where r, r' > O.
If I = 0 for example

( 2.24)

Substituting Eq. (2.12), the solutions of the one-dimensional harmonic oscillator,


in Eq. (2.6), one has analogously:

(2.25)

H2n+1(e) = (-nIt (2n + 1)!2e [1F1 ( -nj ~j e)] .


Substituting this expression into Eq. (2.24) and using Legendre's duplication
formula 2

n.
'r( n + 3/2) = (2n 2+2nI)!
+1'
-..Ii
The Eq. (2.24) can be written in a form similar to Eq. (2.25),
namely
00
Hn (rib) Hn (r'lb)
(2.26)
n=1, 3,5

The Eq. (2.26) involves a sum only over odd n terms, for which Hn(O is zero if = o. e
Thus 8(r - r') in Eq. (2.26) vanishes if either r or r' is zero. Also, an additional
factor 2 arises in Eq. (2.26) as compared to Eq. (2.24) because x, x' extend from
-00 to +00 whereas r, r' extend from 0 -+ 00.
Using the Hille-Hardy formula3

+ 1)y} /2) ~
( xy)a/2 exP (-2({x + r(n + a:, + 1) 1 F1 (-n., a: + 1., x) 1 F1 (-n·, a: + 1., y) t n
r a:
~
n=O n.

_ r
-l-t
a/2
- - exp ({I2
l+t}) I (2(x yt)1/2)
- -(x + y)--
1-t 1-t a'

one can sum Eq. (2.23). Here

Ia(x) = i-a Ja(ix) = J ~:-1


7r ja-1/2(ix).
26 Chapter 2
Thus Eq. (2.23) can be cast into the form:

_ 1 _. {2(rr/)1/2 r(21+1)/4 (_ [(r2 + r/2) ~]) ( 2rr't 1/ 2 )}


8(r r ) - ~~ b2 1 _ t exp 2b2 1 _ t 11+1/2 b2(1 - t) ,
(2.27)
where (r, r' > 0), which can be written in terms of spherical Bessel functions:

{4rrl t- I/ 2 ([(r2 + r/2) 1 + t]) . (i2rrltl/2)}


~~ i l b3Vi (1 _ t)3/2 exp -
1 •
8(r - r ) = 2b2 1 _ t Jl b2(1 - t) .
(2.28)
If one makes the substitution f = (1 - t)b2 , this becomes:

or

For I = 0, since4

But as f -+ 0,
2rr/) 1 (2rr/)
sinh ( - f - -+ 2exp - f - ,

and Eq. (2.31) reduces to

uC( r-r') = - 1.1m {2 1


-exp (r2+r/2)}eXP(2rrljf)
-
Vi <-+0 ..fi f
/
= - 1.1m -exp {(r-f r )2}}
2
1 {I
Vi ....0 ..fi
(2.32)
an obvious generalization of Eq. (2.15).

EXAMPLE 2.8
Verify Eq. (2.26) is consistent with Eq. (2.2) if f(r) = Hm(rjb)exp( -r2 /2b 2), (modd
According to Eq. (2.2)
The Delta Function, Completeness, and Closure 27
Multiplying Eq. (2.26) by Hm(r /b) exp (-r2 /2l?) (modd), and integrating both
sides over r one obtains

::.) (_r/2)?2exp(-r/2/2b2)",Hn(rl/b) tx) (:) (:) (_r2) (:)


Hm ( b exp 2b2 - .Ji n~d 2nn! 10 Hn b Hm b exp b2 d b '
(2.33)
where integration and summation have been interchanged.
But the integrand in Eq. (2.33) is always even. Hence the r.h.s. of Eq. (2.33) can
be written

exp(-r/2/2b2) '" Hn(rl/b)joo (:) (:) (_r2) (:)


.Ji L..J 2n ' _ Hn b Hm b exp b2 d b
7r nodd n. 00

exp (_r,2 /2b 2) Hn (r' /b) m _ (r'2) (r')


= .Ji
7r
2:
nodd
2n , Dnm 2 m!.Ji - exp - 2b2 Hm -b '
n.
which is identical with the 1.h.s. of this equation.

Since completeness enables one to expand any arbitrary function in terms of a


complete set, it can also be used to describe what happens if the potential of a
system suddenly changes without the wave function undergoing any modification:

EXAMPLE 2.9
Suppose a particle is in the ground state of the potential:

Suddenly the potential changes to

One wishes to find the probability the particle will be in any (say the ground) state of
this new potential. What is involved in this case is a displacement of the equilibrium
position (from x to Xl) and a change of frequency (from Wo to wd of the potential,
as illustrated in Figure 2.5.

The wave function of the system is:

1 )
<Po (X, bo) = ( bo";;
1/2 (X2 )
exp - 2b5 .
28 Chapter 2
The potential changes frequency where:

and is displaced so the complete set of states which now describes the system is
¢n (x - Xl, bd. To find the probability one merely uses Eq. (2.3).

L: an¢n (x -
00

¢o (x, bo) = Xll bl ),


n=O
v(x)

----------~~----------------------~~~----------.- X
Xi

Figure 2.5: Change in potential in Example 2.9.

i:
where

ao = L: ¢o (x - XI. ~) ¢o (x, bo) dx = ~ exp (-[;:6])


X exp (_ [(X ;b~d2])dX = b62b+obbil exp -
([ xi
2 (b6 + bi)
]) ( 2.34)

1.e. laol 2 = b~:b~i exp ( - [b6 ~ bi])


an expression which, as expected, is symmetric in bo, bl . Thus for instance if Xl =0,
laol 2 = 0.8 if bl equals either bo/2 or 2 boetc.
This procedure applies equally well if one changes the form of the potential.
The Delta Function, Completeness, and Closure 29

EXAMPLE 2.10
Consider a particle wave function:

(1/4) J15/b 5 W- x 2 ), Ixl < b,


1jJ(x) = {
0, Ixl > b.

[ This involves the potential Vex) - {


( 1i2/4mb2) [(5x2 - b2)/(X 2 - b2)]
' Ixl < b, 1
(d. Example 13.1) -
00, Ixl > b.
The potential is suddenly changed to

0, Ixl < b,
Vex) = {
00, Ixl > b.
What is the probability the particle is in the ground state of this potential?

The normalized ground state of this new potential is just

fl fl
Vb cos kx = Vb cos 2b '
7rX

where

Hence

I.e.

ao = -8v15
3 - '" 0.9993,
7r
thus laol 2 = 0.9986 '" 1,

as it should be since the wave function forms are very similar (d. Example 13.1,
Figure 13.1), as too their energies:
30 Chapter 2
EXAMPLE 2.11
Consider a particle bound by the potential:

-Ivai c5(x).
The potential suddenly changes to:

0, Ixl < b,
V(x) ={
00, Ixl > b.
Find the probability the particle will be in one of the even-parity states of this
new potential.

Here (cf. Eq. (13.7)

if;o( x ) -- JlVolm
h2 exp (_IValmlxl)
h2 ,

and one wishes to evaluate the the overlap:

an = l b
-b
J'Valm
-,;,z- exp
(IValmlxl) {l
h2 Vb cos
(2n + 1)7rX d
2b x

= 2 (IValm)1/2 [be (_IValmx) (2n+1)1rx d


h2b Jo xP h2 cos 2b x. ( 2.35)

Evaluating this integral one obtains:

an = 2 (IVaJm) 1/2 [ IValm/h2 + (_l}n [ (2~ + 1}1r/2b] exp (-IValmb/h 2) ]


hb [ IValm/1i2] + [ (2n + 1)1r/2b ]2
As b --+ 00, an --+ o.
There is no probability the particle will be in one of the odd-parity states of this
potential, namely:

if;m(x) = Vb . (m1rx)
(l sm -b- .

EXAMPLE 2.12
A particle is originally in the ground state of the well

00, x < 0, x > b,


V(x) = {

0, 0 < x < b.
The Delta Function, Completeness, and Closure 31
Suddenly the wall at x = b is shifted to x = 2b. Find the probability the particle
will be in the ground state of the new potential which results from this shift.
The wave function is originally:

4>o(x) = {
{iib sin (7fX/b), °< x < b,

0, x < 0, x> b,
and one is interested in the overlap:

rb f2 . 7rX {1. 7rX


ao = Jo V"b sm T V"b sm 2b
d
x

Thus

The original energy is


E = n, 2 7r 2 /2mb 2 •
The energy of the new ground state is n,2 7r 2 /8mb 2 while that of the new first-excited
state is n, 2 7r 2 /2mb 2 , which is just the original energy. There is therefore a 32/97r 2
probability the energy will be less than before.
The probability the energy will be unchanged is related to the overlap:

rb f2 . 7rX {1 . 7rX d 1
Jo V"bsmTV"bslllT x = V2.
Thus the probability the energy is unchanged is 1/2. The probability the new energy
is more than the original energy is

1- (
32
97r2
1)
+ 2" .

One can also use complete set expansions in related problems. For instance:

EXAMPLE 2.13
Given a particle with wave function:

J3/2c (I-lxi/c), Ixl < c,


1jJ(x) = {
0, Ixl > c.
32 Chapter 2
It is placed in a well
0, Ixl < c,
Vex) = {
00, Ixl > c.
Find the probability the particle will be in the various states of this well.
As can be easily verified, '¢( x) is properly normalized. However it has a discontin-
uous derivative at x = 0, implying a potential which is not even piecewise continuous
at that point.
The wave function '¢( x) can be expanded in terms of the complete set of eigen-
functions of the well it is now placed in i.e.

All the bn 's are zero since '¢( x) is an even function, whereas the

1 . (
;::;Slll
yC
n + 1) -7I"X
c
,
s

are odd functions. Meanwhile

an = l 1 ( 21) -,¢(x)dx=
c

-c
;::;cos n+-
yC
7I"X
C (n
v'6
+ 1/2)
2
71"2

Thus laol 2 = 0.986 etc.

EXAMPLE 2.14
Verify the sum of the probabilities la n l2 in Example 2.13 is unity.

From Example 2.13:


2 6 96
lanl = (n + 1/2)471"4 = (2n + 1)471"4
But (d. Eq. (17.18) ),
00 1 71"
~ (2n + 1)4 = 96 .
Hence, as expected

The Eq. (4.22) of Chapter 4 lists some other representations of the delta function.
These can be compared to Eqs. (2.9), (2.13), (2.15), (2.17), (2.19), (2.23), (2.25), and
(2.30), which were discussed in this chapter.
The Delta Function, Completeness, and Closure 33
References

1. Erdelyi et al., Higher Transcendental Functions, McGraw-Hill (1953), v. 2, p.


194.

2. Op. cit., v. 1, p. 5.
3. Op. cit., v. 2, p. 189, v. 3, p. 272.

4. Op. cit., v. 2, p. 79.

5. Op. cit., v. 2, p. 193.

6. Op. cit., v. 2, p. 195.

7. 1. B. W. Jolley, Summation of Series, Dover (1961), p. 16.

8. Op. cit., p. 32.


9. E. Merzhacher, Quantum Mechanics, Wiley (1970), p. 60.
CHAPTER 3

Momentum Space

Working in momentum space involves taking the Fourier transform of the eigen-
function 'Ij.;(x, t) of the Schrodinger equation. Thus if:

1
<p(p, t) ==..j'i; Joo exp (ipx)
-00 -T 'Ij.;(x, t) dx, (3.1)

it follows from the delta-function property of Eq. (2.19):

~
271'
Joo exp (iY(X 1i- Xl))
-00
dy = 1:( _
1i u x
')
x,

1
that
1 1i exp (ipX)
T <p(p, t) dp. (3.2)
00
'Ij.;(x, t) = ..j'i; -00

The function <p(p, t) is called the wave function in "momentum space". Assuming 'Ij.;,
and <p are normalizable (i.e. vanish at ±oo so one can integrate by parts and drop
surface terms), it can readily be shown that:

"'()
p", P t = -1-
, ..j'i;
1
00

-00
exp (ipx)
- - -1i o'lj.;(x, t) dX'
1i i ox '

0 <p(p, t) =
-T1i op 1
..j'i; 1 00
-00 exp (ipX)
-T x'lj.;(x, t) dx.
More generally for any operators A(p), B(x) :

A(p)<p(p, t) = vk 1: exp ( - i~X)A (~!) 'Ij.;(x, t) dx;


(3.3)

B (-~ ~) <p(p,t) 1
..j'i; 1 00
-00 exp (ipx)
-Ii: B(x)'Ij.;(x,t)dx.

Similarly
Momentum Space 35

(3.4)

Also
84>(p, t) _ _
8
t
1
- /?L
v2~
1 00

-00
exp
(_ iPx) ~.I'( ) d
t
It
8 'f' X, t x.
t
(3.5)

Thus

[-p2
2m
+ V (h
--;-- + -;--
z 8p
h
z at
8) 8] 4>(p, t)
(3.6)
2 2
00
-t: - 2m
$11-00 exp (iPX)[ h 8x2
8 + Vex) + ih8]
8t '1f;(x, t)dx.

But since the integrand on the r.h.s. of Eq. (3.6) vanishes, '1f;(x, t) being a solution of
the Schrodinger equation, the momentum-space function 4>(p, t) satisfies an analogous
equation:
p2
[2m (h8) h8]
+ V -i 8p + i 8t 4>(p, t) = 0, (3.7)

known as the Schrodinger equation in "momentum space" .


Similarly, expectation values can be equivalently evaluated in momentum space
since, using the complex conjugate of Eq. (3.2) and Eq. (3.4):

L: '1f;*(x, t)n(x)'1f;(x, t) dx

= ~ 1",=00 [P'=OO ¢>*(p', t) exp (_ i~x)dP' n(x)'1f;(x,t) dx = 1t,21"'=00


V 2~ h ",=-00 Jp'=-OO It 2~1t "'=-00

X 1~~~: ¢>*(p', t) exp ( - iP~X) dp' 1:~: exp C~X)n ( -~ ;p) ¢>(p, t)dpdx
(3.8)

1:~:1~~~: ¢>*(p',t)n(-~ ~) ¢>(p,t) [2~ 1:~: exp C(p~p')x) d;]


X dp'dp:;;1 = h 8) ¢>(p, t) Tdp
IP=oo ¢>*(p, t) n (--;--8 .
It p=-oo Z P It
36 Chapter 3
One also notes that if 'IjJ(x, t) is normalized ¢(p, t) is also normalized since, using
Eq. (3.8) with n = 1 :

100 'IjJ*{x, t)'IjJ{x, t)dx = 100


-00 -00 ¢*{p, t)¢{p, t)r;
dp
.

The above formalism generalizes to three dimensions by replacing

n abyn-;-'\7
--a etc.
i x z
Six simple problems follow which illustrate the usefulness of the momentum rep-
resentation and the fact that for certain potentials it is easier to work in momentum,
rather than coordinate space.

EXAMPLE 3.1
Consider the case of a free particle (V{ x) = 0). In momentum space the Schrodinger
equation, (Eq. (3.7) ) for this system is:

(3.9)

I.e.
¢(p) = 8 (~_ v'2;:E) .
vk J:
Since generally
'IjJ{x) = ¢(p) exp C~X)d (~) ,
if one suppresses the time variable in Eq. (3.2), in this case

'IjJ(x) = v'211i" exp (.v'2mE


z n x) '
as it should be.

EXAMPLE 3.2
Consider a particle in the momentum-dependent potential

V{x) = a~~
z ax
= ap.

In momentum space the Schrodinger equation (Eq. (3.7) ) for this potential is:

{:: + ap - E} ¢(p) = 0, (3.10)


Momentum Space 37
i.e.
p - -ma: + Jm 2 a: 2 + 2mE ,
<jJ(p) = A8(t - ~) + B8(t - P;) , { p12 -=
-ma: - Jm 2 o: 2 + 2mE ,

and as above
1/J( x) =
A
y'2; exp
(iPIX) B
T + y'2; (ip2X)
exp -11,- .

EXAMPLE 3.3
Obtain the exact solution of the problem
Ax, x > 0, (A> 0),
V(x) = {
00, x < 0,
(illustrated in Figure 3.1), with the help of the momentum representation.

v (x)

~----------------------~~x

Figure 3.1: Potential in Example 3.3.

In momentum space the Schrodinger equation (Eq. (3.7) ) for this potential is:

{ -p2 - A -;--
11, d }
- E <jJ(p) = 0 (E > 0), (3.11)
2m z dp
1.e.
38 Chapter 3
Integrating this expression yields:

i {p3 } {<p(p)}
Ali 6m - Ep = In C '
I.e.

<p(p)=cexp(~li {:: -EP}) A, E >0. (3.12)

The constant C can be obtained by normalization i.e. requiring

l: <pE(p) <PEI(p) t = li(E - E') = ICI 2 l: exp (i {:li - :~} p) t·


This implies ICI 2 27rA = 1, where one has used Eq. (2.19).
Thus to within a phase

V27rA Ali 6m
(i
- -Ep }) .
<p(p) = -1- exp - {p3 (3.13)

Since generally

'IjJ(x) = v'2i
1 1 <p(p)
00
-00 exp (ipx)
T dr;" (P) (3.14)

in this case

'IjJ(x) = -1-
27rvA
l P=oo
p=-oo
exp ( - i {p3
- - Ep }
Ali 6m
+ -iPX) d (P)
li
- .
li
(3.15)

For this problem the wave function 'IjJ(x) satisfies the boundary condition 'IjJ(0) = 0,
since V(O) = 00. Hence for this example

=
1
27rvA li
[jOO cos [p3/ 6m
-00
- EP] dp + .joo.sm [p3/ 6m
Ali
- E P] dp] .
Ali
2 -00

This implies the even integral,

roo [p3/6m - EP] dp = 0, (3.16)


10 cos Ali
since the integrand of the integral

J OO
-00

sm
[p3/ 6m - EP] dp
Ali
Momentum Space 39
is odd, making the latter integral automatically zero.
Defining u == pj(2mAfi)1/3, Eq. (3.16) becomes

(2mAfi)1/3 [Jo[00 cos (u"3 - E


3
( 2m
A2fi2 )1/3 ) du = -Ii q, ({
U
2m }1/3)]
-E A2fi2 = 0,

where q, ( -E {2mj A2fi2} 1/3) is the Airy function.


The energy eigenvalues E in this problem thus satisfy the condition that:

2m
q, ( -En { A2fi?
}1/3) = o. (3.17)

For reasonably large negative arguments

. (21 1
q, ( X ) -+ Ix11
1/ 4 sm 3" x
3/ 2
+ 4"71") .
Hence in this limit:

n = 1, 2, 3... ,

I.e.

En = (n - ~) 2/3 (A~2) 1/3 (2~) 2/3, n= 1, 2, 3... , (3.18)

which can be compared with

E =n 2 / 3 ( - -
n m
A2fi2) 1/3 (32\1'2'
~
)2/3

the Wilson-Sommerfeld result (Eq. (1.14) ).

EXAMPLE 3.4
Obtain the exact solution for a particle in the potential

Aj x, x > 0 (A < 0),


V(x) = {
00, x < 0,

using the momentum representation. This potential is illustrated in Figure 3.2.


In momentum space the Schrodinger equation for this potential is:

p2 A }
{ 2m + -(fiji) (djdp) - E cp(p) = 0, (E < 0), (3.19)
40 Chapter 3
I.e.
_ (iA/,h - p/m) dp = d¢(p) = _ d (p2/2m + lEI) + (iA/h) dp
E - p2/2m ¢(p) p2/2m + lEI p2/2m + lEI·

V( x)

A------------------------x

Figure 3.2: Potential in Example 3.4.

Integrating this expression yields:

1 { ¢(p) (lEI + p2/ 2m )} __ ilAI J2m (p)


n C - h lEI arctan v2mlEI '
I.e.

where 1¢(p)1 2 vanishes at ±oo and is square normalizable.


Thus C can be determined by normalization i.e.

1 00
-00
1"'( )1 2d (E.)
<f' P h
= 1= IC121°O
-00
dp/h
(lEI + p2/2m)2 .
Hence:
Momentum Space 41
To within a phase one therefore has:

1-0000 ¢(p) (ipx)


Generally
1
1/;(x) = -.,fFi exp T (P) d ~ .

In this case 1/;(0) = 0, since V(O) = 00.


Hence

0-- IEI3
1 (- -
- 7r 2mTi2
)1/4100 1 ([IAI~m
-00 lEI + p2/2m cos - Ti -arctan
lEI
{p}]
J 2m lEI

.. [IAI
-z sm lET arctan {p
TV{2ffi J lEI }]) dp.
2m
The integral

is automatically zero since the integrand is odd. Thus the even integral

roo cos ({IAI/Ti} J2m/IEI arctan {pI J 2m lEI })


Jo lEI + p2/2m dp,
must be zero in order for 1/;(0) to be zero.
Integrating this expression one obtains

Ti. IAI 2m 7r
Ti. AI
TAT sm [ TI ~
lET () "/2] = IAI sm [~]
2m
10 T lET '2 . (3.22)

]
Hence

sin [ 11~7r Jf~ = O.

The energy eigenvalues IEnl of this problem thus satisfy the equation
42 Chapter 3

n = 1, 2, 3... ,

or

(3.23)

With IAI = ahc this reduces to the well-known Bohr result and corresponds to
the energy of the state uno(r) (as well as of the states unl(r) since the energy levels
for hydrogen are independent of I, an effect known as "accidental" degeneracy) of the
hydrogen atom.
Substituting Eq. (3.23) into Eq. (3.21) one obtains a general expression for the
wave function in the momentum representation:

'" ( ) _ (2h,Bn/1r)1/2exp(-2inarctan{,BnP}) nh
'l'n P - 1 + (,Bnp)2 , ,Bn == mlAI ' n = 1,2,3 .... (3.24)
Using this expression one readily obtains:

= I: ¢:(p)p2m+1¢n(P) d (~) = 0, m = 0,1,2, ... ,


m 21AI2
n 2 h2 ,
(3.25)

\/2Pm
2
) = -En.

These two equations allow one to evaluate .6p~, the "variance" of p. In particular

(3.26)

Similarly one can easily obtain2:

(3.27)

h4 2
(X2) = 2" (5n 2 + 1) m~A12 . (3.28)

These results imply that .6x n , the square root of .6x~, the variance of x:

(3.29)
Momentum Space 43
Combining Eqs. (3.26) and (3.29) one obtains:

h h
t1xt1p = 2"v'n 2 + 2 > 2"' (3.30)

consistent with the uncertainty relation (cf. Chapter 4).

EXAMPLE 3.5
Solve exactly the problem of a particle in the potential

V(x) = Alxl, -00 < x < 00.


EXAMPLE 3.6
Solve exactly the problem of a particle in the potential

A
V(x) = !xi ' -00 < x < 00.

The potentials of Examples 3.5 and 3.6 are symmetric potentials i.e. if one plots
V(x) vs. x, the potential for x :::; 0 is the mirror reflection (about the V(x) axis) of
the wells for x > 0 in Examples 3.3 and 3.4 (Figures 3.1 and 3.2). One can proceed
here using methods similar to those used in Examples 3.3 and 3.4. However, instead
one can compare (cf. Eqs. (1.12) and (1.10) ) the solutions of the problem of the
standard harmonic oscillator potential restricted to x > 0 i.e. V (x) = AX2, 0:::;
x < 00, V(x) = 00, x < 0, to the oscillator extending over all space i.e. V(x) =
Ax2, -00 < x < 00. The former potential's allowed energies are

En = (2n - ~) htf., n= 1, 2, 3 ... ,

and correspond to odd-parity solutions (see Eq. (10.12) ), while the latter's allowed
energies are

En = (n - ~) htf. ' n= 1, 2, 3 ... ,

and both odd and even parity solutions are allowed. One can thus obtain the solutions
in the latter case from the solutions in the former by letting n -+ n/2.
Similarly if one considers the infinite well V(x) = 00, x < 0, x > a, V(x) =
0, 0 < x < a, its solutions are the odd-parity wave functions tjJ( x) = )2/ a sin kx
(where ka = mr) i.e.
44 Chapter 3
whereas if one considers the infinite well V(x) = 00, x < -a, x > a, V(x) =
0, -a < x < a, its solutions are tf;{x) = V1/asinkx (odd parity), and tf;{x) =
V1/a cos kx (even parity), where ka = mr or (n -1/2)11", n = 1, 2, 3... , i.e.
2
E = h k 2 = ~ (~)2 11"2,
2 n = 1, 2, 3... ,
2m 2m 2 a
Again the solutions for the latter case can be obtained from the former by letting
n -+ n/2.
A third example which illustrates the fact that this procedure may be applied
generally is the potential weIll

-110 cosh- 2 ax, x> 0,


Vex) = {
00, x < O.

The (odd-parity) solutions for this case are:


(s-2n+1)/2
tf;(X)= ( 1-e ) 2Fl(1-2n,2s-2n+2;s-2n+2;{1-0/2),

with corresponding energies

1 + 8mVo
a h
]2 , n = 1, 2, 3... ,
2 2

while the symmetric potential

V(x) = -Vo cosh- 2 ax, (all x),

has solutions
(s-n+1)/2
tf;(x)= ( 1-e ) 2 Fl(1-n,2s-n+2js-n+2j{1-0/2),
where
1 8mVo n = 1, 2, 3 ... ,
+ a2 h2

(with the symbols s, and edefined as


s -2
= ~ (-1 + 8mVo )
1 + a2 h2 , e== tanh ax ).

Thus in this case as well one obtains the solutions for the symmetric well by letting
n -+ n/2 in the solutions for the well which extends only from 0 < x < 00.
Momentum Space 45
The solutions for Example 3 are

En = (n - D2/3 (A~2r/3 (2~r/3


= (2n - ~r/3 (A~2) 1/3 (4~) 2/3, n= 1, 2, 3....
These correspond only to the odd-parity solutions of Example 3.5. By analogy with
the above three problems one thus expects that the odd- and even-parity solutions
for Example 3.5 have energies:

(3.31 )

an expression which for large n, n rv (n - 1/2) agrees with the Wilson-Sommerfeld


result (Eq. (1.13) ).
Similarly the solution for Example 3.4 is

and one expects that the energy in Example 3.6 is just

n = 1, 2, 3... , (3.32)

and that this includes both even- and odd-parity solutions. This result agrees pre-
cisely with the Wilson-Sommerfeld result (Eq. (1.15) ).

EXAMPLE 3.7
Obtain the ground-state wave function in coordinate space for the Hamiltonian
in Example 3.4 starting with the momentum-space wave function Eq. (3.24).
From Eq. (3.24) one notes that the ground-state function (n = 1) is:

'" ( ) =
'l'IP
(2h (.l )

7r1-'1
1/2 exp (-2i arctan {flIP})
1 + (fllp)2 , (3.33)

Using Eq. (3.2) and suppressing the time variable:

~1 (x) 1 Ii
= v'2i 100
-00 ¢1(P) exp (ipx)
T d (P)
1i,
46 Chapter 3
Substituting Eq. (3.33) in this equation

1/J ( ) - (2/i
1 X -
(3 )1/2
-:; 1
1
V2i n
1 00

-00
exp (ipx//i - 2 arctan {(31P}) d
+
1 (filP)2 p.
With the change of variable u == arctan /3IP this becomes

1/Jl (x) = ~ ((3~/i f/2 1:/: exp {i (/i~1 tan u - 2U) } du .


Writing the exponential in terms of sines and cosines, only the cosine term survives
because of parity considerations when one integrates from -1r /2 to 1r /2, i. e.

2 ( 1 )
1/Jl(X) = -; (31/i
1/2 1:/2/2 cos (/i~1 tan u - 2u) du.
Evaluating this expression3 one obtains finally:

1/Jl(X) = 2 ((3~/ir/2 xexp (-/i~J, (3.34)

a wave function which one can easily show to be normalized.


With this expression one can for example calculate:

(x)
= (1)3 10roo exp (2X)
4 (31/i
3 3(31/i 3/i2
- /i(31 x dx = -2- = 2mlAI '

(X2) = 4 (f3~/ir 1 exp (-


00
:;J X4 dx = 3 {f31/i}2 = m~~~12 '
results consistent with Eqs. (3.27), (3.28) (in the case n = 1).

References

1. 1. Landau et I. Lifchitz, Mechanique Quantique, Mir (1966), p. 94.

2. I. S. Gradshteyn and I. M. Ryzhik, Tables of Integrals, Series and Products,


Fourth (English) Edition, prepared by A. Jeffrey, Academic Press, New York
(1980), p. 292.
3. Op. cit. p. 405.
CHAPTER 4

Wavepackets and the Uncertainty Principle

The properly normalized free-particle wave function is:

'ljJp(x, t) = 1 exp {.Z (PX


J2i Et)}
h - t; (4.1)

(cf. Eq. (2.18) ).


One problem with this function is that it has no spatial localization, i.e. though
the momentum is precisely known (in other words tlp = 0), tlx = 00, where tlA
is rigorously the square root of the variance of A (see Eq. (3.26) ), or more loosely
speaking represents the uncertainty in A.
This shortcoming may be easily removed by constructing a wavepacket

where A(p') is a function concentrated about p' = p. Thus if A(p') is a delta function,
(cf. Eqs. (2.1), (2.2) ):

A(p') = 8 (pI ; p)

Eq. (4.2) reduces to Eq. (4.1). The function

A( p') exp {iE(pl)t


---'h- }

is in fact the wave function in momentum space (d. Eq. (3.2) ).


Any detailed functional form A(p') gives the same general properties to w(x, t),
namely localizes the particle, (i.e. ensures that tlx =J 00).

EXAMPLE 4.1

.fi
Suppose

A(p') = exp ( _Ip' ~ pI) , (4.3)


48 Chapter 4
a normalized function peaked about rI = p, and illustrated in Figure 4.1.
At first sight it appears this gives an uncertainty in momentum

The gain however is that W(x, t) is now localized. Thus

W(x,O) = ~
28
7f
lp'_oo
pl =oo (Ip' - pi +T
exp-
iP'X)
8 (d{) ,
or

W(x,O) = ft.
837f exp Ct:PX ) x + h /8
2
1
2 2 •
(4.4)

A ( pi) = J~ e -I p' - p I/6'

____~--~~--~-+--~~~~------~~pl
p-26' p-6' p p+6' p+26'

Figure 4.1: The momentum distribution in Example 4.1.

[ The normalization of W(x,O) can be confirmed by noting roo dx 2 = ~ .]


Loo (x 2 + 1) 2

The amplitude of the function given by Eq. (4.4) (ignoring the phase) is drawn
in Figure 4.2. There is a spread about x = 0 given approximately
by tJ..x = 2h/8. Thus tJ..x tJ..p '" 2h/8 28 = 4h.
A more accurate estimate of tJ..x tJ..p is possible since
Wavepackets and the Uncertainty Principle 49

A02 = (02) _ (0}2 h {O} = J \lI*O\ll dr = J A*(p)OA(p)dp


~ , were - J \lI*\lI dr J A*(p )A(p )dp .

Using the function of Eq. (4.4), {x} = 0, while in this case:

21i31x=oo x 2 dx 1i 2
.6.x2={x2)= 63 7r "'=-00 (x2+n2j6 2)2 = 62 •

Figure 4.2: Amplitude of \lI(x,O) in Example 4.1.

Similarly
( )=
P
!!.l
P '=oo,
fJ p'=-oo P exp
(_ 21p' -
fJ
pI) (dP')
fi
=
P,
while

62 .
<P2) = P2+"'2 I.e. D..p =
6
V2 '
and
fi 6 fi
D..x D..p = -.M = M rv D.71fi, (4.5)
6y 2 y2
in this case. Thus the uncertainty in the position multiplied by the uncertainty in the
momentum of a localized wavepacket is of the order of fi. Other A(p)'s yield similar
results.
50 Chapter 4
EXAMPLE 4.2
Consider the case

J31i/28 3 [8 - Ip' - pI], Ip' - pi < 8,


A(p') = (4.6)
0, Ip' - pi> 8,
a normalized function peaked at p' = p. This A(p') is illustrated in Figure 4.3.
At first sight it appears this gives an uncertainty /1p '" 8. A more accurate estimate
can be carried out by evaluating

(p 2) = ~ lP'=P+li(8 _ I ' _ 1)2,2


283 p'=p-li P P P
(d1iP') = P2 + 1082 .
Thus
8
/1p = v'IO .
In this case

1 f3i"l P'=P+li,
W(x,O) = v'2-iV W p'=p-li (8 -Ip - pl)exp T (dT·
(iP'X) P')

A (pi>

A (pi) = 0 Ipl_pl >s


= J ~~3 (6-l pl_pl > Ipl_pl<6

r-____~------+------+--------------~~pl
p-s p

Figure 4.3: The momentum distribution in Example 4.2.


Wavepackets and tbe Uncertainty Principle 51

Evaluating this integral one obtains:

\IT(x,O) = -2
1 ~8
~t exp ~
(iPX) sin 2 (8x/21i) (4.7)
"It /l (8x/21i)2

[The normalization of this particular \IT(x,O) may be easily verified using the result

1o
00 sin4 u 7r]
- - du=-.
u4 3
The amplitude of the function of Eq. (4.7) (ignoring the phase) is drawn in Figure
4.4.
There appears to be a spread in x of approximately 21i7r / 8, giving an approximate
6.x 6.p ,...., (21i7r / 8) X 8 = h .

..LJ36
2 lTh

2liTf 2li1l"
-5 -8-

Figure 4.4: Amplitude of the wave function 1lJ(x, 0) in Example 4.2.

For a more accurate 6.x one evaluates:

(x) ~ ~ ~ JOO sin 4 (8x/21i) x dx - 0


4 7r 1i -00 (8x/21i)4 - ,

~ ~ ~ JOO sin4 (8x/241i) x2 dx = 61i 2 joo sin 4 u du


4 7r 1i -00 (8x /21i) 7r8 2 -00 u2
52 Cbapter 4

6tl,2 [jOO sin 2 u _ ~ JOO sin 2 2u ] _ 3n 2


= "2
7ru -00 U
2 du 4 -00 U
2 du - u
"2 .

Hence,

v'3 n v'3 n 8 [3
~x = -8-' and ~x~p = -8-JTIj = ViO n = 0.55n,

in this case, a somewhat smaller result than that obtained in Example 4.1.

[ The integral 1 sin:~du i


00
= is needed to obtain the above result. ]

Obviously W(x, t) of Eq. (4.2) also satisfies the free-particle Schrodinger equation
just as Eq. (4.1) does, namely:
n2 f)2 n f)
- - II 2 W(x, t) = --;- ll.I. W(x, t), (4.8)
2m uX Z UL

independent of the detailed form of A(p') since

1
v'2i 1 A(p') (p,2
00
-00
dp' = 0,
2m - E(p') ) exp { Ii [p'x - E (p') tl } T

if
p2
2m = E(p).
Further, independent of the form of A(p), since it is the wave function in momen-
tum space, one can write:

A(p') = 1 JOO W(x', 0) exp (-T


v'2i -00
iP'X') dx'. (4.9)

Substituting Eq. (4.9) into Eq. (4.2), and dropping the prime for the p's one
obtains:

W(x, t) = 1:&'=00
:&'=-00
W(x', O)~
27r
l P=oo
p=-oo
exp (i [Cx - x')p _ E(P)t]) dnPdx'.
n n (4.10)

Defining

G(x, x', t) ~1°O exp(i[(x-X')P_ECP)t]) dp


- 27r -00 n n Ii
C4.11)

1 )..*( ')).. C)
-00
00
'Pp x 'Pp x exp
{_iECP)t}dP
n n'
Wavepackets and the Uncertainty Principle 53
where the </>p(x )'s are the properly normalized free-particle wave functions (cf. Eq.

i:
(2.18) ), one can write Eq. (4.10) as follows:

1I1(x,t) = 1I1(x',0)G(x,x',t)dx'. (4.12)

The function G(x, x', t) of Eq. (4.11), known as the free-particle Green's function,
can be evaluated explicitly by integrating Eq. (4.11):

,
G(x, x, t) =
(m
27riht
)1/2 exp (m(x
-
_ X')2)
2iht . (4.13)

This function also satisfies the Schrodinger equation Eq. (4.8) and for t ----+ °
limG(x,x',t) ----+ 6(x - x'),
t-+O

(Eq. (4.13) having in this case the delta function form of Eq. (2.15» as it must
since for t = 0,
111 (x, 0) = JW(x',O)G(x,x',O) dx'.
A useful formula in calculating reflection and transmission times for wavepackets,
(which is a result independent of the details of A(p') but assumes it is peaked about
p' = p) involves expanding E(p') about p = p' in Eq. (4.2)

, p,2 p2
E(p) = 2m = 2m
dEep')
+ ~p'
I '
(p - p) + ....
p'=p

Keeping only linear terms (i.e. assuming only values about p' = p are important)
yields:

1I1(x, t) = _1_ fP'=oo


Jp~'=-oo h 2m d p ,p =p dp',p =p h
I
A(p') exp {~ (p' x _ L t + dE(~') tp _ dEep') t p')} dp' I
V'2(ff
£,7f

= _1 exp
~
{_~
h
(Lt _
2m
dEep')
dp'
I
p'=p
t p)}

X 1~:~: A(p') exp {*p' (x - d~(~')


p p
I p'=p
t)} d{ .
Thus
W(x, t) = exp (i</» 111 (x - d~(~') I
P p'=p
t, 0) . (4.14)
54 Chapter 4
The Eq. (4.14), which is independent of the details of A(p') and is a good approx-
imation if A(p') is peaked about p' = p, implies the wavepacket \J!(x, t) moves with

a speed:
dE(p')
-d-'-
I .
p,2
. Smce E(p ) = 2m'
dE (p'), p'
= m == Vg . -a;p-
p p'=p
This is known as the group velocity of the wavepacket and is just the classical
speed of a free particle.

EXAMPLE 4.3

J21ro3fi
Given
A( ') = sin 2 {(p' - p)jo} (4.15)
P {(p' _ p)jO}2 '
find \J!(x, 0).
One can proceed as in the previous examples or note that since, according to Eq.
(4.2) (where if one uses k rather than p, the equations look more symmetric)

\J!(x, 0) = .~
v21r
1 A(k')
00

-00
exp (ik'x)dk', (fik = p),

if

\J!(x, 0) = exp (ikx)¢>(x)

¢>(x)= 1
tn=
v21r
1 A(k')exp{i(k'-k)x}dk'
-00
00
(4.16)

where according to Eq. (4.9)

A(k') = .~
v21r
1 00

-00
\J!(x', 0) exp (-ik'x')dx' = ~
v21r
1 00

-00
¢>(x') exp {-ix'(k' - k)}dx'.

Hence

A(k + k') =
v21r
1
tn= 1 ¢>(x')
00

-00
exp (-ix'k')dx' = 1
tn=
v21r
1 ¢>(x')
00

-00
exp (ix'k')dx',

if ¢>(x) = ¢>(-x), or

A(k'+k)= .~1°O ¢>(x'-x)exp{i(x'-x)k'}dx'. (4.17)


v21r -00

Comparing Eqs. (4.16) and (4.17) one notes that if one identifies A(k') with a
particular (even) function ¢>( x' - x) one can then identify ¢>( x) with the corresponding
A(k + k').
Wavepackets and the Uncertainty Principle 55
In Example 4.2 for instance,

4>(x) =~ [38 [sin (8X/21i)]


2Y;r;, 8x/21i
2
and A(k+k') = J23 ! [8-Jk'J1i].
v

Hence if one identifies A( k') with

~ [38{sin{(k'- k)8/21i}}2
2y;r;, (k' - k)8/21i '
one obtains the corresponding

directly from A(k' + k).


Letting 8 __ 2Ti2 / 8, if

A(p') =J 3Ti {sin {(p' - P)/8}}2 ,


27r8 (p' - p)/8
then
\If(x 0) = exp (ipx)
, Ti
~4YT
{3i [2 _ ~]
Ti· (4.18)

EXAMPLE 4.4
Consider the case
A(')
p
e3
= J2---;-Ti 1
(p' --p""-')2"'-+-e-=-2 . (4.19)

For this distribution of momenta:

On the other hand the wave function at time t = 0, i.e.

\If(x, 0) = ~JeTi
7r
1 00

-00
exp (ip'x/Ti) dp'
(p'-p)2+e 2 Ti

Je Ti
exp
(iPX) .!.
Ti 7r
1 exp(p'{i(p'- p)2- p)x/Ti}
00

-00 +e
d( , _ )/Ti
p p 2

2.Je 3 Ti ex (ipx) roo cos (p' - p)x/Ti d( , _ )/Ti


7rTi 2 P Ti Jo ({p'_p}/Ti)2+(UTi)2 P P .
56 Chapter 4
. Hence

2 W (ipx) roo cosux du ([ (iPX) (elxl)


W(x,O) = ;Ytl exp T Jo u2 + (e/n)2 = Y"h exp T exp -r;- .
(4.20)
For this wave function

(x) = 0,

I.e.
n
~x~p= V2'
The form of these distributions is given in Figures 4.1 and 4.2 with the appropriate
identifications. The symmetry between the A(p') and W(x, 0) in this example and the
W(x,O) and A(p') in Example 4.1 can be understood in the light of the remarks in
Example 4.3.
What is of some interest is that one can extract a representation of the delta
function from this example, namely:

6(u) = .!.lim~ 1 . (4.21 )


'K ...... Ot (u/t)2+ 1
Thus if one defines

one notes that as


e --+ 0, A'(p') --+ 6 (p, ; p) .

This follows from the observation that when W(x,O) is evaluated using A'(p) rather
than A(p), then
W(x 0) = _1 exp (iPx) exp
, y"h n (_~)
n
and as e --+ 0 this expression for W(x, 0) becomes just a plane wave

1 (ipx)
y"hexp T .
This simplification occurs whenever a delta function 6( {p' - p} In) is substituted for
A(p') in Eq. (4.2), since then W(x, t) in this equation reverts to a plane-wave expres-
sion. Hence the A(p') above is a particular delta-function representation in the limit
Wavepackets and the Uncertainty Principle 57
e - t O. The procedure thus involves essentially determining the right overall factor

by which to multiply A(p') such that in some limit the resulting w(x, 0) reduces to a
plane wave. The A(p') multiplied by this factor is then a delta function in that limit.
By a similar analysis one can extract several other delta-function representations
from the A's used in this Chapter's examples. Some of these are listed in Eq. (4.22):

(i) 8(u) = (l/2)limf..... o (l/f) exp (-lui/f) (Ex. 4.1)

(ii) 8(u) = (1/4)limf..... o (l/f)(l + \ul/f)exp(-Iul/f) (Ex. 4.6)

(iii) 8(u) = (1/'rr) limf-+ OO f [sin (W)/W]2 (Ex. 4.3)

(iv) 8( u) = (4/37r) lim. ..... oo f [sin (w)/ W]3 (Ex. 4.5)

(4.22)
(vi) 8(u) = lim.-+o (1 -Iul/f)/f, lui < f (Ex. 4.2)

(vii) 8(u) = (3/2) limf-+o (l/f) (1 -lul/f)2, lui < f

(viii) 8(u) = (2/7r)limf.....o (l/f) (U 2/f2 + 1)-2 (Ex. 4.6)

= (1/8) limf.....o (l/f) (3 - u 2 / (2) , 0 < lul/f < 1

(ix) 8(u) = (1/16) lim......o (l/f) (3 -lul/f)2, 1 < lul/f < 3 (Ex. 4.5)

=0, lul/f> 3

(Ex. 4.11)

(xi) 8(u) = (2/7r)lim f ..... oo f exp (-w)[sin (w)/w] (Ex. 4.10)


58 Chapter 4
EXAMPLE 4.5
Consider
A(p') = V201i [sin (p - p') / 8] 3
(4.23)
Ibr5 (P' - p)/5
Obtain the corresponding W(x,O) by a straightforward calculation.

W(x,O) = _1 J 201i
.j2; 1l1r8
100

-00
(sin(pl- P)/8)3 exp (ipIX)d (pI) .
(pl-p)/8 Ii Ii

pex ,0)

--3~~~----~~---O+----h~--~--~~~------~X
-T -0 ~ T

Figure 4.5: Plot of W(x,O) in Example 4.5.

°< Ixl < n/8,


W(x,O) = J58/881i exp(ipx/Ii) (1/2) [3 - 8Ixl/Ii]2, n/8 < Ixl < 3n/8,

0, Ixl > 3n/8.

Hence1 :

W(x,O) = J~~~exPC~X)[~(3_8~~2)],
Wavepackets and the Uncertainty Principle 59
(4.24)

Expression (4.24) is plotted in Figure 4.5.


Besides localizing \fI(x, t) the A of Eq. (4.23) is proportional to another represen-
tation of the delta function. In particular

u!:()
x = -311"4 l'1m
<-+00
E
(Sin
--
W)3
lOU
(4.25)

Similarly one may obtain representations of the delta function which involve higher
powers of sin (w) / w,
e.g.

limE (SinW)4
-3
--
211" .-+00 lOU '

etc. with the general expression for 1/ > 0 being:

1/ (-l)k(1/ - 2k)V-1
N(I/)-l = 2,,-1 ""'
L.,
O<5:.k<,,/2 k!(1/ - k)!

a result which reduces to Eq. (2.17) for 1/ = 1, and Eq. (4.22) (iii) - (iv) for
1/ = 2, 3, 4.

In turn the \fI(x, 0) of Example 4.5 is related to the delta function:

= "81 .
}!..~8 3[( 832 -x 2)] , 0<
1
Ixl < 7i '

8(x) =
1
16
}~~ 83[(~ _ Ixl) 2] , 7i1 < Ixl < 7i3 ' (4.26)

3
= 0, x> 7i .
Higher powers of other delta function expressions, for instance of

8(u) - ~ lim ~ 1 and 8(u) = lim 1 -Iul/E , (lui < E),


- 11" <-+0 E (u / ff +1' <-+0 E
60 Chapter 4
can also be expressed as delta functions.
This is illustrated in Example 4.6 for the former of these expressions.

EXAMPLE 4.6
Consider:

A(p') =
hf(2v + 1) ((pi ~ P)2 + 1) -1'-1/2
(4.27)
Eft f(2v + 1/2)
a normalized 2 function peaked at p' = p.
Find the corresponding '1f (x, 0).
From standard tables 3 one obtains:

'1f(x, 0) = 2f(2v + 1) E ( Ixlc)V K (Elxl) (ZPX)


ft f(2v + 1/2)1i {f(v + 1/2)}2 21i v T exp T '
(4.28)
which is proportional to a delta function if E -) (X).

If v = 1/2 this reduces to Example 4.4.

If v = 3/2,

A(p') ~ J~~~ [(" ~ p)' + r, (4.29)

fs If;
and
'1f(x, 0) = {liE + I} exp (_ f~l) exp
ZPX) .
(T (4.30)

If one replaces V16h157rf by 21iI7rf in Eq. (4.29),

'1f(x, 0) -) {1{ TIxlf + 1} exp (flxl)


V2; -T exp
ZPX)
(T '
°
and as f -)

'1f(x,O) -) a exp C~X),


implying Eq. (4.22) (viii) is a delta-function representation. Similar arguments
starting from Eq. (4.30) lead to the corresponding delta function Eq. (4.22) (ii).
Generally:

8(u) -
- \hTT(v + 1/2)
1 lim
<-0 f
~ (M) K (M) (v>
2E
v
l/ f '
1) ,
--2

is a generalization of Eq. (4.22) (i) - (ii).


Wavepackets and the Uncertainty Principle 61
Also,

is a generalization of Eq. (4.22) (vi), (vii).


And,

8() r(v + 1/2) r [1 1 ]


u = .J1rf(v) f~ ; (u 2 /t 2 +1),,+1/2 '
is a generalization of Eq. (4.22), (viii).

EXAMPLE 4.7
Find
\l1(x, t) if
A(p') = 1% exp ( _Ip' ~ pI),
and assuming the linear approximation (previously mentioned in this Chapter, d.
Eq. (4.14) ):

, p,2 p2 dE(p') I ' dE(p') I '


E(p) = 2m = 2m + -d-'- (p - p) + ... ""' E(p) + -d-'- (p - p). (4.31)
p p'=p p ,
p=p

Substituting Eq. (4.31) into Eq. (4.2) (and defining u == p' - p) :

'T'(
'¥ x t
,
)-
-
Ml°O
-
27r8 -00
exp {IP' -
8
pi + -p
Ii
iE(p)t
i ,x - - --
Ii
i dE(p')
-
Ii
--
dp'
I ('p - P)}
p'=p
t (dliP').

J2:8n exp [~{px - E(p)t}] [1 00


exp ( -u {~ - i; + ~Vgt}) du
+ 1 exp ( -u {~+ i; - ~Vgt}) dU]
00

\l1(x,t) J 7r!1i exp[~ {px - E(p)t}] 1°O exp (-~) cos {* (x - vgt)} (4.32)

X du= ~
C3li3 (
7ru
1
(x - Vgt) + Ii
2 2/
82
[in
) exp t{px-E(p)t}. ]
62 Chapter 4
If t = 0 this reduces to Eq. (4.4) of Example 4.1. Eq. (4.32) shows that to the
extent the linear approximation Eq. (4.31) is valid, the wavepacket moves forward at
a speed Vg , but its form does not change, i.e. there is no spreading. To get spreading
one must keep at least quadratic terms in the expansion for E(p').

EXAMPLE 4.8
Show that
alGI 2+ ~~ {G*aG _ aG* G} = 0 (4.33)
at ax 2mi ax ax '
where G(x,x',t) is the Green's function for a particular system (i.e. relates W(x,t)
to W(x,O) according to Eq. (4.12) ).
Since G(x, x', t) satisfies the time-dependent Schrodinger equation (possibly with
a potential Vex), assumed real),

a
1i2 2 ] 1i
[- - - - + V(x) G = --:--G,
a
2m ax 2 z at

and
[ -~~
2m ax
+ V(X)] G* = ~~G*.
2 z at
Premultiplying the first of these equations by G* and the second by G one obtains,
after subtracting that

Hence

~~
ax 2mz
{G*~G
ax
- G~G*} + ~IGI2 = o.
ax at
An equation identical to Eq. (4.33) is obviously satisfied when G is replaced by
W, and G* by W*. This latter equation is known as the Continuity equation for the
current density

and the probability density

namely
(4.34)
Wavepackets and tbe Uncertainty Principle 63

I:
EXAMPLE 4.9
Show that
G*(x,x',t)G(x,x",t) dx = 8(x' - x"), (4.35)

where
G(x, x', t) = ; , <p~(X')<Pn(X) exp ( -i E;t), (4.36)

(cf. Eq. (7.3) ). (This expression for G is the counterpart of Eq. (4.11) if the basis
one is using is discrete rather than continuous.)

I:
Hence show additionally that

8(x - x') 8(x - x") dx = 8(x' - x"). (4.37)

Substituting Eq. (4.36) into the integral on the 1.h.s. of Eq. (4.35) and rearranging
the order of the summations and integration one obtains:

'"
n,n'
x exp (iEnt)
~ <Pn (') -r,,- <Pnl x exp (iEnd)
* (") --r,,- _ <Pn*()
00
1
x <Pnl ()
x dx
00

= '" X exp (iEnt)


).. (')
~ o/n -r,,- <Pnl
* (") -T 8nnl
X exp (End)
n,n'

= L <p*(x")<Pn(x') = 8(x' - X").


n

where one has assumed that the <p's satisfy the standard orthonormality and closure
conditions, (cf. Eq. (2.6) ).
Eq. (4.37) can be obtained by considering the case t goes to zero in Eq. (4.35)
since then G(x, x', t) and G*(x, x', t) go to 8(x -x'), while G(x, x", t) goes to 8(x -x").
The same result, Eq. (4.35) also follows if

( ') 1 <Pk*(')X cPk ()x


G x, x, t =
00
-00 exp (iE(k)t)
--r,,- dk, (4.38)

(d. Eq. (7.3) ), since then Eq. (4.35) becomes

1:~: 1:~: cPk{X)cPk(X') exp {iE~k)t}


x 1~~~: cPdX)cPkl(xll)exp{ iE~k')t} dk' dk dx

= 1:~: cPk(X') exp { iE~k)t} dk 1~~~: cPkl(X") exp { iE~k')t } dk' (4.39)
64 Chapter 4

smce i:~: ¢>Hx)¢>d x ) dx = 8(k- k'),

while 1:~: ¢>'k(X')¢>k(X) dk = 8(x-x'), (d. Eq. (2.7) ).

EXAMPLE 4.10
Consider the momentum distribution

A'(') 20:" (
p = -exp -0:"
Ip' - pI) sinO:"(p' - p)/n
7r n (Y(p' - p)/n .

Obtain the corresponding \Ii (x, 0).

Substituting3 in Eq. (4.2) with t = 0,

\Ii (x, 0) -1-


v'2i
100

-00
20:"
-exp
7r
( -0:" Ip' - pI) sin O:"(p' - p) /n exp -:t:P'x
n O:"(p' - p) /n n
(i) (dnP')
= ~ exp C~X) ~ arctan C;2).
As 0:" -+ 00 one notes that \Ii(x, 0) -+ a plane wave.
Hence

· A'(')
11m Ip' - pI) sin (O:"(p' - p)/n) =
p = 1·1m -20:" exp ( -0:"--- u$: (p'
---p) . ( 4.40)
"'-->00 "'-->00 7r n O:"(p' - p) /fi fi

The functions A'(p') and \Ii (x, 0) (ignoring the phase of the latter) are illustrated
in Figures 4.6, 4.7. One notes that
fi7r
f),p rv - , while f),x rv 20:"-J2, I.e. f),p f),x rv 2-J2 fi7r = -J2 h.
0:"

EXAMPLE 4.11
Consider the momentum distribution

A '(')
p = -0:" arctan {21i2}
2( )2.
27r 0:" p' - P

Obtain the corresponding \Ii(x, 0).


Substituting the above A'(p') in Eq. (4.2) with t = 0, in this case yields 4

1 (ipx) exp(-x/O:") . (/ )
\Ii ( x,O ) = v'2i exp 1i x/O:" sm x 0:" . (4.41 )
Wavepackets and the Uncertainty Principle 65
From this latter equation it is clear that:

. a
hm -211" arctan
{2fI,2}
2( )2 = b
(pI _ p)
-·t- . (4.42)
01---+00 a p' - p It

ACp')

_ _ _- L_________~b-~~----L---L---L---~~~L------ p'
p _1'fn p _ 11" p + 11_"
~ 20c ~

Figure 4.6: The momentum distribution A'(p') in Example 4.10.

.---~ I
~ .. X

Figure 4.7: The resulting IW(x, 0)1 in Example 4.10.


66 Chapter 4
TABLE 4.1.
The A(P')'S and the corresponding W(x, o)'s discussed in Chapter 4. The A's and
W's are normalized according to:

A(pl) : (I~oo A(p')2 (~) = 1) W(x,O) : (I~oo Iv;(x, oW dx = 1) Exam

If exp (_Ip' - p1/8) 1* (x 2+fi2/t52fl exp(ipx/fi) (Ex.

~ [8 - Ip' - pI] , IP' - pi < 8 tJ!{ (8x /2fi)-2sin 2(8x /21i)exp( ipx /Ii) (Ex.
0, Ip' - pi> 8

I!£ {(p' - p)/8r 2sin2 {(p' - p)/8} tl¥ [2 -11] exp (ipx/fi) , Ixl::; 2fi/8 (Ex.
0, Ixl > 2fi/8

J2S;1i. {(p' _ p)2 + 82 r 1


If exp(-8Ixl/fi) exp(ipx/fi) (Ex.

vm°<
if
[t(3-8 2x 2/fi 2)] exp(ipx/fi),
lxi/ii < 1/8

J121°!;S [(p'- p)/8r3 sin3 {(p- p')/8} ..ffJ [i (3 - l]


8I x l/ fi exp (ipx/fi) , (Ex.
if 1/8 < lxI/ii < 3/8

0, if lxi/ii> 3/8

1i.r(2v+l) [(pl_ )2/82+1rv-1/2 2Q2v+l) S (Ex .


r(2v+1/2)
liV7r p , ..;:; r(2v+1/2)1i.{r(v+1/2)}2 X
(v> -1/2) (lxI8/2fi)" Kv (8Ixl/fi) exp (ipx/fi)

J¥A [{(pI - p)/8}2 + 1]:2 If If {lxI8/fi + I} exp (-8Ixl/li)x (Ex.


(v = 3/2) exp (ipx/fi)
Wavepackets and the Uncertainty Principle 67
References

1. Erdelyi et al., Tables of Integral Transforms, McGraw-Hill (1954), v. 1, p. 19.

2. 1. S. Gradshteyn and 1. M. Ryzhik, Tables of Integrals, Series and Products,


Fourth (English) Edition, prepared by A. Jeffrey, Academic Press, New York
(1980), p. 295.

3. Erdelyi et al., Tables of Integral Transforms, McGraw-Hill (1954), v. 1, p. 11.

4. Cp. cit., p. 29.


CHAPTER 5

Uncertainty Principle

and
Ground-State Energies of Quantum-Mechanical Systems

Consider a particle moving subject to a potential

V{x) = Alxl n , -00 < x < 00.


Classically the particle's energy is:

(5.1)
But tlxtlp rv n (cf. Chapter 4), i.e.

n
tlp rv tlx

and one expects p2 2:: {tlp} 2 = {1i, / tlx }2 since generally p is expected to be at least
of the order of tlp.
Hence one can write for the ground-state energy E :

Requiring that tlx is of the order of 2x one obtains

E 1i,2 A (tlX)n
= 2m {tlx }2 + 2 '
where tlx is assumed greater than zero.
The choice of !:::.x which minimizes E is such that:

8E. 1i,2 A {A }n-l


8 {!:::.x} = 0 I.e. 0 = - m {!:::.x } 3 + n 2n uX ,

implying
{tlxr+2= 1i,22An.
nm
Uncertainty Principle and Ground-State Energies ... 69
For this choice of .6.x

I.e.

n +2 { 'li 2 A2/n }n/(n+2)


E>
- 2(3n+2)/(n+2) - nm- (n =I- -2). (5.2)

EXAMPLE 5.1
Consider the case n = 2, -00 <x< 00 .

E = ~ {'li2A}1/2 = _1 {'li2A}1/2
222m J2 m
The exact ground-state energy of this system is also

_1 {'li2 A}l/Z
J2 m

EXAMPLE 5.2
Consider the case n = 1, -00 <x< 00.

E = _3_ {'liZ A Z }1/3 '" 0.94 {'liZ A2 }1/3


2 5/ 3 m m

This can be compared to the ground-state energy of this system (for large argu-
ments of the relevant Airy function (d. Eq. (3.31) ) namely'" 0.89 {'li2 A2 1m f/3.
Meanwhile a variational calculation (Example 10.3) yields

EXAMPLE 5.3
Consider the case n = -1, -00 < x < 00.

This agrees exactly with the ground-state energy of the system (d. Eq. (3.32) ).
70 Chapter 5
EXAMPLE 5.4
Consider the case n = 1/2, -00 <x< 00.

E
5/2 {n2 A4 }1/5
2 7/ 5 m/2
= _5 {~}2/5 {~}2/5
212/ 5 7r fiiTi
0.599 {
hA2
~
}2/5
y2m

This is comparable to the analogous Wilson-Sommerfeld ground-state result (d.


Eq. (1.17) ):
15
En:1 = { 32
}2/5 {
hA2
fiiTi rv 0.739
}2/5
hA2
fiiTi
{ }2/5

If
E =p2/2m+Axn
only for x > 0, and V(x) = 00 for x < 0, the above derivation must be modified in
that Llx rv x and

E= n + 2 {n2 A2/n }n/(n+2)


(n i- -2). (5.3)
2 nm

EXAMPLE 5.5

n: f/2 ,
If n = 2, x ~ 0,

E= ~ { ~~ f/2 = Vi {
compared to the exact ground-state energy in this case (d. Example 8.1),

n2 A}1/2
Eground =.J:4.5 { -;:;:;:

EXAMPLE 5.6
If n = 1, x ~ 0,
Uncertainty Principle and Ground-State Energies ... 71
which can be compared to the variational calculation for this problem (d. Example
1/3
10.2) Eground < 1.86 [ A2h,2 Jm ] and the lowest energy of this system for large
1/3
arguments of the relevant Airy function, namely 1.84 [ A 2h,2Jm] (d. Eq. (3.18) ).

EXAMPLE 5.7
If n = -1, x 2:: 0

which agrees exactly with the ground-state energy of the system (d. Eq. (3.23) ).
One should emphasize that Eqs. (5.2) and (5.3) are very rough estimates of the
ground-state energy of quantum-mechanical systems, but are nonetheless convenient
if one is interested in a result which is of the right order of magnitude.

EXAMPLE 5.8
If n = 1/2, x 2:: 0,

E= ~ {h,2 A4 }1/5 rv 1.65 { h,A 2 }2/5,


4 m/2 v'2ffi
as compared to the Wilson-Sommerfeld result for this problem (d. Eq. (1.18) )

E1
= {15}2/5
16
{~}2/5
v'2ffi rv
2.03
{~}2/5
v'2ffi

EXAMPLE 5.9
Consider a particle moving in the attractive potential:

IVoIE (5.4)
V(x) = - (2 + E2)·
11" X

This potential is illustrated in Figure 5.1.


Using the arguments at the beginning of this Chapter,

and
oE
ox =
72 Chapter 5

v (x)

Figure 5.1: Potential in Example 5.9.

The choice of b..x which minimizes E is thus:

I\X}2
{ L...l = 40'(;2 h
werea=
4-a

Therefore

(5.5)

As E ~ 00

E mm
. ~ [ 11101
--+11,
7rE

This is consistent with Example 5.1 since for E ~ 00,

V(x) ~ C + Ax2, where C = _I~:I, and A = ~; .

One can also obtain energy estimates for states other than the ground state pro-
vided one makes some assumption about b..x b..p for these states. For instancel one
may assume b..x b..p '" qn, which is not however generally true. Thus for the hydrogen
problemb..rb..p=nJq2+2/2 (d. Eq. (3.30) ),andifq»J2, b..rb..p"'qn/2.
Since this involves a further approximation, the resulting calculations for excited
Uncertainty Principle and Ground-State Energies ... 73
states will be even less reliable than the ground-state energy estimates made above.
With this further approximation the excited energy estimates can be obtained from
the above results by replacing Ii by qli in the relevant expressions.

References

1. H. A. Gersch and C. H. Braden, Am. J. Phys. 50 (1) 1982, p. 53.


CHAPTER 6

Free Particles Incident on Potentials,


Time Delay, Phase Shifts, and
The Born Approximation

When quantum-mechanical particles are incident on a potential, one is in the


first instance interested in the fraction transmitted through the potential, and the
fraction reflected by it. One therefore calculates the probability of reflection and the
probability of transmission.
In detail, if one writes for a particle of energy

that the wave function on the left side of a one-dimensional potential is:

'¢L(x) = Aexp (ikx) + Bexp (-ikx), (6.1)

and the wave function on the right side is

'¢R(X) = Fexp(ikx), (6.2)

this choice implies the particle is incident on the potential from the left.
Additionally

R = B I A is the reflection amplitude, with


IRI2 the reflection probability, and T = FIA (6.3)
is the transmission amplitude, with ITI2
the transmission probability.

The poles of the transmission amplitude correspond to the allowed bound states
for that particular potential, and continuity considerations require

(6.4)
In addition an incident wavepacket can be written (cf. Chapter 4) as

Win (x, t) = ~ JA(k') exp [i {k'x - E(k')tlfi}]dk', (6.5)

and based on this the reflected and transmitted wavepackets are:


Free Particles Incident on Potentials, ... 75

Wref(x, t) = ~ JR(k')A(k') exp [-i {k'x + E(k')tjh}]dk', (6.6)

t) = ~ JT(k')A(k')
and
Wtrans(x, exp [i {k'x - E(k')t/Ti}]dk', (6.7)
respectively.
Using standard procedures (d. Eq. (4.14) ) one can then obtain from Eqs. (6.5)-
(6.7) the time required for reflection off (the so called 'time delay') and transmission
through a particular potential.

EXAMPLE 6.1
Discuss the problem of particles incident on a potential

V(x) = Vac5(x).
If
'ifJL(x) = Aexp(ikx) + Bexp(-ikx),
(6.8)
Fexp (ikx),
continuity of the wave function at x = 0 implies:
A+B=F. (6.9)
V( X)

Figure 6.1: Potentials in Example 6.1, if Va < 0, or Va > O.


76 Chapter 6

This potential is discontinuous at x = o. Thus, using the Schrodinger equation,

1i 2 d2
- - - d2'I/J(X)
2m x
+ Vo5(x)'I/J(x) = E'I/J(x),
and integrating across x = 0,

1i
- 2m
2
j'
_, d (d'I/J)
dx + j'_, Vo5(x)'I/J(x) dx = j'
E _, 'I/J(x) dx.

In the limit as E ~ 0 this becomes:

lim{d'dx
,-+0
I/Jl _d'I/Jl
dx x=-€
}=2m2Vo 'I/J(0),
1i
(6.10)
x=€

I.e.
ik(F-A+B)-
- 2mVo
1i 2 F • (6.11)

Solving Eqs. (6.9) and (6.11) simultaneously yields:

exp (i5r )
(6.12)

while,

(k1i2/mVo) exp {i (5r + 7f/2)} k1i 2 }


VI + k21i4/ m2V02 -arccot { mVo

(6.13)

In this case
1
IRI2 k21i4/m2V02 +1 '
(6.14)
k21i4/m2Vi
ITI2 =
k2n4/m2V02 + 1 '

and as expected /R12 + /T/2 = 1, consistent with Eq. (6.4).


The poles of the transmission amplitude satisfy (ik11?) /(mVo) = 1,
I.e.
(6.15)
Free Particles Incident on Potentials, ... 77
Thus there is only one pole here and this is physically meaningful when Yo < o.
To get reflection and transmission times one must construct wavepackets:

dk'

(6.16)

= - -1- J A(k')k"h 2jmVo


..j2; V + k,2Ti jm2Va2
exp [.z (k' x - --t
Tik,2
2m
£)] dk'
+ Ut
1 4

exp (iO) Wtrans (x - Vgt + ~~: Ik'=k' 0) ,


where
Vg = Tik'i '
m k'=k
and 1;, 'f/, and 0 are phases. If at t = 0, x = 0 for the incident wavepacket, in the
reflected wavepacket
1-
x = 0 at t = - d8,r I = Tref ,
Vg dk k'=k
while in the transmitted wavepacket

dk' I
= -1 d8 = Ttrans
t
x = 0 at t .
Vg k'=k
Thus

(6.17)

As k - t 00, Tref, Ttrans -t 0, the classical result.


78 Chapter 6
EXAMPLE 6.2
Discuss the problem of particles incident on the potential
Vo8(x+a), x < 0,
V(x) = {
00, x> O.
If
'ljJL(x) Aexp{ik(x + a)} + Bexp{-ik(x + a)}, x<-a,

G sin kx, -a < x < 0 (consistent with 'ljJ(O) = 0),

v (X)

Figure 6.2: Potential in Example 6.2.

continuity of the wave function implies


A +B = -Gsinka. (6.18)
The discontinuity of the derivative of 'ljJ at x = -a (see Eq. (6.10) ) implies
. 2mVo
kG cos ka - zk(A - B) = ~(A + B). (6.19)

Thus
B _ -2mValli 2 -kcotka-ik _ ('C)
R - - 2 - - exp lor ,
A 2m Valli + k cot ka - ik
Free Particles Incident on Potentials, ... 79
(6.20)

Dr = 2arccot ( cot ka + 2mVo)


h2 k .

One notes IRI2 = 1. In fact B I A in this case is both the reflection amplitude and
the transmission amplitude since there is no transmission beyond x = O.
The allowed negative energies satisfy the equation:

2mVo .
-,;r + k cot ka = zk. (6.21)

Defining

u". -= k -- V2mE -h2 -


-V2m1EI
Z h2 ,

Eq. (6.21) becomes


2mVo
~ + K, coth K,a + K, = 0,
for allowed bound states. Also the particle is reflected from x = -a at

As k becomes very large

2a { csc2 ka } 2a
7 ref -+ Vg 1 + cot 2 ka = Vg ,

the classical value.

In three-dimensional systems where particles of mass m and energy E = h 2 P 12m


are incident on potentials VCr) of finite range a, the radial wave function for r > a
can be written:

ul(r) = RI(r) = AI {jl(kr) - tanDI{k)nl{kr)}, (6.22)


r

where jz, nl are the usual spherical Bessel and Neumann functions and DI( k) is the
"phase shift" for a particular "wave" or angular momentum I component of the wave
function.
In terms of these phase shifts a standard derivation 1 shows the scattering am-
plitude, the three-dimensional analogue of the transmission amplitude in Eq. (6.3)
IS:
1
k:L
00
fk(O) = eiSz(k)sinDI(k)PI(cosO), (6.23)
1=0
80 Chapter 6
while the (scattering) cross section

satisfies the relationship:


41T
= k2 ~J21 + 1) sin2 Ez(k).
00

a(k) (6.24)
z=o
As in the one-dimensional case, for a given potential, the poles (for each I) in the
scattering amplitude it gives rise to determine its allowed negative energies (for the
corresponding angular momentum).
The phase shifts Ez(k) may be obtained exactly, by requiring that the wave function
(and its derivative, if V has no infinite discontinuities) in the region where there is
a potential, match smoothly onto the external wave function Eq. (6.22) (and its
derivative) at r = a, where a is the "range" of the potential.
Alternatively at high incident particle energies, or for weak potentials one may
have recourse to the partial-wave Born approximation:

tan Ez(k) = --2-


2mk
n
1 j?(kr)V(r)r2dr,
0
00
(6.25)

or else use the (first) Born approximation, which for central (l independent) potentials
reduces to
iB(k') = - :~ roo
sinkk'r' V(r')r,2 dr " (6.26)
n Jo 'r'
(with k' = 2ksin(}/2), where the differential cross section is:

:~ = If(k'W, (6.27)

while the total cross section a is

a(k) = JIf(k'WdO '" JliB(k')1 2 dO. (6.28)

EXAMPLE 6.3
Consider a particle scattered off the three-dimensional potential

V(r) = VoE(r - a).


In this case because the wave function must be well behaved at the origin, and nz
blows up at r=O, one has

uz(r) = Bzrjz(kr), 0 < r < a,

= Azr Uz(kr) - tan Ez(k)n/(kr)} , a<r< 00.


Free Particles Incident on Potentials, ... 81
For angular momentum 1 = 0 if one matches the two pieces of the wave function
at r = a one obtains:
Bo sin ka = A o" sin (ka + 80 ), (6.29)
cos vo
while the discontinuity of the derivative of uo(r) at r = a (see Eq. (6.10) ) requires:

Ao 2mVo .
--,,- cos (ka
oos~
+ 80 ) - Bo cos ka =- k
~
2 Bo sm ka. (6.30)

The ratio of Eq. (6.30) to Eq. (6.29) is:

2mVo
cot (ka + 80 ) - cot ka = k~2 ,

I.e.

~2k )
cot 80 =- ( cot ka + 2m Vo csc 2 ka

V( r )

a
~---+-------------r

Figure 6.3: Potential in Example 6.3.

or
sin80 (k) = _ 1 ,
(esc ka h/1 + 2A oot ka + A2 csc2 ka
where
82 Cbapter 6
The 1=0 contribution to the scattering amplitude (cf. Eq. (6.23) ) is

A(O) = ~exp {i60 (k)} sin60 (k),


which has poles when

kh? ) 2
1+2 +(
k'h,2 )
( 2mVo cotka 2mVo csc2 ka = 0,

i.e. when
k1i2
( 1 + 2m Vo cot ka
)2 = - (2mkfi2Vo )2 '
which is in agreement with the result of Eq. (6.21), as it must be.
For higher l's different (additional) bound states result. The I = 0 contribution
to the scattering cross section is:

O'(k) = 47r
(P csc 2 ka) { 1 + (k1i 21m Vo) cot ka + (k1i2 12m Vo f csc ka}
2

If a ---t 0 sin60 (k) ---t -2mVoka 211i2 ---t 0, and with it the 1= 0 contribu-
tion to the scattering amplitude fk(O) and the cross section O'(k) which goes to
47ra 2 (2mVoalfi2r ---t O.
From the expression above for sin60 (k) one notes that 60 (k) depends on the sign
of Vo. Moreover, if k is large, i.e. A is large (where A = {h2k} I {2mVo}), then

viI + 2A cot ka + A 2 csc2 ka '" A csc ka


and one can write that approximately:

1
sin 60 ( k) '" --::----:----:-- = (6.31)
csc ka A csc ka
while the partial-wave Born approximation I = 0 phase shift may be easily evaluated:

tan 60 ( k) '" -2mk


7Jo.2 (ka) Vo a2 =
in agreement with Eq. (6.31), as it should be.
The first Born approximation for this potential (Eq. (6.26) ), can also be evalu-
ated:
f B -- -t;2
2m {OO sink'r' ,2 Vr C(' ) d ' _ 2mVoa. k'
Jo -y;;;:tr 0 u r - a r - -~ sm a,
Free Particles Incident on Potentials, ... 83
yielding a differential cross section (d. Eq. (6.27) ),

As a-+O, (da)/(df!) -+ {(m2V02a2){2ak sin (61 /2))2} / {1i4k2 sin2(61 /2)} -+ (2m Voa 2/1i 2f
with a(k) -+ 41l' (2mVoa 2/1i2f, in agreement with the partial-wave results.
EXAMPLE 6.4
Consider a particle scattered off the three-dimensional potential:

Obtain the relevant expressions for l=O (S wave) scattering.


For
r > a, uo(r) = Au" sin (kr + 80 ),
cOS Vo

For
0< r < a, uo(r) is the solution of the equation:
1i2 J2 21i2 }
{ - 2m dr2 + 2mr2 - E uo(r) = O. (6.32)

The solution of Eq. (6.32) which does not diverge at the origin is Borjl(kr).
V( r )

,
,

,
______-+____________ ~ __ ~ __ r
a

Figure 6.4: Potential in Example 6.4.

Hence one can easily find the 1 = 0 phase shift for this system by requiring
continuity of the wave function and its derivative at r = a :
84 Chapter 6

= cos~:(k) sin{ka + t5o(kH


Aok
= cos80{k) cos{ka + 80 (kH.

Thus

where
. () sin p cos p
Jt P = - - - - - ,
p2 P
I.e.
80 ( k) = arccot [: In {pjt (PH]
P
I - ka.
p=ka

Evaluating this one obtains an explicit expression for 80 (k) :

tant5o(k) = -ka {tanka + cot ka - (1/ka)2 tanka}.


tan ka + cot ka - l/ka
As k - t 00, tan 80 ( k) - t -ka, the infinite barrier result, since for an infinitely
repulsive barrier of radius a,

0= sin (ka + 80 ), I.e. 80(k) = -ka. (6.33)

If a is very small (i.e. the potential has a very short range) or k (i.e. the energy)
is small such that the product ka is small,

tant5o(k) '" -ka {ka + (l/ka -ka/3) - (1/ka)2 {ka + (ka)3/3}} = _ ka .


ka + l/ka - ka/3 - l/ka 2

In the Born approximation the total differential cross section for this potential
may be obtained by first evaluating the scattering amplitude:

If one allows a - t 00 this integral can be performed analytically and one may
write:
Free Particles Incident on Potentials, ... 85
The cross section in this case (a ~ 00) for scatttering between () = ()1 and () = ()2
1S

71"2 r02 sin () d() ~ 271" r02 2 sin (() /2) cos (() /2) 2d( () /2)
4k2 271" }01 sin 2 (() /2) = 4k 2 }01 sin 2 (()/2)

= 2;3 {In sin (0 2 /2) - In sin (01/2)}

= 271"3 In { sin (0 2 /2) }


k2 sin (Od2) .

EXAMPLE 6.5
Consider a particle scattered off the potential:

-lVol + mw 2 r2/2, 0 < r < a,


V(r) = {
0, r > a,
which is purely attractive or partly attractive and partly repulsive depending on the
values of Vo, w and a (see Figure 6.5).
v (r )

V( r )

101 -ivol
Figure 6.5: Potential in Example 6.5.

Consider in particular the special case


3 n2 p
E = -IVoI + -nw
2
= -2m .
86 Chapter 6
At this energy
mwr2)
uo(r)=Nrexp ( ----v;:- , 0< r < a.
Matching the internal and external wave functions at r = a yields:

Naexp(-mwa 2/2n) = ~tan{ka+8o(k)}


N exp ( -mwa 212n) - N (a 2mw In) exp ( -mwa 212n) k
a
- 1- mwa 2/n .
Hence
ka
tan {ka + 8o( k)} =
1-mwa 2 In '
I.e.
ka + (mwa 2In) tan ka - tan ka
=
1 - mwa 2In + ka tan ka
tan8o(k) (6.34)
ka + (2ma 2/3n 2)(E + IVoI) tan ka - tan ka
1 - (2ma2/3n2)(E + IVoI) + ka tan ka
If w is large the potential looks like an infinitely repulsive barrier of radius r = a
and the scattering takes place at high energy since E = -IVoI + 3/2 nw.
In this case tan8o(k) rv -tanka, i.e. 8o(k) = -ka, exactly the infinite repulsive
barrier result Eq. (6.33).
If w is small this problem reduces to scattering at low energy off a square well of
depth IVoI (since small w implies small E). For square-well scattering:
(k) _ (kIK)tanKa-tanka
- 1 + (k I K) tan K a tan ka '
C
tanuo

where
2m
K= -,;: {E + lVol} ,
and the potential V = -lVol has a range 0::::; r ::::; a.
At low energy

k k { (Ka)3} kK 2a3 ka32m


KtanKarv K Ka+-3- =ka+-3-=ka+31i2{E+IVoI},

hence for square-well scattering at low energy:


Free Particles Incident on Potentials, ... 87
Eq. (6.34) agrees with Eq. (6.35) in this limit.
For E ~ 0, sin8o(k) rv tan8o(k) ~ 2ka3 m IVoI/31i 2 , and a(k) ~ 47ra 2 (2a 2 m IVoI/31i 2
The total scattering amplitude for this potential in the Born approximation is
f.
easily obtained:

=
_ 2m
'h
2
la
0
sin (k'r') {
k'r'
-
Ill,0 I + -21 mw 2,2
r
} ,2 d ,
r r

2mlVoI fk1a m 2w2 fk1a


f(k') = 'h 2k,3 Jo (k'r') sin (k'r') d(k'r') - k,5'h 2 Jo (k'r,)3 sin(k'r') d(k'r')
2mlVoi .
= 'h 2k,3 {sm (k'a) - k'a cos (k'a)} -
m 2 w2
k,5'h 2 {[3(k'a)2 - 6] sin (k'a) + k'a [6 - (k'a?] cos (k'a)} .

For small k

consistent with Eq. (6.35) in this limit.

EXAMPLE 6.6
Evaluate the differential and total scattering cross section for the potential:

V( r)
V(r) = Voexp (-2r22 ) ,
a

L-____ ~====~ ________ ~_r

Figure 6.6: Potential in Examples 6.6, 6.7.

in the Born approximation.


88 Chapter 6

- Va
-2m
li 2
- 1
0
00
sin k'r'
---exp
k'r'
- - r ,2 dr '
(r'2)
2a 2

= _ m Vo
2h 2 k'
lr
'=oo
r'=-oo
sin k'r' exp (_~) d {rI2}
2a 2

= _ mVo
211,2 k'
Iml r
'=oo
r'=-oo
eXP{ik'r' _ ~}d{r'2}
2a2

=
mVo (k,2 a 2 ) T
'=oo
- 2h2k' exp --2- 1m r'=-oo exp - 2az
I
((r' - ik'a 2)2)
dr.
{,2}

Defining u == r' - ik'a 2 , r' = u + ik'a 2, dr' = du, dr'z = 2 (u + ik'aZ ) duo
With this substitution:

= _ mVo exp (- k,2 a 2 )


2h 2 k' 2
l u
=oo
u=-oo
exp (-~)
2a
2
2k'a 2 du

fB(k')

I.e.
.,fiii m Vaa 3 exp (-k'2 a 2 /2)
fB(k') = liz (6.36)

The differential scattering cross section in this approximation is:

dO' _ 2?rm Vo2 a 2 6 (_k,2 a2),


dn - 1i4 exp

and the total cross section:

O'(k)

Thus
Free Particles Incident on Potentials, ... 89

O"{k) =

(6.37)

a result independent of the sign of Va. This problem may also be solved using Cartesian
coordinates.

EXAMPLE 6.7
Using the partial-wave Born approximation find the S-wave (I = 0) phase shift if
a particle is scattered off the potential of Example 6.6, namely:

V{r) = Vaexp ( - ;:2)'


Substituting in the partial-wave Born approximation, Eq. (6.25):

-2 Vo
-2m -
1i k
1 2kr (r22a ) dr
0
00
sm

exp - -
2

- - - 1 {l-
mVa r2) dr
00
cos2kr}exp ( - -
1i k
2 0 2a 2

(6.38)
_ mVo [ roo exp (_~) dr _ Re roo exp {2ikr - ~} dr]
1i k
2 Jo 2a 2 Jo 2a
2

1i2kVa V'2
- ma -2 a 2)} .
{if" { 1 - exp (k2

If ka ~ 1 Eq. (6.38) may be approximated by

Do(k) ""' - maVa f!.. 2k2a2 and fk ""' Do(k) = _ ma3Va .J2; .
1i 2 k V'2' k 1i2
Correspondingly if only 1=0 contributes to the cross section, using Eq. (6.24):
90 Chapter 6
which agrees with the result Eq. (6.37) of Example 6.6, in the limit k -+ o.
Analogously the approximate scattering amplitude obtained above, namely
ma3 Vo
+k '"
J. ---.j'2;
1i,2 ,

is consistent with the Born approximation result Eq. (6.36) if ka <t:: 1 since in this
limit one can write for Eq. (6.36):

fB(k') = _...& ;:;voa 3


exp ( _ k':a 2 )

...& mVoa3
1i,2

EXAMPLE 6.8
Find the Born-approximation scattering amplitude if a particle is scattered off
the potential:

This potential is illustrated in Figure 6.7.

V( r )

r
0[2
Figure 6.7: Potential in Examples 6.8, 6.9.

In the Born approximation:

h(k') = - a~~~,Im 1')0 exp (ik'r') exp ( - ; ; ) r,3 dr'


Free Particles Incident on Potentials, ... 91

Defining u == r' - ik'a2 , and taking the imaginary part of this integral yields:

m v;0 k ,2 a
---exp ( - - -
2) 00 , 2 2 ,3 6
{{3ka u - k a }exp - -
( U 2·) du
~~~ 2 h 2~
fB(k') (6.39)

Hence the differential cross section is:


d _
...!!...
dO -
m
2TT2 6
"0
2li4
a 7C' (3 _ k,2 a2)2 exp (_k,2 a2),
and the total cross section is
~W~~l~
a= - 04
Ii 0=0
exp {-4k2a 2 sin2 (0/2)} [3 - 4k 2 a2 sin2 (0/2)] 2 d(cosO).
A straightforward, though rather lengthy calculation yields in this case

EXAMPLE 6.9
Using the partial-wave Born approximation find the S-wave (1 = 0) phase shift if
a particle is scattered off the potential in Example 6.8, namely:

Var2
V(r) = 2a 2 exp - 2a 2
(r2 ) •

According to the partial-wave Born approximation the 1 = 0 phase shift is:

x 2ika2)2) ~ d(_r )}
{...fi4 _exp (-2k 2a2)Reh{OOexp (_ (r - ~ ~ ~a

21i 2Va
= - ma {if { 1 -
k V'2 ( 1 - 4k 2a 2) exp (-2k2a 2)} . (6.41 )
92 Chapter 6
If ka <: 1

and

which agrees with Eq. (6.39) in this limit.

EXAMPLE 6.10
Evaluate the differential and total scattering cross section for the potential:

v; r
V(r) = _o_exp (r2
- -)
,j2 a 2a 2

in the Born approximation. As in the previous examples,

-
2 mvo
TT

,j2 an?
1° 00' sm k" r , r
- - - r exp - -
k'r'
(,2 ) r
2a 2
,2 d
r
'

- 2mVo
,j2 ak'n?
1m 1°
00
e x{'k"
pzr--
2a 2
r,2} r ,2dr ,
- 2mVo exp (k,2
,j2 ak'1i 2
- --a2) 1m
2
1° 00
exp ( - - 1 {r' - zk
2a 2
.,a 2}2)'r2 dr.,
Defining u == r' - ik' a2 , r' = u + ik'a 2 , r,2 = u 2 + 2ik'a 2u - k,2 a4 •
With this substitution:

h(k')

Hence the Born approximation differential cross section is:

and the tofal cross section in this approximation is:


Free Particles Incident on Potentials, ... 93

EXAMPLE 6.11
In the preceding problem,

V(r) =
v; r
a~exp
(r2
- 2a
) 2

evaluate the S-wave phase shift using the partial-wave Born approximation.

tanoo(k) = _ 2mk foo sin 2 (kr) Vor exp (_~) r 2 dr


1i 2 Jo Pr 2 a../2 2a 2

~~:k 10 00
{1- cos (2kr)} exp ( - ;;2) (;;2)
d

= - ~~:k { I - exp (-2k 2a2 )Re 1000


exp ( - 2U;2 )2(U + 2ika2) d (2: 2) }

Thus one obtains

-1/2 V{r)
Vo e " ~--

<l $0
Figure 6.8: Potential in Examples 6.10, 6.11.
94 Chapter 6
EXAMPLE 6.12
Evaluate IB(k') if V(r) = Vornexp(-ar).

- Vo
-2m -
fi?
10
00 sin (k'r') r In exp ( -ar') r ,2 dr ,
k'r'

-2mVo
- - Im
ti 2 k'
1 0
00
t r , - ar ') r m+1d'
exp (·k' r

2mVo Im(
--2-
tik'
1.k) +2
a-z'n
10
00
exp ( -u ) u n+l du

2mVo(n + I)! .
ti2k'(a 2 + k,2)Hn/2 1m exp {z,B(n + 2)} where ,B == arctan (~)

2 2mVo(n
( +2)I)! / sm
. [(n+2)arctan (k')
-
]
ti k' a 2 + k' Hn 2 a

Special cases of interest include: (i) n = 0 i.e. one has a Yukawa potential
V(r) = Voexp(-ar).
In this case

IB = -ti2k'~~~ k,2) sin [2 arctan (~)] = -ti2(:7~;'2)2


(ii) n = -1 i.e. one has a shielded Coulomb Potential V(r) = Voexp (-ar)/r.
In this case
2mVok' 2mVo
IB = -ti2k'(a2 + k,2)1/2(a2 + k,2)1/2
(iii) a -+ 0, in which case the potential is some power of r, V(r) = Vorn.
In this case
2mVo(n + I)! .
IB = - ti2k,n+3 sm (1 + n/2)7r
for integral n, and if n is not an integer,

EXAMPLE 6.13
Find the reflection time for a particle incident with energy E on a step barrier:

0, x < 0,
V(x) = {
Vo, x > ° (Vo > 0, Vo > E).
Free Particles Incident on Potentials, ... 95
For x <0
'IjJ(x) = Aexp (ikx) + Bexp (-ikx);
while for x >0

'IjJ(x) = Cexp(-'I}x),
where
k = J2;2E, 'I}(k) = J2m (~2 - E) = J2~;o - k2 .
Matching boundary conditions at x = 0 one obtains:

B = A [k - ~'I}] .
k + ~'I}
Constructing wavepackets:

Winc(X, t) = JA(k') [i {k'X - E(:')t}] dk',


exp

Wref(x, t) = JB(k') [-i {k'X + E(:')t}] dk'


exp

j A(k,)k'-i'l}(k') [-'{k'
k' + i'l}(k') exp ~
E(k')t}]
x + -n,
dk'

= j A(k') exp [-i {k'x + E(:')t + 6(k')}] dk',

where

-n,2k2 6(k) = arctan [


E(k) = -2-' 2'1}(k)k 2
(k)] .
] = 2 arctan [ ~k
m k2 - {'I}(k)}
Using stationary phase arguments (cf. Eq. (4.14) ) one obtains:

If the particle reaches x = 0 at t = 0, it leaves x = 0 at

T = -:9 d~~~') Ik'=k = ~~


96 Chapter 6
SInce

hk
Vg =- .
m

If 'fJ -t (Xl, i.e. Vo - t (Xl, then T - t O. In this limit B = -A. If k =


= /mVo/1i2 , T is
'fJ an extremum. Moreover one notes that d8/dk is singular at
k = /2m Vo/h 2 ,i.e. 'fJ = 0, for which case T = (Xl.

EXAMPLE 6.14
Consider the two-step barrier:

V(x) = !0,

Vo,
VI,
x < 0,

0<x <

x> a.
a,

A particle is incident on this barrier from the left with energy E. Assume E >
Va, E> VI (See Figure 6.9.) Find for what values of a the transmission is a minimum
and for what values of a it is a maximum.

V (x) V( x)

---+-----'---t----x
d a

Figure 6.9: Potential in Example 6.14.


Free Particles Incident on Potentials, ... 97
The wave function can be written as follows:
+ Bexp (-ikx), x < 0,
j
Aexp (ikx)

'¢(x) = C exp (iklX) + D exp (-iklX), 0 < X < a,

Fexp (ik2X) x> a,


where
k _ /2mE
- ",2'

Requiring continuity of '¢, '¢' at x = 0, a, implies:


F 2exp(-ik2a) .
=
A (1 + k2 I k) cos kl a -
i ( kd k + k2 I kd sin kl a '
4
1~12 =
(1 + k2/k)2 + ({ki - kD IP + kUki - 1) sin2 kIa'
F 2exp (-ik2a) .
=
B (1 - k2lk) cos k1a + i (kdk - k2/k 1) sin kIa'
4
I~( = (6.42)

Thus

But

and

Tmin - v~ ~{ 4 }
(1 + V{E -
2 ,
max E Vi} IE) + {Vo(Vo - Vi)} / {E(E - Vo)}

where the upper value is for Va > Vi and the lower value for Vo < Vi .
This example shows that the condition kl a = mr does not necessarily yield max-
imum transmission. This is the case if Va > Vi. The opposite is the case if Va < Vi.
In this case the transmission is a maximum if kl a = ({2n + I} /2) 7r.
98 Chapter 6

EXAMPLE 6.15
Consider a particle scattered off two identical delta-function potentials, namely:

V(x) = -Ivai {8(x - a) + 8(x + an.


Find the scattering amplitude and its poles.
For convenience one may choose

Aexp{ik(x+a)}+Bexp{-ik(x+an, < x < -a,

1
-00

'lj;(x) = C exp {ik(x + a)} + Dexp {-ik(x + an, -a < x < a,

Fexp {ik(x - an, x> a.

The wave function is continuous at x = ± a and its derivative satisfies equations


analogous to Eq. (6.10), namely:

lim{d'lj;j _ d'lj;j }=_2ml;01'lj;(x=±a).


5.....0 dx x=±a+5 dx x=±a-5 h

Hence, defining
_ m/vol
1I=~,

one obtains that:

A+B C+D

A(1+2ill) +B(-l +2ill) C-D

C exp (2ika) + D exp (-2ika) F

Cexp(2ika)-Dexp(-2ika) = F(1-2ill),

which imply

A (1 - ill)C - illD

C = exp (-2ika)(1 - ill)F

D ill exp (2ika) F.


Free Particles Incident on Potentials, ... 99
Solving for F / A one obtains
F
{(1 - i//)2 exp (-2ika) + //2 exp (2ika) }-1
A

= { (1 - i2//) cos 2ka + i(2//2 - 1 + 2i//) sin 2ka} -1 .

The scattering amplitude is thus


Fexp (-ika)
Aexp (ika) = ~ exp (-2ika)
exp (-2ika)
=
[cos 2ka - 2// sin 2ka + i {(2//2 - 1) sin 2ka - 2// cos 2ka}]
(6.43)
exp {-i(2ka - 8)}
=
J{cos 2ka - 2// sin 2ka}2 + {(2//2 - 1) sin 2ka - 2// cos 2ka}2 '
with
(' = -arctan [(2// 1) tan 2ka - 2//] .
u
2 -

1 - 2// tan 2ka


If a --4 0 Eq. (6.43) reduces to Eq.(6.13) as it must (but with Vo replaced by
2Vo), namely T = 1/(1- i2//), 8 = arctan (2//).
The poles of Eq. (6.43) occur when:
" 1 - i2//
- dan2ka = 2 2 2" (6.44)
// -1+ w
This equation is satisfied if E < 0 since then

mlVoI
// = --:=-U:. "
1i 2 i",
Eq. (6.44) then reduces to the equation for the allowed bound states
2f -1
tanh2",a= 1 - 2 f + 2 f 2 '

which is satisfied when either


tanh ",a = 2f - 1 (6.45)
or
(6.46)
100 Chapter 6
V( X)

-0 0
------.-----~----_T----------. X

Figure 6.10: Potential in Example 6.15.

Eq. (6.45) corresponds to even-parity bound states while Eq. (6.46) to odd-parity
bound states.
If a -+ 0 Eq. (6.46) has no solution, while Eq. (6.45) has only one solution,
namely

2c: -1 = 0 i.e. c: = ~, that is lEI = m(~~1)2


which is just Eq. (6.15), with Va replaced by 2Va, while if a -+ 00, both Eq. (6.45)
and Eq. (6.46) reduce to c: = 1 which is just Eq. (6.15). These results are as expected
since in the former case the delta-function potentials coalesce, while in the latter they
essentially uncouple.
If k becomes large, v tends to zero and 8 to 2ka. This means if the incident
wavepacket is

i.e. x = -a at t = 0, then the transmitted wavepacket Wtrans is

i.e. x = a at t = 2a/vg which as expected is the classical transit time for crossing
these two potentials.
Free Particles Incident on Potentials, ... 101
References

1. E. Merzbacher, Quantum Mechanics, Wiley (1970), p. 234 f.


CHAPTER 7

Heisenberg Representation

Starting with the expression for the expectation value of an operator Os in the
Schrodinger representation (at some time t) :

(O}t = L: W*(x, t)OsW(x, t) dx, (7.1)

and the fact that generally (Eq. (4.12) )

W(x, t) = L: G(x, x', t)W(x', 0) dx', (7.2)

where for a free particle,

G(x,x',t) = 1 001
--exp
V2i
- 00
(iP'X') 1
- - - --exp
fi V2i
(iP'X)
-
fi
exp (iE(P')t)
- - - - d (p,)
fi
-
fi
whereas for a particle in a potential (which is assumed time-independent) (7.3)

G(x,x',t) = f<p:(x')<pn(x)exp(-i~nt),
n=O

one can go from the Schrodinger to the Heisenberg representation.


The procedure involves first noting that (for a free particle) Eq. (7.3) can be
written

G(x, x', t) = 1 00

-00
_1 exp (_ iP'X') exp (_ iH(X)t) _1 exp (iP'x)d (p,)
V2i fi fi V2i fi fi

fi - 00 V2i
1
exp (_ H(X)t) 00 _1 exp (_ iP'X')_l exp (iP'x)d (p,)
fi V2i fi fi

iH(X)t)c(
exp ( ---fi- U x - x
') ,

(d. Eq. (2.19) ).


Similarly, for particles subject to a time-independent potential, Eq. (7.3) can be
written

iH(X)t)
G(x,x',t) = ~<p:(x')exp ( - - iH(X)t)
f i - <Pn(x) =exp ( - - f i - 8(x-x'),
Heisenberg Representation 103
(d. Eq. (2.6) ), i.e. generally

G(X,X', t) = exp ( - iH(X)t)


- h - S(x - x'). (7.4)

Substituting Eq. (7.4) back into Eq. (7.2) yields:

iH(X)t)
\II (x, t) = exp ( - - h - \II (x, 0). (7.5)

Thus Eq. (7.1) can be rewritten

(O)t = 100
-00 \II * (x,O)exp (iH(X)t)
- h - Osexp ( - iH(X)t)
- h - \II(x,O) dx, (7.6)

I.e.
(7.7)
where
OH -= exp (iH(X)t)
- h - 0 s exp (iH(X)t)
--h- (7.8)

and the subscripts Sand H stand for the operator in the Schrodinger and Heisenberg
representations respectively.
In Eq. (7.7) one thus has simpler wave functions than in Eq. (7.1) i.e. only
\II(x,O), independent of the time, but more complicated operators OH rather than
Os.
The operator OH can be seen to satisfy the differential equation:

(7.9)

SInce

dOH
- i
- = -exp (iHt)
- (HO s - OsH)exp (iHt)
- - +exp (iHt)
- aDs
- - exp (iHt)
--
dt h h h hat h'
and where

{A,B}=AB-BA, - H =exp (iHt)


aO- - aDs
- - exp (iHt)
--
at- hat h'
To obtain Eq. (7.9) one must insert exp (-iHtjh) exp (iHtjh) = 1 between
H and Os. One assumes here that H of:. H(t), i.e. that the Hamiltonian is time
independent, in which case

-
HH-exp (iHs(x)t)H
h sexp (iHs(X)t)
- h -- H s·
104 Chapter 7
If 0 = x or P and H =p 2 /2m+ V(x),
then
dXH i PH
Tt=r;{HH,XH}= m' (7.10)

dPH i aVH
Tt = h {HH' PH} = - ax = FH , (7.11)

and combining Eq. (7.10) and Eq. (7.11) one obtains:

(7.12)

The standard Heisenberg representation results Eqs. (7.10) - (7.12) look exactly
like the corresponding classical expressions for velocity and force. In fact Eq. (7.12)
looks just like Newton's second law. This analogy has formal merit.
A word of caution is however in order here. Both PH and XH are quantum-mechanical
operators and are more complicated than their classical analogues. Thus

XHPH ~ PHXH· (7.13)

h
Rather since {XH,PH} = ,
z
h
XHPH = PHXH - -;- .
z
(7.14)

EXAMPLE 7.1
Show that generally

but rather that

d(02)H = d(OH)2 = 20 dOH {dOH 0 }.


dt dt H dt + dt' H

From Eq. (7.9)

d(OH)2 = ~ {H (0
dt h H, H
)2}

if (0 8 )2 does not depend on t explicitly, where


Heisenberg Representation 105

exp C~t)O~exp (_ i~t)

exp C~t) Os exp ( - i~t) exp C~t) Os exp (_ i~t)

and

Hence
. .
d0 2 z z
-.!L
dt
1i {HH, OH} OH + Ow:;;, {HH' OH}

dOH 0 0 dOH
T t H+ HTt

l.e.
(7.15)

EXAMPLE 7.2
Show
d(X2)H = d(XH)2 = 2x dXH ~
dt dt H dt + mi .
From Eq. (7.15)

But from Eq. (7.10)


dXH PH
dt m
Thus
dXh dXH 1 dXH 11,
- =2XH-+-{PH XH} =2XH-+- (7.16)
dt dt m' dt mi'
where one has also used Eq. (7.14).
106 Chapter 7
One notes in passing that one readily gets Ehrenfest's expression from Eqs. (7.7),
(7.9) since:

1-00
00 w*(
x,
0) dOH W( 0) d = / dOH)
dt x, x \ dt t
d (O)t =
dt

EXAMPLE 7.3
Evaluate

d () d /1 d]OO *( ) 1
dt VH t = dt \2'mw XH t = dt -00 w x,O 2'mw xH W x,O dx,
2 2) 2 2 ( )

for a particle in the ground state of a simple harmonic oscillator potential:

1
Vex) = 2'mw 22 1
x, i.e. where VH (x) = 2'mw 22
XH .

From Eq. (7.16)


dXh _ 2x dx H !!:...
dt - H dt + mi '
therefore
dxh) = -2100 ¢o(X)XHPH¢O(X)
( -d * dx + -n, ..
t t m -00 m%

i: ¢~(X)XHPH¢O(X) i: ¢~(x)
But

dx = x [~!] ¢o(x) dx
= -in,]00 * mwx
-00 ¢o(x)x-n,-¢o(x) dx

2
--;-
]00 mw 2 x 2
¢~(x)-2-¢o(x) dx
zw -00
n,

where one has used the fact that

¢o(x) = N exp ( ----u:-


mwx2)
'
Heisenberg Representation 107
and that
iH(X)t) ( iwt) 4>o(x),
exp ( - - h - 4>o(x) = exp -2

etc.
Hence

(d;; ) t = - i~ + i~ = 0,
and the expectation value of the kinetic energy of the ground (and by a similar
calculation of any) state of a simple harmonic oscillator is independent of t.

EXAMPLE 7.4
Generalize the result of Example 7.3 for any state of the simple harmonic oscilla-
tor.
Defining
±=
aH - ~ 1 [p H ±.zmwxH, 1
v2m
one can quickly show, using Eq. (7.13) that {a1f, all} = -nw, and that

XHPH = 2~i [(a1f)2 - (all)2 + {at, all}]

Thus

jdX'k)
\ dt t

-.zmw1_1
-00
00
4>~(x) [(at)2 - (alif - nw] 4>n(X) dx +~
mz
h h
= --. +-. =0,
mz mz

since (allY operating on 4>n(x) produce states orthogonal to 4>n(x). For non-diagonal
matrix elements (i.e. min), it may happen that (d/dt) (nl mw 2x'k/2Im}t i 0
(d. Example 7.7).

EXAMPLE 7.5
Show
dTH _ PHFH _1_ {F }
dt - m + 2m H, PH . (7.17)
where TH is the kinetic-energy operator in the Heisenberg representation.
108 Chapter 7
From Eq. (7.15)

d(PH )2 = d(p2)H = 2 dPH {dPH }


dt dt PH dt + dt' PH .

But from Eq. (7.11)

Therefore:
dp2
d: = 2PHFH + {FH, PH},
or
dTH _ PHFH _l_{F }
dt - m + 2m H, PH .
Thus for example if VH = mw2x~/2,

-dPH
dt
aVH
= ---
ax
= FH = -mw 2 XH ,

and

EXAMPLE 7.6
Evaluate
d
dt (x}t
for a particle in the ground state of a simple harmonic oscillator potential:

Using Ehrenfest's expression one obtains:

d{x}t
dt
= (dXH)
dt
= i{{H
Ii
})
H, X H t·
t

From Eq. (7.10):

i ({HH' XH }}t (~)t = ~ (PH}t


i:
=

~ ~~(x)exp (i~t)~ :x exp (- i~t)~o(x)dx


i: i:
=

= i~ ~~(x) [- m;x] ~o(x)dx = iw ~~(x)x~o(x)dx = 0,


since J::'oo l~o(xW xdx = 0 from parity considerations.
Heisenberg Representation 109

EXAMPLE 7.7
Evaluate

d \2
dt 11 2 21) t =
/ 2'mw xH 0 d
dt 1 4>2
00
-00
1 2
*( x ) 2'mw 2
xH4>o ()
x dx,

where <Po{x), <P2{X) are the ground and second excited states of the simple harmonic
oscillator:
1
VH{X) = 2'rnw XH .
2 2

One notes, using the results of Examples 7.3-4 and the fact <Po{x), <P2{X) are or-
thonormal, that

EXAMPLE 7.8
Evaluate
d
dt (TH}t
for a particle in the ground state of a simple harmonic oscillator potential:

One wishes to evaluate:


110 Chapter 7
Using Eq. (7.17) and

mwx2)
W(x,O) = 4>o(x) = Nexp ( -'ih '

= _w 2 1 4>~(X)pHXH4>O(x)dx + 2i
00

-00
'hw2

= _w 2 1 00 'ha x4>o(x)dx +
4>~(X)-:--a
'hw 2
-2.

+ -. 1
-00 Z x Z
nw21°O 2w mw 2x 2 'hw2
+ -2.
00
= --.- 4>~(x)4>o(x)dx 4>~(X)-2-4>0(x)dx
Z -00 Z -00 Z

'hw 2 2w ('hw) 'hw 2 _ 0


= --z-· +T T +2i- .

Combining the results of Examples 7.3 and 7.8,

in this case, where

P'i:,
H H = 2m 1
+ '2mw xH = TH + TJ'
2 2
vH, an d 'T,(
'I" x,
0) = '1'0 X.
A.. ( )

Thus the expectation value of this Hamiltonian does not change with time. This
is not surprising since H does not depend explicitly on time and generally using
Ehrenfest's expression:

What is more surprising is that (TH}t and (VH}t do not change with time. This
is because one is evaluating the time evolution of the expectation or "average" value
of these operators, not of the operators themselves.
CHAPTER 8

Two-, Three-, and N-, Versus One-Dimensional Problems

The one-dimensional Schrodinger equation for a particle in a potential Vi (x) is

(8.1)

where
L: l1l!n(xWdx = l.
The radial equation for a particle in a three-dimensional radial potential Vj(r) is

(8.2)

where
Unl(r) I
1 = 0, 1, 2, ... , 1l!nlm(r, (), 4» = - - Ym«(),4», -I ~ m ~ I,
r

1 unl(r?dr
and
00
= l.

If 1 = 0 and Vi(x) = Vj(x) Eqs. (8.1) and (8.2) are identical. However,1l!n(x)
in Eq. (8.1) extends from -00 ~ x ~ 00 while Unl(r) must be zero at r = 0 in order
that Unl(r)Jr be finite at the origin. In addition Unl(r) extends only from r = 0 to
r = 00. A consequence of the above is that some solutions acceptable for Eq. (8.1)
are not acceptable for Eq. (8.2). Radial solutions of Eq. (8.2) with potential Vj(r)
are identical to solutions of the one-dimensional problem Eq. (8.1) if:

Vi(x) = 00, x < 0,


¥() h 1(1+1)
2
V1(x) = 3x + 2mx 2 ' x> O. (8.3)

If
V,()-¥() h2 1(l+1) x> 0,
lX-3 X + 2mx 2 '
and moreover Vi(x) = Vi(-x) then the odd-parity solutions (i.e. those solutions
which vanish at x=O) of 1l!n(x) (n = 1, 3, 5 ... ) are identical to Unll(X)J...ti (n' =
0, 1, 2, ... ,1 = 0, 1,2, ... ).
112 Chapter 8
EXAMPLE 8.1
Consider the potential Vi. (x) = (h 2 /2mb4 )x 2 • For this potential the solutions ofEq.
(8.1) for the energy En are En = (n+ 1/2)h 2 /mb2 (n = 0, 1,2 ... ). The lowest energy is
;,,2 /2mb2 and the ground-state wave function wo(x) = {l/(bJ"i )} 1/2 exp (_x 2 /2b 2).
Consider now the potential ¥J(r) = (;,,2 /2mb 4 )r2 , and the case 1 = O. The min-
imum energy Eoo for a particle in this potential is 31i? /2mb 2 • Examine why this is
so.
The three-dimensional radial function uoo(r) cannot be proportional to (the even
function) Wo(r) since wo(x) does not vanish at x = o. The second energy eigenvalue
for the one-dimensional problem is E1 = 3h 2 /2mb2 , with (odd) eigenfunction

2
WI (x) = ( b.ji
)1/2 (i) exp (2 )
-;b2 .

This is acceptable as the lowest eigenfunction of Eq. (8.2) (to within an overall factor)
since WI(O) = o. The normalization must be modified however since:

while

Thus
uoo(r) = (bV: (7;)
)1/2 exp (-;~).
Similarly all odd-parity solutions W3(X), ws(x) are acceptable (see Example 10.3)
solutions since for these solutions W(O) = o. Thus

etc. and
E(n-1)/2,o = (n + 2"1) mbh 2 '
2
n = 1, 3, 5 ... ,
i.e.

EXAMPLE 8.2
Consider a particle in the three-dimensional well:

0, 0< r < a,
V3 (r) = -IVoI, a < r < b,
0, b < r < 00.
Two-, Three-, and N-, Versus One-Dimensional Problems 113
Compare this and the analogous one-dimensional system:

0, x < a,
Vt(x) = -IVoI, a < x < b,
0, b < x < 00.

V( r) V( X)
J

~-----aT-----~~b----~r --+------a~-----r.b----~-x

Figure 8.1: Diagram in Example 8.2


The allowed 1 = 0 (negative) energy levels are easy to obtain. In detail the
solutions of Eq. (8.2) which satisfy the boundary conditions at the origin (uno(O) = 0)
are

uno(r) = Asinhkr, 0< r < a, (k = V2m1EI/h2 )


uno(r) = B sin (Kr + 6), a < r < b, (K = V2m(IVoI-IEI)/h 2 )

uno(r) = C exp (-kr), b < r < 00.

Matching the wave functions and their derivatives at a and b yields

tanh ka tan (Ka + 6)


=
k K
t.e.
K
tan(Ka+6) = Ttanhka, (8.4)
and
tan (Kb + 6) 1
K =-"k'
1.e.
K
tan(Kb+6) = -T. (8.5)
114 Chapter 8
One can rewrite Eq. (8.5) as
K tan(Ka+h)+tanK(b-a)
tan {K a + h + K (b - a)}
= - k = 1- tan (K a + 15) tan K (b - a)
(K/k) tanh ka + tan K(b - a)
= 1- (K/k) tanh ka tan K(b - a) .
Thus the energy levels for this system are obtained by solving the transcendental
equation
K (K/k)tanhka+tanK(b-a)
(8.6)
-k= l-(K/k)tanhkatanK(b-a)'
If a = 0 Eq. (8.6) reduces to -K/k = tanKb.
If a ---+ 00, b ---+ 00, c == b - a, Eq. (8.6) reduces to:
K K/k+tanKc
-k = 1- (K/k) tan Kc .

These results can be compared with the energy levels for a particle in the one-
dimensional well given in this example. The solutions of Eq. (8.1) for this potential
are
Aexp (klxl), x < a,
\li'(x) = Bsin(Kx + 15), a < x < b,
Cexp (-kx), x> b.
Matching boundary conditions one obtains

1 tan(Ka + 6)
"k= K (8.7)

tan(Kb+h) 1
K --"k' (8.8)
One can rewrite Eq. (8.8) as

tan (Ka + 15) + tanK(b - a)


tan {(Ka + 15) + K(b - an K
= - k =1- tan (K a + 15) tan K (b - a)
Kjk + tanK(b - a)
=
1 - (Kjk) tanK(b - a) .

Thus the energy levels for this system are obtained by solving the transcendental
equation
K K/k + tanK(b - a)
(8.9)
- k = 1 - (K j k) tan K (b - a) .
Two-, Three-, and N-, Versus One-Dimensional Problems 115

This is not equal to Eq. (8.6) except in the limit

a -+ 00
c == b-a
b -+ 00

where c is a constant, in which case the boundary condition at the origin is unimpor-
tant.

The two-dimensional Schrodinger equation for a particle of mass M in a potential


V2(p) can be written:

(8.10)

where

,T, ( "-) = Wnm(p) exp (±imcP)


~nm P,o/ vp "fi1r' m = 0, 1, 2, 3... ,

and

Here wnm(p) must be zero at p = 0 so that wnm(p) will be finite at the origin.
vp
Comparing Eqs. (8.2) and (8.10) and noting that

indicates that if one has a solution Unl(r) of Eq. (8.2), then wnm(p) is Unm -l/2(P).
One notes both u(O) and w(O) must be zero, as too u(oo), and w(oo), (if the particle
is bound) i.e. u(r) and w(p) have identical boundary conditions.

EXAMPLE 8.3
Suppose one has a particle of mass M in a two-dimensional Coulomb potential
well:
Ze 2
V2(p) = --4- .
'!rEoP
Find the corresponding negative energies and wave functions.
116 Chapter 8
Since, as standard texts report, for the three-dimensional well
Ze 2
Va(r) = --47rEor
- ,

(2Kr)I+lexp(-Kr) [ K(n+l)! ]1/2


Unl(r) = (21+1)! n{n-I-l)! tFt{-n+l+li21+2i2Kr), (8.11)
and
Z2a 2 Me 2 ZMea e2
---::---, K = - - , a = - - , n=l, 2... , 1=0, 1, ...n-l,
2n 2 nn 47r€one
one can immediately write:
_ (2Kp)m+1/2 exp (-Kp) [K (n + m -1/2)!] 1/2
Wnm{P) - (2m)! n (n _ m _ 1/2)! X

(8.12)
1Ft (-n +m + ~i2m+ 1;2Kp) ,
where

Enm =
Z 2 a 2 Me 2
2n 2
K = ZMea,
nn n = 2"
1 3 1
2' ... , m = 0, 1, ...n - 2' .
Here n must be a positive half-integer so that -n + m + 1/2 is a negative integer
or zero, making 1Ft a terminating series. Otherwise wnm{p) blows up as p - 00.
Thus
, ....
9

EXAMPLE 8.4
Suppose one has a particle of mass M in a two-dimensional oscillator well:
1
V;(p) = 2'MW 2p2.
Find the corresponding energies and wave functions.
Since, as standard texts report, for the three-dimensional well:
1
Va ()
2 2
r = 2'Mw r ,

= {2r (n + 1 + 3/2) }t/2 r l+ exp (_r2 /2b2)


l

b3n! bl r(l + 3/2)

x 1Ft (-nil+~; ~),


(8.13)

Enl = (2n+l+~)nw, b=J);w' n=O, 1,2... ,1=0,1,2... ,


Two-, Three-, and N-, Versus One-Dimensional Problems 117
one can immediately write

wnm(p}-
_ {2 (n + m)!
b2 n.I
}1/2 pm+1/2 exp
bm
(_p2 /2b 2)
I IFI
(_. . p2)
n,m+1'b2'
m.
(8.14)

Enm = (2n+m+1)'liw, b=J:;w' n=O, 1, 2 ... , m=O, 1, 2 ....

EXAMPLE 8.5
Suppose one has a particle of mass M in a two-dimensional well:

0, 0 < p < a,
V=
00, a < p < 00.
Solve this problem.
To find the eigenvalues and eigenfunctions of this system one notes that for the
analogous three-dimensional case the standard result involves spherical Bessel and
spherical harmonic functions:

0, a <r < 00,

where k= J2ME
'li 2 '
with allowed energies corresponding to values of k such that jl(ka) = O.
For the two-dimensional case therefore

Wnm(p, </J) (X pl/2jn_l/2(kp) exp (±im</J),


with allowed energies when k is such that jl_l/2(ka) = O. One notes that

Hence these results may be written

Wnm(p,</J) = CnJn(kp)exp(±im</J},
with allowed energies when the cylindrical Bessel function In(ka) = O.
118 Chapter 8
EXAMPLE 8.6
Treat the finite square-well system in two dimensions:

-IVoI, 0< P < a,


v=
0, a < P< 00.

by analogy with the standard three-dimensional results if E < 0:

\}11m, int.(r,O,</» = Azjl(Kr)Y;',(O,</», 0 < r < a,

\}11m, ext.(r,O,</» = Blhfl)(ikr)Y;',(O,</», a<r < 00,

2m (IVoI-IEI)
K= 1i,2

By analogy

\}Inm int.(P,</» ex pl / 2 jn_I/2(Kp)exp(±im</», 0 < r < a,

But

Therefore

\}Inm ext.(P,</» = FnH~I)(ikp)exp(±im</», a < P< 00.

Matching the wave functions and derivatives at p = a gives one the appropriate
quantization condition for the energy.

The less frequently encountered, but all the same interesting, N-dimensional prob-
lem (N > 3) of a particle of mass m in a potential VN(r), which is a function only of
the radial distance
2
r = ( Xl + X 22 + X32 + ... X N2 )1/2 ,
(or of N one-dimensional particles in such a potential) can also be written l in a form
similar to Eq. (8.2), namely:

{ _~~ (T7 () 1i,2(l + N/2 -


2m dr 2 + VN r +
3/2)(l + N/2 -
2 2
mr
3/2 + I))} '/'nl-I. ( ) _
r -
E -I. (
nl'/'nl r ,
)

(8.15)
Two-, Tbree-, and N-, Versus One-Dimensional Problems 119
where
w(T)
cPnl(r)
= r(N_l)/2S1m e,
( )

Slm(e) are polynomials of degree 1 in e, and the index m runs over the set of values
A(N, I), where

{(21 + N - 2)(1 + N - 3)!} / {1!(N - 2)!}, 1 ;?: 1,


A(N,I) =
1, 1 = o.

The radial function cP in Eq. (8.15) is normalized according to

1000
cP!l(r)dr = 1. (8.16)

One notes that Eq. (8.15) reduces to Eqs. (8.1), (8.10) and (8.2) if N = 1, 2 and
3 respectively.

EXAMPLE 8.7
Suppose one has a particle of mass M in an N-dimensional Coulomb potential
well:
Ze 2
1;2(p) = --4-' { r = ( Xl2 + X 22 + X32 + ... XN
2 )1/2} •
1rfor
Find the corresponding wave function and energy levels.
Making a direct comparison of Eqs. (8.2), and (8.15) enables one to generalize
Eq. (8.11) (by making everywhere the identification 1-+ 1+ N/2 - 3/2):

(2Kr)I+N/2-1/2exp(-Kr) [Kr(n + 1+ N/2 _1/2)]1/2 X


(21 + N - 2)! nr(n -1- N /2 + 3/2)
1 F1 ( -n + 1 + N /2 - 1/2; 21 + N - 1; 2K r), (8.17)

where
K= ZMca, e2
0:=--,
nn 41rfonc
n = N/2 -1/2, N/2 + 1/2 ... , 1 = 0, 1, ... n - N/2 + 1/2.
Here increasing the dimension N by two units is equivalent to increasing 1 by one
unit in the wave function but not in the expression for the energy. As N increases
the minimum energy increases since
120 Chapter 8
EXAMPLE 8.8
Solve the N-dimensional Schrodinger equation for a particle of mass m subject to
a potential
1
V3 (r) = 2mw2r2.

Here one generalizes Eq. (8.13) (with the identification 1--+ 1 + N/2 - 3/2):

2r(n+Z+N/2)}1/2rl+N/2-1/2eXp(_r2/2b2) ( N r2)
tPnl(r) = { b3n! lJ+N/2-3/2r (1 + N/2) IFI -n; 1+ "2; b2 '
(8.18)

Enl = (2n + 1+ ~) nw, b= J!, n = 0, 1, 2... , 1 = 0, 1, 2....

One notes here that as N increases the minimum energy increases since

Eoo = (~) nw.


Also of interest to note is that increasing N by two units is equivalent to increasing
Z by one unit in both the energy and wave function expressions.

It is clear that the results in Examples 8.7 and 8.8 are consistent with the N =2,
3 results in Examples 8.3, 8.4. Increasing N effectively pushes the particle away
from the origin since it increases the numerator of the positive I/r 2 term on the
l.h.s. of Eq. (8.15). Finally one also notes that whereas the level spacing of the
N-dimensional harmonic oscillator in Example 8.8 is nw independent of N, the level
scheme for the N-dimensional Coulomb potential is more involved. In particular the
spacing between the ground and first excited state is :

Z 2a 2m2 { Z 2a 2 mc2 } Z 2 a 2mc2 N


2(N/2 + 1/2)2 - 2(N/2 -1/2)2 = 2 (N2/4 - 1/4)2 ,
i.e. for large N this goes as 1/N 3 •

EXAMPLE 8.9
Suppose a particle of mass M, and energy E, is in an N-dimensional potential:

0, 0 < r < a,
V=
00, a < r < 00.
Solve this problem.
Two-, Three-, and N-, Versus One-Dimensional Problems 121
By analogy with Example 8.5 one has

A lr 3!2-N!2jl+N!2_3!2(kr)Slm(e), °< r < a,


1l1 n1m (f) = (8.19)
0, a < r < 00.
where k = .j2ME/h2 .
As for the allowed energies, these are given by the requirement:

j/+N!2-3!2(ka) = 0, ( or Jl+ N!2-1(ka) = O. )


If N is odd one seeks the zero's of successively higher-order spherical Bessel func-
tions as N increases (i.e. if N=5 the lowest energy is obtained by requiringjl(ka) = 0,
as opposed to the requirement jo(ka) = 0 for the case N = 3).
If N is even one seeks the zero's of successively higher-order cylindrical Bessel
functions as N increases (i.e. if N =4 the lowest energy is obtained by requiring
J 1 (ka) = 0, as opposed to the requirement Jo(ka) = 0 for the case N=2).
This in turn implies that as N increases the minimum allowed energy also in-
creases since the smallest zero of the Bessel function (whether spherical or cylindrical)
increases as the order of the Bessel function increases.

EXAMPLE 8.10
Treat the finite square-well system in N dimensions by analogy with the standard
three-dmensional results given in Example 8.6.
By analogy with the results in Example 8.6 one obtains:

0< r < a,
(8.20)
I+N/2-3/2 (·k)S
B Ir3!2-N/2h(1)~· (t) ,
trim." a < r < 00.
where
K=
Matching 111, 111' at r = a gives the quantization condition on the energy, while
the normalization condition:

gives a further relation between At. and B l , hence one can also obtain the wave
function exactly with no undetermined parameters.
122 Chapter 8
EXAMPLE 8.11
Consider a particle in the potential well:
VCr) = -IVoI, 0< r < a,
fi2['(l' + N - 2)
2mr2
r > a, (8.21)

h
were r = ( Xl2 + X 22 + X32 + "'XN
2 )1/2

Find the allowed energies assuming E < 0 and that [ = O.
V( r)

1\2t 'CI,I + N-2 )


2ma2.

----~r-----~~a----------------------~-- r
-Ivol
Figure 8.2: Potential in Example 8.11
For the standard three-dimensional case,
Windr) = Ajo{I<r)Yo°{O, ¢), 0 < r < a,
Wext.(r) = Bh}!){ikr)lt{O, ¢), r > a,

~here I< = J2m (IVoI- IEI)jfi2 , k = J2mlEI/fi2 ,


I.e.

uindr) = Crjo{I<r) , 0< r < a,


uext.(r) = Drh}!)(ikr), r > a.
By analogy for general N :
Uint.{r) = Crj(N-3)/2(I{r) = Fr1 / 2JN/2_1(I<r), 0< r < a,
Uext.{r) = Drh}'~(N_3)/2(ikr) = Gr l / 2 H/\2N/2_I(ikr), r > a.
To find the allowed energies one requires continuity of u(r), u'(r) i.e.:
Caj(N-3)/2{I<a) = Dah/'+(N_3)/2{ika),
Two-, Three-, and N-, Versus One-Dimensional Problems 123
I.e.
(ajar) {rj(N-3)/2(Kr) }La (ajar) { rh n(N_3)/2(ikr) }La
j(N-3)/2(Ka) = h~n(N_3)/2(ika)
or
~ in {rj(N-3)/2(Kr) }L=a = ~ in {rh~n(N_3)/2(ikr) }L=a .

References

1. C. Miiller, Lecture Notes in Mathematics, Spherical Harmonics Springer (1966),


No 17.
CHAPTER 9

'Kramers' Type Expressions, The Virial Theorem,


and Generalizations

Consider a (bound) particle moving in a central three-dimensional potential V (r) =


ArP • The (radial) differential equation for unl(r) = r Rnl(r), (where the complete wave
function ¢(r, 0, ¢) = Rnl(r)Y;'/(O, ¢) ) is

d? {2m
dr2 unl(r) = fi2(V(r) - E) + 1(l+1)}
r2 unl(r), (9.1)

with boundary conditions that Unl(O) = 0 and unl(r) ---+ 0 as r ---+ 00.
Assuming Unl (r) is a real function one can readily show, integrating by parts
(where all integrals are from r = 0 to r = 00), that given a constant k,

- 2"k ( r 10-1) == -2"k JUnl ()r r 10-1


Unl (r) dr = JdUnl(r)
dr r Unl r 10 ()
dr, (9.2)

(provided rku~l(r) ---+ 0 both when r ---+ 0, and when r ---+ 00), while

JdUnl(r)
dr r
10 dUnl(r) d
dr r
-
-
k(k -1) (10-2)
2 r -
JUnl ()r r10 d?unl(r)
dr2
d
r

__ 2_
k +1
JdUnl(r)
dr
rkH d2u l(r) dr
dr2'
n

(provided (i) rkdu~l(r)Jdr---+O, (ii) krk-lu~l(r)---+O, and (iii) (r kH J(k+1»)


(du n l(r)Jdr)2 ---+ 0, when r ---+ 0 and when r ---+ 00, and that the integrals do not
diverge), i.e.

k(k -1) ( 10-2) _


2 r -
J{ () Unl r r
10 _ _ 2_dunl(r)
k + 1 dr r
kH} d2udr2l(r) dr.
n
(9.3)

Substituting Eq. (9.1) into Eq. (9.3) and using Eq. (9.2) one obtains after
regrouping terms:

k {k2 - (21 + 1)Z} (rk-2) _ 2m {2k + p + 2} (V(r) rio)


2(k+1) h,z k+l
(9.4)
4mE ( 10)
+~ r =0,
'Kramers' Type Expressions, The Vidal Theorem, ... 125
(provided additionally

If one studies in detail the restrictions under which Eq. (9.4) is valid, it is obviously
not valid if the constant k = -1. The constant k need not however be an integer.
If k = 0, Eq. (9.4) reduces to:

E=p+2(V(r)}. (9.5)
2
But E = (H) = (T) + (V). Thus for any quantum-mechanical state Inl} of H,

{nlITlnl} = ~ (n11VInl) , (9.6)

which is just the quantum-mechanical analogue of the classical Virial Theoreml :


- p-
T=2" V.

If k = 1,

-1(1 + 1) (r- I ) - 2m~; 4) (rV(r)} + 4;2E (r) = 0,


I.e.
E = (p + 4) (rV(r)} + ti 2 1(l + 1) (r- I ) 1m (9.7)
4(r) .
Equating Eq. (9.5) and (9.7) one obtains:

(rV(r) = 2(p + 2) (r) (V(r) _ ti 2/(l + 1) (~). (9.8)


p+4 m(p+4) r
If k = 2,
E =. (p + 6) (r 2 V(r) + Ti 2 {41(1 + 1) - 3} 12m
(9.9)
6 {r 2 )
and
/ r2V(r)) = 3(p + 2) / r2) (V(r) _ ti 2 {41(1 + 1) - 3} (9.10)
\ p +6 \ 2m(p + 6)
etc.
Eqs. (9.7) and (9.9) are natural generalizations of Eq. (9.5), while Eq. (9.8) and
Eq. (9.10) can be written as sum rules. Thus,
126 Chapter 9

L (n11V(r)ln'l') (n'l'l r2 In1) L (nllV(r)ln'l) (n'll r2 Inl)


n'I',pnl n',pn
(9.11)
1i {4l(l + 1) - 3}
2

2m(p+ 6)
(where l' = 1 since VCr) is a central potential) with similar expressions for higher
moments.
The above equations apply equally well to one-dimensional problems. For such
problems unl(r) becomes 'l/JN(X), the complete eigenfunction for the problem in ques-
tion, while 1must be set equal to zero. That 'l/JN( x) has different boundary conditions,
namely 'l/JN(X) ---+ 0, as x ---+ ± 00, merely implies the limits in the various integrals
are x from -00 to 00 in the one-dimensional case, rather than r from 0 to 00, and
that the conditions are for instance:

xk'I/J'fv(x) ---+ 0 as x ---+ ±oo etc.

Thus:

k(k - 1) ( k-2) _ 2m(2k + p + 2) ( kV() 4mE / k) _ 0 (9.12)


2 x 1i2(k + 1) x x + 1i 2 \x - .

EXAMPLE 9.1
Consider the one-dimensional harmonic oscillator: Vex) = mw 2x 2/2, (i.e. p = 2)
where EN = {N + 1/2} 1iw.
Obtain expectation values of various powers of x.
With the above substitutions in Eq. (9.12) one obtains:

(9.13)

Substituting k = 0, 2, 4 in this equation yields:

(Nlx4IN) = 8!~:2 { 4 (N + ~) + I}2 (9.14)

(Nl x6 IN ) = 12~:w3 {N+~}{6(N+~r +7.5}


etc.
'Kramers' Type Expressions, The Viria} Theorem, ... 127

EXAMPLES 9.2, 9.3


For two important three-dimensional problems:

V{r) =
Q1l,e
r
En = 1
--Q
2
2
me-
n2
2 1 ( where Q
e2
= - - '" -
41rE{)1l,e
1)
137 '

and

1
V(r) = _mw 2r2,
2
En! = (2n + 1+ 3/2)1l,w,
obtain recursion relations.
Substituting into Eq. (9.4) one gets for the first of these potentials, V(r) = Qnc/r,
Kramers' well-known formula:

k{P-(21+1)2} Irk-2) ~ (2k+1) I k-l)


4{k+1) \ + 1i, k+1 \r
(9.15)

while for the potential V(r) = mw 2r2/2:

k{P-(21+1)2} Irk-2) m 2w 2 {k+2}/rk+2)


4(k+1) \ + 1i,2 k+1 \
(9.16)

EXAMPLE 9.4
Show that if
Ax4 x> 0,
V(x) = { '
00, x < 0,
then
(nlxV(x)ln) = 1.5 (nlV{x)ln) (nlxln).
In this case °
< x < 00 in the integrals of Eq. (9.12) though this is a one-
dimensional problem.
Using Eq. (9.12) with k = 1,
p+4
-2- (xV(x)} = 2E (x).
128 Chapter 9
But from Eq. (9.5) E = {(p + 2)/2} {V}, (whether the system is one or three
dimensional);
therefore

(xV(x)} = 2(p + 2) (v(x)} {x}. (9.17)


p+4

If p = 4,
(xV(x)} = 1.5 (V(x)} (x).

EXAMPLE 9.5
Given the potential:
V(x) = ax\
show that:
(9.18)
Substituting directly into Eq. (9.10), with p = 4, 1=0,
31i 2
(r2V(r)) = 1.8 (r2) (V(r)} + 20m'
The same result applies to the one-dimensional case with r -+ x.
If V{x) = ax 4 for all x, the integrals extend over all x.
If
ax4 x> 0,
V(x) = { '
00, x < 0,
the integrals extend over x 2:: 0.

EXAMPLE 9.6
Given
ax, x> 0,
V(x) ={
00, x < 0,
show that:

= 5'6 (nlxln) 2 , (9.19)

i.e. that ~x! = (nlx2In) - {nlxln}2 = ~ (nlxln)2, and that:

9 31i 2
= "7 (nlx2In) (nix In} + 14ma . (9.20)
'Kramers' Type Expressions, Tbe Virial Tbeorem, ... 129
Substituting the value p = 1 in Eq. (9.17) ,
6
5' (ax) {x},
(xax) =
which yields Eq. (9.19), while substituting p = 1, I = 0, r -+ x in Eq. (9.10) one
obtains:

which yields Eq. (9.20).

EXAMPLE 9.7
If V(x) = alxl, show that:

lo
oo 18100 4>n(x)x 4>n(x) dx 100 <Pn(x)x<Pn(x) dx + -
4>n(x)x 3 4>n(X) dx = -7
0
2
0
31i- .
28am
2

(9.21 )
Substituting p = 1, I = 0 in Eq. (9.10), and taking into account that <Pn(X)2 =
4>;,(-x) since <Pn(X) has a definite parity (because V(x) = V(-x) in this case), one
immediately obtains Eq. (9.21).

EXAMPLE 9.8
Obtain expressions for Xrms = J{Nlx2IN) , if p = 1, 2, i.e. V(x) = A1x, or
V(x) = A2X 2.
If p = 1, one has from Eq. (9.17):

(Nlx2IN) = 5!r (NIAIXIN)2 ,


I.e.

#s ! .
Therefore
xrms = (9.22)

If p = 2, one has from Eq. (9.14):

( Nlx2IN) = ~ (N +~)
mw 2
= _2
mw 2 2
nw (N +~)
2
E
2 (mw 2 /2) .

Therefore
(9.23)
130
Chapter 9
where (NlxIN ) = 0 for both of the above potenti als.

EXAM PLE 9.9


Show that if p = 2,

En! = {41(l + 1) - 3} 1i 2 (nllr- 4Inl)


4m (n1Ir-2Inl) (9.24)

Substit uting k = -2 into Eq. (9.4) immed iately yields E. (9.24).


Note this result
is indepe ndent of A, the constan t of V(r). Since E > 0, and
both (nllr-4 Inl) ,
(n1Ir-2Inl) are positive, this implies 41(1 + 1) > 3, i.e. 12: 1 for the
integra ls in Eq.
(9.24) to be convergent.

EXAM PLES 9.10, 9.11


Show

(9.25)
if p = 1,
and

(9.26)
if p= -l.
Substit uting k = -3/2 into Eq. (9.4) yields Eq. (9.25), while the
substit ution of
k = -1/2 into Eq. (9.4) yields Eq. (9.26). Both these results are
indepe ndent of A.
EXAM PLE 9.12
Obtain the two-dimensional analogu e of Eq. (9.4).
From the results of Chapte r 8 one makes the identifi cations m
-+ M (for the
mass), r -+ p, and 1 -+ m - 1/2. The result is:

k(P- 4 m2 )/ k_2)_ 2M{2 k+ P +2}/ V() k)


2(k+1 ) \p 1i 2 k+1 \ pp
(9.27)
4ME /
+--,;? \p
k) = 0,
where, assumi ng Wnm (p ) is real,

Ifk=O
E= P;2 (V(p)) ,
'Kramers' Type Expressions, The Viria] Theorem, ... 131
while if k = 1
E = (p + 4) (pV(p») + -n? (m 2 - 1/4) (p-I) /M .
4 (p)
EXAMPLE 9.13
Obtain Kramers' relation for

V(p) = Mw 2 p2/2, E = (2n + m + l)liw.


Substituting in Eq. (9.27):

k (k2 - 4m 2) / k-2)
4(k+1) \p
_ M 2w2 {k +
li 2 k+1
2} \p/ k+2)
(9.28)

+
2Mw(2n+m+1) /
Ii \p
k)-O
- .

With k = 0:
(9.29)
while if k =2:

(9.30)

etc.

EXAMPLE 9.14
Obtain the analogue of Eq. (9.4) if

V(r) = Vt(r) + V2 (r),


where
Vt(r) = AIr P\ V2 (r) = A 2 r P2 •
Carrying through this calculation as in the derivation of Eq. (9.4) one obtains for
this case:

(9.31 )

If PI = P2 as expected this reduces to Eq. (9.4).


132 Cbapter 9
If k = 0, Eq. (9.31) yields:

E = (PI + 2) (Vi(r)} + (P2 + 2) (l!2(r)} ,


2
I.e.
(T) = PI (Vi) + Pa (l!2) .
2

The interesting and useful thing here is that using Kramers' techniques one obtains
the values of, or relations between various matrix elements, not only without having
to evaluate these matrix elements, but without even having to obtain the associated
wave functions.

References

1. H. Goldstein, Classical Mechanics, Addison-Wesley (1950), p. 69.


CHAPTER 10

Upper Bounds and Parity Considerations

Consider a system with a Hamiltonian H such that

(10.1)
If .,pt( a, (3, ... ) is a normalized 'trial' wave function with parameters a, (3, ... , one
can define the integral

E(a,{3, ... ) == J .,p;H.,pt dr. (10.2)

Since the Wn's constitute a complete set one can expand .,pt(a, (3, ... ) in terms of
the Wn's in Eq. (10.1) which are assumed normalized in what follows. Thus
00

.,pt(a,{3, ... ) = LCnWn, (10.3)


n=O

where the fact that .,pt and Wn are normalized implies

(10.4)
n=O

Substituting Eq. (10.3) in Eq. (10.2) yields

E(a,{3, ... ) = JEc:W:HEcmWmdx


n=O m=O
00 00

= L \cn1 2 En ~ Eo L Icnl2 = Eo,


n=O n=O

using Eq. (10.4).


Hence
E(a, (3, ... ) ~ Eo. (10.5)
This result enables one to use any trial wave function and in addition optimize,
i.e. choose parameters a, {3 etc. which minimize E( a, (3, ... ) for that particular trial
wave function by requiring that BE(a, {3, ... )/Ba = 0, etc. In this way one gets upper
bounds to the ground state of any quantum mechanical system.
The only restriction on .,pt(a, (3, ... ) is that it obeys the same boundary conditions
as the eigenfunctions of H. Otherwise the assumption Eq. (10.3) is not valid. One
134 Chapter 10
also of course assumes the quantities one works with are well enough behaved that
one can interchange summations and integrations in the expression for E( 0'., (3, ... ).

EXAMPLE 10.1
Consider the system H = T +V where

V X> 0,
(10.6)
00, x < o.
The exact ground-state wave function for this system and the corresponding ground-
state energy are in fact known:

( 1 ) 1/2 2x (X2 )
b.,fo b exp - 2b2 '

(10.7)
n,2
Eground exact = v'2.25 mb2 •

--------------~----------------~--x

Figure 10.1: Potential in Example 10.l.


Consider the trial wave function

(20'.)2n+1
'if;t= xn exp (-O'.x),
~-..:.....,-:--
(2n)!

where nand 0'. are parameters, and n > o.


Performing the integral in Eq. (10.2) one obtains
Upper Bounds and Parity Considerations 135

The condition
oE O. l' 4 (4n2-1)(n+1)
ocr = Imp Ies cr = 2b4 '

while oE/on = 0 implies 4cr4 b4 = (4n + 3)(2n -1)2,


I.e.

1 (2n + l)(n + 1) 17,2 1 (2n + l)(n + 1) 1i 2


E(n)
= 2 2(2n - 1) mb + 2'
2 2(2n - 1) mb2
(10.8)
(2n+1)(n+1) n?
=
2(2n - 1) mb2 '

with optimal n = (1 + V6 )/2.


If n = 1, E(l) = J3 n,2/m b2. If n = 1.5 or 2, E(1.5) = E(2) = v'2.5 n,2/mb 2 ,
and if one chooses the optimal n = (1 + V6 )/2 one obtains E {(1 + V6)/2} =
v'2.47 n,2/mb2, which is quite close but slightly larger than the exact result Eq. (10.7)
(as it must be).

EXAMPLE 10.2
Consider the system H = T + V where
Ax, x> 0, (A> 0)
Vex) = {
00, x < O.
V( x)

--------------~-------------------x
Figure 10.2: Potential in Example 10.2.
136 Chapter 10

If one uses the trial wave function

where b is a free parameter,

E(b) = 31i 2 + 2Ab . (10.9)


4mb2 .,fi
The requirement 8Ej8b = 0 implies

This yields

Consider the trial wave function

(2a)2n+I n
(2n)! x exp(-ax)

where n and a are arbitrary non-negative parameters.


This function, when substituted in Eq. (10.2) yields

1i2a2 (2n + l)A


Ea,
(n) =
(2n -1)2m
+ -'--~--'--
2a
The condition
3 (4n2-1)Am
8E = 0 implies a = 2 ,
8a 21i
while the condition
3 Am(2n -1)2
8E = 0 implies a = 1i 2
8n
Substituting in E(a,n) yields
Upper Bounds and Parity Considerations 137

= ~ (2(2n + 1)2)1/3 (1i2A2) 1/3 + ~ (2(2n + 1)2)1/3 (1i2A2) 1/3


E(n)
4 (2n-1) m 2 (2n-1) m
(10.10)

= (
27(2n + 1)2)1/3 (1i2A2) 1/3 with optimal n= 1.5.
32(2n -1) m '

If
( 1i2 2) 1/3 (1i2 A2) 1/3
n = 1, E(I) = (1.5)5/3: '" 1.97 --:;;;-

If

n=2, E(2) = (~225) 1/3 (1i::2) 1/3 '" 1.92 (1i::2) 1/3
Finally if one chooses the optimal n = 1.5

It is interesting to note that even with the best two-parameter trial wave function
'I/J~2) one does not do as well as with the one-parameter trial wave function 'I/J~I). In
other words increasing the parameters in one's wave function does not necessarily
result in lower energies, i.e. better results.
Consider the trial wave function

2 )1/2 ( 2)
'I/J~3\ x) = ( b.,fi exp - ;b2

with b as a free parameter. For this trial wave function,

1i 2 Ab
E(b) =- b + v7r
4m 2
,-'
The constraint
2.,fi
BE
fib = o·Impl'les b3 = 1i2mA .
Substituting this value of b into E(b) implies

27 )1/3 (A 1i
E(bopt'1mal) = ( -167r -m
2 2) 1/3
'" 0.81
(A- 1i2) 1/3
2

m
138 Chapter 10
This is much less than 1.86 (A 2h2 /mf/:
the best value obtained with the other
two trial wave functions! But this wave function does not satisfy the boundary con-
ditions and hence is unacceptable. In particular it is not zero at x = O. Hence this
particular result is wrong!
The lowest energy for large arguments of the relevant Airy function is El = 1.84 (A2h2/mt
(see Eq. (3.18) ).

EXAMPLE 10.3
Consider the system H = T +V where

Vex) = Alxl (A> 0) for all x.


If P, the so called 'parity' operator is such that when operating on any function
f(x)
P f(x) = fe-x),

Then
PH(x)tfJ(x) = H(-x)tfJ(-x) = H(-x)PtfJ(x) = H(x)PtfJ(x)
for this case since for the V of this example Vex) = V( -x),
V (X)

-----------------&--------------------x
Figure 10.3: Potential in Example 10.3.

while quite generally


T(x) = T(-x).

Hence in this case H(x) = H( -x). Thus for this problem (indeed whenever Vex) =
V(-x) )
[PH(x) - H(x)P]tfJ(x) = 0, I.e. [P, H] = O.
Upper Bounds and Parity Considerations 139
But this has consequences on the eigenvalues and eigenfunctions of H. Assuming
[H, Pj = 0 consider the Schrodinger equation Eq. (10.1) :

Premultiplying by P

Assuming the system is non-degenerate (see Chapter 12), if Wn is an eigenfunction


of H and at the same time PW n is an eigenfunction of H with the same eigenvalue
En, this implies

(10.11)
i.e. PWn(x) is proportional to Wn(x).
But Eq. (10.11) is just the eigenvalue equation for P. Thus whenever [H, Pj = 0,
the eigenfunctions of the Hamiltonian are also eigenfunctions of P!
Consider now the eigenvalue problem for P. If Po are the eigenvalues of P,

Pf(x) = Pof(x)
= f(-x) (10.12)
P 2 f(x) = p~f(x) = f(x).

Hence Po = ± 1.
Thus in Eq. (10.11) A = ± 1, and

Thus whenever, as is the case in this problem, V(x) = V( -x) and the system is
nondegenerate, the eigenfunctions of H have either even parity or odd parity.
Going through the derivation for upper bounds Eqs. (10.1) - (10.5), one sees that
in this case Eq. (10.5) applies independently to the odd- and even-parity solutions
since Eq. (10.3) will be an expansion either in terms of the even or the odd eigen-
functions 'lin. Hence working with odd-parity trial wave functions one gets an upper
bound to the lowest odd-parity state energy and similarly for even-parity trial wave
functions an upper bound to the lowest even-parity state energy.
A simple (though inadequate, since its derivative is not continuous at x = 0)
even-parity trial wave function for Example 10.3 is:

J3/2b3 (b - Ixl) , Ixl < b,


'1Pt(x) = {
0, Ixl > b,
140 Chapter 10
with free parameter b, yielding,

The condition
8E+ O· Ii b3 = 121i2
7ib = Imp es Am
(where one uses the representation of the delta function

d?
dx21xl = 28(x) ).

For optimal b,

~ (3A 21i 2) 1/3 + ~ (3A21i2)1/3


4 2m 2 2m
~ (A 21i2) 1/3 '" (A 21i2) 1/3
= 122/3 m 0.86 m

A better even-parity trial wave function one can use is

~t(x) = Jb~ exp (-;;2)'


with free parameter b, yielding

If
8E+ b3 = 1i2../i .
7ib = 0, 2mA
For optimal b,

= ~ (A21i2) 1/3 + (A 21i2 ) 1/3


E;round exact :::; E+(boptimal) 2 21Tm 21Tm

= (~:) 1/3 ( A~2) 1/3 '" 0.81 ( A~2) 1/3

An odd-parity trial wave function one can use is

~t(x) = ( bv'i
2 ) 1/2 ~ exp (2
- ;b
) 2 '
Upper Bounds and Parity Considerations 141
with free parameter b, yielding

If

and for this optimal b :

= (3h 2A2)1/3 +2 (3h 2A2) 1/3


41rm 41rm

= (!~r/3 (h~2r/3 ~ 1.86 (h~2r/3


One thus has roughly determined two energy levels with the help of parity con-
siderations in this case. One notes
E- (boptimal ) = 1.86 ~ 2.30,
E+ (boptimal) 0.81
as opposed to
E- ground exact = 1.84 rv 2.07
E+ground exact 0.89
(c£. Eqs. (3.18), (3.31) ).

EXAMPLE 10.4
Consider the system H = p2/2m - Vo6{x), Vo > O.

V( x)

Figure 10.4: Potential in Example 10.4.


142 Chapter 10

Estimate the energy of this system using as trial wave function

'ljJt(X) = ( mw)1/4
h7r exp
(mwx2)
-21i
(the ground-state wave function of the one-dimensional harmonic oscillator), with
free parameter w. One obtains directly

hw (mw)1/2
E(w)=--Va -
4 h7r
Requiring
oE(w) = 0 implies ~ _ Va (~)1/2 _1__
ow 4 2 h7r W1/2 - 0,
I.e.
4mVo2
Woptimal = 7rh3 .
Hence for this trial wave function,

mVo2 _ (~)1/2 (mVi) 1/2


7rh2 Vo h7r 2 7rh3

m Vo2 2mVo2 m Vo2


= 7rh - 7rh 2
2 =- h 27r '

vs. the exact ground (and only) state and only energy of this system (cf. Eq. (13.8»:
mV,2
Eground exact -- - 2h2 .°
One again notes
mv,2
E (Woptimal) = -0.318 h20 > Eground exact = -0.5 t/ '
mv,2

as required by Eq. (10.5).

EXAMPLE 10.5
Consider the system H = p2/2m + V (x)
where

mw2x2/2 + a/x 2 , x 2: 0, (where a is a positive or negative constant),


Vex) = {
00, x < o.
Estimate the ground-state energy of this system using as trial wave function:
Upper Bounds and Parity Considerations 143

(mw')3/4 2x (mw'x 2)
(1)
'l/;t (x) -
_
T '/r1/4 exp --v;- .
(This is the lowest eigenfunction of H' = p2/2m + V'(x) where
x 2:: 0,
V'(x) = {
00, x < 0,
with eigenvalue 3nw' /2.)
Using this trial wave function,

E(1)(w') = ~nw' + 2mw'a + (w2


2 n W,2
_ 1) ~nw'
4

3 /{ 8ma} 3 nw2
4"nw 1 + 3n2 + 4"7 .

[This result can easily be obtained by rewriting

x + -xa2 + ( W,2
1 2 1
p2
H = -2m + -mw
2
,2 2
-w - )
1 -mw x
2
/2 2

and noting that


1 o
00 01.(1)*( )
'f/t
X
a
x ""2
01.(1)( )
'f/t X
d _ 2mw'a .
X - -~- ,

Jo[00 ol't(l)*(X)
'f/
~mw'2
2 x2 ol't(l)(X)
'f/
dx = ~
2 (~nw')
2 c. Eq. (9 .5) ) . 1
(f
From the constraint

one obtains
, w
w optimal = V1 + 8ma / 31£
~2 •

Hence

E(l) (w' optimal) 3 nwV1


-4 + 8ma/3n2 + ~1iwV1
1
+ 8ma/3n?
= ~nwV1 + 8ma/31i 2 .
If a = 0, E(1) (W~Ptimal) = 3nw/2, as expected for this trial wave function.
144 Chapter 10
If a = n /m, the Hamiltonian above admits of an exact solution since n2/mr2 =
2
2n2/2mr2 is then just the centripetal term in the three-dimensional simple harmonic
oscillator (V(r) = rnw 2r2/2) Hamiltonian, for angular momentum quantum number
1 = 1. The exact energies in this case are (2n + 1 + 3/2) nw, n = 0, 1, 2... i.e. the
lowest exact solution is Eground = 5hw/2.
In this case

E (1) ( Woptimal
I ).J33
= -2- i:
nW 'V 2.873 nw > 2.5 nw,

as expected.
One may use instead:

(2,8)2n+l n
(2n)! x exp (-,8x),

as trial wave function.


For this value of a,

For optimal ,8

R • = (~)1/2 (2n -1)n(2n + l)(n + 1))1/4


,voptImal n 2( n + 4) ,
one obtains

E(2) (~. n) = nw (2n + l)(n + l)(n + 4)


optImal> 2n(2n - 1)

which has values 2.556 nw and 2.535 fiw for n = 3 and 4 respectively, while for optimal
noptimal 3.74,
'V

E(2) (,8optimal, noptimal) 'V 2.533 nw > 2.5 fiw.

As expected one again gets an energy greater than the exact energy, but ¢F) (x) is
seen to give a better upper bound than ¢(l)(x).
If a = 3n2/m, 6fi 2 /m, 10fi 2/m etc. again one knows the exact solution which
one can compare with the upper-bound results that arise in these cases.
Thus for a = 3fi 2 /m, and 1 = 2, Eexact ground = nw/2, while

E(l) (W'optimal) = 4.5 nw > 3.5 fiw.


Upper Bounds and Parity Considerations 145
For general a

,82 (n1i? + 4am) mw2(2n + l)(n + 1)


2m(2n - l)n + 4,82 '

8E(2) (,8, n) ,8 (nfi2 + 4am) mw2(2n + l)(n + 1)


=
8,8 m(2n - l)n 2,83

and
a . _ Jmw [(2n + 1)(2n -l)(n + 1)nfi 2] 1/4
t"Optlmal - 1i 2(n1i2 + 4am)
Thus

~ {(2n + l)(n + 1)(n1i2 + 4am) }1/2


2 2n(2n - 1)
~ {(2n + l)(n + 1)(n1i2 + 4am) }1/2
+ 2 2n(2n - 1) ,
= w {(2n + l)(n + 1)(nfi2 + 4am) }1/2
2n(2n - 1)

For a = 0 this reduces to Eq. (10.8)

E(2) (a. ) _ fi {(n + 1)(2n + 1) }1/2


t"optlmal, n - w 2(2n -1)

For a = 3fi 2/m

E(2) (,8. ) _ fi {(n + 1)(2n + l)(n + 12) }1/2


optlmal,n - w 2n(2n -1)

For this value of a, noptimal rv 5.74 and

E(2) (,8optimal, noptimal) rv 3.52 fiw.

If
~~3\X) = C~f/2 (m;'r/4x2exp (- m;~x2)
is used as trial wave function,

E(3)( ')
w
= (!..- 2am) fiw'
12 + 3fi2 + ~4 fiw
2
Wi ,
146 Chapter 10
with
, w
Woptimal = J7/15 + 8exm/15h 2
Substituting this value of w',

If
E(3) (W'optimal) = 5/2 hw = Eexact ground,

which implies that ~P)(x) is the exact ground-state wave function for this value of
ex, while if

_ 3h 2
a - m' E
(3) ( ' .
WoptImal - 2
) _ ~J155
3 hw
rv ~_
3.59 hw > 2 hw - E ground exact·

Thus ~~3\ x) gives a better upper estimate for Eground than ~?) (x) while the best
upper limit for this case is given by ~}2).

EXAMPLE 10.6
Consider the system H = p2/2m + V( x), where
-ahe/x, x 2': 0,
V(x) = {
=, x < 0,
with ground-state eigenvalue of E ground exact = -mc 2 a 2 /2.

Estimate the ground-state energy of this system using as trial wave function:

~)
~t (x) -
_ 2 1
b ( byf1r )
1/2 x exp - 2b2
(2 ) X

Performing the integral in Eq. (10.2) one obtains:

Imposing the condition

8E(1) (b) .. b . _ 3hyf1r


8b = 0, ImplIes optImal - 4mea .
Upper Bounds and Parity Considerations 147
Hence

4
= -_mc2a 2 '" -0.424 mc2a 2
37r
> Eground exact'

as it must be. This result is quite close to the exact ground-state energy. Other trial
wave functions do not do as well.
Thus, if instead one uses:

(2) _ 2 J2
2 (X2 )
'!f;t (x) - b2V~ x exp -2b2 '

one obtains

71i2 41ica
E(b) = 12mb2 - 3b..,fi .

Imposing the condition

8E(2)(b) b . _ Th..,fi
8b = 0, implies optimal - 8mca .

Hence
16 32
= _mc2a 2 - _mc2 a 2
2b 2b
16
= -217rmc2a2 '" -0.243 mc2a 2.

If one uses:

one obtains

Imposing the condition

8E(3) (b) .. 3amc


8b = 0, ImplIes boptimal = ~ .
148 Chapter 10
Hence
3 3
= _mc2 a 2 _ -mc2 a 2
8 4

3
= _gmc2a 2 = -0.375 mc2a 2•

Though both E(2) (boptimal) and E(3)(bopt imal) are above Eground exact,
neither is as close to it as is E(1)(boptimal)'

A fourth trial wave function:

.1,(4)( ) _
'f't X -
V2
b2 x
3/2 (x2 )
exp - 2b2 '

yields

and imposing the condition

8E(4) b . _ 51i
--m;- = 0, results in optImal - 2macV; .

Hence for this trial wave function:

EXAMPLE 10.7
Consider the Hamiltonian:

(all x).

Given the trial wave function:

tPt(X,w') = 1
7r 1 / 4
(mw')
T 1/4
exp
(mw'x
----y;-
2)
,
evaluate E (w') and find the minimum upper bound to the ground-state energy with
this trial wave function for the case a = 1.
Upper Bounds and Parity Considerations 149
Evaluating I.,p; H.,pt dt, one obtains:
liw' liw 2 3aliw3
E(w') = 4" + 4w' + 4W,2 ,

hence
8E(w') Ii liw 2 6anw 3
---::a:-'-w""'"-"- = '4 - 4W,2 - 4w,3
The condition 8E(w')/8w' = 0 yields the result:
W,3w'
----6a=0.
w3 w
Thus if a = 1, w' = 2w, and E(2w) = 13/16 nw = 0.8125 nw.
This can be compared to the exact answer (obtained by numerical methods) in
this easel =0.803771... liw .
If one substitutes w' = w then E(w) = liw/2 + 3aliw/4, which is the zero (nw/2),
plus first order (3anw/4) perturbation contribution if one writes H = Ho + V with
Ho = p2/2m + mw 2x 2/2, V = aliw (Jmw
/Ii x) 4, a decomposition which for a = 1
results in a much worse upper bound E(w) = 1.25 liw, and a diverging perturbation
series 2 (for all values of a).

EXAMPLE 10.8
Consider the Hamiltonian (cf. Example 10.7):

H= : : +~mw2x2+anw(J~w X)4 (all x).

Use the trial wave function:

J3/21? (b -Ix!), Ixl < b,


.,pt(x, b) = {
0, Ixl > b,
to evaluate E(b). Find the minimum energy for this trial wave function in the cases
a = 0, 1.
By a straightforward calculation (cf. Example 10.3)

3li2 mw2~ am 2w3 b4


E(b)
= 2m~ + ----w-+ 351i
8E(b) 3li2 mw 2b 4am 2w3 fi3
8b - ml? + 10 + 35ft
The optimization requirement is :

8au6 + 7u 4 - 210 = 0, where u= J~w b.


150 Chapter 10

If a=O, t h e con d ·,
ItIon ---a;;-
8E(b) = O'ImpIies:

b = (30)1/4 J mw
Ii rv 2.340 Jmw
Ii .

Hence

E (boptirnal) = f[; liw rv 0.55 liw VS. 0.5 Tiw, the exact result.

If a = 1, boptirnal rv 1.645 Jm~' hence E (boptirnal) rv 0.899liw.

Similarly

E(b = 2.0 Jm~) rv 1.032 liw,

E(b = 1.7 J!) rv 0.902 liw,

E(b = 1.4 J~w) 0.973 hw, rv

as compared to 0.803 hw, the exact (numerical) ground state energy.

EXAMPLE 10.9
Consider the Hamiltonian :

(all x).

Given the trial wave functions:

'IjI~1) (x,w') = ( mw')


1i1r
1/4 (mw'x 2 )
exp -----v;- ,
Ixl < b,

Ixl > b,
find E(l)(w',a), E(2)(b,a) and the minimum values for E(1), E(2) for arbitrary a.
Upper Bounds and Parity Considerations 151
By a straightforward calculation (d. Example 10.7) one obtains:

nw' nw 2 a15nw 4
= 4 + 4w' + 8W,3
aE(1) (w', a)
ow' 4" - 4W,2 - 8W,4

-
The minimization condition is thus:

2 (~) 4 _ 2 (~r 45a = 0, I.e.


,
Woptimal = w
J1 + VI + 90a
2 .

For example if a = 8/9, then W'optimal = J5 w, and


11 v~
E (1) ( w, optimab 8 /9) = 30 5 nw ~ 0.820 nw.

Similarly

and
OE(2)(b a)
ab' = 0, implies 5au8 + 7u 4 - 210 =0 where u
Jmw
= b T'
I.e.
uoptimal
. = Jmwn (-7 + V49lOa+ 4200a )1/4
If a = 8/9, boptimal rv 1.57 Jmw/n, and E(2) (boptimab 8/9) rv 0.910 nw. This
indicates 'I/'~1) is a better trial function than 'l/'F).

EXAMPLE 10.10

Consider H = p2/2m + Alxl n , -00 < x< 00.

Evaluate Emin for 'l/'t(x, w) = (mw/n7r// 4 exp (_ m;nx2)

(d. Eq. (5.2) where the analogous result is obtained using the Uncertainty Principle.)
One obtains:

nw (n + 1) A (~)n/2
+ _1 r
E(w, n)
4 Vi 2 mw
oE(w, n) ~ _ ~r (n + 1) A (~)n/2 (~)n/2+1
ow 4 2Vi 2 m w
152
Cbapter 10
Hence
. _ (2nr ([n + 1]/2) An(n-2)/2) 2/(n+2) and
WoptImal - ~ mn/2 '

~ (nnnr ([n + 1]/2) A) 2/(n+2) + ~ (nnnr ([n + 1]/2) A) 2/(n+2)


E (Woptimal, n) 2 ~ (2m)n/2 n ~ (2m)n/2

(10.13)
= (nnr ([n + 1] /2) A( n + 2)(n+2)/2) 2/(n+2)
~ (nm)n/22n+1

n2 A) 1/2
while if n = 2, E (Woptimab 2) = ( 2m = Egroun d exact, etc.

EXAM PLE 10.11

Consid er H = p2/2m + Alxl n, -ex; <x< ex;.

J3/2b3 (b - Ixl), Ixl < b,


Evalua te Emin for '¢t(x, b) = {
0, Ixl > b.
One obtains

2 6Abn
E(b, n) -3n- + -;---;--~--:-:--;-
2mb2 (n + l)(n + 2)(n + 3)
----::-;-

8E(b, n) 3n 2 6nAbn- 1
8b --m-b3 + (n + l)(n + 2)(n + 3)
and
b . _ ((n+1 )(n+2 )(n+3 )ti 2 )1/(n+2)
optImal - 2nAm
Thus

3(n+2)/2 Atinn ] 2/(n+2)


E(boptimal, n) [
(2m)n/2 (n + l)(n + 2)(n + 3)
2 [ 3(n+2)/2 Atinn ] 2/(n+2)
+ ;;- (2m)n /2(n+ l)(n+2 )(n+3 )
Upper Bounds and Parity Considerations 153
= [3(n+2)/2 A1in(n + 2t/ 2 ]2/(n+2)
(2nm)n/2(n + 1)(n + 3)

E(boptimab 1)

E( boptimal, 2)

vs. Eground exact(n = 2)


etc.

EXAMPLE 10.12
x < 0,
Given H = T + V(x), V(x) ={ 00,
Axn , x> o.
Find E(b, n) and E(boptimab n) if 'l/Jt(x, b) = 2/b (1/bVi)1/2xexp(-x2/2b2).
One obtains

E(b, n) 31i 2 2Abn f (n +


4mb2 + Vi 2 '
3)
8E(b, n) 3h 2 2nAbn- 1 (n +
-2mb3 + Vi f -2-
3)
8b
The requirement 8E(b, n)/8b = 0 implies
3h2 Vi ) 1/(n+2)
(
boptimal = 4nmAf([n + 3] /2)

Thus
3n/ 2Ahnnf([n + 3] /2)] 2/(n+2)
E(boptimab n) [
m n / 2 2 n Vi

+ ~ [3 n/ 2Ahnnr([n + 3] /2)] 2/(n+2)


n m n / 2 2n Vi

[
3n/ 2Ahn(n + 2)(n+2)/2f([n + 3] /2)] 2/(n+2)
(nm )n/22n Vi

E (boptimab 1) G~r/3 (A~2r/3 ,


154 Chapter 10

Ah2)1/2
E(boptimal, 2) = Eground exact(n = 2) (
3 2m '

etc.

EXAMPLE 10.13

find E(boptimal) where

.jlib cos (7rx/2b) , -b < x < b,


'¢;t(b, x) = {
0, Ixl > b.
By a straightforward calculation one obtains:

oE(b)
fib
Hence

and

( 7r2/68 _1)1/2 (7r 2/6 _1)1/2 (7r2/6 _1)1/2


E (boptimal) = hw + 8 hw = 2 hw '" 0.568 hw.

EXAMPLE 10.14
Consider the system (cf. Example lOA)

Estimate the energy of this system using as trial wave function:

J3/2b3 (b -Ix!), Ixl < b,


'¢;t(x, b) = {
0, Ixl > b.
Upper Bounds and Parity Considerations 155
One obtains:
311. 2 3Vo
E(b) = 2mb2 - 2b'
8E(b) 311. 2 3Vo
fib = - mfi3 + 2b2 .

This implies boptimal = 211.2 1m Yo.


Hence
E(b . ) - 3mVo2 _ 3mVo2 __ 3mVo2 _ -0375 mVo2
optImal - 811. 2 411. 2 - 8h 2 - . h2'
This is to be compared with Eground exact = -0.5 mVo2 /ti?,
and E (Woptimal) = -0.318 m Vo2 /h 2 in Example 10.4.

EXAMPLE 10.15
Consider the system (d. Example 10.3)
p2
H = 2m + Vex), Vex) = Alxl , for all x.

Estimate the energy of this system using as trial wave function:

f1ib cos (1rX/2b) , Ixl < b,


¢t(x, b) = {
0, Ixl > b.
By a straightforward calculation one obtains:
h2 7r 2 Ab
E(b) = 8mb2 + 27r 2 [7r 2 - 4] .
Requiring 8E(b)/8b = 0 implies

io2 4 )1/3
b- ( n7r
- 2Am (7r - 4)
2

Hence

This result can be compared to the results 0.81 ( A2'/i2/m ) 1/3 and 0.86 ( A 2h2/m )1/3
of Example 10.3.
156 Chapter 10

One notes in passing3 that for all the Hamiltonians discussed in this Chapter of
the form
H=T+Axn,
the Virial Theorem (cf. Eq. (9.6) )

(T) = (n/2) (V),

is also satisfied if

etc. provided b is chosen to minimize

This situation arises in Examples 10.1-4, 6, 10-15.

References

1. F. T. Hioe and E. W. Montroll, J. Math. Phys. 16 1975, p. 1945.

2. C. M. Bender and T. T. Wu, Phys. Rev. 184 1969, p. 1231.

3. H. A. Mavromatis, to be published.
CHAPTER 11

Perturbation Theory

Consider a system whose Hamiltonian H contains a 'perturbation' V such that

H= Ho+ V. (11.1 )
If one knows the eigenfunctions (cPn) and eigenvalues (En) of H o, i.e.

(11.2)
and wishes to find the eigenfunctions (1/Jn) and eigenvalus (En) of the Schrodinger
equation:

(11.3)
where 1/Jn reduces to cPn as V --+ 0, one can formally expand each eigenfunction 1/J in
terms of the complete set cP :

(11.4)
m=O

where the cP's are assumed orthonormal, i.e.

One can then rewrite Eq. (11.3) as


00

L (H - E)amcPm = 0, (11.5)
m=O

where for notational simplicity the subscripts n have been omitted from the 1/J's and
E's.
Premultiplying Eq. (11.5) by cP; and integrating over the relevant variables one
obtains:
00

L (Hpm - E Opm) am = 0, p = 0, 1, ... , (11.6)


m=O

where
158 Chapter 11
and
Vpm == J4>;V4>m dr = (4)plVl4>m). (11. 7)

Writing Eq. (11.6) in matrix form one has:

Hoo-E

Vio H11-E Vi2 =0.


"V:!o

The eigenvalues E of Eq. (11.6) are the exact energies of the system Eq. (11.3)
and the corresponding coefficients am when substituted into Eq. (11.4) give the
corresponding eigenfunctions.
Consider for simplicity the special case when Eq. (11.6) is a 2 x 2 matrix. The
determinant of this matrix must be zero i.e.

E2 - E (Hoo + H11 ) - VOl Vio + HooH11 = o. (11.8)


Hence

(Hoo + H11 ) +_ (Hoo - H 11 ) ( 1 + 4VOl Vio/{Hoo - H11 )2 )1/2


E =
2

(11.9)
{Hoo + H11 ):!: (Hoo - H 11 ) (1 + 2V(n Vio/(Hoo - H11)2 + ...)
- 2

Keeping only the first two terms in this expansion results in the following expansions:

Eo
IT
flOO + HooVOl-VioHu = Eo + VOO + -
TT
- -V(n-Vio- - -
fO + Voo - fl - ViI
V, volVio
= Eo + 00 + (fO - fl)( 1 + (Voo - Vid/(fo - fl) ) ,
(11.10)
H 11 + Vio VOl = f1
Vio VOl
+ V;11 + ---::-:--"-~':---::-::,-
Hu - Hoo f1 + ViI - fO - VOo
= fl Viovol
+ V;11 + ..,..--~--,----,-----:---:-:---~
(fl - fO) { 1 + (ViI - VOO)/(fl - fo) )
Upon expanding the denominators in these expressions one obtains:
Perturbation Theory 159

VOl Vlo Y(ll Vll Vlo VOl Vlo


Eo = fo +
TT
voo + -- + (
fo - f1 Eo - f1
)2 -
TT
Voo (
fO - f1
)2 + ... ,
(11.11)
VlOY(ll
El = fl + Vll + -
TT
- + VlOVoOY(ll2 - TT
vll
VlOVo1
2 + ....
f f1- O (f1-fO) (f1- fO)
From the 2 X 2 matrix one can also obtain expressions for tP by solving for the
eigenvectors ao and aI,

= { E - Hoo } { E - Hoo }
tPo ao <Po + VOl <PI -+ <Po + VOl <PI ,
(11.12)

tPl = al {E-VloHll <Po + <PI } {E -VIOHll <Po + <PI } ,


-+

if in addition one requires {<P011f10} = {<P111f1l} = 1, i.e. that the 1f1'S are 'cross-
normalized' functions.
Substituting Eo and El of Eq. (11.11) into Eq. (11.12) yields:

(11.13)
'" <Po VOl <Po Voo Vln <Po VOl Vll
1f1l = '1'1+--+
fl - fo (fl - fO)
2- (fl - fo) 2+····
One can generalize the results in Eq. (11.11) as follows:

= fn + Vnn + L
00 v.nm v.mn +
(11.14)
mIn En - fm

{
~
L..J
VnmVmpVpn _ V.
nn L..J
~ VnmVmn
2
}
+ ... ,
m,#n (En - Em) (En - fp) mIn (fn - fm)
which is the standard Rayleigh-Schrodinger expansion for the exact energy to third
order in V. The first term on the r.h.s. of Eq. (11.14) is called zero-order. The term
on the r.h.s. of Eq. (11.14) involving V once (the second term) is called first order.
The term on the r.h.s. involving V twice (the third term) is called second order, and
terms involving V thrice (the fourth and fifth terms) are called third order, etc.
160 Chapter 11
One can generalize the results in Eq. (11.13) as follows:

(11.15)

which is the standard Rayleigh-Schrodinger expression for the exact wave function
to second-order in perturbation theory when one has cross-normalized functions 'ljJn
i.e. (¢nl'ljJn) = O. Fourth and higher-order terms and third and higher-order terms
in Eqs. (11.11) and (11.13) respectively may be obtained by keeping more terms (in
expanding the radical) in Eq. (11.9). The usefulness of Eqs. (11.14) and (11.15) in
turn depends among other things on whether the expansions converge.

EXAMPLE 11.1
Consider a particle subject to the Hamiltonian:

Ixl ::; a,
Ixl > a.
Potential Energy V ( x )

_ _ _ _ _ _ _ _ _ _ _ _- L _ _~~~---L---------- __ x
-a a

Figure 11.1: Potential in Example 11.1.


One can rewrite this Hamiltonian as:

H=Ho+V,
where
Perturbation Theory 161
Ixl < a,
Ixl > a.
To first order in V one then has

where En, <Pn are the standard eigenvalues and (real) eigenfunctions of the one-
dimensional infinite harmonic oscillator.
This problem illustrates one important limitation of perturbation theory. It pre-
dicts here an infinite, discrete set of energies En. However, if E > mw 2 a2 /2 the
particle is no longer bound, but rather is free i.e. can have any energy!
In fact only if E « mw 2 a2 /2 (i.e. (n + 1/2) « mwa 2 /2n) does one expect
to get reasonably accurate results using perturbation theory.
Classically for this system it makes no difference what the potential is for x greater
than Xmax where Xmax = J2E/mw 2 • The quantum-mechanical treatment of this
problem, however, shows that the potential for x > Xmax also affects the particle.
If a --t 00 the exact energies of the system go to En. For finite a however, the exact
energies En are less than En since Vnn < O. Roughly speaking this is because there is
more likelyhood the particle will be in the region x > a for the potential V = mw 2 a 2 /2
than for the more repulsive potential mw 2 x 2 /2. On the other hand the more the
particle spreads, the bigger its wavelength A, and smaller its energy since k '" 1/)",
while E '" k 2 •

EXAMPLE 11.2
Consider a particle in the potential

Ax, x> 0,
v(x) = {
00, x < O.
Suppose one wishes to know the ground-state energy of a particle in this potential.
The Hamiltonian of the system is

H = T+Ax, x> O.
But one can rewrite H as
162 Chapter 11
where

(x> 0),

and b is a parameter.
With this decomposition ofthe Hamiltonian one can treat (Ax - { Ii? /2mb4 } x 2 )

(x > 0), as the perturbation V.


Moreover, for this choice of Ho,

and

The first-order energy contribution is:

One particular choice for b makes the first-order energy contribution zero, namely

b3 = b3 = 3h2v:;r
1 - 8Am .

vex)

--------------~~~-----------------x

Figure 11.2: Potential in Example 11.2.

With this choice for b, to first order


Perturbation Theory 163

(1i2A2) 1/3 (1i2A2) 1/3


Eo = ( -24)1/3
11"
-m- I'V 1.97 - -
m
The second-order energy contribution is:

f: VanVnO = f: (0IAx-1i2x2j2mb4In)2(nIAx-1i2x2j2mb410),
(11.16)
n~O Eo - En n=1, 2... -2nn /mb 2

smce
En = (2n +"23) mb1i 2
2 •

One can evaluate Eq. (11.16) using

etc.
Considering the first (n = 1) term in Eq. (11.16) one has

( 0 lAX - ::~: 11) = ~ ~ - ~ 2~b2 = - ~ ~ .


With the above choice for b the first term of Eq. (11.16) is:

2A2b2/311" = _~ (~)1/3 (1i?A2) 1/3 = -0.06 (n2A2) 1/3 (11.17)


-21i2/mb2 8 811" m m

All additional second-order terms are also negative, hence add to the magnitude
of the second-order result, but should be small compared to Eq. (11.17). Indeed the
next second-order term (n = 2) is:

1\0 lAx _1i2x2j2mb412)12 l-bAjJ3Q1r - 01 2 =


-41i 2jmb2 -41i 2jmb2

= 1
__ (~)1/3 (A 21i2) 1/3 = -0.0015 (A21i2) 1/3
320 811" m m

which is considerably smaller than Eq. (11.17).


Hence to second order
164 Chapter 11
which can be compared with two variational approach results for this problem, (Ex-
ample 10.2):

With the choice b = ~ the third-order energy contribution comes from:

f (0 lAx - 1i 2x 2j 2mb4 1 n) (n lAx - 1i2x2j2mb4Ip) (p lAx - 1i2x2j2mb41 0)


n,p;loO (-2n1i2jmb 2) (-2p1i2jmb2) ,
(11.18)
since this choice for b makes the second third-order term zero (see Eq. (11.14) ).
Taking only the term In) = Ip) = II} in Eq. (11.18) one obtains an approximate
third-order result:

1(0IAx-1i2x2j2mb411)1\IIAx-1i2x2j2mb411) = (2b 2A2j31r) (3bAj.,fi-71i 2j4mb2)


41i4jm 2b4 41i4 jm 2b4
A 2 1i 2 ) 1/3
= -0.02 ( - -
m
which is smaller than the second-order correction.
Thus to this approximation

If one instead uses


b3 = b~ == (3v'K 1i 2) j4mA
which minimizes the energy to first order (see Example 10.2), one obtains

Eo = 1.86 (1i2 A2 jm f/3


to first order and no contribution in second and third order from the state n = 1.
The exact result (valid for large arguments of the relevant Airy function (d. Eq.
(3.18) ) is:

E = G) 2/3 ( A~2) 1/3 C~) 2/3 = (8~) 2/3 ( A~2) 1/3 ~ 1.84 ( A~2) 1/3

(3:::)
With the first choice for b

{ namely the choice b1 = 1/3 which makes Voo = o} ,


Perturbation Theory 165
the exact ground-state wave function to second order in perturbation theory becomes:

(11.19)
'" cPo (bt) + (0.153 + 0.048) cPt (bt) = cPo (bt) + 0.201 cPt (bt) ,
if one includes only the first excited state in Eq. (11.15).

_~ ________-==-t_ X
b

Figure 11.3: Wave function component cPo(b) in Example 11.2.

~1 (b)

~-~-~--+----~~~X

Figure 11.4: Wave function component cPt(b) in Example 11.2.


With the second choice for b,

{~ = c:::) t/3 which minimizes the energy to first order, i.e. makes 8Hoo /8b = o} ,
166 Chapter 11
the matrix element VCn is zero so there is no contribution to the exact ground-state
wave function from f/J1 for this choice of b :
(11.20)
to second-order in V.
The Eqs. (1.19) and (11.20) are less dissimilar than they look since they involve
different b's with b1 '" O. 794 ~. Thus though f/Jo (bd of Eq. (11.19) is more compressed
than f/Jo (b 2 ) of Eq. (11.20), this is compensated for by the small admixture of f/J1 (b 1 )
in the former expression as can be seen by considering Figures 11.3 and 11.4.

EXAMPLE 11.3
Consider a particle subject to the Hamiltonian:

0::; p ::; a, where p = ..jx2 + y2,


(11.21 )
p > a.
To find the ground-state energy of this sytem using perturbation theory one can
write
1
H = T + '2mw2l + V(p) = Ho + V(p),
where the perturbation

o ::; p ::; a,
V= {O,
-mw2p2J2, p> a.

Potential Energy

~_----L- _ _ _ _;

a
Figure 11.5: Potential in Example 11.3.
The ground-state wave function for Ho is (see Eq. (8.14)):

( p2)
Woo(p) = ( b2)1/2
2 p1/2 exp - 2b2 '
Perturbation Tbeory 167
with b = Jh/mw, and EO = hw.
To first order in V one then has

Eo = fiw - -21
b2 a
00
(1-mw
2
2p2)
fiw - fi; { _ exp ( _ p2 / b2) (1 + ~:) [ } (11.22)

fiw _ fi; {1 + m~a2} exp ( _ m~a2)

which is expected to be accurate if fiw < < mw 2a2.


In second order the first state which contributes to the energy is WlO(P) at energy
3hw yielding the contribution:

(laOO WlO(P) (-mw 2p2/2) woo(p) dp)2 (_ {fiw/2} {exp (_p2 /b 2) (1 + p2/b2 + p4/b4) I:;"}) 2
-2fiw -2fiw

Hence approximately:

(11.23)

Obviously this analysis is incorrect if E > mw 2 a2 /2 for reasons similar to those


mentioned in Example 11.1.
168 Chapter 11
EXAMPLE 11.4
Consider a particle subject to the Hamiltonian:

H = T + Alxl P (all x) . (11.24)


This may be written:

where
Alxl P - n2 x2
__
2mlfl
can be treated as the perturbation V. With this decomposition the ground state
energy of this sytem can be immediately calculated using first-order perturbation
theory, with

Ho = T + 2!2b4 x 2, Eo = 2:b2 and <Po(x) = (b~ r!2 exp (- ;;2) .


If p is odd one has

(oIAlxl P- __ 2
2mb4
n2 x I)
0 = -AbP
Vi
(P---1) ! - -n-
2
2 .
2 4mb
The choice
11'+2 = n2 Vi
4mA([p -11/2)!
makes the first-order energy zero, and yields

E_ 2(2-p)!(2+p) {(P - 1), A ;"P }2/(P+2)


(11.25)
o- 1l"l/(p+2) 2 ' mP/2

If P is even one has

lolA p- ~ mw2 21 0) = AbP(p-l)!! _ ~


\ x 2 x 2P/2 4mb2 .
The choice
;,,22P/ 2- 2
11'+2 = _-,-_....,..,.,.
mA(p-l)!!
makes the first-order energy zero and yields:

Eo =
1
2(2p-2)/(p+2)
( ;"P
(p - 1)!1 m P/ 2
A) 2/p+2
(11.26)
Perturbation Theory 169
These results may be combined:

E
o
= 1
2(71- 2)/(71+ 2)
(r ([P + 11/2) liP A) 2/(71+ 2)
7r 1 / 2 m 71 / 2 for arbitrary p. (11.27)

These expressions may be compared with a qualitatively similar result, namely


Eq. (5.2). If instead one chooses b so that the ground state energy will be a minimum
to first order (cf. Chapter 10) i.e.

(0 IHo + VI 0) = (0 IT + AlxlPI 0) = (OIHIO)


is a minimum, for odd p this implies

bP+ 2 _ "h 2 ..,fi


- 2mAp {(p - 1)/2}! '

while for even p

li 2271 / 2- 1
bfJ+2 = .
mAp(p-1)!!
For p odd in this case

p+2 (Ali 71 {(p _1)/2}!)2/(71+2)


Eo '" 7r 1 /(p+2) 2(271+2)/(71+2) (pm )71/2 '

while for p even

p +2 (Ahp(P _1)!!)2/(71+2)
Eo '" 2(371+2)/(71+2) (pm )71/2
These results may be combined yielding (d. Eq. (10.13) ):

p+2 (Ahpr ([P + 111 2)) 2/(71+2)


Eo = 2(271+2)/(71+2) (pm )71/2 7r1/2 (11.28)

for arbitrary p. For p = 2 Eqs. (11.27) and (11.28) both yield fiw/2 which as expected
is the right result (where A = mw 2 /2).

EXAMPLE 11.5
Consider the Hamiltonian

H=Ho+V(x,y),
where
170 Chapter 11
while
V(x, y) = >..mw 2xy .
Suppose one wishes to find the (non-degenerate) ground-state energy of this system.
The unperturbed ground-state energy of this system is nw with corresponding
wave function <Po(x)<Po(Y) (cf. Eq. (12.4) ).
The first-order correction to this energy is zero since

>..mw2 JJ<p~(x) <p~(y) xy <Po(x)<Po(Y) dx dy = O.

The only term which contributes in second order because of the nature of this par-
ticular V is <Pl(X)<Pl(y) at energy 3nw.
Hence the second-order contribution to the energy is

(>..mw 2)21{<po(x)<po(y) xy <Pl(X)<Pl{y))1 2 = _ 4>..4\1 ",\mw 2x 2\ '" )\2 = _ >..2nw


-2nw 2nw \'1'0 2 '1'0 8'
in agreement with the exact result (see Example 12.9 ),

>..2nw
Eo = 21 nw { ~ + v'1'+1" } = nw - -8- + .... (11.29)
The exact normalized wave function for this state (see Example 12.9) is
'l/Jo = <Po (X, w = WI) <Po(Y, w = W2),
where

while
mw)I/4
<po{e,w) = ( 1i1r exp
(mwe)
--u '
1.e .

• 1. (
'1'0
)
X, Y
= (mw) 1/2
n1r
(1 _ A\2) l/S exp (_ mwvr=-x
21i
X2)
exp
(_ mwv'fD Y2)
21i '
or

(11.30)

x exp ( -
mw{ Jf='141i+ v'f+1' )y2)
exp
(mw( vT='1'21i- v'fD )x y ) .
Perturbation Theory 171
If ,\ « 1 one can write Eq. (11.30) as

tPo(x, y),..., ( 1 -"8 (mw x + y2)) exp (mwxy,\)


,\2) <Po(x,w}<po(y,w} exp 16h'\2 (2 - 2h
(11.31 )
>..mwxy >..2 { mw (2 2)
<Po(x,w)<Po(y,w) ( 1- 2h +"'8 -1 + 2h x + y

On the other hand, to second order, perturbation theory yields for the wave
function:

tPo(x, y)
(11.32)

smce

(<P2(X)<PO(Y) IVI <P1(X)<P1(Y)} = >"hw


.,fi' (<P2(X)<P2(Y) IVI <P1(X)<P1(Y)) = >..hw,
and where N = 1 if one assumes tPo is a cross-normalized function i.e.
(<Po(x)<Po(Y) ItPo(x,y}) = 1,

while N ,..., (1 + >..2/16 + ... )-1/2 if one assumes (as in the case of Eq. (11.30) ) that

(tPo(x,y) ItPo(x,y)} = 1.
The Eqs. (11.31) and (11.32) are identical to order >..2 if one substitutes
N,..., 1 - >..2/32 in Eq. (11.32), since

<Pl(X) = V~
T-h- x <Po(x) and <P2(X) 1 (
= -.,fi 1-
2mw x2) <Po(x).
-h-
172 Chapter 11
EXAMPLE 11.6
Consider the Hamiltonian

p2
H= -+v(x)
2m
where

Voc5(x), -a ~ x ~ a,
vex) = {
00, Ixl > a.
To obtain the exact result one writes the Schrodinger equation as follows:

(Ho - E) tjJ(x) = -Voc5(x)tjJ(x),

where
p2
Ho = 2m' Vex) = Voc5(x) and x ~ 14
Thus
1
tjJ(x) = A {Voc5(x)tjJ(x)} , (11.33)
E - p2/2m
where the p2 operates on the term in the brackets { } .
vc x)

Figure 11.6: Potential in Example 11.6.


Using closure (d. Eq. (2.6) ) one may insert the set In) of eigenfunctions of the
potential:

0, -a ~ x ~ a,
V= {
00, Ixl > a.
Perturbation Theory 173
between
1
and {l/oo(x)~(x)}
E - p2/2m
in Eq. (11.33) (where the odd eigenfunctions of this set do not contribute to
(n Ivoo(x)1 ~(x)) ) . One then obtains:

~(x) = - f: In} (nll/oo(/x) I~(x)} = l/o~(O)


E - n 2 h 2 32ma 2
f:
4>n(X)4>/~(O) 2 . (11.34)
E - n 2 h 2 32ma
n=1,3,... n=l,3, ...

Substituting the value x = 0 in Eq. (11.34) one obtains:

~(O) = VO~(O) ~ <Pn(O)<p~(O) ,


n=B, . . E - n 2 h2 /32ma 2
I.e.

SInce
n = 1, 3, 5... ,

I.e.

h2 00 1
32 voma = nodd
fT L f3 - n 2 (11.35)

where
f3 = 32ma
2E
h2 •

If one plots the two sides of this equation one obtains Figure 11.7, where the
intersections correspond to acceptable values of E.

I
{3-n2 h2
::?;
1,3, ... 32 '{,rna
Vo > 0

16 20 25
P
h2
32Vo ma Vo < 0

Figure 11.7: Exact energies in Example 11.6, graphical solution of Eq. (11.35).
174 Chapter 11
If j3 = q2 + 8, (where q = odd integer)

h2 1 8 ,...., 32VQma
32VQma ,...., S ' h2 '

I.e.
j3 _ 2 32VQma
- q + h2 '

or
h 2q2 VQ Va
E=--+ - = f . q +-.
2 32ma a a
A more accurate calculation yields:

h2 1 odd 1 1 odd 1 odd 1


32Voma = L:
S + n'i-q q2 _ n2 L:
+ 8 ,...., S+ n'i-q q2 _ n2 - 8 L:
(2
n'i-q q - n
2)2'

I.e.

implying

where
odd 1 h2 odd 1
Cl = L (2
n'i-q q - n 2)2'
C2=
32Vroma
- ~
L..J q2 - n2
n'i-q

Thus

Since

(11.36)
Perturbation Theory 175
which is just the perturbation result for Eq (q=odd) to third order since if r, s are
odd (<Pr IVI <p$) = Vo/a for this particular V and set of eigenfunctions of H o. For the
complimentary set of zero-order eigenfunctions,

1 . qrrx
<pq(x) = Va sm 2a' q = 2, 4, ... ,

there is no correction to the unperturbed energy Eq from this particular perturbation.


As regards the exact wave function, from Eq. (11.34):

But if (<pq 11jI(x)) = 1, (i.e. the wave function is 'cross-normalized') this implies:

Vo1jl(O) 1 _ 1
Va E - q2h2 /32ma 2 - .

With this constraint:

q2h2 ) odd <Pn(x)


1jI(x) = <pq(x) + (E - 32ma2 ~ E _ n 2h2 /32ma 2 • (11.37)

Substituting Eq. (11.36) into Eq. (11.37),

Vo{
Vo odd
1jI(x) = <pq(x)+- 1 + - L 1
+ ... L
} odd <Pn(x)
,
a a mi-q Eq - Em ni-q (Eq - En) {I + Vo/(a [Eq - EnD + ... }
I.e.

1jI( X)

(11.38)
VO)2 (VO)2
+ (-a
odd <Pn(x) odd <Pn(x)
L
n,mi-q (Eq - Em) (Eq - En)
- -a L
ni-q (Eq - En)
2+···'

which is just the perturbation result for 1jIq to second order in V.


If q is even

1jIq(x) = <pq(x) (where <pq(x) = ~-a sm. -


qrrx
2a
, q = 2, 4, ... ) ,
176 Chapter 11
i.e. the perturbation also does not affect the unperturbed wave functions of the
system in this case. The reason for this is because for q even, <Pq(O) = 0, i.e. the
system does not feel the presence of V when it is in these states.
Explicitly, to first order, the wave function described by Eq. (11.38) is

.,p(x) = _1_ cos q7rX + Yo ~ ( 1/ va) cos (mrx/2a) (q, n odd).


va 2a a ~ (q2 - n 2 )(h2 /32ma 2 )
(11.39)

For this particular Hamiltonian one can also obtain the exact eigenfunctions of
this system in a form which unlike Eq. (11.37) does not involve infinite sums. This
can be done by writing

.,p(x) = A sin k(a - x), 0 < x < a,

.,p(x) = B sin k(a + x), -a < x < O.


This combination of eigenfunctions of Ho satisfies the requirement that
.,p(a) = .,p(-a) = O.
Additionally continuity of .,p( x) at x = 0 implies:

A sin ka = B sin ka i.e. A = B.

However, the derivative of .,p(x) has a discontinuity at x = 0, where the condition


that must be satisfied is:

d.,p
dx I - d.,p
dx I = 2m Yo ·/'(0)
;,,2 '1/
(f
C.
E q. (6.10 )) .
x=€ X=-E

'II (x)

----~----------~--------~----~x
-0 o

Figure 11.8: Two posssible even eigenfunctions of Example 11.6.


Perturbation Tbeory 177

Figure 11.9: Plot of transcendental Eq. (11.40).


This implies
- Ak cos ka - Ak cos ka
2mVo A sm
= --,;,r- . ka,
or

1i 2 k
tanka = - - - . (11.40)
mVo
Thus the even-parity solutions 1jJ( x) are given by:

1jJ(x) = Asink(a -lxI), -a ~ x ~ a,


where k is obtained by solving Eq. (11.40).
Normalization of 1jJ(x) additionally requires

IAI 2ja
-a sin2 k(a -Ixl) dx = 1,

or

1jJ(x) = sin k(a -Ixl) , (11.41)


Ja {I - (sin2ka) /2ka}
which is illustrated in Figure 11.8.
Plotting tan ka and -1i 2 ka/mVoa of Eq. (11.40) vs. ka (see Figure 11.9) shows
that (for q odd) ka = q7r/2 + 8 where 8 is small.
Thus tan ka = - cot 8 = -1i2k/mVo or tan 8 '" 8 = mVo/1i2k,
and
178 Chapter 11
i.e. E = fl, 2 k 2 /2m "" E.q + Vo/a (cf. Eq. (11.36) ).
Hence one can expand the exact wave functions Eq. (11.41) in a Taylor series about
ko = q7r/2a where k - ko = 2mVo/fI,2q7r, and q is odd.
Thus (to within an unimportant overall phase)

(11.42)
(Vo/a){(l - lxi/a) (q7r /4ya) sin (q7rlxl/2a) - (1/4ya ) cos (q7rX/2a)}]
+ q2 h 2/32ma 2 '

as opposed to Eq. (11.39).


By comparing the terms in Vo/a of Eqs. (11.39) and (11.42) one incidentally
obtains the relationship:

(1 -lxi/a) (q7r/4) sin (q7rlxl/2a)

- (1/4) cos (q7rx/2a)

or
(11.43)
(q/2) ((7r/2) - u) sinqu

- (1/4) cos qu

where u = 7rlxl/2a, u:::; 7r/2.

EXAMPLE 11.7
Consider the Hamiltonian:

= Ho+V,
where
Ho = T + ~mw2x2 and V = Voexp ( _ :2).
Perturbation Theory 179
v

Vo if Vo>O

Figure 11.10: Perturbation in Example 11.7.

To first order the ground-state energy of this system is:

Vc 100
Eo
nw b.,fir
= 2+ -00 exp (X2) (X2)
- b2 exp --;- dx since ¢o(x) = ( b.,fi
1 ) 1/2 (X2 )
exp - 2b2 '

where b = J'fi/mw .
Thus,

Eo
hw Yo
2 + bft {1/b2+ 1/fV/2
100
-00 exp (_x 2{b21 + ~1 }) dx { b21 + ~1 }1/2 ,
I.e. (11.44)
fiw Va
= -+----:-=
2 {1+b2/f}1/2'

One notes that if Yo - VtJl/7rf and f - 0, this reduces to the correction to


the unperturbed energy introduced by a delta-function potential VtS(x) since one
representation of the delta function is

Sex) = 1 I¥
r.;lim - exp - -
v7r <-+0 f f
(x2) (see Eq. (2.15) ).

EXAMPLE 11.8
Consider the Hamiltonian
p2 1 afiw ( )
H = - + -mw x +
2 2
x ~ 0 .
2m 2 (Jmw/fi xf
180 Chapter 11
Treating

as a perturbation, study the energy of the ground state of this system.


The unperturbed ground-state energy Eo = 3nw/2. The first-order energy E~l) ==
Voo is:
(00 1
anw 10 <p~(x) 2 <Po(x) dx = 2anw,
o (J111MJ/n x)
where, defining u == Jmw/n x :

<Po{x) = (m;) 1/4 )/4 uexp ( _ ~2),


<PI (x) (~wr/4 .j62'Jr 1/ 4 u (2u 2 _ 3) exp ( _ ~2),
<P2(X) (~w) 1/4 y'301'Jr1/ 4 U (4u4 _ 20u2 + 15) exp ( _ ~2),

<P3(X) = (~w) 1/4 ~ 'Jrl/4 U (8u 6 _ 84u 4 + 210u 2 -105) exp ( _ ~2).

The second-order energy E~2) == L~lo YOm VmO/(cso - fm) is:

E(2) _ I < 01Y11 > 12 I < 01Y12 > 12 I < 01Y13 > 12
o - -2nw + -4nw + -61iw + ...

This is a slowly converging series. The exact result in this case can be obtained by
examining the three-dimensional simple harmonic oscillator system (cf. Eq. (8.13) ).
In detail:

Eground exact = (1 + ~)
2 hw '" (32" + 2a - 4a 2 3+...) hw,
+ 16a
where the series expansion converges for a < 1/8 ). Thus the second-order series
implies:
Perturbation Theory 181

EXAMPLE 11.9

Solve the Schrodinger equation for a particle which is confined to a two-dimensional


square box whose sides have length L and are oriented along the x and y coordinates
with one corner at the origin.
(a) Find the eigenvalues and eigenfunctions.
(b) If a small perturbation

V(x,y) = L2 Vi8 (x -~) 8 (y -~)


is introduced, using first-order perturbation theory find the change in the ground
state energy.
(c) Compare this with the ground-state energy obtained using the trial wave function
(d. Chapter 10):
?jJ(x,yhrial = Nx(L - x)y(L - y) .
By straightforward methods one obtains:

Thus if
EO=Eo+Voo,

h2~2 h2~2
,/.. ,/.. ( ) 2. ~x . ~y
EO = En = 2mL2 { 12 + 12} = mL2 ' 'f'0 = 'f'00 L sm L
x, Y = Ism ' (11.45)

and

4 2
Voo = -LVi
L2
ii
00
L L . 2 -~x sm
sm
L
~y ( x - -
. 2 -8
L
L)
4
8 (y - -L)
4
dx dy = Vi.

Combining terms:
n2~2
Eo = mL2 + Vi. (11.46)

Normalizing the trial wave function in (c):

<?jJtriar! ?jJtrial) = 1 implies N = ~~ .

Hence:
182 Chapter 11

Thus
(11.47)

Comparison with Eq. (11.46) shows this is larger than to + Voo, the result to first
order in perturbation theory.

EXAMPLE 11.10
Consider the Hamiltonian:

where
Yrel. (Xl - X2) = ~mwi (Xl - X2)2 •

Find the exact energy of this sytem and the ground-state energy to second order
in perturbation theory, assuming V = Vrel. (Xl - X2)'
To obtain the exact energy one recasts the Hamiltonian in a more transparent
form:

where
= + X2
M == 2m, X C.M. -
Xl
2 ' /1 == m/2, xrel. == Xl - X2'

l.e.
H = ( P22M 1
C.M. + -Mw 2X 2 ) + (p2reI. + +_11
1 {w 2 + w2} X2 )
2 0 C.M. 2/1 2'- 0 I reI. .

Hence:

N, n = 0, 1. .. (11.48)

1 1 WI 2 1/2 1 1 { 1 WI 2 1 WI 4 }
-nwo + -nwo { 1 + (-) } "-'-nwo + -nwo 1 + - (-) - - (-) + ...
2 2 Wo 2 2 2 Wo 8 Wo

nwo 1 (w2)
+ -n .2. 1
- -n (W4)
-{ + .... (11.49)
4 Wo 16 Wo
Perturbation Theory 183
In the perturbation theory treatment of this problem one writes

To first order one thus has:

From parity considerations the term XIX2 is seen not to contribute to this integral.
Thus

Eo = nwo + 2~5 {~nwo + ~nwo} = nwo + ~n (::) (11.50)

The second-order contribution:

:E 1(001 mwi (Xl - X2)2 /41 nln2)1 2


nl, n2 - (nl + n2) nwo

=
1(00 I mwixi/4 120)1 2 + 1(00 I mwix~/4 102)1 2 + .:...:.......-'--.!.-~:..:..-~~
1(00 I mwi2xlx2/4 111)1 2
-2nwo -2nwo - 2nwo
-~n (wt)
16 w5
Thus to second order:

1 (W2)
Eo = nwo + -n -1.. - 1 (w4)
-n ~ , (11.51)
4 Wo 16 Wo

and as expected this expression agrees with the first three terms of Eq. (11.49).

EXAMPLE 11.11
Consider the Hamiltonian:
p2
H = 2m +v +V = Ho +V

Ixl < a,
v = { 0,
(Xl, Ixl > a,
184 Chapter 11
v = { (Va/f) [I-Ixl/f], Ixl < f, f < a, (Va < 0) ,

0, Ixl > f.
Find the ground state of this system to second order in perturbation theory. Examine
the limit f ---+ o. (Note that: lim<-+o (I/f) (1 -lxi/f) = o(x) cf. Eq. (4.22) (vi) ).

One notes (cf. Example 11.6), that the ~ zero-order functions and energies
are:
1 mrx n 2 h2 h2
rPn(x) = va cos 2;;' En = 32ma2 ' n = 1, 3, 5 ..., I.e. fl = --.
32ma 2
Also the first-order contribution:

V;11 = -2Val« 1- - X) cos2rrxd


- x= -Va{I- + I-cosrrf/a} <-+0---+ Va
af 0 f 2a a 2 (rrf/a)2 a

Figure 11.11: Potential in Example 11.11.

As to the second-order contribution to the energy, the odd zero-order functions:


(1/va) sinmrrx/2a, m = 2,4 ... do not contribute from parity considerations. Thus:

where

VIm = 2Va
fa
r (1 _ .:) (cos rrx2a cos mrrx)
Jo f 2a
dx

= Va [1 - cosrr(m -1)f~2a +1 - cosrr(m +1)f~2a] <-+0---+ Va [~+~] = Vo .


a (rr(m - 1)f/2a) (rr(m + I)f/2a) a 2 2 a
Perturbation Theory 185
Thus the second-order term is:

32mV02 [01: ({I - cos 7r(m - I)E~2a}


h2 m=3,5... {7r(m-l)E/2a}

+{1- cos7r(m + I)E~2a})2 (_1_)].


{7r(m + I)E/2a} m2 - 1

As E --+ 0,

Hence, to second order, as E --+ 0, these expressions reduce to the results of


Eq. (11.36).
CHAPTER 12

Degeneracy

One-dimensional systems
The Schrodinger eigenvalue equation, Eq. (12.1) for the simplest possible case,
namely a one-dimensional system:

h2 d2 )
{ - 2M dx2 + V(x) 1/;(x) = E1/;(x) (12.1)

involves a second-order differential equation. Hence there should in general be two


solutions 1/;1, 1/;2, for each energy E. Such a two-fold 'degeneracy', as it is called, does
not usually arise however. It is removed by the boundary conditions.

!~~ol'
Consider for instance a particle in the well:

x < 0,

V(x) = 0:::; x:::; a,


0, x> a,
illustrated in Figure 12.1, where E < O.
In the region 0:::; x :::; a one has two solutions:

1/;(x) = A sin Kx or BcosKx, K=


2M {IVoI - lEI}
h2
while in the region a < x < 00:

1/;(x)=Cexp(-kx) or Dexp(kx), k-
-
J2MIEI
h2 •

However, the wave function in this case satisfies the boundary condition that it van-
ishes at x = 0 where the potential is infinitely repulsive, as well as at infinity since
the particle is localized in the well. Hence

B cos K x and D exp (kx )

must be discarded. One thus has only ~ acceptable solution which, together with
its derivative, must be made continuous for all x > 0, in particular at x = a. This in
turn imposes a quantization condition on the energy namely:
Degeneracy 187

K
tanKa = --.
k

vex)

______________~~-----ar_-----------4~X

Figure 12.1: A one-dimensional potential producing non-degenerate eigenfunctions.

Two-dimensional systems
Consider the case where V = V(p), p2 = X 2+y2. The radial Schrodinger eigenvalue
equation for this system is (d. Eq. (8.10)):

;,,2 d2
{ - 2M ;,,2(m2 -1/4)}
dp2 + V(p) + 2Mp2 wnm(p) = Enmwnm(P), (12.2)

and the total wave function of the system is:

( A.) _ wnm(p) exp (±im¢»


Xn±m P'l' - pIt2 .,.fii

Independent of the details of Wnm (p) (unless m = 0) there are here at least two
normalized solutions for each energy eigenvalue, namely:

Wnm{p) exp (im¢» Wnm(p) exp (-im¢»


and (12.3)
pIt2 .,.fii pl/2 .,.fii
where
10 00
w!m(P) dp = 1,
because Eq. (12.2) depends on m 2 rather than m. One way this m degeneracy can
be removed is by inserting a term in the interaction which depends on the variable
¢>, for instance the term aL", = a (n/i) a/a¢>, which adds the linear term ±anm to
the operator in square brackets in Eq. (12.2), and distinguishes and differentiates
between the +m and -m cases. Additional degeneracies may arise depending on the
188 Chapter 12
details of the potential V(p) provided more than one n, m combination corresponds
to a given energy.

EXAMPLE 12.1
Consider the two-dimensional oscillator potential V (p) = M w2p2 /2. Examine the
degeneracies which arise in this case.
Solving the problem in cylindrical coordinates yields:

Enm (2n + m + 1) tiw, n, m = 0,1..., (d. Eq. (8.14) )

= Ntiw N= 1, 2 ....
A simple tabulation shows the degeneracy increases as the energy increases, in
fact is N, where Ntiw is the energy of the system. These results are tabulated in
Table 12.l.
It is of interest to compare these with the degeneracies which result if one solves
the same problem in Cartesian coordinates. This particular potential can also be
written V(x, y) = Mw 2 (x 2 + y2) /2 and has the Cartesian-coordinate solutions:

En",ny = (n x + ny + 1) tiw = En", + Eng = Ntiw,

where

1 1 (1) 1/2
2n/2 (n!)1/2 b,Ji Hn
(X) (X2 )
b exp - 2b2 '
(12.4)

Table 12.2, which is similar to Table 12.1 shows that the number of degeneracies
for a given energy are, as they must be, basis independent.
In case there is no degeneracy (in the ground-state) the wave functions are iden-
tical. Thus if E = tiw:

xoo(p, c/» = woo(p)


V2i p1/2 =
1
V2i
(2)1/2 (p2)
b2 exp - 2b2

( 1)
bJ7r
1/2 (X2) ( bJ7r
exp - 2b2
1) 1/2
exp - 2b2
(y2 )
=
(1) (p2 )
by'i exp - 2b2 .
Degeneracy 189
However in the cases there is degeneracy the wave functions are not identical. Thus
the E = 2 nw states are

XO±1(P, ¢) = W01 (p) exp (±i¢)


p1t2 .;2-ff =; b2 exp (p2
( 1 ) 1/2 P
- 2b2 exp (±z¢)
) .

in cylindrical coordinates while in Cartesian coordinates:

vJ01(X,y) = (b~r/:xp (-;:2) ~ (b~r/2 2: exp (-~:) = (;r/2:2 exp (-;~)


and

TABLE 12.1
Energy levels in Example 12.1, using cylindrical coordinates

Partial Total
Enm n m
Degeneracy Degeneracy

1nw 0 0 1 1

2nw 0 1 2 2

1 0 1
3nw 3

0 2 2

1 1 2
4nw 4

0 3 2
190 Chapter 12
TABLE 12.2
Energy levels in Example 12.1, using Cartesian coordinates

Ensny N n", ny Degeneracy

11iw 1 0 0 1

1 0
21iw 2 2

0 1

2 0

31iw 3 1 1 3

0 2

3 0

2 1
41iw 4 4

1 2

0 3

~lO(X,y) = (~r/2 ~ exp (-::2)'


which are obviously not identical with XO±l(P, tjJ).
Degeneracy 191
Despite this one can quickly confirm that

"-)
(p, 'I' .,pl0(X, y) ± i.,p01(X, y)
XO±1 = V2
Generally:

00

Xnm(P,¢) = ~ an"ny.,pn"ny(x,y) , (12.5)


n3:tny

where the sum is over all degenerate states at the same (in this case n",+ny = 2n+m)
energy as Xnm(P, ¢) and reciprocally:

.1. ( ) _ X01(P, ¢) + XO-l(P, ¢)


'1'10 X, Y - V2 '
.1. ( ) _ X01(P, ¢) - XO-1(P, ¢)
'1'01 x,y - iV2 '
i.e.
00

.,pn"ny(X,y) = ~bnmXnm(P,¢) , (12.6)


n,m
with analogous restrictions on n, m in this sum.

EXAMPLE 12.2
Discuss the degeneracies for a particle in a two-dimensional Coulomb potential
(cf. Eq. (8.12) ).
The energy for this system is

1 3
n= 2' 2····
The degeneracy is:
n-1/2
2n = 1 + L: 2,
m=1
since aside from m = 0 all other terms are two-fold degenerate, while m varies from
o to n -1/2.
Three-dimensional systems
Consider the case V = VCr). The Schrodinger eigenvalue equation is then (see
Eq. (8.2) ):
192 Chapter 12

where (12.7

This equation has at least a (21 + I)-fold degeneracy because Eq. (12.7) is
independent of the integer m which can take all values from -1 to 1. There may
be additional degeneracies which depend on the details of V(r), i.e. on how many
different n, 1 combinations yield the same energy Enl •

EXAMPLE 12.3
Discuss the degeneracies of the three-dimensional simple harmonic oscillator.
For the three-dimensional simple harmonic oscillator the energies are given by

Enl = (2n + 1+ 3/2)1iw


(see Eq. (8.13)), in spherical coordinates and

in Cartesian coordinates.
One can easily verify that if E = (N + 1/2)1iw the degeneracy is:

{ N(~ + I)} _ fold = 1=1:4..521 + 1),


if N is odd or:
N-1
= L (21+ 1),
1=1,3,5...
if N is even, where N = nx + ny + n z + 1 = 2n + 1+ 1 also corresponds to the energy
level. Thus N = 1 is the ground state, N = 2 the second or first-excited state, etc.

EXAMPLE 12.4
Discuss the degeneracies of the three-dimensional Coulomb potential. For the
three-dimensional Coulomb problem
Degeneracy 193
(see Eq. (8.11)) and the degeneracy is:
n-1
2:(21 + 1) = n2•
1=0

This considerable degeneracy in the three-dimensional Coulomb system is due to


the fact that 'accidentally' in this case the energy does not depend on 1. Thus this
system has a so called 'accidental degeneracy' in addition to the degeneracy due to
the fact that the energy does not depend on m. In fact the Nth energy level of the
Coulomb system with a degeneracy N 2 has approximately twice the degeneracy of
the three-dimensional simple harmonic oscillator's (N(N + 1)/2 '" N 2 /2 for large
N) Nth energy level. Similarly the degeneracy of the Nth energy level of the two-
dimensional Coulomb potential which is 2N - 1 '" 2N for large N as compared
to that of the two-dimensional harmonic oscillator potential which is N. The two-
dimensional Coulomb potential also has an 'accidental degeneracy' in that the energy
is completely independent of m.
In the three-dimensional problem one may remove the (2/ + 1) degeneracy of each
state for example by adding a term of the form aL z to the Hamiltonian, and hence
a term ±anm to the bracketed operator on the 1.h.s. of Eq. (12.7).
If one has two- (or more) particle systems the degeneracies are generally quite
numerous.

EXAMPLE 12.5
Consider the system

and one can immediately write a product solution:

or
1/1~~2m2(1, 2) = ~ {Y,!:l (I)Y,!;2(2) ± Y,!:l (2)Y'!;2(1)} ,
(where the 1/ V2 is required for normalization purposes and the latter function is
symmetric (or anti symmetric) under interchange of the coordinates 1, 2 ).
For a given II, 12 there are generally (211 + 1) (212 + 1) symmetric wave functions
and the same number of anti symmetric wave functions where for symmetric functions
1/1(1,2) = 1/1(2,1) while for antisymmetric wave functions 1/1(1,2) = -1/1(2,1). In the
special case it = 12 there are only (21 + 1)2 wave functions at the energy C2n21( 1+ 1)
of which 21(21 + 1) are symmetric and (21 + 1) are antisymmetric.
These results can be easily illustrated by listing the wave functions for II = 1, /2 =
1; h = 1, l2 = 2.
194 Chapter 12
(a) h = 1, 12 = 1, E = 4C'h2,
1/>(1,2) = Yl 1(I)Yl1(2); Yl 1(1)Yi1(2); Yl 1(1)Yd(2); Yo1(I)Yl1(2); Yo1(I)Yd(2);
or
Yl(1 )1-;1(2);
(1/../2) [1-;1(I)Yo1(2) ± 1-;1 (2)Yci (1)] ;
Yo1(I)Yo1(2);
1/>(1,2) =
(1/../2) [1-;1(I)Y!1(2) ± 1-;1 (2)Y!1(1)] ;
(1/../2) [Y_.\(I)Yo1(2) ± Y!1(2)Yci(I)];
y1-1 (l)y1-1 (2)-,
or equivalently

1-;1(1 )1-;1(2);
(1/../2) [1-;1(I)Yo1(2) + Yo1(1)1-;1(2)];
(1/../6) [2Yo1(I)Yo1(2) + Yl(I)Y!1(2) + Y!l (1)Yl (2)] ;
(1/../2) [Y!1(I)Yo1(2) + Yci(I)Y!1(2)];
yl-1 (l)yl-1 (2).,
1/>(1,2)=
(1/../2) [1-;1(I)Yo1(2) - Yci(1)1?(2)];
(1/../2) [1-;1(I)Y!1(2) - Y!1(1)1-;1(2)] ;
(1/../2) [Yo1(I)Y!1(2) - Y!1(I)Yo1(2)] ;

(1/v'3) [-Yo1(I)Yo1(2) + 1-;1 (1)Y!1(2) + Y!1(1)Yl(2)] ;

i.e. there is a ninefold degeneracy. (This latter combination corresponds to the set
of eigenfunctions 1/>~ot. of the total angular-momentum operator

L = L(I) + L(2) where Itot . = 2, 1, 0,


corresponding to six symmetric, and three antisymmetric combinations d. Eq. (15.20)
and Table 15.3, Chapter 15. Five of these states are degenerate (symmetric) eigen-
functions of 1 = 2, three degenerate (antisymmetric) eigenfunctions of 1 = 1 and one
(a symmetric) eigenfunction of 1 = 0.)
(b)lt=l, 12=2, E=8C'h2,

1/>(1,2) = Yl 1(1)Y;2(2); Yll(1)Y~2(2); Yl 1(1)Y;1(2); Yll(1)Y~1(2); Yl 1(1)Yo2(2);


Yo1(I)Y;2(2); Yo1 (I)Y;1(2)j Yci(I)Yo2(2);
Yl1(2)Y;2(I)j Yll(2)Y~2(1); Yl 1(2)Y;1(1); Yll(2)Y~1(1); Yl 1(2)Yo2(1);
Yo1(2)Y;2(1); Yo1(2)Y;1(1); Yd(2)Yo2(1);
Degeneracy 195
or
(1/V2) [Yl 1(1)Yl2(2) ± Yl 1(2)Yl2(1)] ;

(1/V2) [Yo1(1)Yl2(2) ± Yo1(2)Yl2(1) ] ;

(1/V2) [Y~1(1)Yl2(2) ± Y~1(2)Yl2(1)] ;

(1/V2) [Yl 1(1)Yl1(2) ± Yl 1(2)Yl1(1)] ;


1/1(1,2) =
(1/V2) [Y~l (1 )Yl2(2) ± Y~l (2)Yl 2(1)] ;

(1/V2) [Yl 1(1)Yo2(2) ± Yl1(2)Y~(1) ] ;

(1/ V2) [Y01 (1 )Y:f1 (2) ± Yo1(2)Y:f1 (1) ] ;

(1/V2) [Yo1(1)Y02(2) ± Yo1(2)Y02(1) ];


or equivalently, 1/1(1,2)

(1/V2) [1';2(l)"l';?(2) ± 1';2 (2)Yl (1)] ;


(1/..{6) [1';2(1)Y01(2) ± Yl(2)Yo1(1)] + (1/}3) [y;.z(1)y/(2) ± y;.z(2)Y/(1)];
(2/vT5) [Yl(1)Y01(2) ± Y?(2)y;N1)] + (1/V5) [Yo2(1)1';.1(2) ± Yo2(2)Yl(1)]
+ (l/VW) [1';2(1 )Y!l (2) ± Yl(2)Y!1 (1)] ;
()3/10) [Y~(1)Yo1(2) ± Yo2(2)Yo1(1)] + (1/00) [Y!1(1)Y/(2) ± Y!l (2)1';.1 (1)]
+(1/00) [1';.2(1)Y!1(2) ± Yl(2)Y!1(1)];
(2/vT5) [Y!1(1)Yo1(2) ± Y!1(2)Yo1(1)] + (1/V5) [Y~(1)Y.~1(2) ± Y~(2)Y21(1)]
+(l/VW) [Y!2(1)1';.1(2) ± Y!2(2)1';.1(1)] ;
= (1/..{6) [Y!2(1)Yo1(2) ± Y!2(2)Yo1(1)] + (1/}3) [Y': 1(1)Y!1(2) ± Y':1(2)Y!1(1)] ;
(1/v'2) [Y': 2(1)Y!1(2) ± Y': 2(2)Y!1(1)];
(1/}3) [1';2(1)Yo1(2) ± Yl(2)Yo1(1)] - (1/..{6) [1';.2(1)1';.1(2) ± 1';.2(2)1';.1(1)];
(1/..{6) [Yl(1)Y!1(2) ± 1';2 (2)Y!1 (1)] + (l/vTI) [Yl(1)Yo1(2) ± 1';.2(2)Yo1(1)]
-(1/2) [Yo2(1)1';.1 (2) ± Yo2(2)1';.1 (1)] ;
(1/2) [1';.2 (l)Y!l (2) ± Yl(2)Y!1(1)] - (1/2) [Y': 1(1)1';.1(2) ± Y!1(2)Y/(1)];
-(1/..{6) [Y!2(1)Y?(2) ± Y!2(2)Yl(1)] - (l/vTI) [Y':1(1)Yo1(2) ± Y!1(2)Yo1(1)]
+(1/2) [Y~(1)Y!1(2) ± Yo2(2)Y!1(1)] ;
-(1/}3) [Y':2(1)yl(2) ± Y': 2(2)Y01(1)] + (1/..{6) [Y!1(1)Y!1(2) ± Y': 1(2)Y!1(1)] ;
196 Chapter 12
(J3/1O) [Y;2(1)Y!1(2) ± Yi(2)Y!1(1)] - (J3/20 ) [Y?(1)Y;l(2) ± 1';.2 (2)Ycl (1)]
+ (l/VW ) [Yo2(1)1';.1(2) ± Yo2(2)1';.1 (1)] j
()3/20) [l}2(1 )Y!l (2) ± l}2(2)Y!1 (1)] - ()2/10 ) [yi(1)}'~N2) ± Yo2 (2)Y;l(1)]
= +()3/20 ) [Y~l (1)1';.1(2) ± Y~1(2)Yl(l)] j
()3/1O) [Y~2(1)1';.1(2) ± Y~2(2)1';.1(1)] - ()3/20 ) [Y~1(1)Yo1(2) ± Y~1(2)Yo1(1)]
+(l/VW) [Yo2(I)Y!1(2) ± Yo2(2)Y!1(1)] j
(The last combinations involve eigenfunctions .,p~ot., of the total angular momentum
with ltot . = 3, 2, 1, cf. Eq. (15.20) and Table 15.3, Chapter 15. )
Thus there is a thirty-fold degeneracy, comprising seven degenerate symmetric eigen-
functions with 1 = 3, five with 1 = 2 and three with 1 = 1, and the same number of
antisymmetric eigenfunctions for each of these l's.
One can see why the wave functions above can be written as either symmetric or
antisymmetric combinations by the following considerations:
Given P12 the exchange (or 'permutation') operator, where

the Hamiltonian H(l, 2) in this problem commutes with P12 , i.e.

[Pl2 , H(1,2)] = O.

But the eigenvalues ,\ of Pl2 are ±1 since if

then
Pi277(1,2) = ,,\P1277(1,2) = ..\2 77 (1,2) = Pl2 77(2,1) = 77(1,2),
i.e. ,\2 = 1, hence..\ = ±l.
Since in this example Pl2 and H(l, 2) commute, the eigenfunctions of H can also
be written as eigenfunctions of Pl2 , i.e. as combinations which have eigenvalues +1
or -1 under particle interchange.
One can remove the degeneracy of this system for instance by adding the term

to the Hamiltonian. With this additional term the Hamiltonian no longer commutes
with the exchange operator Pl2 (unless al = a2) and different projections have differ-
ent energies so there is no degeneracy. If al = a2 = a there is still 'exchange degener-
acy', though states with different projections are no longer degenerate. For example
in the case It = 12 = lone then has one (symmetric) state with energy 4Ch 2 + 2an,
two states (one symmetric and one antisymmetric) with energy 4Cn 2 + an, three
Degeneracy 197
states (two symmetric and one antisymmetric) with energy 4Ch 2, two states (one
symmetric and one antisymmetric) with energy 4Ch 2 - ah, and one (symmetric)
state with energy 4h2 - 2ah. In the case 11 = 1,12 = 2, one has two states (one
symmetric and one antisymmetric) at energy SCh 2 + 3ah, four (two symmetric and
two antisymmetric) at energy SCli 2 + 2ali, six (three symmetric and three antisym-
metric) at SCli 2 + ali, six (three symmetric and three antisymmetric) at SCli 2 , six
(three symmetric and three antisymmetric) at SCli 2 - ali, four (two symmetric and
two antisymmetric) at SCli 2 - 2an, and two (one symmetric and one antisymmetric)
at SCli 2 - 3ali.
EXAMPLE 12.6
Consider the Hamiltonian:
1 1
H ( 1,2 ) =T1 +T2 +2"Mw
22 22
r1 +2"Mw r 2 •

Discuss the degeneracies of this two-particle system.


We note that this H(1,2) commutes with the permutation operator, and can
immediately write down:

Enl>n2 = (2nl + 11 + ~ + 2n2 + 12 +~) nw.


The degeneracies of the lowest three energy levels are shown in Table 12.3. One notes
that except for the ground state at 3hw, the degeneracies are quite numerous, increase
rapidly as the energy increases, and that symmetic and antisymmetric combinations
of the resulting wave functions can easily be formed.
Thus for E = 4liw :

where the subscript ± stands for symmetric and antisymmetric wave functions re-
spectively and the 1/.../2 is inserted to ensure wave function normalization.

EXAMPLE 12.7
Consider a particle subject to the Hamiltonian:

H = In (1 + ~ (L! + L!)).
Discuss the energy levels of this system and their degeneracies.
One uses the identity:

to obtain
198 Chapter 12
Hence one deduces:

where -1:S m :s 1, 1 = 0, 1,2, ... , and m is an integer.

The first few levels and their degeneracies are illustrated in Table 12.4.
One notes that one can write for Elm, Elm = In (1 + na), n = 0, 1, 2 ... , with
different degeneracies for different values of n.
For small a,

Adding a term proportional to L z will remove some or all the degeneracy depend-
ing on the coefficient in front of this additional term.

Consider an energy level p which is say j-fold degenerate. If nil n2, n3, n4, ... nj
label the corresponding degenerate states, it may not be possible to use the pertur-
bation expansion Eq. (11.14) directly. In particular if the perturbation V connects
:s
any of the j degenerate states say n = nl and m = ni where 1 < i j, i.e. Vn1ni =

TABLE 12.3
Degeneracies of Hamiltonian in Example 12.6

Energy Degeneracy nl n2 11 12 ?f>(1,2) ml m

3nw none 0 0 0 0 Roo (1 )Roo (2) You(1 )YoU(2) 0 0

41iw 3+3 0 0 1 0 1tol(1)lloo(2)l'~1(1)l'oO(2) 0,±1 0


=6-fold 0 0 0 1 Roo(1)Rol(2)l'oO(1)l'~2(2) 0 0,::1

9+5+5 0 0 1 1 1tol(1)Rol(2)l'~1(1)l'~2(2) 0,±1 0,:1


+1+1 0 0 0 2 Roo(1)Ro2(2)Yo°(1)l'~2(2) 0 0,±1
5liw =2 I-fold 0 0 2 0 1to2(1)Roo(2)l'~1(1)Yo°(2) 0,±1,±2 0
0 1 0 0 lloo(1)Rlo(2)l'~(1)Yo°(2) 0 0
1 0 0 0 R lO (l )Roo(2)l'~(1 )l'oO(2) 0 0
Degeneracy 199
TABLE 12.4
Degeneracies of Hamiltonian in Example 12.7

Partial Total
Energy 1 m
Degeneracy Degeneracy
0 0 0 1 1
In(1+0:) 1 ±1 2 2
1 0 1
In (1 + 20:) 3
2 ±2 2
In (1 + 30:) 3 ±3 2 2
In (1 + 40:) 4 ±4 2 2
2 ±1 2
In (1 + 50:) 4
5 ±5 2
2 0 1
In(1 +60:) 3
6 ±6 2
In(1+70:) 7 ±7 2 2

(nl IVI ni) =f:. 0, there will be terms in this expansion with a zero energy denominator
lOnl - tOn, and a non-zero numerator Vnln ,. In this case Eq. (11.14) will consequently
be divergent. The way out of this difficulty is to use the fact ~ee Eqs. (12.5), (12.6) )
that one can take different combinations of the j degenerate functions, in particular
a combination for which all Vn,n,1 = 0, i =f:. if, where i, if S j. To do this one
diagonalizes V in each subspace of degenerate states. The resulting eigenfunctions
will have only diagonal matrix elements of V.
If one therefore uses this (rather than the original) basis in Eq. (11.14) it will not
diverge since terms with zero denominator also have zero numerator. Moreover the
eigenvalues of these diagonalizations are to first-order just the energies of each level
p since Ep = lOp + Vpp to this order and only the diagonal elements contribute to Vpp
in this new basis.
In the simple case of two degenerate states nl and n2 connected by V, the nor-
malized combination one must choose (which uncouples if 0: = 0) is:

(12.8)

with an 0: which makes (n~ IV In~) = 0, namely:


200
Cbapte r 12

2 Viz
tan 2a = V, V;' where Viz = (nIl V Inz) etc., (12.9)
11 - zz
(provid ed Viz = VZl )'
If one has three degene rate states nl, nz, and n3 connec ted by
V, a normal ized
combin ation (which uncoup les when a, 13 = 0) that one may choose
is:

(12.10)

(provid ed Viz = \121, Vi3 = V31 , \123 = V3Z ),


where (n~IVln~), (n~IVln~), and (n~IVln~) are zero.
These constra ints imply:

tan2a = tana =-11,


\123
13
(12.11)
tan 213
2 (Vi3 cos a + VZ3 sin a)

EXAM PLE 12.8


Consid er a Hamilt onian involving the two-dim ensiona l oscillat or:

where one wishes to find the energy of the second excited state.
More specifically, if
Vo = 0, EN = (N + l)1iw, and one is interes ted in the state with N

!
= 2.
Clearly the Cartesi an basis is conven ient here, where the basis
states for N = 2
are
12l) = ¢z(x)¢o(Y),

12z) = ¢l(X)¢I(Y),

12 3 ) = ¢o(x)¢z(y).
(cf. Eq. (12.4) ).
Using Eq. (12.11) , tan 2a = 2Viz/ (ViI - \l2z) = 0, tan a = \l23/Vi3
= 0, i.e. a = 0,
since ¢1(0) = o.
Degeneracy 201
Additionally
2Vi3 2<p~(0)<p~(0)
tan 2(3 = Vil _ Va3 = <p~(0)<p~(0) _ <p~(0)<p~(0) = 00,
i.e. 2(3 = 7r /2, (3 = 7r / 4.
Thus:
1
= v'2 {<P2(X)<PO(Y) + <PO(X)<P2(Y)}

12;) = <Pl(X)<Pl(Y) (12.12)

and the energy to first order is:

Vomw
31iw + (2~ 1V 12~) = 31iw + 2Vo<p~(0)<p~(O) = 31iw + ~

E 31iw + (2;1 V 12~) = 31iw (12.13)

31iw + {2~ IV 12D = 31iw .


Two of the three basis states thus remain degenerate states to first order in pertur-
bation theory.
This problem can also be solved exactly by graphical techniques. Thus the
Schrodinger equation for this system can be written:

(Ho - E)1/J(x,y) = -110 Sex) S(y) 1/J(x,y).


Hence:
1
1/J(x,y) = -110 Ho _ E S(x} S(y) 1/J(x,y).
Using closure one may insert a complete set of states which are eigenfunctions of
Ho (namely 2::n .. ,ny <Pn.. (x)<Pny (y) ), between the operator 1/ (Ho - E) and Sex) 8(y),
obtaining:

1/J(x,y) = -110 Ho~E L l<Pn .. (X)<Pny(Y)) (<Pn,,(X)<Pny(y)IS(x) 8(Y)I1/J(x,y))


n,x,ny
202 Chapter 12
Inserting the values x, y = 0 into this expression yields:

-v,1o = 'L.J
nSt
" ( (n +n +l)1iw-E
ny :r: y
1 ) 4>n.,(O)4>ny(O),
2 2

I.e.

1 4>~(0) 24>~(0)4>~(0) {24>~(0)4>~(0) + 4>~(0)}


Yo = 1iw - E + 31iw - E + 51iw - E + ....

This expression is illustrated graphically in Figure 12.2.


Consider the energy level close to 31iw, E 31iw + f. Approximately for this level
f'V

_~ _- 24>~(0)4>~(0).I.e. f -
./..2(0)./..2(0)
_ 2V,0'1'0 '1'2 ,
V,o -f

and E = 31iw + 2Vo4>~(0)4>~(0), in agreement with the degenerate perturbation theory


result of Eq. (12.13).

Figure 12.2: Graphical solution for energy levels in Example 12.8.

EXAMPLE 12.9
Consider the Hamiltonian:

H= Ho+ V,
where:
Degeneracy 203
which involves the two-dimensional oscillator as the unperturbed Hamiltonian Ho in
this decomposition, and where one wishes to find the energy of the first excited state
(i.e. if ,X = 0, EN = (N + 1)1iw, where N=1).
Clearly the Cartesian basis is convenient here, with basis states:

Ill} = <P1(X)<PO(Y) (cf. Eq. (12.4) ),


{
11 2 } = <PO{X)<P1(Y)·
Using the results given by Eq. (12.9),

tan2a = 2 Vi2 = 00,


Vi1 - V22
since Vi1 = V22 = 0, whereas

i.e. a = 7r /4.
Thus

(12.14)

and the energy to first order is:

= 2fiw + ,XMw2 (<Po(x) I x 2 1<Po(x))


1
= 2fiw + 2'x (<Po(x) I 2"Mw2x2 I<Po (x)} = 2fiw + 'xfiw/2.
(12.15)

= 2fiw - 'xfiw /2.

This problem can also be solved exactly since:


204 Chapter 12

H = ~M (v; + v~) + ~MW2 (X2 + 2AXY + y2)

= ~M (vi- + v}) + ~MW2 (1 - A)X2 + ~MW2 (1 + A) y2,


where
1 1 ~
X = y'2(X - y), y = y'2(x + y), and ve == dt .

Written in terms of X, Y the Hamiltonian decouples and is seen to have eigen-


values:

E NxNy = (Nx + ~) nw~ + (Ny + ~) nw~ (12.16)


To order ,\ one thus has:

= 2nw+ Anw
2

E 2 ~w _ Anw
10 = n 2'

etc. in agreement with the degenerate-perturbation theory results of Eq. (12.15)


above.
Likewise the exact wave functions corresponding to these energies are:

'ljJ10 = <PI (X, W = WI) <Po (Y, w = W2)


and (12.17)
'ljJol = <Po(X, w = WI) <PI (Y, W = W2)
where
Wl=W~,
I.e.

1/;10 = ~ (~;) 1/2 (1 _ ,\2f/8 exp ( M W 7 X2)2X (~;) 1/2 (1- ,\)1/4

X exp (
M wv'f+'"1
21i
Y2)
1
y'2 {<Pl(X)<PO(Y) - <PO(X)<Pl(Y)} + O(A),
Degeneracy 205
and

results which are consistent with Eq. (12.14) to within an unimportant overall phase.

EXAMPLE 12.10
Consider the Hamiltonian:
h2 d2
H = - 21 d(p + Vah' ( ¢ - ¢o)

(Here 121r 6 (¢ - ¢o) d¢ = 1, 121r f(¢) h' (¢ - ¢o) d¢ =f (¢o) ).

Treating _h 2/2I d2/d¢2 as Ho, there are two possible pairs of degenerate eigen-
functions, either:

or:
'if;l = (1 exp(in¢),
y?;; and 'if;2 = (1 exp(-in¢),
y?;;
(n n
each pair having eigenvalues En = 2 h2 ) /21, where must be an integer so that
'if;(¢ + 211") = 'if;(¢) , i.e. is single valued. Consider cases for which n "# o. (The
case n = 0 gives the non-degenerate ground state for which the wave function is a
constant.)
Choice 1: For the choice:

Va 2 Va Va. 2 -I..
Vil = -;- cos n¢o ; Vi2 = 211" sin 2n¢o ;
TT
V22 = -
11"
sm n'l'O ,
and
(2 Va/211") sin 2n¢o 2
tan 2a = 2
(Va/1I") cos 2 n¢o - (Va/1I") sin n¢o
= tan
-I..
n'l'O, I.e. a = n¢o.
206 Chapter 12
Thus

In~} = ~(cosn</>cosn</>o+sinn</>sinn</>o) = ~cosn(</>-</>o),

In~} = ~(-cosn</>sinn</>o+sinn</>cosn</>o)= ~sinn(</>-</>o),


and

€n + -
Va 1211" sin2 n (</> - </>0) 0 (</> - </>0) d</> = €n ,
7r 0
a result independent of </>0 ,i.e. of the choice of original axis orientation.
Choice 2: For the choice:

InI} = ~exp(in</»,
v27r
and In2} = ~exp(-in</»,
v27r
one obtains:

ViI = ~; V22 = ~; Vi2 = ~ exp (-i2n</>0) (i.e. V2; = Vi2) .


Hence one cannot use Eq. (12.9) which assumes V12 = V2I .
Instead one may go directly to the expression (cf. Eq. (12.8) ):

En = tOn + (n~ IV In~) = tOn + V11 cos 2a + V22 sin 2a + ~ (V12 + V21 ) sin 2a

= fn+ V1I +
2
V22 V1I - V22
+ 2
2
cosa+
V12 + V2I . 2
2 SIna.

Additionally requiring (n~1 V In~) = 0 yields:


(ViI - V22) sin 2a = (Vi2 - V21) + (Vi2 + V2I ) cos 2a ,
from which one obtains:

[(ViI - V22 ) (Vi2 - V2I) ± (V12 + V2I) V(V1I - V22)2 + 4V12V2I ]


tan2a
[- (V12 - Y:n)2 ± (V11 - V22) V(V1I - V22)2 + 4V12V21 ]
sin2a
=
cos2a
Degeneracy 207

( Vll -V22 )( V12 -V21 ) ± ( V12+V21 h/( Vll -\I22 )2+4 V12V21
( Vll - V22 ) +( V12 +V2l )

Hence:

(12.18)

The Eq. (12.18) is quite general. In fact it involves the solutions of the eigenvalue
equations which result from diagonalizing the Hamiltonian H in the space of two
degenerate states nt, n2, i.e.

=0.
"21 H22 - En

Substituting in Eq. (12.18) the values of Vib "22, and Vi2 for the choice 2 one
obtains:
En
Va
= En + -, En = En ,
1f
which as expected are identical with the energies resulting from the choice 1 for the
degenerate basis, since measureable results should not depend on what degenerate
combination of states one chooses.

EXAMPLE 12.11
For the Hamiltonian in Example 12.10 calculate the second-order energy contri-
bution.
Using the choice 1 for the degenerate states one obtains:

(m IYo 8 (e/> - e/>o) I n~) = Va


1f
cosme/>o, Va
1f
sinme/>o,

depending on whether m is (1/ -Ji ) cos me/>, or (1/ -Ji )sin me/>. Also
{m I Yo 8 (e/> - 4>0) In~} = o.
Thus:
208 Chapter 12
a result, which like the first-order expression is independent of rPo, i.e. of the choice
of axis orientation. Thus:

which can easily be summed for a particular n. Thus if n = 1,

Hence

h 2 /21,
2h2 /1+ Vo/7r - 31V02/8h 27r 2 + ... ,

etc.

EXAMPLE 12.12
The 2S1/ 2 and 2p1/2 states of the hydrogen atom are separated by an energy fj.E
known as the Lamb shift (where fj.E/h =10 9 Hz). Consider these two states in
the presence of an electric dipole field Vd = (Ame/h) z. Neglecting the other energy
levels of hydrogen, calculate exactly the two new eigenstates of the system and the
corresponding eigenvalues. Compare these results with the eigenvalues obtained using
non-degenerate and degenerate perturbation theory. (N.B. Consider carefully the
parity of the integrals involved.)
Defining:

me
V = hZ,
one obtains

+A(1 1V1 1}-E


[
t1
A (11V12) 1[a 1= 0,
1

A (21V11) 102 + A (21V12) - E az


in agreement with Eq. (11.6).
Degeneracy 209
But {IIVII} = {21V12} = 0 from parity considerations.
Therefore:

Also
a2 = E1 - f1 _ (-(f1 - (2)/2 + -)(f1 - (2)2/ 4 + ,\2 (IIV12}2 )
a1 ,.\ {IIV12} - ,.\ {IIV12}
and

etc.
If
"\{IIV12} « (f1- f2)/2,
2{IIV12}2
E1 rv f1 + ,.\ + ... ,
f1 - f2
and

Also
,.\ (21VII)
a2 rv aI,
f1 - f2
etc., in agreement with non-degenerate perturbation theory for E and 1/J.
If ,.\ {IIV12} » (f1 - (2) /2, f1 rv f2 , i.e. the Lamb shift is much smaller than the
effect of the electric dipole field, and

and a2/a1 rv 1,
in agreement with the degenerate result (cf. Eq. (12.9) ) for the case tan2a = 00,
1.e.

a = i, in which case II'} = ~ {II} + 12}},


and
E1 = (1'1 H II') = f1 + ,.\ (IIV12) .
210 Chapter 12

EXAMPLE 12.13
Consider the same Hamiltonian as in Example 11.9, namely the Schrodinger equa-
tion for a particle which is confined to a two-dimensional square box whose sides have
length L and are oriented along the x and y coordinates with one corner at the origin,
and additionally subjected to a small perturbation:

but invesigate the first-excited state rather than the ground state.
In this case there is a two-fold degeneracy, namely:

<P1(X,y) = (2/L) sin ('lrX/L) sin (27ry/L) ,


{
<P2(X,y) = (2/L) sin (27rX/L) sin (7ry/L) .
The unperturbed (or zero-order) energy of these two states is:

h 2 7r 2 5h 2 7r 2
101 = 102 = 2mL2 (12 + 22) = 2mL2 .
Also one can easily show:

Thus setting up and diagonalizing the perturbation in this degenerate basis:

=0.

with the resulting eigenvalues E~l) = 0, or 4 Vi .


Thus:

EXAMPLE 12.14
Consider the same Hamiltonian as in Example 11.10, namely
the Hamiltonian:

H = IT
nO + v:, h
were IT
nO = 2m
p~ p~ + 2"1 mwo2 (2
+ 2m 2)
Xl + X2 1 2 (Xl - X2 )2 ,
an d V = 4"mw1

but investigate the first-excited rather than the ground state.


Degeneracy 211
In this case there is a two-fold degeneracy, namely:

tP01(X,y) = 4>O(X)4>t(Y)
{
tPlO(X,y) = 4>t(X)4>O(Y).
The unperturbed (or zero-order) energy of these two states is:

Setting up and diagonalizing the perturbation in this degenerate basis, after a bit of
calculating one obtains:

1iw~ / 2wo - E~t)


(
-fiwU4wo

with resulting eigenvalues E~1) = 3fiwU4wo, and fiwU4wo.


Thus:

3fiw~ fiw 2
El = 2fiwo + - -, or E1 = 2fiwo + _ 1 .
4wo 4wo
This may be compared to the exact energies for these states:

E01 =
3
-fiwo
1
+ -fiwo 1+
(W- t )2 '" 21iwo + -fiw~ ,
2 2 Wo 4wo
(12.19)

= 1 3
-2fiwo + -2fiwo 1 +
(W- t )2 '" 2fiwo + -
3fiw~
-.
Wo 4wo

EXAMPLE 12.15
Find the minimum degeneracy in the N = 4,5 dimensional systems for "central"
potentials:
VN(r), r = Jd + r~ + r~ + ...rh.
Using Eq. (8.15), the number of m values for a given 1 is:

{(21 +N - 2)(1 + N - 3)!}/{1! (N - 2)!}, 12::1,


A(N,/) ={
1, 1 = o.
Thus
A (5, I) = (l + 1)(/ + 2)(21 + 3)
6
212 Chapter 12
In other words, for N = 4 the degeneracy is at least [(1 + 1)2]-fold for the state <Pnl(r),
while if N = 5 the degeneracy of such a state is at least [(1 + 1)(1 + 2)(21 + 3)/6]-fold,
as compared to the (21+1) minimum degeneracy for the three-dimensional system. In
addition to the above degeneracies there may be additional "accidental" degeneracies
for such systems (d. Examples 12.16 - 12.18).

EXAMPLE 12.16
Find the degeneracy of the Mth energy level of the four-dimensional harmonic
oscillator.
The allowed energies of this system are given by (d. Eq. (8.18) ):

E = (nl + n2 + n3 + n4 + 2) fiw = (2n + 1+ 2) fiw (ni' 1 = 0, 1,2 ... ).

One thus has a degeneracy

cr+ 2 = M(M + 1)(M + 2) , (12.20)


6
where M = nl + n2 + n3 + n4 + 1 = 2n + 1 + 1, corresponds to a particular level.
Hence the first (M = 1) level (ground state) is non degenerate, the first-excited state
(M = 2) is four-fold degenerate etc.
The result Eq. (12.20) can be obtained by summing (d. Eq. (8.18) )
M-l
for M odd E (1 + 1)2,
1=0,2,4

M-l
and for M even L (l + 1)2,
1=1,3, S

where one has used the results of Example 12.15.

EXAMPLE 12.17
Find the degeneracy of the Mth level of the five-dimensional harmonic oscillator.
The energy of this system is (d. Eq. (8.18) ):

E = (nl + n2 + n3 + n4 + ns + 5/2) fiw = (2n + 1 + 5/2) fiw (ni,l = 0, 1,2 ... ).

One thus has a degeneracy

C M +3 _ M(M + 1)(M + 2)(M + 3)


4 - 24 '
(12.21)

where M = nl + n2 + n3 + n4 + ns + 1 = 2n + 1 + 1, corresponds to a particular


level. Hence the first (ground) state M = 1 is non-degenerate, the first-excited state
is five-fold degenerate etc.
The result Eq. (12.21) can be obtained by summing (d. Eq. (8.18) ):
Degeneracy 213

~l (l + l)(Z + 2)(21 + 3)
for M odd
L..J
1=0,2,4
6 '

M even ~1 (I + 1)(1 + 2)(21 + 3)


an d £or L..J 6 '
1=1,3,5

where one has used the results of Example 12.15.

EXAMPLE 12.18
Extrapolate the results of Examples 12.16, 12.17 to the Mth level of the N-
dimensional harmonic oscillator.
In the notation of Examples 12.16, 12.17 the degeneracy of the Mth level of the
three-dimensional harmonic oscillator is Cr+! = M(M + 1)/2 where M = n1 + n2 +
n3 + 1 = 2n + 1 + 1, while that of the Mth level of the two-dimensional system is
cf! = M ( d. Example 12.1), where M = n1 + n2 + 1 = 2n + m + l.
For the N-dimensional oscillator the degeneracy of the Mth level where M = n1 +
n2 + n3 + ... nN + 1 = 2n + 1 + 1, and E = (M + (N - 2)/2 )1iw, is

M+N-2 (M +N - 2)!
(12.22)
CN _ 1 = (N - l)!(M - I)!

(Thus if N = 1 c1:- 1 = 1.)


The result Eq. (12.22) can be obtained by summing:

for M odd 11
1=0,2,4
(21 +N - 2)(1 + N - 3)!
I!(N - 2)! '

and for M 1:
even 1=1, 3, 5
1
(21 +N - 2)(1 + N - 3)!
l!(N - 2)!
CHAPTER 13

The Inverse Problem

Usually in Quantum Mechanics one is given the potential of a system and one
wishes to find the energy levels and eigenfunctions if the particle is bound, or the
wave functions and hence the amount of scattering off this potential if the particle is
free. Sometimes however, one knows the wave function (or phase shifts) and wishes
to find the potential which produces this wave function (or these phase shifts). In
the literature this is known as the 'inverse' problem.
Consider the one-dimensional eigenvalue equation:

",2 d2 }
{ -2mdx 2 + Vex) w(x) = EW(x).

Solving for Vex) yields:

-00 <x < 00. (13.1)

EXAMPLE 13.1
Find Vex) for:

0, Ixl > a. (13.2)

One evaluates
1 d2 w(x) 2
W(x) ~ x2 - a2 '
Ixl < a.
From Eq. (13.1) this implies
",2
Vex) = E + m (2
x -a 2)'
Ixl < a. (13.3)

To find Vex) and E individually and unambiguously, one must impose an addi-
tional condition.
Suppose one is also given the information that (WIVIW) = O.
This implies
The Inverse Problem 215
and

Ixl < a,
(13.4)
= 00, Ixl > a.
Figure 13.1 is a plot of the wave function in this example (and Figure 13.2 of
the resulting potential) and of the very similar function (in this region), cp(x) =
VIla Cos7rx/2a. _1_

1~ V':(x).~~(a2-x2)
~'/

/
/'
/
/
'/
V
__L -________________ r-________________ ~x

-0 o
Figure 13.1: Wave function in Example 13.1 (full line) and the similar
wave function cp(x) = VIla cosn;/2a (dotted line).

Other conditions, for example that V(O) = 0, or E = 0 merely shift the energy
reference.
v(x)

Figure 13.2: Potential in Example 13.1 if (wWlw) = o.


216 Chapter 13
Thus if
r,,2 r,,2 X2
V(O) = 0, then E = - 2 and V(x) =
ma ma2 x2 - a2
, Ixl < a,
while if
r,,2
E = 0, then V(x) = m(x 2 _ a2)' Ixl < a.
EXAMPLE 13.2
Suppose
Ili(x) = V2Ct 3 xexp(-Ctlxl), (13.5)
and one wishes to find the potential which results in this wave function.
One must evaluate l/lli(x) cPlli(x)/dx 2. Using the representation

6(x) = ~ cPlxl
2 dx 2
for the delta function one obtains
1 d21li(x) 2 2Ct
Ili(x) dx2 = -2Ct6(x) + Ct - j;f .
Hence
2 2 r,, 2Ct [ 1
V(x) = E + r,,2m
Ct ]
- --;:;;- j;f + 6(x) ,

I.e.
V(x) = _ r,,:Ct [I!I + 6(X)] '. E = _ r,,;:2 . (13.6)

!Cx)

Figure 13.3: Wave function and potential in Example 13.2.


The Inverse Problem 217
The potential here is a sum of a one-dimensional Coulomb potential and a delta-
function potential (at the origin). As W(x) is odd the delta-function potential does
not affect this wave function's energy which is just that due to the Coulomb potential
(see Eq. (3.32) with n=2). This example is illustrated in Figure 13.3.

EXAMPLE 13.3
Suppose
w(x) = vaexp (-alx!), (13.7)
find Vex), E.
Evaluating
1 d2 w(x) 2
w(x) dx2 =a - 2a8(x),

(see also Example 13.2).


Thus

A suitable choice here is


fi,za
Vex) = --Sex). (13.8)
m
This problem is illustrated in Figure 13.4.

v (x)

Figure 13.4: Potential and wave function in Example 13.3.


218 Cbapter 13
EXAMPLE 13.4
Suppose
25 / 4
w(x) = af (1/4)
(13.9)

find V(x),E.
Evaluating

V(x)=E+ ma2 2n,2 (X)2


~ ((X)4
4 ~ - 3).

The identification

(13.10)

is acceptable.
This problem is illustrated in Figure 13.5.
fIx)
2 15/ 4
afc'/4)

Figure 13.5: Potential and wave function in Example 13.4.

If the form of the potential is known, the detailed parameters may sometimes
be deduced in a straightforward way from the scattering data. This is illustrated in
Example 13.5.
The Inverse Problem 219
EXAMPLE 13.5
Given a beam of particles of mass m and energy E incident on a step potential
Vo. If n is the fraction of the particles reflected, find the height of the potential.
This problem is illustrated in Figure 13.6, where E must be greater than Vo since if
E < Vo, n = 1.
v (x)

Figure 13.6: Form of the potential in Example 13.5.


For

x < 0 \}i(x) = Aexp(ikx) + Bexp(-ikx), while for x > 0 \}i(x) = Fexp(ik'x),

where
k- J2mE
- n,2'
Matching the wave function and its derivative at x = 0 one obtains

IBI2
A
= n = (~)2
k+k"
I.e.
v; _ [-8n + 4(1 + n)y'n] E (13.11 )
0- (1-n)2 '
where as n -+ 1, Vo -+ E, while as n -+ 0, Vo -+ O. If n = 0.36, Vo = 0.9375E, etc.
Three-dimensional inverse problems follow along the same lines.

EXAMPLE 13.6
Suppose the wave function describing a spinless particle of mass m, in a short-
range central three-dimensional potential, is:
220 Chapter 13

W(r,8,4» = A {exp (-ar) ~ exp (-{3r)} , (13.12)

where A, a, f3 are constants and a < f3 .


From this data find the angular momentum of this particle, the energy of this
state, and the potential which results in this wave function.
Firstly:

and
L z ( 4»w(r, 0, 4» = 0 = mlliw(r, 8, 4».
Thus I, ml are zero. Indeed this particular wave function has no 8, 4> i.e. no angular
dependence.
Also
uno(r) = rRno(r) = A {exp (-ar) - exp (-{3r)}.
Since the Schrodinger equation for uno(r) (see also Chapter 8) is like that for a
one-dimensional problem one can use Eq. (13.1) and

V(r) = E + ~_1_d2uno(r) = E + li 2 {a2exp(-ar) - f3 2exp(-f3r)} .


2mu no(r) dr2 2m {exp(-ar) -exp(-{3r)}
For large r, V(r) goes to zero, i.e.

(since (3 > a), or

Thus generally
V r _ ~ [ (a 2 - (32)exp(-(3r) ]
(13.13)
( ) - 2m exp ( -ar) - exp ( - (3r) .
For small r,

V(r) '" Ji2 (a - (3)(a + (3) exp (-(3r) = _ li 2 (a + (3) exp (-f3r) .
(13.14)
2m ({3 - a)r 2mr
The potential Eq. (13.14) is a 'shielded' Coulomb potential, i.e. it looks like a
Coulomb potential for small r but for large r it goes to zero much faster than the
Coulomb potential.
If 1 =f 0 one easily generalizes Eq. (13.1) with the result:

V/(r) =E _ li 2 /(1 + 1) + ~_l_~Unl(r), r ~ o. (13.15)


2mr 2 2mu n l(r) dr2
The Inverse Problem 221
This generalization is illustrated in Example 13.7.

EXAMPLE 13.7
If
Unl(r) = Nr2 exp (_ar2) ,
obtain V,(r) as a function of 1.
One quickly finds:

1 d2un l(r)
i.e.,
Unl(r) dr2 =
Using Eq. (13.15),

Thus if I = 0, a reasonable identification is:


5a1i 2
E=--
m

With the more familiar oscillatior notation a = mw /21i :

E _ 51iw 1i
= -21mw 2 r 2 + -mr
2
- 2 ' Vo(r ) 2 .

If 1 = 1,
E = 51iw
2 '
If 1 = 2,
E _ 51iw 1 2 2 2h2
- 2 ' V2(r ) = -2 mw r - --2 '
mr
etc.

Though to solve the inverse problem exactly in quantum-mechanical scattering


theory one must use the rather involved Gel'fand-Levitan-Marchenko equation!, it
turns out that it is possible2 to invert one of the simplest approximations in quan-
tum mechanical scattering theory, namely the partial-wave Born approximation (Eq.
(6.25) ), without too much trouble. This expression relates a particular integral of
222 Chapter 13
the scattering potential Vi(r) in the angular momentum 1 channel to the phase shift
SI( k) which this potential produces, namely:

1i,2 SI(k) {'XI (j ( ]2 () 2


- 2m-k- '" 10 I kr) Vi r r dr. (13.16)

(Various authors use sin SI( k) or tan SI( k) instead of SI( k).)
This approximation is known to be fairly good for channels where the potential
is "weak" in some sense, i.e., the phase shifts produced are small.
In Eq. (13.16) k = .j2mE/1i,2, where E is the energy, and m the mass of the
scattered particle. The function il(kr) is the usual spherical Bessel function.
Using the formula derived in reference 2:

O(p) (/+1il(P»)2 = sin(2p), (13.17)

where
d
O(p) == ( dp2
)1 dpd ( dp2d )1 '
one can extract the potential Vi(r) from Eq. (13.16). To do this one pre-multiplies
both sides of Eq. (13.16) by p/+2 and then operates on the resulting expression with
the operator O(k) of Eq. (13.17), (where if one writes p = kr, O(k) = r41+10(p) ).
Hence one obtains:

d)1 d ( d)1 (2/+2{ 1i,2 SI(k)})


( dk2 dk dk2 k - 2m-k- = 10[00.sm(2kr)Vi(r)r21+1 dr. (13.18)

which is just the Fourier-sine transform3 of (1/2)Vi(r)r 21+1.


Multiplying both sides of Eq. (13.18) by (4/1r) sin(2kr') and integrating over k
one then obtains

[00 [( d )1 d ( d )1 ( 21+2 { 1i,2 SI(k)})] . ( ,


Vi (r ') -_ 1rr,21+1
4
10 dk2 dk dk2 k - 2m -k- sm 2kr )dk,
(13.19)
since (see also Eq. (2.19) )

~ [00 sin(2kr) sin(2kr')dk =


1r 10
.!.1r 10[00 {cos 2k( r - r'} - cos 2k( r + r'}} d2k

= S(r - r') - S(r + r').


For the cases 1 = 0, lone thus obtains:

Vo(r') = .i.. [00 [~(k2 {_~ SoCk)})] sin(2kr')dk, (13.20)


1rr' 10 dk 2m k
The Inverse Problem 223

') 4 [00 [( d ) d ( d ) ( 4 { n,2 Dl(k) })] . ( ')


Vl (r = 7rr,310 dk 2 dk dP k - 2m-k- sm 2kr dk, (13.21 )

and so on for larger 1.


Examples 13.8-11 illustrate this approximate inversion technique.

EXAMPLE 13.8
Given
Vo, r < a,
Vo(r) ={
0, r > a.
Use Eq. (13.16) to obtain the Born-approximation 1 = 0 phase shifts for this potential
and then assuming these phase shifts use Eq. (13.20) to retrieve the potential.
From Eq. (13.16)

_~ Do(k) IV Vr {....!!:...- _ Sin(2ka)}


2m k 0 2P 4P .

Hence
k2 (_~ 80 (k») IV Vo {~ _ sin(2ka)}
2m k 2 4k '
and
~ {k2 (_~ 80 (k»)} IV Vo (sin(2ka) _ 2acos(2ka») .
dk 2m k 4 k2 k
Substituting this expression into Eq. (13.20) and integrating over k :

Vo IV ~~ Vo roo (sin(2ka) _ 2acos(2ka») sin(2kr) dk


r 7r 4 10 k2 k

= Vo { roo sin(2ka) sin(2kr) dk _ a {OO sin {2k(r - an + sin {2k(r + an dk}


r7r 10 k2 10 k

= -Vo (7r
-2a -7ra ) =0, r > a,
r7r 2

= -Vo (7r
-2r) = Vo, r < a.
r7r 2
224 Chapter 13
EXAMPLE 13.9
Given
Va(r) = Va exp ( -ar 2 ).
Use Eq. (13.16) to obtain the Born-approximation 1 = 0 phase shifts for this potential
and then assuming these phase shifts use Eq. (13.20) to retrieve the potential.
From Eq. (13.16)

_~ 60 (k) rv Va f!. (1 _ exp {_ k2 } ) ,


2m k 4k2V~ a

~ {k2 (_~ 60 (k))} rv Va f!. kexp {_ k2}.


dk 2m k 2aV~ a
Substituting this expression into Eq. (13.20) and integrating over k :

Va(r) ~ roo Va f!.exp {_


7rr 10 2a V~ a
P} ksin(2kr) dk

= Va exp ( -ar2 ).

EXAMPLE 13.10
Given
Va(1-r/a), r < a,
Va(r) ={
0, r > a.
This potential is illustrated in Figure 13.7.

~------------~-- ___ r
a

Figure 13.7: Form of the potential in Example 13.10.


The Inverse Problem 225
Use Eq. (13.16) to obtain the 1 = 0 phase shifts for this potential and then, assuming
these phase shifts use Eq. (13.20) to retrieve the potential.
One has
_~ bo(k) = -Vo loa (1- -r).sm 2 kr dr
2m k k2 0 a

= Yo [~_ (1 - cos2ka)] .
2k2 2 4k2a
Thus
= Yo [~_ (1-COS2ka)]
2 2 4k2a '

!:... [k2 {_~ bo(k)}] = Yo [_ sin2ka


4 k2 +
(1- cos2ka)]
Pa .
dk 2m k
From Eq. (13.20)

Vo(r) = :; 10'' ' [- si~~ka + (1- ~~: 2ka)] sin2kr dk

= Vo [_ (00 sin 2ka sin 2kr dk + ~ (00 (1 - cos 2ka) sin 2kr dk]
7rr io k2 a io k3

= Vo [-7ra+ ~7ra2] = 0, r > a,


7rr a 2

= Vo [-7rr+ ~7rr(2a-r)] = Vo (1-~), r > a,


7rr a 2 a
where one uses 4

(00 sin 2 ka sin 2kr _ { 7ra2/2, r> a,


io k3 dk-
o 7rr(2a - r) /2, r < a.
EXAMPLE 13.11
Given
= Vir2 exp (-ar 2).
Vi(r)
Use Eq. (13.16) to obtain the Born-approximation 1 = 1 phase shifts for this potential
and then assuming these phase shifts use Eq. (13.21) to retrieve the potential. From
Eq. (13.16)

_~ b1(k)
2m k
rv Vi ~.!...
4 Y;' k4
{I + ~_
2a
(1 + 3k2 + k4) exp {_
2a a2
k2}}.
a
226 Chapter 13
Hence

Substituting into Eq. (13.21):

It)(r) .i.- {CO It) f!. (15k _ lOP + 2k S ) exp (_ k 2 ) sin(2kr)dk


7rr3 10 4 V~ 203 a4 as a
= It)r2 exp (-ar 2).

References

1. F. Calogero and A. Degasperis, Spectral Transform and Solitons, North-Holland,


Amsterdam, (1982).

2. H. A. Mavromatis and A. M. AI-Jalal, J. Math. Phys. 31 1990, pp. 1181-1188.

3. A. ErdeIyi et al., Tables of Integral Transforms (Bateman Manuscript Project),


McGraw-Hill, New York, (1954), V. 1, p. 63.

4. 1. S. Gradshteyn and I. M. Ryzhik, Tables of Integrals, Series and Products,


Fourth (English) Edition, prepared by A. Jeffrey, Academic Press, New York
(1980), p. 422.
CHAPTER 14

The Dalgarno-Lewis Technique

In treating quantum mechanical systems which do not admit of exact solutions


for their energy levels and corresponding wave functions one has at one's disposal,
as mentioned in a previous chapter, the variational approach (Chapter 10). This,
however, is restricted to the ground and possibly first excited state (if the Hamilto-
nian commutes with the parity operator). For other states (and even for the ground
state if one requires greater accuracy or additional information against which to jux-
tapose the variational results) one must fall back on the straightforward application
of perturbation theory (Chapter 11) which is generally tedious and at best only ap-
proximate. This is because, for other than the first-order results, using perturbation
theory, one has to evaluate infinite sums in each order (d. Eqs. (11.14), (11.15) ).
These sums one generally approximates by selecting only a few of the terms which
arise to that order. One does not usually feel very comfortable with this truncation
procedure however since one is then uncertain whether or not one has obtained a
good enough approximation to the correct energy, or wave function.
In this context there does however exist a technique, first pointed out by Dalgarno
and Lewis 1 ,2,7 which in some cases allows one to do away with tedious summations
and gives exact answers to a given order.
The basic equation which defines the state-dependent operator involved, Fn , is:

(14.1)

where the Hamiltonian H of a given system is broken up into H o + h and the


eigenfunctions and eigenvalues of H o are ~n and EiO) = En, respectively. Moreover
Ei1 ) is just the first-order energy term for such a decomposition of the Hamiltonian
namely Ei1 ) = (~nlhl~n). The matrix elements obtained from Eq. (14.1) namely
(~ml{Fn' H o} I~n) are consistent for the diagonal case since, using the Hermiticity of
Ho,

and for non-diagonal matrix elements

(m # n) , (14.2)

(if one assumes (~nlhl~m) is real).


228 Chapter 14
From Eq. (14.1) Fn is seen to be dimensionless. Moreover if Ho commutes with
the parity operator (i.e. <P n, <Pm have either even or odd parity), and h has definite
parity, one sees from Eq. (14.2) that Fn has the same parity as h.

One-dimensional Systems:

Evaluation of the commutator in Eq. (14.1), for a one-dimensional Hamiltonian


H o, whose potential energy term does not involve the momentum, yields:

CIi n d2Fn + 2 dFn dCIi n = 2m(h _ E(l»)CIi n = 2..~ (CIin2dFn). (14.3)


dx 2 dx dx '11,2 n CIi n dx dx

I:
Integrating:
CIi n2(x) d~;x) = 2h~ 1 [h(x) - EA1)] CIi~(x) dx,
X

and if CIin(a) = 0, a further integration leads to the closed-form expression:

(14.4)

where K is an arbitrary constant.


One notes from Eqs. (14.1), (14.2) and (14.4) that Fn is clearly generally state
dependent. Moreover, as can be seen from Eq. (14.1), Fn is determined only to within
an arbitrary constant K, since {K, Ho} = O. Thus K in Eq. (14.4) can be chosen to
be zero for convenience. In Eq. (14.4), a is also a constant (which usually turns out
to be 7r/2, 0, 00, etc.) The choice of this lower limit a is however more problematic
than the choice of K since this results in an additional variable term for Fn, namely
Dr CIi n(x)-2dx where D is the integral - (2m/h2) r(h(x) - E~1»)CIin(x)2dx. The
correct choice for a may, however, be determined by noting that a must be such that
CIin(a) = 0 while Fn must satisfy Eq. (14.2) for any m # n. Another way to see why
there is this uncertainty is to note from Eq. (14.3) that dFn/dx is undetermined to
within a variable D/CIi n(x)2, since

i.e. Fn is undetermined to within D ICIi n(x)-2dx.


The usefulness of Fn(x) is obvious if one considers for instance the second-order
energy term (cf. Eq. (11.14) ):

= E v..m Vmn =E (CIinlhICIim) (CIimlhICIin)


m¢n En - Em m¢n En - Em

= E (CIinlhICIim) (CIimlFnICIin),
m¢n
The Dalgarno-Lewis Technique 229
and using closure:

1 = E I~m} (~ml +
m
Jdk I~k) (~kl,
(where the second term on the r.h.s. arises if Ho also involves continuum states):

(14.5)

a result which involves only two, or only one additional integral if for some reason
(e.g. parity considerations) Ei1) = 0, as opposed to an infinite number of matrix
elements.
Similarly the wave function to first order (cf. Eq. (11.15) ) is:

(14.6)

where N = 1 if one requires only that \lin is a cross-normalized function, i.e.

(~nl\lln) = 1, and N = (1+AF;)-1/2,

(where the variance of F n , AF~ == (~nlF~l~n) - (~nlFnl~n}2) if one requires

The third-order energy is:

m,pim

-Ei1) L (~nlFnl~m) (~mlFnl<lln) .


m:;n
Using closure here also, one obtains:

+2Ei1 ) (<IInlFnl<lln}2 - Ei1) (<lin IF; I<lin)


= (<IInlFnhFnl<lln) - 2Ei2) (<lin IFni <lin} - Ei1) (~nIF;I<IIn) , (14.7)

a result which needs only two additional integrals for its evaluation. As concerns the
expectation value of an operator 0, using Eq. (14.6), and assuming (<lin 10Fni <lin) is
230 Chapter 14
real,

(14.8)

where in this case N 2 = (1 + LlF; + ... )-1 . Some examples which illustrate the effi-
cacy of this technique are given below, while a concise listing of some of the Dalgarno-
Lewis expressions is given in Table 14.5.

EXAMPLE 14.1
Consider H = p2/2m + V, where

-anc/x + >.mw2x 2/2, x 2:: 0,


V= {
00, x ::; 0,

and >. is a dimensionless constant,

while a = e2/(47rfonc) is the fine structure constant a "" 1/137 which occurs fre-
quently in atomic physics. Treating p2/2m - anc/x as the unperturbed Hamiltonian
Ho, obtain the Dalgarno-Lewis function Fo(x). Use this to obtain the energy of the
ground state to third order, the wave function to first order, and two upper bounds
to the energy of this system.
For the partition Ho = p2/2m - anc/x, h = >.mw2x 2/2,

T
<I>o(x) = 2 J(amc)3 xexp (amcx)
--n- ,
and

E~1) = >. (<I>ol~mw2x21<I>o) = >'~(a::c2) 2a2mc2,


i.e. to first order in perturbation theory

(14.9)

One notes this series involves the (square of the) ratio of the energy-level parameter
(viXnw) for the Hamiltonian H' = p2/2m + h to the energy-level parameter for Ho,
The DaJgamo-Lewis Technique 231
(namely o?me2 ), and as ~xpected the bigger w, the more important the repulsive
first-order correction to the energy of this system.
Substituting in Eq. (14.4) with the convenient choice a = 00 one obtains:

which one can easily verify satisfies Eq. (14.1).


One readily obtains:

Hence substituting in Eq. (14.5),

E~2) = _129 A2{~}4Ot2me2


16 Ot2me2 '
a result again directly proportional to (in this case the fourth power of) w.
To obtain the exact energy to third order one must evaluate the two additional
integrals: '

4095 3{ fiw }6 2 2 55 2{ nw }4
(C)O 1FohFo I)
C)O = 16A Ot2me2 Ot me, and <C)o 1Fo21 C)O ) = '2A Ot 2 me2 •

Substituting into Eq. (14.7) one obtains

5451
Eo(3) _- -32 \3{
1\
fiw
-2--2
}6Ot2me2.
Ot me
Thus

The wave function to first order is:

where N = 1, or
232 Chapter 14
If one uses as trial wave function in Eq. (10.2), CI>o, the ground-state wave function
of Ho in the decomposition H = Ho + h, one obtains:
Eo = Eexact ground state :::; E~O) + E~l). (14.11)
If one instead uses as trial wave function in Eq. (10.2), Wo from Eq. (14.6),
i.e. the wave function correct to first order (with N = (1 + ~FJ)-1/2), one obtains:
E(2) E(3)
E E < E (O) + E(l) + 0 + 0
0= exact ground state - 0 0 1 + Lm#o Vom VmO / EO
(-)Em
2
E(2) E(3)
= E(O)
o
+ E(l) +
0
+ 0
1 + ~FJ .
0 (14.12)

All the terms needed to evaluate Eq. (14.12) are known in this case, yielding
Eexact ground state :::; E~O) + E~l)

2a
->.?o?mc {~}4
2 mc2
{129/16 - (545LX/32) [1i.w/a 2mc2]2}.
1 + (319.\2/16) [1i.w/a 2 mc2]4
(14.13)

Both Eqs. (14.9) and (14.13) are upper bounds to Eexact ground state ,and
Eq. (14.13) is a better upper bound than Eq. (14.9) if [VX1i.w/a 2 mc2 t< 86/1817.

EXAMPLE 14.2
Consider the system

H =p2/2m+ V

x ~ 0,

x:::; 0,

where .\ is a dimensionless constant.


Obtain the energy of this system to second order in perturbation theory using the
Dalgarno-Lewis technique.
It is reasonable to partition H into

p2 1
Ho = -
2m
+ -mw
2
2 2
x
'
and
Then

CI>o(x) = {7r/4~~4exp(-m;;2),
CI>l(X) ~{:;r/4 {2[v'T xf - 3v'T x} exp ( - m;;2) ,
The Dalgarno-Lewis Technique 233
)..ft2 E~l)
Eo(O) = ~ftw E~l) = 2)"ftw, and h(x) = -mx2 = 2mwx 2/ft .
2 '
If ).. = 1, E~O) + E~l) = 7ftw/2, as opposed to the exact answer in this case (see
Example 10.5 ), namely Eo = 5hw/2.
Evaluating the Dalgarno-Lewis function:

Fo(x) = J'" 1 { {'" ( mwx2) }


x 2 exp(-mwx 2/fi) la W(x)x 2 exp - - f i - dx dx

where
W{x) = ~~ {h{x) - E~l)}.
With the convenient choice a = 0 which satisfies <Po (a) = 0, and using6

Fo{x)

which satisfies Eq. (14.1) as one can easily verify. Substituting into Eq. (14.5) yields:

2[E~1)]2 {
10 10
00 00 }
Eo(2) -- Vifiw exp (_u 2 ) In u du - 2 u 2 exp (_u 2 ) In u du

[ (1)] 2
= 2 Eo {_ Vi (-y + 210 2) + 2Vi (-y + 21n2) _ Vi}
Vifiw 4 8 2
[ (1)] 2
= -~
fiw
= -4)..2fiw '
(where'Y rv 0.5772157, and one has used I;' exp (-ou 2 ) In u du = - (Vi/4)[ 'Y +
In 40 ] 0- 1 / 2 and derivatives ofthis expression with respect to 0 ). If).. = 1, E~2) =
-4fiw, and E~O) + E~l) + E~2) = 3fiw/2 + 2fiw - 4fiw = -fiw/2.
Substituting into Eq. (14.7) (using the expression 1000 exp (-ou 2){ln u)2du = (..;:;/8)
(b + In 40]2 + 7r 2/2)0- 1 / 2 and derivatives of this expression with respect to 0 ) one
additionally obtains
234 Chapter 14
Finally substituting into Eq. (14.12):

3 -4,X21iw + 16,X31iw
Eo = Eexact ground state::; '21iw + 2'x1iw + 1 + ,X2(7f2/2 _ 4) .

For'x < 1/4 this is a better upper bound than E~O) + E~l).
That the choice a = 0 is the correct one in this case can be verified by checking
that

(~olhl~l) = 2~ = (~olFol~l) .
fO - fl v6
Using this expression one can also compare:

E~2) = -4,X21iw = (~olhl~l) {~llhl~o}


fa - fl
+... = _ (~+
3
...) ,X21iW.
These results can be compared to be exact ground-state energy (cf. Example 11.8):

2 + v'1 + 8'\ ) (3
Eexact ground state = ( 2 1iw'" '2 + 2,X - 4'\
2
+ 16'x3 -... ) 1iw.

EXAMPLE 14.3
Consider H = Ho + h, where
p2 1 2 2 ;mw x
= ,X1iwYT
Ho = 2m + 2'mw x , h (all x).

Evaluate Fo(x), E~2), etc.


Since

as can easily be shown from parity considerations. A straightforward calculation


yields
Fo(x) = -,X [J~w x] ,
and

E o(2) -__ ,X21iw (3)


2 ' Eo = 0, \]! =N { \ Jmw
1 - /\ T x
}
~o.

One can also evaluate the exact ground-state energy in this case Eo = nw /2 -
,X 21iw /2 since
The Da,lgarno-Lewis Technique 235

~
H=-+-{/
2m
~
2
~
n
2
x+>.} _>.2_=_+_ x+
~
2
~
2m
~
2
{ If } h
->.
mw
2
~
_>.2_,
2

i.e. En = (n + 1/2) hw - >.2~/2.

EXAMPLE 14.4
Given H = Ho + h, where
p2 1 2 2
Ho = 2m + '2mw x ,
Find Fo(x), E~2), etc.
From Example 14.3 Ci>o and E~O) are known
E~l) = >.nw / Ci>o Imw2x21 Ci>o) ~ = >'hw ~nw ~ = A1i.w .
\ 2 nw 4 nw 2

= _ >.2 hw >.3 h
Fo(x) = -2">. [mw]
T x2, EO(2)
4 '
E(3) _
o - 4 w.
The exact ground-state energy in this case can easily be shown to be:

Eo = ~hwy'(l + 2>.) = ~hw {I + >. _ ~2 + ~3 + ...}.


EXAMPLE 14.5
Given H = Ho + h, where
p2 1 3/2
h = >'hw { ~w }
2 2
Ho = 2m + '2mw x , x3 (all x).

Obtain Fo(x), and E~2).


Using the expressions in Example 14.3 one easily sees E~l) = 0 from parity con-
siderations, while,
Fo(x) = ->. {[7f/2 ~3 + [~wr/2 x}, and E~2) = - l1S>'\w.
EXAMPLE 14.6
Given H = Ho + h, where
p2
Ho = 2m + 2"mw
1 2 2
x, and h = >'hw
Jmw
T x (all x ).
Obtain F1(x), E~2) etc.
Here
236 Chapter 14
One obtains E~l) = 0 from parity considerations.

Fl(X) = -). {Vrm:w [h


T x - V;:;;:;; , I} and El
(2) ).2
= (CPIIFhICPl) = -2"1iw,
which turns out to be the only non-zero correction to E 1 , while EP) =0 for this
system (cf. Example 14.3). .

EXAMPLE 14.7
Obtain F1(x), and E~2), if

Ho = :~ + ~mw2x2, and h = )'1iw {~W} x2 (all x).

Here

CPl(X) = {T
mw }3/4.,fi Xexp (mwx2)
7r1/4 ---v;:- , and E(l)
1
-- )'~1i
2 w.

In this case

mwx2
Fl(X) = -).--v;:- ,
(the same result as for Fo{x) of Example 14.4), an expression which can easily be
shown to satisfy Eq. (14.1). Using F1{x) one obtains E~2) = _).2 3Tiw/4, which agrees
with the).2 term in the expansion of the exact first excited-state energy (31iw/2) (1+
2).)1/2 in powers of ). (cf. Example 14.4). (The Fn{x)'s for this system if n > 1 are
more complicated than Fl(X) = Fo{x). )

A useful property of the Fn function is that it is additive. If for a given unperturbed


Hamiltonian H o, the perturbations hI, and h2 individually result in first-order energies
E!(l) , E;Y) and lead to Dalgarno-Lewis functions F~ and F~ respectively, then, for
the case h = hI + h2' the defining Eq. (14.1) is satisfied by Fn = F,! + F~ since
Ei1 ) = E!(l) + E!(l). Thus one can sometimes use results from simpler problems in

more complex situations.

EXAMPLE 14.8
Given

Find Fo(x), F1(x) etc.


Combining the results of Examples 14.3 and 14.4,
The Dalgarno-Lewis Technique 237

Fo(x) =
and from Examples 14.6 and 14.7,
-AlJ~W x - ~2 {J~w x r,

E(2) = -nw {A~ + A~}


o 2 4 '
I.e.
Eo nw (1 + A2 -
="2 {2Al +"2A~} +... ),

in agreement with the exact energy results:

I.e.
I, nWAi
Eexact ground state = 2nwy 1 + 2A2 - 2(1 + 2A2) ,
3, nWAi
Eexact 1st excited state = 2nwy 1 + 2A2 - 2(1 + 2A2) ,
expanded to order A2 •

EXAMPLE 14.9
Given the Hamiltonian:

H = - - - +A
p2
--,
anc n2
x 2:: o.
2m x 2mx 2
Treating h = An 2 /2mx 2 as a perturbation, study the energy of the ground state of
this system. The unperturbed ground-state and first-excited state wave functions are

iP o(x)=2V FIT-n-J
amc] 3
xexp
( amcx) ,iPl(X)= ~ [~]3 X(I- :.~) exp(-~)
--n- 2 n 2 n 2n'
and the unperturbed ground-state energy is
238 Chapter 14

One immediately evaluates the first-order energy contribution:

The exact energy if ). = 2 is -a.2mc2 /8, which (E~O) + E~1) being an upper bound)
is less than (-1/2 + 2)0:2mc2 = 30: 2mc2/2.
Using Eq. (14.4) (with a = 00, for which value 4>o(a) = 0) one obtains Fo in this
case:

F.o(x ) = ).~
Ii x + ),1n 20:mcx
Ii .

Hence evaluating Eq. (14.5) in this case yields:

where one uses the integral

1o
00 'Y In 0:
exp ( -o:u) In u du = - - - -
0: 0:
,

and derivatives of this expression with respect to 0:, and the constant 'Y '" 0.5772157.
Using Eq. (14.7) one can also evaluate E~3) with the help of the integral

and derivatives thereof with respect to 0:. This calculation yields:

It is interesting to confirm the choice of a = 00 is the correct one in this case.


Thus
(4)0Ihl4>1) = -). 29 / 2 = (4>0 IFol 4>1) .
1"0 - 1"1 33
One can then also compare
The Dalgarno-Lewis Technique 239
Finally one can compare the first-, second- and third-order contributions to the exact
ground state in this case:

EXAMPLE 14.10
Given the Hamiltonian H = Ho + h, where

Ho = :~ + ~mw2x2, and h = .\nw{ J~w xr (all x).

Obtain Fo(x), E~l), E~2), E~3) etc., noting that

<llo(x) = { mw}1/4
7rh
(mwx2)
exp - ' j j t .

By using

lo w exp (-w )dw = exp (_u L: (k )" '


u
2 2 2)
00 2k-lU2k+l

o 2 + 1 .. k=l
2k-2U2k+1
(-w )dw = 3exp (_u L:
[U 00

10 w 4 exp 2 (k )'"
2)
o k=2 2 + 1 ..

(results which may be obtained via integration by parts) in Eq. (14.4) one obtains

(14.14)

With the help of Eq. (14.14) one can then obtain:

{<llolhl<llo} = ~.\nw, (<lloIFol<po) = -196.\' (<PolF~l<Po) = ~~~A2,


(<llolFohl<po) =- 1:45.\2nw, (<PoIFohFol<po) = 2150521: .\3 nw.
Thus the ground-state energy to third order can be written:
1 3 21 2 333 3
Eo = -nw
2
+ -.\nw
4
- -.\
8
nw + -.\
16
nw.
On the other hand substitution in Eq. (14.11) yields:

Eo ~ {~ + ~.\ } nw, (14.15)


240 Chapter 14
while substitution in Eq. (14.12) yields:

< {I 3} (14 - 11L\) 3).2


Eo - '2 + 4'). hw - (1 + 39).2/32) 16 hw. (14.16)

If ). = 14/111, then to third order Eo = (66/111) hw = 0.594 hW, while at the


same time the exact ground state is less than or equal to this value.
If ). > 14/111, the third-order contribution is greater than the second-order con-
tribution, and Eq. (14.16) gives a worse approximation than Eq. (14.15) for the
ground-state energy. However, for)' < 14/111, Eq. (14.16) gives a better upper
bound to the ground-state energy than Eq. (14.15).

EXAMPLE 14.11
Given the Hamiltonian H = Ho + h, where
Ho = ;~ + ~mw2x2, and h = )'hw {j";,.w x} 4 , (x ~ 0).
Obtain Fo(x), etc.
Using
(U
10 w6 exp (-w 2 )dw = 15 exp (_u 2 ) E
00 2k-3U2k+1
(2k + I)!! '

in addition to the integrals in Example 14.10, and substituting in Eq. (14.4) with
a = 0 one obtains:

(14.17)

Hence
E(l) _ 15)'h
o - 4 w,
Evaluating
(q>olhlq>l) = _~). = (q> IF. Iq> ).
-2hw 2V6 0 0 1

This verifies that a = 0 (for which value q>o(a) = 0) is an acceptable choice and
one has:

EXAMPLE 14.12
Given the Hamiltonian H = Ho + h, where
The Dalgarno-Lewis Technique 241
Obtain Fo(x), E~l), E~2), etc.
By straightforward calculations one obtains:

o - ~An
E(l) -
2 w, Fo(x)
rmwx }2 ,
= -2"). {VT
Evaluating

{q,olhlq,l} = ).nw (q, I{jmw x}21 q, )


-2nw -21iw 0 . Ii 1
= {q, 0 1F.1q,
0 1
} = __3 A.
2V6
Hence:
= {q,olhlq,l} ~q,llhlq,o} + ... = 29 ).2[-2nw] _ ....
_~).2Iiw
4 -2 w 4
Thus in this case all the second-order contribution comes from the first-excited
state. The exact energy in this case can be obtained by writing

~ 1 ~ 1
H = - + -mw 2(1 + 2).)x2 = - + _mw'2x 2.
2m 2 2m 2
For this Hamiltonian, Eo = 3Iiw'/2, where w' = w(1 + 2).)1/2. Thus
3 3 3 2 3
Eo = -nw + -)'liw - -). nw + -).3 nw - ....
2 2 4 4

EXAMPLE 14.13
Consider a one-dimensional system with Hamiltonian:

H :m +
2
Vex)
2
m e 3
Vex) = ).--x (Ixl < b), (14.18)
Ii
= 00 (Ixl > b),
where). is a dimensionless constant.
The obvious perturbation-theory decomposition of this system, (which cannot
be treated exactly analytically) is:
p2
Ho =
2m
hex) ).me2 {mc
fiX } , (Ixl < b), (14.19)
00, (Ixl > b).
Here
242 Chapter 14
and E(O) = (n + 1)2n211"2
n 8mb2
One immediately obtains

(14.20)

since h(x) has odd parity. (Using Eq. (14.7) one also obtains Ei3) = 0, in this case
since h and Fn have the same parity.) Substituting into Eq. (14.4) for Fo(x),

Fo(x) = 2m >.m 2 c3 [2b] 3 j1rX/2b -1-1Y' «('cos 2 (') d('dy'


n2 n 11" cos 2 y' 1ra/2b

= SA ( ~: )' rOb "",' y' {(~' + ~(2(',", 2(' + eo, 2(')}1:~/" dy'
= mcb)3j1rX/2b {(y' - 2"1) sec y' + 2y' tan y' + 1
4>. ( n1l"
2

r
2

-sec2 y' ( {;: + {;:} sin [7rba ] + cos [;a/b1) } dy'. (14.21)

Here an ambiguity arises because there is no obvious choice for a. Two choices
suggest themselves, namely a = 0, and -b. The first choice however does not satisfy
the condition cp( a) = O. The second choice (a = -b) results in:

Fo(x) = -2>'2 {mCb}3


- - (X2
{7r - - 1) tan -7rX + -x} (14.22)
7r n 2 b2 2b b'

as opposed to

if a = O. This ambiguity arises because in Eq. (14.3) dFo(x)/dx in this case is


uncertain to within D sec2 (7rX /2b), i.e. Fo( x) is uncertain to within the variable
(2bD / 7r) tan (7rX /2b), with D an unknown constant. To determine the constant D
one notes for instance that a condition on Fo is

(14.23)

Evaluating this expression shows that the choice of Eq. (14.22), namely a = -b is
the correct one. Once one has evaluated this expression one can also immediately
The Dalgarno-Lewis Technique 243
display the first term contributing to the second-order energy, namely,

E (2) _ («polhl<<pl) («pllhl<<po)


o - (0)
Eo -El
(0) + ...
= -3511"6).2me2
213 {meb}4
T +... (14.24)

One may additionally confirm that «(«PO\FOI«P2) being zero from parity considera-
tions),
(14.25)

Thus
213 ).2mc2 [mCb] 4 _ 215 ).2me2 [mCb]4
I53 5 11"6 h 3555 11"6 h

=
,\2mc2 [mCb]
-(33.7119 + 0.0432 ... )---;;:s T 4 (14.26)

-33.7551 A;;;e
\2 2 [
m;b] ,
4

versus the exact second-order contribution from Eq. (14.5):

E o(2) -
-

showing that almost all the correction in this case comes from the first-excited state
and incidentally that :
212 {I 32 42 }
11"2
(1 - 11"2) 22
15 = 5"" (1 X 3)5 + (3 X 5)5 + (5 X 7)5 + (7 X 9)5 + ... (14.28)

is a very quickly converging series relation (giving 5 digit agreement after three terms,
and 6 digit agreement after four terms) which to the knowledge of the author is not
quoted in the literature3 , 4.

EXAMPLE 14.14
Consider a one-dimensional system with Hamiltonian:
p2 o.he
H = ---+V(x)
2m x
,\ 0. 2me2(20.me) = 1, 2... ),
Vex) = -h-x n , (x ~ 0; n (14.29)

= 00, (x < 0),


244 Chapter 14
where A is a dimensionless constant. The obvious perturbation-theory decomposition
of this system, which cannot be treated exactly analytically (See however Example
14.15 where the problem can also be solved analytically if n = -1.) is:

afie
Ho =
2m x
h(x) =
\
AO:
2
me2(2ame)n
-fi-x , (x ;::: 0; n = 1, 2 ... ), (14.30)
00, (x < 0).

Here

ame)3/2 ( ame 2 a 2m 2 e2) (amex) E(O) __ a 2me2 1


(
iP 2 (x)=2"""'3h x 1- 2x """'3h+ 2x 27fi2 exp -31i' q - 2 (q+1)2·

One immediately obtains:

Eo(l) -_ 4\(ame)3
A -- a 2 me2la (2ameX)n
OO
- - - exp (2ameX)
- - - - x 2dx
fi 0 fi fi
(14.31 )

SUbstituting into Eq. (14.4) for Fo(x) with the choice a = 00, for which value
iPo(a) = 0,

Fo(x) =

(14.32~

where use was made of the expression5

J xnexp(-x)dx = -exp(-x) E
n n!
(n _ k)! x n - k
The Dalgarno-Lewis Technique 245
If one substitutes Eq. (14.32) in Eq. (14.1) one sees that this latter equation is
satisfied by the Fo(x) obtained above. Also one can confirm that the choice a = 00
in Eq. (14.32) is the correct one by verifying (for various n's) that for this value of a:

(14.33)

E o(2) --

(14.34)

whereas the exact second-order result from Eq. (14.5) is:

E(2)
o
= _,\2a 2 mc2 (n
23
+ 2)! I:
k=O
2(2n +3 -k)! - (n + 2)!(n + 3 - k)! .
(n+2-k)!(n+1-k)
(14.35)

A feeling for the rate of convergence may be obtained by comparing the exact
second-order result for the case n = 1, where Fo(x) = -'\[amcxJn.J2,

E~2) = (<I>o\hFo\<I>o) - E~l) (<I>o\Fo\<I>o)


_15,\2a2mc2 + 9,\2a2mc2 = _6,\2a 2mc2, (14.36)

versus the first two terms in the second-order expansion in this case, namely

Thus for n = 1 the first term is little over half the total second-order value which
also includes a contribution from the term

since there are also continuum states <I>k in this basis. As noted previously Eq. (14.5)
also automatically includes this contribution. For n = 3 however one obtains E~2) =
-42,780,\2a 2mc2 from Eq. (14.35), versus the first two second-order terms, Eq.
(14.34) rv -37, 883.15,\2a 2mc2 - 2, 432.74,\2a 2mc2, indicating that the convergence
improves with increasing n.
246 Chapter 14

EXAMPLE 14.15

Consider the Hamiltonian in Example 14.14


p2 exne
H=2m---;;-+V(x) (14.37)

but with

Vex) = )..ex 2 me2 Cex;ex)-1 (x ~ 0), (14.38)


= 00 (x < 0).
Assuming hex) = Vex), the same unperturbed energies and eigenstates arise as
in Example 14.14 and
\ 2 2
E(l) _ "ex me
o - 2 (14.39)
Additionally, with a = 00 (for which value ~o(a) = 0)

Fo(X) = )..ex 2me2


2ex 2me2
J 2Otm C:C/fi 1
y/2 exp (_y')
jyl (e' _(/2)
00 2
exp ( _(') de'dy'

).. J2Otmcx/fi 1
- - dy'
2 2
).. exm ex
= (14.40)
2n
where hFo = )..2ex 2 me2 /4 is a constant in this case. This yields:
E~2) = (~olhFol~o) - E&l) (cJ>olFolcJ>o)
)..2 ex 2me2 3)..2ex2me2
=
4 8
)..2ex2me2
= (14.41)
8
Using Eq. (14.7) one also obtains

E~3) = (cJ>o lFohFol cJ>o) - 2E~2) (cJ>o lFo I cJ>o) - E~l) ( cJ>o IF~ I cJ>o)

3 3 3}
>. 3 ex 2 me2 { 16 + 16 - 8' = O.

In this case one can also trivially solve this problem exactly by writing the
Hamiltonian as
The Dalgarno-Lewis Technique 247
where a' = a(1 - A/2), which has exact ground-state energy:

in agreement with the result obtained by the Dalgarno-Lewis approach above, veri-
fying also that the choice a = 00 was the correct one. One may also write:

indicating the series is slowly convergent in this case.

EXAMPLE 14.16

Consider the Hamiltonian H = H o + h, where

p2 1 2 2
Ho=-+-mw x (all x).
2m 2 '

Evaluate F1 (x), EI2), etc.


One notes that:

CP1 (x) = {7} 3/4 ~/: exp ( - m;;2) , and E~l) = 0, from parity considerations.

Substituting in Eq. (14.4) one obtains:

F1(X) = -A [~(IT X)3 +2 (IT x) (IT x) -1]


-2

as one can easily verify by substitution in Eq. (14.1), and by noting that

{CP11 h Icpo} = 30 A = {cp IF, Icp }


EIO) _ E~O) 4 1 1 0·

Finally using Eq. (14.5) one calculates:

71 2
E1
(2)
= {CP11 hF1 ICP1} = -SA nw.
248 Chapter 14

EXAMPLE 14.17

Find F2(X) for

p2
Ho= -+-mw x
1 2 2
and mw
h(x) = ).}iw { r;:x 2} .
2m 2

Hence obtain E~2).


One notes

1
<P2(X) = ( 2";; )1/2 (mw)1/4
T [2 (mw)
r;:X2 - ] (mwx2)
1 exp -~ .

Direct substitution yields

Evaluating Eq. (14.4) in turn yields:

).[mw2 2 ]
F2(X) = -"2 r;:x - 2(mw/n)x2 -1 '

a result which may be verified by substitution in Eq. (14.1).


Hence

while
E2 exact = 2"5 nw .J1 + 2), rv nw [52" + 2").
5 5 2 + ....
- 4). ]

EXAMPLE 14.18

Find F2(X) for

Ho = ;~ + ~mw2x2 and h(x) = ).nw {J~w x}.

Hence obtain E~2).


One notes

1
<P2(X) = ( 2";; )1/2 (mw)1/4
r;: [2 (mw)
r;:X2 - ]
1 exp
(mwx2)
---v;- .
The Da,lgarno-Lewis Technique 249
From parity considerations one concludes that E~l) is zero.
Evaluating Eq. (14.4) in turn yields:

F2 {x} = -..\
rmw x -
[VT 4~x
2 {mw/h}x 2 -1 '
1
a result which may be verified by substitution in Eq. {14.1}.
Hence in this case

On the other hand

If one wishes to extend the above Dalgarno-Lewis techniques to fourth or higher


order this can be done if one defines some additional operators.
Thus

(14.43)
+ [E(l)] 2 ""' hnihin _ E(2) ""' hnih in
n L...J .3 n L...J t.2 '
i"#n tn. i"#n m

where for any operator 0, Oij == (<I>;j0I<I>j) , and tij == ti - tj .


In particular up to fifth order, two more operators G n , In are needed in order to
simplify Eq. (14.43), and the corresponding fifth-order expression. The equations
these two operators satisfy are:

I.e.

and

I.e.
(mIn).
250 Chapter 14
(One easily confirms that the operators Gn , and In are consistent for the diagonal
case (m = n). ) The equivalent integral expressions, analogous to Eq. (14.4) for Fn,
are:

IX <Pn~y)2 {21i~ l
and
In(x) = Y
[Fn«() - [Fn]nn] <Pn«()2d(} dy.

The first term in Eq. (14.43) for E~4) (assuming hij = hji ) can thus be written

-{[Fn]nn}3hnn - 2[Fn]nn[hGn]nn + 2[Fn]nnhnn[Gn]nn,


while the second term becomes:

The third term reduces to

and the fourth term becomes:

Thus

Ei4) = (lGnhFn]nn - [Gn]nn[hFn]nn + {[Fn]nn}2 [Fnhlnn

-Eil) {[Fnlnn}3 - 2[Fnlnn[hGnlnn + 2Eil} [Fnlnn[Gnl nn )


(14.44)
-2Eil) {[FnhInlnn - [Fnhlnn[Inlnn - [Fnlnn[h1nlnn + Eil) [Fnlnn[Inlnn }

+ {Ei1)r {[InFnlnn - [Inlnn[Fnlnn}

By a similar manipulation the wave function to second order, Eq. (11.15) can be
cast in the form:

(14.45)
The Da.lgarno-Lewis Technique 251

EXAMPLE 14.19
Obtain the fourth-order contribution to the energy of the ground state of the
system with the Hamiltonian in Example 14.10, assuming the same decomposition
H=Ho+h.
After some calculating one finds:

(14.46)

while
Io(x) = 16~w [{ ~xr +9{ ~xr]· (14.47)

Using the expression for Fn(x) obtained in Example 14.10, one evaluates seven
additional straightforward integrals:

Three-dimensional Systems:

The difficulty with three-dimensional systems is that generally one cannot write
down a closed-form expression analogous to Eq. (14.4) for the three-dimensional
counterpart of Fn. One must therefore fall back on Eq. (14.1).
If
252 Chapter 14
and VCr) is purely central, then

If, in addition, the perturbation

h(r, (), 4» = h(r)Y,!:,((), 4»,

then the counterpart of Fn can be written

The commutator relationship in this case becomes:

which is generally a complicated expression. In the simple case [ = 0 (and conse-


quently also m = 0) however, this reduces to:

which in turn reduces to Eq. (14.3) if [' = o.


EXAMPLE 14.20
Given V(r) = -ahe/r. The ground-state unperturbed wave function is:

UlO(r) = ame]
2VFIT-h-J
3
(amer)
rexp --h- ,

while

1 [2 Fame]5 2
IT-h-J r exp (amer)
1 2 2 1
U21(r) = "4V"3V --v;- , and t I
2 me-
= --a n2·

a
n

Suppose
h = 'xa 2me2 [a~ez] = 'xa2me2a~crYr}((),4».
This potential corresponds to the perturbation introduced by a constant Electric
Field Eo in the z direction since such a field is equivalent to a potential -zIEol. Find
E 10
(2)

For this h one can easily verify that

f 10l ( r ) = _A amer
Ii
(~
2h + 1)'
Tbe Dalgarno-Lewis Tecbnique 253
satisfies Eq. (14.48), where [' = 1, m' = O.
The first-order energy E~~) for (the ground state of) this system is zero due to
parity considerations. Hence

(2)
ElO = ( <1>100 IhFnlI' I<1>100 ) = - 27 >. 20: 2mc2.
1611"

To verify that nothing is amiss here one can evaluate:

(<I>lool>.0:2mc2 (o:mcr/fi) Yl(O, ~)I<I>2lO)


- (1/2) 0:2mc2(1 - 1/4)

= ( cP 10o l- >. o:~cr (o:;cr + 1) Y~(O, ~)1<I>2lO)


219/ 2
= 311/ 2 .,fi>'·

This expression in turn is related to the first term in the second-order energy expan-
sion for El~). Thus:

EXAMPLE 14.21
Consider the same unperturbed Hamiltonian as in Example 14.20, i.e.:

V(r) = _ o:fie .
r

If

verify that
f 10l' ( r ) = ->. [o:mer]" (_1_~ ~)
fi I' + 1 .", + I' .
Note that Example 14.20 is just a special case of this result when 1'=1.
This result may be verified by direct substitution in Eq. (14.48) with m' = 0,
where, as in Example 14.20, El~) = 0, and

d
dr {lnu 10(r)} =
(1;:- - o:me)
-.",- .
254 Chapter 14
Here

= ->.2a?mc2~ [00 u 2 /' [ 1 u + .!.] 4u 2 exp (-2u) du


471" 10 (1' + 1) l'

=
>.2a-.2mc2(21' + 2)! (2l'2 + 51' + 2) (14.49)
22/'+471" (l' + 1)l' .

EXAMPLE 14.22
Given the unperturbed Hamiltonian

and

h= h >'1iwyrmw
V{3 T Z = >.1iwYo1 (0, <I»yrmw
T r.
· E(2)
H ence 0 b trun 00'
The relevant unperturbed ground state here is:

uoo(r) =
2
71"1/4
[mw]3/4
T rexp
(mwr2)
----v;- ,
while

U01 (r) = V/8


3;i72 [mw]
T 5/4
r2 exp
(
-
mwr2) ,
---v;-
and the unperturbed energies are

(0)
Enl = ( 2n + 1+ 2"3) 1iw.
Whence it is easy to show E~~) = (<Pooolhl<pooo) = O.
One can easily verify:
fJo(r) = ->.J7 r.

Hence

E~~) = (<Pooo I>.J~w rYo1 (0, <1» 1iw ( ->'Yo1 (0, <I»J~w I
r) <Pooo)

= _ \ 231iw (14.50)
A 871" •
The Da,lgarno-Lewis Technique 255
Also direct evaluation confirms that:

= -f[;'\.
This expression in turn is related to the first term in the second-order energy
expansion for E~~). Thus:
_,\23nw = (cJ>ooolhlcJ>OlO) (cJ>01olhlcJ>ooo)
8~ -nw

= _,\23hw.
811"
Thus in this case only one term in the infinite series contributes to this sum, since

EXAMPLE 14.23
Consider the same unperturbed Hamiltonian as in Example 14.22, but let

h = ,\hwYci' (0, <1» [Vrmw ]11


T r , (1' = 1, 2... ).
Verify that
f~~(r) = -~ [Jm; r]11
Note that Example 14.22 is just a special case of this result when ['=1.
This result may be verified by direct substitution in Eq. (14.48) where just as in
Example 14.22
d
dr {In ulO(r)} = ;:- - -11,- .
(1 mwr)
As in the previous three examples E~~) =0 due to the integral over 0, while,

= _ ,\ 2 1iW J:.--"'!" {'X> u 211+2exp (-u 2)du


[' 4~.fi 10

,\21iWr([' + 3/2) (14.51 )


=
21'~.fi
256 Chapter 14

If both the unperturbed potential V(r) and the perturbation h(r, B, </J) = h(r)
are central, the problem of obtaining F reduces essentially to the one-dimensional
case, since then Fn -+ fnl(r), and the commutator relation becomes

I.e.
[Un! (r ) d2fnl(r) + 2Unl(r)dfn!(r)] = 2m (h( )_E(1») ()
dr2 dr dr 1i2 r n! Un! r

and with the choice unt(a) = 0,

fn!(r) - _ JT u;k) t1" iT [h(r) -


1 {2m a
(1)] 2
En! Unl(r) dr } dr. (14.52)

If n' -# n

ii/l' iimml roo Unll/(r)


10 r
fnt(r) Unl(r) r 2dr (E~~) -
r
E~~l,) = {n'l'm'lhlnlm}

I.e"
(14.53)

EXAMPLE 14.24
If
H.o = (J;? + ~mw2r2 and h(r) = >'1iw {";;' r2},
2m 2 '
Evaluate E~~), where

u01(r) = Vf8{mw}5/4
3;I72 T (mwr2)
r2 exp --v;- , and

Directly one obtains:

=
8>.1iw roo (2) 6 8>.1iw 15y'7r
31r 1 / 2 Jo exp -u u du = 31r 1/ 2 ~
The Dalgarno-Lewis Technique 257
Substituting in Eq. (14.52) and using the integrals given in Examples 14.10, 14.11
one obtains for this case :

Hence

E 01
(2) -
-
>.mw r 2) >.'hw {mw
JofOO ( -"2T T r 2} U 2 ()
lO r dr -
(1)
E01
>.mw r 2) U 2
JofOO ( -"2T ( )dr
lO r

4>.2hw foo 10>.2hw foo () 5


= - 311"1/2 Jo US exp (-u 2)du + 37r1 / 2 Jo u6 exp _u 2 du = -4">.2 hw .

The exact solution here is:

Eexact g.s. (/=1)

The main results of this chapter are summarized in Tables 14.1, 14.2, 14.3 and
14.4.

References
1. A. Dalgarno and J. T. Lewis, Proc. Roy. Soc. A233 1955, p. 70.
2. C. Schwartz, Ann. Phys. 6 1959, p. 156.
3. L. B. W. Jolley, Summation of Series, Dover Publications (1961), p. 65.
4. 1. S. Gradshteyn and 1. M. Ryzhik, Tables of Integrals, Series and Products,
Fourth (English) Edition, prepared by A. Jeffrey Academic Press, New York
(1980), p. 7.
5. Op. Cit., p. 92.
6. Op. Cit., p. 306.
7. H. A. Mavromatis, Am. J. Phys. 59 (8), 1991, p. 738.
258 Cbapter 14
TABLE 14.1: One-Dimensional Systems

Interaction Perturbation h( x) Fo(x) Example

V(x) = ~mw2x2 AttW [t¥ x] -A [t¥ x] 14.3

(all x) >.1iw [t¥ xr -~ [t¥ xr 14.4

Anw [t¥ xr -~ {[t¥ xr +31¥ x} 14.5

Anw [I¥ xr -~ {[~ xr +3 [t¥ xr} 14.10

V(x) = ~mw2x2 Anw [~ X]-2 2Aln [~ x] 14.2

(x ::: 0) Anw [~ X]2 xr


-~ [t¥ 14.12

= 00 (x ~ 0) Anw[~Xr -~{[~ xr +5 [~ xr} 14.11

V(x) = -ancx- 1 Ali; [t¥ xt _~


6
{mw2 x3 + 3w2 X2}
IicOi ~ 14.1

(x ::: 0) Ali; [~ xr 2 ~ { [2"';.= x] + 21n [201;:'C x] } 14.9

= 00 (x < 0) Aa2mc2 [201;:'C x r -


).(n+2)! En-1
2
1
k=O (n+2-k)!(n+1-k)

(n::: 1) X [201;:'C X]n+1- k 14.14

Aa2mc2 [201;:'C X]-1 ~ [201;:'C X] 14.15

V(x) =0 (Ixl < b) Amc2 [~cx] ~ [!!!£br [{ x2b2 - 1} tan ~


"Ii
~]
2b + "b 14.13

= 00 (Ixl > b)
The Dalgarno-Lewis Technique 259
TABLE 14.2: One-Dimensional Systems

Interaction Perturbation h(x) Fl(X) Example

V(x) = ~mw2x2 Ahw{/¥ x} -A [{/¥ x} - {/¥ xrl] 14.6

(all x) Ahw{/¥ xV -~{fF xV 14.7

Ahw{/¥ xr -A [~{/¥ xr +2{/¥ x}

-2{/¥ xrl] 14.16

TABLE 14.3: One-Dimensional Systems

Interaction Perturbation h(x) F2(X) Example

V(x) = ~mw2x2 Ahw{/¥ x} F2(x) = -~ [ m;x2 - 2(mw/~)x2_1 ] 14.17

(all x) Ahw{/¥ xV F()-


2x - - A[~
T x- 4y,nwzt;X]
2(mw/li)xLl 14.18
260 Chapter 14
TABLE 14.4: Three-Dimensional Systems

Interaction Perturbation h(x) n [ [' F~~(r, 0, </J) Example

V(r) = _a:c Aa2mc2a~crYo1(O, </J) 1 0 1 - >. £!1!!£


Ii
r{£!1!!£
Ii 2
1: + 1}

14.20
xYo1 (O, </J)

(r 2:: 0) >.a 2mc2 [a~cr]


I'
1 0 [' ->. [~rll'
Ii
{~_r_
Ii 1'+1
14.21
xYJ'(O, </J) +t}Yd'(O,</J)

V(r) = mw;r2 Xhw/¥ rYOl(O, </J) 0 0 1 ->./¥ rYo1(0, </J) 14.22

(r 2:: 0) Anw [/¥ r)" Yci' (0, </J) 0 0 [' -~ [J¥ r)" Yd'(O, </J) 14.23

>.nw [J¥ rr 0 1 1 -~ [/¥ rr 14.24


The Dalgarno-Lewis Technique 261
TABLE 14.5: Some Dalgarno-Lewis Expressions

k == h - (cJ>nlhlcJ>n) (( cJ>nlhlcJ>n) = 0)

Fn == Fn - (cJ>n\FnlcJ>n) (( cJ>nIFnlcJ>n) = 0)

(N = 1 - (cJ>nlwn) = 1)
Wn = N [1 + Fn + ... J cJ>n
(N- 2 = 1 + (FJ) + ... _ (wnlw n) = 1)

Ei2 ) = (cJ>n IkFnlcJ>n)

En(3) = ( cJ>n\FnhFnlcJ>n
• •• )

(wnIOlwn) = N 2 ( (cJ>nIOIcJ>n)
(N- 2 = 1 + (1';) + ...)
+2 (cJ>nIOFnlcJ>n) + (cJ>nIFnOFnlcJ>n) + ...)
CHAPTER 15

Angular Momentum and Coupled States

In working with the orbital angular momentum operator:

L == r x p,
the major difference between classical and quantum mechanics is that whereas clas-
sically the vector product of L with itself is zero, i.e.
Lx L = 0,
in quantum mechanics
Lx L = iliL,
or the commutator:

[L;, L j ] == L;Lj - LjL; = iliLk , (i, j, k = x, y, z or cyclic permutations). (15.1)

This is because of the non-commutivity of x and p in quantum mechanics i.e. due to


the fact that the commutator
Ii
[PO" a] = -:-,
z
(a = x, y, z).
As expected, if Ii ~ 0, one goes to the classical limit.
Thus, one cannot simultaneously have eigenfunctions \]! of say Lx and Ly:

(15.2)

(15.3)
since operating on Eq. (15.2) with Ly and on Eq. (15.3) with Lx and subtracting,
one obtains

(the eigenvalues Ax, Ay being commuting numbers), a result which is inconsistent


with Eq. (15.1).
One can however find simultaneous eigenfunctions of L2 = L . L = L; + L~ + L;
and say Lz since despite Eq. (15.1)
Angular Momentum and Coupled States 263
This follows since [Lz 2, Lz] is trivially zero and the sum of the two commutators
[Ly Lz) + [Lz Lz)
2, 2, (though not the individual commutators) is zero. Thus one can
simultaneously solve the pair of equations:

(15.4)

mhY~(O, 1». (15.5)

With this choice of operators (as opposed say to L2, Lx),


Y~(O, 1», the simultaneous
eigenfunctions of Eqs. (15.4), (15.5) are the spherical harmonics, where 1 = 0, 1, 2 ...
and m = -1, -(1-1), -(1- 2), ... -1,0,1, ... 1- 2, 1-1, 1. The spherical harmonics
are well-known functions which have been extensively studied. Table 15.1 lists the
first few of these Y,;,(O, 1»'s.
The spherical harmonics constitute a complete set and satisfy the usual closure
and completeness relationships. Thus analogously to Eq. (2.6) one may write:
00 I
S(cosO-cosO') 15(1)-1>') = L L Y;;/(O', 1>')Y,;,(O, 1», (15.6)
I=Om=-1
and a two-dimensional generalization of Eq. (2.3) applies, namely for 0 ::; 0 ::; 7r,
o ::; 1> ::; 27r:
00 I
1/;(0,1» = L L almY~(O, 1», (15.7)
I=Om=-1
where analogously to Eq. (2.5)

aim = ["'=21r [e=1r Y;;,z (0,1> )1/;( 0, 1» sin OdOd1>. (15.8)


J",=o Je=o
Integrating Eq. (15.6) over 1> and noting that

(21+1)(l-m)! mi.
47r(l + m)l (-1) Pm ( cos 0) exp (zm1»,

(15.9)
pJ( cos 0) PI ( cos 0),

yields:

00 I [21r 1 00
S(cos O-cos 0') = L L Jo y::.I(O,1»Y~(O',1>') d1> = - L(2l+1)PI(cos O)PI(cos 0') .
1=0 m=-I 0 2 1=0
264 Cbapter 15
Moreover if ()J = 7rj2, defining u == cos(), in the range -1:::; u:::; 1 one obtains:
1
2 L:(21 + I)P1(u)P1(0)
00

c5(u) = . (15.10)
1=0
One notes that only even l's contribute to this infinite series since P1(0) is zero
for odd Z. The result of truncating the sum in Eq. (15.10) at [ = 12 is illustrated in
Figure 15.1, which looks very similar to Figures 2.1-4 .

6 (u)

-1.

Figure 15.1: Eq. (15.10) terminated at l = 12.

EXAMPLE 15.1
Show that the delta function in Eq. (15.10) is properly normalized, i.e. that
f~1 c5( u) du = 1.

1 c5(u) du
-1
1 1 00

= - L:(21 + 1)
2 1=0
1 PI(U) du PI(O)
1

-1
1 00 2
= - L:(2Z + 1) -Z- c510 PI(O) = 1,
2 [=0 2 +1
where one uses the useful orthogonality property of Legendre polynomials namely
that

-1
1
1 P[(U)PI'(U) du = -2[2 c51l',
+1
and the fact that Po(u) = 1. One notes that no matter at what [ one truncates
the series in Eq. (15.10), the normalization condition is always satisfied by this
expression.
Angular Momentum and Coupled States 265

EXAMPLE 15.2
Consider '1/;(8, ¢» = N cos 28 sin ¢>.
Evaluate the normalization constant N and then expand this function in terms
of the complete set of spherical harmonics.
Firstly, to obtain N one evaluates:

Thus, to within a phase, N = J15/(147r) .


To expand

{!f; 5
-cos20sin¢>
141r
= L L
00

1=0 m=-I
I
almY~(O,¢» , (15.11)

one must evaluate:

aim = JJY:/(0,¢»J1~~ cos20sin¢>sinOdOd¢>.

One notes that cos 20 sin </> is an odd function of 0, </> (since cos (2 {1r - O} ) sin (</> + 1r)
= - cos 20 sin </» so only odd-parity Y,!, functions can contribute to this integral. Since
the parity of the spherical harmonics (i.e. their behavior as x -+ -x, Y -+ -y, Z -+
-z, or equivalently 0 -+ 1r - 0, </> -+ </> + 1r) goes as (_1)1, this implies that one gets
contributions to aim when yl, y3 ... Additionally one can write

(i</» - exp (-i</»


.
Sln
</>
= exp
---::~"""'--'------':~-'-"-
2i
and since as one can see from Eq. (15.9) the </> dependence of Y±~(O, </» is exp (=fim</»
while
r21r
Jo exp(-im</»exp(in</» d</> = 21r Dnm ,
266 Chapter 15
TABLE 15.1
Spherical Harmonic Functions

1 m Y;'(B, </»
I I
I 0 I 0 l/Vii
1 0 J3/(47r) cosB

1 ±1 -=fJ3/(81r) sinBexp(±i</»

2 0 J5/(161r) (3cos 2 B -1)

2 ±1 -=f J15/(81r) sin Bcos Bexp (±i</»

2 ±2 J15/(321r) sin2 Bexp (±2i</»

3 0 J7/(161r) (5 cos3 B-3 cos B)

3 ±1 -=fJ21/(641r) sin B(5 cos 2 B-1) exp (±i</»

3 ±2 J105/(327r) sin2 BcosBexp (±2i</»

3 ±3 -=f J35/(647r) sin 3 Bexp (±3i</»

this implies only spherical harmonics with m = ±1 contribute to this integral. Thus
the first non-zero contributions in Eq. (15.11) are a1±1.
From Table 15.1

a1±1 = JJ -=f ( ~)1/2.


87r
B (·"I..)V 15
sm exp -=fZ'l' 147r cos
2B (exp (i</» - exp (-i</») . B dBd,l
2i sm 'I-

3
-=f ( 87r
)1/2 V{T5 1 f"
14; 2i Jo (2 cos B-1) sin B dB
2 2

x 121< exp (-=fi</>){exp (i</» - exp (-i</>)) d</>


Angular Momentum and Coupled States 267
3)1/2 {15 1 ( 7r) 37ri {5
= + ( 87r yu; 2i -'4 (±27r) = -T6Y"7·

A similar evaluation of a3±1 yields:

Thus

{15 . 37l"i {5 (1 37ri {5 1)


YU;cos20sm<l>=-16Y"7 ~ +Y-l +3?:Y"2 ~ +Y- 1 + ... ,
(3 3)
which accounts for
457r 2 457r 2 457r 2 457r 2 .. . .
1792 + 1792 + 2048 + 2048 '" 93% of the total probabIlIty m thIs case.

Example 15.2 involves an infinite sum. Sometimes however one only has a finite
number of terms in one's expansion.

EXAMPLE 15.3
Consider .,,(0, <1» = N sin 0 cos 20 sin <1>.
Evaluate the normalization constant N and then expand this function in terms
of the complete set of spherical harmonics.
Firstly, to obtain N one evaluates:

1 N 2 r"'=27r r lJ =7r cos2 20 sin3 0 dO sin2 <I> d<l>


J",=o JIJ=o
N2102'' (1 - ;s 2<1» d<l> 10"(1 - 2 sin2 0)2 sin3 0 dO
2 76
= N 7r 105 .

Thus to within a phase N = V105/(767r) . One also notes .,,(7r - 0, <I>+7r) = -.,,(0, <1»
i.e. only odd l's will contribute to this expansion. The sin <I> term further restricts
the expansion to m = ±1.

One notes here that with the help of Table 15.1 one can write:

(exp(i<l»~;XP(-i<l»)
a
.,,(0, <1» = NsinO(2cos 2 0-1)

N sin 0 (5 cos 2 0 _ 1) _ ~) (exp (i<l» ~;xp ( -i<l» )


268 Chapter 15

= ~ {exp (i¢) - exp (-i¢)} sinO (5cos 2 0 - 1)

_ 3~ {exp(i¢) -exp(-i¢}} sinO


10z
= Ni {~ ~ (~3(O, ¢) + Y!l(O, ¢)) - ~ J6; (~l(O, ¢) + Y2 1 (O, ¢))}

4i f£ (~3(O, ¢) + Y!l(O, ¢)) - ~i fJ (~l(O, ¢) + Y2 1 (O, cfo)),


which accounts for
16 16 63 63
95 + 95 + 190 + 190 = 100% of the probability,
i.e. one obtains a complete expansion with four terms in this case.

The orbital angular momentum operator is a special case of the general class of
angular momentum operators:
J X J = inJ,
for which one can write, analogously to Eqs. (15.4), (15.5):

(15.12)

Jz'IjJ!n j = mjn'IjJ!nj' (15.13)


where j = 0, 1/2, 1, 3/2, 2, 5/2, ... i.e. :
1/2, 3/2, 5/2, ...

0, 1, 2, 3 ...
and for a given j, -j :::; mj :::; j where mj changes in integral steps.

If one has two angular momenta which commute i.e.

J(1)xJ(2) = 0 = { J(2)xJ(1) }

(15.14)

which is the case say if J(l) operates on space 1, while J(2) operates on space 2, then
the sum of these two angular momenta the "total angular momentum operator"
Angular Momentum and Coupled States 269

J == J(I) + J(2) (15.15)


is also an angular momentum operator i.e. :

[J;,Jj] == JJj - JjJi = ittJk , (i,j, k = x, y, z or cyclic permutations). (15.16)

EXAMPLE 15.4
Verify that J defined in Eq. (15.15) is an angular momentum operator if Eq.
(15.14) is satisfied.
Consider

[J."Jy] = [J.,(I), Jy(I)] + [J.,(2), Jy(2)] + [J",(I), Jy(2)] + [J",(2) , Jy(l)]


in [Jz(l) + Jz(2)] = inJz , (15.17)

since summing the k term of Eq. (15.14) including both orderings of coordinates 1,
and 2, yields the result:

Similar arguments apply for other cyclic permutations of Eq. (15.17).

It is of interest to find the eigenfunctions of the total angular momentum operator


J defined by Eq. (15.15). If ¢Jr:..J1 (1) and ¢J;:,.J2 (2) are the eigenfunctions of J(I) and
J(2) respectively, with eigenvalues of the operators J 2(1), J2(2) Jz(I), Jz(2) , the
eigenvalues j(j1 + l)n 2, j2(j2 + l)n 2, mitn, and mi2n, it can be shown that:

J2 [¢Jl(I)¢i2(2)t (15.18)
J

Jz [¢i1(I)¢i2(2)t (15.19)
J

where

[¢jl(I)¢J2(2)t = E (jIjamjl mhli m j) ¢J':'h (1)¢J~i2(2) . (15.20)


J mh (mh)

The j in this equation can have any value in integer steps between IiI - j21 ~ i ~
il + ja,
while mj = mh + mh, and the coupling coefficients (jIj2mitmj2Ijmj) are
270 Cbapter 15
TABLE 15.2
Wigner (or Clebsch-Gordan) Coupling Coefficients

J= m2 = 1/2 m2 = -1/2

j1 + 1/2 it+m+1/2
2i1+1
i1-m +1t.2
2it+1

- i1- m +1t.2 i1+m+1t.2


j1 -1/2 2it+1 2it+1

TABLE 15.3
Wigner (or Clebsch-Gordan) Coupling Coefficients

J= m2 = 1 m2 =0 m2 =-1

(it-mHit- m+1)
j1 +1 (j1+m){j1+m+1)
2(it +1)(2it +1)
(j1-m+1l(it +m+1)
(2it +1)(it +1) 2(j1 +1)(2j1 +1)

- {it +m l{it -m+1l m (it -ml(i1 +m+1l


J1 2jt{j1 +1) 2j1(j1 +1)
:Jit<it+ 1)

j1 -1 (it-ml(it-m+1) - (,it-ml(it+ml (,it +m+1l(it +ml


2j1(2it+1) it(2it +1) 2j!C2j1+1)
Angular Momentum and Coupled States 271
TABLE 15.4
Wigner (or Clebsch-Gordan) Coupling Coefficients

J= m2 = ±3/2

i1 + 3/2 (il±m-1L2}(iI±m+1L2)(jI±m+3/2}
2UI +1)(2it +1)(2il +3)

3(jl±m-1L2)(jI±m+1L2}(jl=Fm +3L2}
i1 + 1/2 T 2il (2jl +1 )(2il +3)

3(iI ±m-1L2}(iI Im+ 1L2l(j1 Im+3L2)


i1 -1/2 2(iJ +1)(2iJ -1)(2iI +1)

(it =Fm- 1/ 2HiI =Fm+1L2l(it =Fm+3/ 2}


i1 - 3/2 T 2it (2jl-1)(2jl +1)

i= m2 = ±1/2

it + 3/2 3(il +m+3L2l(iJ ±m+1L2}(iJ -m+3L2}


2(iJ +1 )(2il +1 )(2il +3)

j1 + 1/2 T (j1 T 3m + 3/2) i1±m+1L2


2it (2it +1 )(2it +3)

iI'fm+1/2
i1 -1/2 - (i1 ± 3m - 1/2) 2(j1+1)(2il 1)(2il +1)

3(iI +m-1L2l(iIIm+1L2l(iI -m-1L2)


it - 3/2 ± 2iJ(2il-1)(2jl +1)
272 Chapter 15
TABLE 15.5: Wigner (or Clebsch-Gordan) Coupling Coefficients (i1 2 m1 m2 Ij m)

J= m2 =±2 m2 = ±1

j1 +2 (it±m-1l(j,±ml(j,±m+1)(j1±m+2}
4(j, +1)(2j, +1)(2j1 +3)(it +2)
(it -m+2l(j1 +m+2l(jdm+1}(j1 ±m}
(2i1 +1)(i1 +1)(2i1 +3)(i1 +2)

I (it±m-1)(j,±m)(j1±m+11(j1 'Fm+2)
j1 +1 =F 2j1 (i1 +1)(i1 +2){2j, +1) =F (j1 =F 2m + 2) (j,±m+1)(idm)
2it (2i1 +1)(i1 +1)(j1 +2)

3(h±m-1Hit±m)(j1'Fm+1Hj,'Fm+2) 3(i1'Fm+1Hh±m)
J1 2(2j, -1)it (j, +1)(2i1 +3) (1=F2m) 2(2i1-1)i1 (i1 +1)(2i1 +3)

j1 -1 (j1±m-1l(i1:Eml(h'Fm+1l(i1:Em+2} (j1:Em+1}(j,'Fm}
=F 2(j1- 1)j, (i1 +1)(2j1 +1) ±(it ± 2m -1) 2j1 (i1 -1)(2i1 +1)(j1 +1)

j1 - 2 (j1:Em-1l(j,:Eml(i1:Em+1l(i1:Em+2} - (i1:Em+1l(i1:Em)(i1'Fm-1)(i1±m-1)
4i1 (j, -1)(2i1-1)(2j, +1) i1 (i1 -1)(2j, -1)(2i1 +1)

J= m2 =0

3(j, -m+2)(j1 -m+1Hi1 +m+2l(h +m+1)


j1 +2 2(it +1)(2i1 +1)(2it +3)(it +2)

3(j1 -m+1 l(it +m+1)


j1 +1 m it{2it +1)(h +1)(i1 +2)

3m 2-hU1 +1)
j1
V(2i1-1)il(i1 +1)(2il +3)

3(h-m)(h+m)
j1 -1 -m
(i1 1)il (2i1 +1)(i1 +1)

3(il-mHh-m-1)(i1 +ml(j1 +m-1}


j1 - 2 2(i1-1)(2it -1)j1(2it +1)
Angular Momentum and Coupled States 273
known as the "Wigner" or "Clebsch-Gordan" coefficients. If mj1 + mj2 f: mj these
coupling coefficients are zero, hence the notation mj1 (mh) in Eq. (15.20). Thus only
the sum over one (either mj1 or mh) of these two projections is independent, and

Some simple angular momentum coupling coefficients are given in Tables 15.2-5.

In some simple cases one may obtain the coupled wave functions and consequently
the corresponding coupling coefficients using symmetry arguments.
Consider for simplicity two eigenfunctions <p~231 (1) and <p~232 (2) each having angu-
lar momentum quantum number j= 1/2 and two possible projections mj = ±1/2.
Assume that they are eigenfunctions of two operators J(I), J(2) which satisfy Eq.
(15.14). There are two possible eigenfunctions for each ofthese two operators, namely
<p!~~(i) and <P~:/2(i), (i = 1, 2). One can thus construct four product states:

l/2( 1)<P1/2
<P1/2 l/2()2 , <P1/2
-1/2 (1) <P1/2
-1/2 () l/2()
2 , <P1/2 1 <P1/2 2 , <P1/2
-1/2 () l/2( 2).
-1/2 (1)<P1/2
The first two of these states are symmetric under interchange of particles 1 and 2,
whereas the third and fourth are neither symmetric nor antisymmetric under such an
interchange. Since J as defined in Eq. (15.15) commutes with the exchange operator
P12 (d. Example 12.5 and subsequent discussion) we should be able to, and in fact
can, immediately construct normalized symmetric and antisymmetric combinations
of these latter states as follows:

(15.21 )

Grouping states by symmetry alone one thus has three orthonormal symmetric
states :

.,p:lsym(l, 2) = <p!~;(1 )<p!~;(2)

.,p~sym(l, 2) = ~ {<p~~;(1 )<P~12/2(2) + q)~12/2(1 )q)~~;(2) } (15.22)

.,p:lsym(l, 2) = q)1/2 -1/2 (2)


-1/2 (1 )l/2
274 Chapter 15
and one antisymmetric state:

If one operates on the symmetric states, Eq. (15.22) with the operator:

(J(l) + J(2»2 = J2(1) + J2(2) + 2 J(l) . J(2), (15.24)

one obtains 1(1 + 1)n 2 times these states, while if one operates on Eq. (15.23) with this
operator one obtains 0=(0+1 )11.2, i.e. one has three symmetric states with quantum
number j=l, and one antisymmetric state with j = O. operating on these "coupled"
states with the operator Jz = {Jz(l) + Jz(2)} one obtains mj = n, 0, -n times the
wave functions respectively for the three states of Eq. (15.22) and 0 times the state of
Eq. (15.23). This justifies the notation for the wave functions in Eqs. (15.22), (15.23).
Thus instead of the four product states above one has obtained four equivalent states,
which however are eigenfunctions of the total angular momentum of this system, the
"triplet" ""L ""~, "":1>and the "singlet" ""g, i.e. 2 ® 2 = 3 + 1. One notes that in
all angular momentum couplings the total number of independent states remains the
same. Performing this calculation requires one use the result:

hW!" = n..j(j =1= m)(j ± m + 1) W!"±l where h = J:c ± iJy , (15.25)

and the fact that:

2J(1) . J(2) = (15.26)

EXAMPLE 15.5
Consider two eigenfunctions cf>~. (1) and cf>~. (2) each having angular momentum
quantum number j= 1 and threel1possible prgjections mj = -1, 0, 1. Assume
that they are eigenfunctions of two operators J(l), J(2) which satisfy Eq. (15.14).
Construct the eigenstates of the total angular momentum J = J(l) + J(2).

There are three possible eigenfunctions for each of these two operators, namely
cf>W), cf>Mi) and cf>:l(i) , (i = 1, 2). One can thus construct nine product states:

cf>~(1)cf>~(2), cf>!(1)cf>~(2), cf>~(1)cf>:1(2), cf>~(1)cf>~(2), cf>M1)cf>~(2),

cf>~(1)cf>:1(2), cf>:1(1)cf>~(2), cf>:1(1)cf>~(2), cf>:1(1)cf>:1(2) .


Angular Momentum and Coupled States 275
The first and ninth of these functions are completely symmetric under interchange
of coordinates 1 and 2. Moreover these are eigenfunctions of Jz = J z(l) + Jz(2) with
eigenvalues 2n, and -2n. Normalized symmetric and antisymmetric combinations
which are eigenfunctions of Jz with eigenvalues n, and -n, are obtained directly:

Finally three orthonormal eigenfunctions of Jz with eigenvalue On can be written


as follows:

v a2 ~ 2(32 [alP~(1)¢~(2) + (3 {¢~(1)¢:1(2) + ¢:1(1)¢~(2)}] ,

~ va 2 ~ 2(32 [-2(3¢M1)¢M2) + a{¢~(1)¢:1(2) + ¢:1(1)¢!(2)}] ,

~ {¢~(1)¢:1(2) - ¢:1(1)¢~(2)} ,

of which the first two eigenfunctions are symmetric while the third is antisymmetric.
To evaluate a, (3 one can operate on one of the two symmetric combinations with
the operator J2, using Eq. (15.24), (15.25) and (15.26):

J2 Va2 ~ 2(32 [a¢~(l )¢~(2) + (3 {¢~(1 )¢:1 (2) + ¢:1 (1 )¢~(2) }]


= .ja2n: 2(P [(4a + 4(3)¢~(1)¢M2) + (2a + 2(3) {¢~(1)¢:1(2) + ¢:1(1)c/>~(2)}]
= J:2+:1;: [2¢~(1)¢~(2) + {¢~(1)¢:1(2) + ¢:1(1)¢~(2)}].
For this to be an eigenfunction one requires:
a
p = 2. For instance if (3 = 1, then a = 2.
Thus the corresponding eigenfunction of J2 becomes:

with eigenvalue 2(2 + 1 )n 2 •


Similarly the second symmetric function becomes:
276 Cha.pter 15
and when this latter function is operated on with J2 its eigenvalue is 0(0 + I)h 2.
Grouping terms by symmetry and their behavior when operated on by J2, and J.. ,
one again has nine functions but now subdivided into in a quintuplet 1li~, 1li~, 1li~, 1li:1l
W: 2, a triplet wL wA, W~l' and a singlet wg. These are an equally good, in other
words complete basis, as the nine product functions above, except that they are
eigenfunctions of J2, J.. , J2(I), J2(2) rather than of J2(I), J 2(2), J.. (1), J.. (2). One
thus has the equivalence 3 ® 3 = 5 + 3 + 1.
In detail:

1li~sym(1, 2) = ¢~(I)¢~(2),
1li~sym (1, 2) = ~ [¢H1)¢M2) + ¢MI)¢H2)]
1li~sym(1, 2) = ~ [2¢MI )¢M2) + {¢HI )¢~1 (2) + ¢~1 (1 )¢l(2) }]
1li:' lsym (1, 2) = ~ [¢:1(I)¢~(2) + ¢~(I)¢~l(2)]
1li:'2sym (I, 2) = ¢~1 (1 )¢:1 (2)

1li~antisym(I, 2) = ~ [¢~(I)¢M2) - ¢~(1)¢~(2)] ,


1li~antisym(1, 2) = ~ {¢!(I)¢~1(2) - ¢~l(I)¢!(2)}
1li~lantisym (1, 2) = ~ [¢:1(1)¢M2) - ¢MI)¢:l(2)] ,

1ligsym (I, 2) = ~ [-¢~(I )¢~(2) + { ¢~ (1 )¢~1 (2) + ¢~1 (1 )¢~ (2) }] .

One thus notes for instance that the Wigner coefficient for

(1 1 0 0120) = 2/../6, etc.

An ambiguity however arises as to the overall signs of the coupling coefficients. Thus
the wave function 1li8sym(I, 2) could equally well be written:

since this function is also normalized and orthogonal to both 1li~antisym(I, 2), and
1li~sym(I, 2). One must thus adopt and stick to a consistent convention.
An observant reader may have noticed that the component ¢MI )¢M2) is missing
from the 1li~antisym(I, 2) wave function. Its absence may be explained by arguing
Angular Momentum and Coupled States 277
that W5antisym(l, 2) is antisymmetric while the component <pMl) <pM2) is symmetric.
On a more technical level the coupling coefficient

(1 1 0 011 0) = O.
This is a consequence of the following three properties of the Wigner coefficients:

(15.27)

anyone of which in this case implies:

(1 100110) = (_1)1+ 1 - 1 (1 1 00110) = - (1 1 00110)

A standard way to get coupling coefficients when symmetry arguments do not


apply is to use Eq. (15.25) on the wave function with maximum projection. This
procedure can be illustrated by a simple example.

EXAMPLE 15.6
Consider two eigenfunctions <p~ )1 (1) and 'TJ~,{2J2 (2), the former having angular mo-
mentum quantum number j1= 1 and three possifile projections, mjl = -1, 0, 1, and
the latter angular momentum quantum number 12= 1/2 and two possible projections,
mh = -1/2, 1/2. Assume that they are eigenfunctions of two operators J(I), J(2)
which satisfy Eq. (15.14). Construct the eigenstates of the total angular momentum
J = J(I) + J(2).

There are three X two = six possible product functions which are eigenfunctions
of J2(1), J2(2) , Jz(I), J z(2):

These states are neither symmetric nor antisymmetric nor does one normally sym-
metrize or antisymmetrize these functions since j = 1 could refer to a Bose particle
while j = 1/2 may refer to a Fermi particle, or more likely the j = 1 refers to the
orbital angular momentum, while the j = 1/2 refers to the spin of a Fermi particle.
The first state above has maximum rhj = mil + mil = 3/2 projection and is
unique. It must therefore correspond to the state W~~~(I, 2).
278 Chapter 15
Thus:
W;~~(1, 2) = cfoi (1 )Tf~~~(2).
(15.28)
But L = L(I) + L(2). Operating with L on the l.h.s and with L(I) + L(2) on
the r.h.s. of Eq. (15.28) one obtains:

liv'3XTW~~~(1, 2) = liv"2X1cfoMl)Tf~~~(2) + n cfoi(I)Tf~1~2(2) ,

If
i.e.

W~~~(I, 2) = ~ cfo~(1 )1l~~~(2) + cfoi(1 )1l~12/2(2) .


Operating on W~~~(I, 2) with L one obtains:

If cfo~1(I)Tfi~~(2) + fa cfoMl)Tf~2/2(2)
i.e. :

W~12/2(1, 2) = .

A further operation by L, this time on W~12/2(I, 2) yields:

Ii v3W~;/2(I, 2) = Ii If cfo:l (1 )1l~2/2(2) + Ii fa ~ cfo~l )Tf~1~2(2)


(1

= Ii v3cfo:l (1 )1]~~2(2) ,
1.e.
W~;/2(1, 2) = cfo~l (1 )Tf~12/2(2) ,
as expected by analogy with Eq. (15.28).
· d a quartet 0f states: .T,3/2
One h ast h uso b tame .T,3/2 .T.3/2 .T,3/2
'£3/2' '£1/2' '£-1/2' '£-3/2.
There remain a doublet of orthonormal states: W~~~, Wl};/2' the former of which must
be orthogonal to W~~~, and the latter orthogonal to W~12/2. These two states will auto-
matically be orthogonal to the other W3 / 2 states since they have different projections
mj. To within an arbitrary phase (hence the same ambiguity mentioned above about
the overall sign of the coupling coefficients which one must treat consistently) these
two states are:

Wi~~(I, 2) = -If cfo~(I)lli~;(2) + fa cfoi(I)Tf~~2(2)


W1/2
-1/2 ( 1, 2 ) = -fa cfo:1(1)1l:~~(2) If cfo~(I)Tf~12/2(2)
+ .
Angular Momentum and Coupled States 279
The equivalence between the above two bases can be expressed as follows: 318)2 = 4+2.

The Eq. (15.20) can be viewed as an expansion in terms of a complete set.


Thus the function [cf>.i1(1)4>i2(2)1~ can be expanded in terms of the complete set
. • 3
<fY.i..11 (1 )4>~.32 (2).

(15.29)

Premultiplying Eq. (15.29) by 4>:'/> (1 )4>::/2 (2) and integrating using the assumed
3 1 32

orthonormality of the cf>.im.'s


3
one obtains:

(15.30)

Comparison of Eqs. (15.20) and (15.30) implies:

EXAMPLE 15.7
Using complete set arguments obtain the inverse of Eq. (15.20) .
One can immediately write:

cf>.i.i. h (1 )cf>.i~h (2) = L


j, mj
b j!n;ai [¢i1 (1 )4>32 (2) t .
J
(15.31 )

Premultiplying Eq. (15.31) by [¢i1 (1 )¢i2 (2)]~, and using the assumed orthonor-
. 3
mality of the [4>i1(1)4>12(2)];"'s one obtains:
3

(15.32)

Since the coupling coefficients can be chosen to be real

(15.33)

where the (mj) in the sum implies there is really no sum over mj since mil +mj2 = mj'

EXAMPLE 15.8
Find orthogonality conditions satisfied by the Wigner Coefficients.
Starting with Eq. (15.20) and using the fact the functions involved are orthonor-
malone obtains:
280 Chapter 15

([qJl(l)qSh(2)J:j l[qJl(l)qJ2(2)]~j,) = 8i,jl8mj,mjl


= L Uti2 m h mj2 Ijmj) Uli2 m h mh Ij'mp).
mh

Thus:
L Uti2 m jl mh Ijmj )(jlhmj] mj2 Ij'mp) = 8j ,jl 8mj ,mjl . (15.34)
mil

Similarly starting from Eq. (15.33) one obtains:

(15.35)

One can consider next the case of three angular momentum operators, J(I), J(2),
J (3). If they operate on different spaces (for instance on the coordinates of particles 1,
2, and 3 respectively), then the total angular momentum is also an angular momentum
operator, namely:
J == J(I) + J(2) + J(3).
Formally the condition for J to be an angular momentum operator is that

J(I) X J(2) = J(I) X J(3) = J(2) X J(3) = 0, (15.36)

in which case J X J = iliJ. One can then examine the problem of finding eigenfunctions
of J assuming one has eigenfunctions of J(I), J(2), J(3), namely ¢#.1' 'T/!;,2' ).,~3'
It is immediately clear that the order of the intermediate coupling is not unique
in this case. Thus one may couple these three eigenfunctions with three different
intermediate orderings:

The different functions which result may however be related to each other. Thus
using complete set arguments one can write:

where the recoupling coefficient is:


Angular Momentum and Coupled States 281
Some of the simple coefficients needed for recoupling three angular momenta are
listed in Tables 15.6-7. In these tables the coefficients given are the so called "U-
coefficients" where:

They involve in all six angular momenta il. i2. i3. J12 • J 23 , J and are related to
the so called "6-j's"as follows:

{ a be} == (-1 ),,+b+c+d U (a b c di e f) .


dc/ v(2e + 1)(2/ + 1)

In terms of U functions:

(15.37)

EXAMPLE 15.9
Obtain orthogonality relations for the U functions using complete set arguments.
Starting from Eq. (15.37) one has:

L U (jl h J hi Jl2 J23 ) U (jl h J hi J~2 J~3) DJ12, J~2 •


J12, Jh

Hence:

:E U (jl i2 J i3i J12 J23 ) U (jl i2 J i3i J12 J~3) = DJ23, J~3 • (15.38)
J12

Similarly one shows:

L U (it i2 J i3i J12 J23 ) U (jl i2 J iJi J~2 J23 ) = DJ12 , Jf2 • (15.39)
J23

EXAMPLE 15.10
Consider three angular momenta with eigenfunctions: cfJ.:.1, "'!;,2' )..~3' where il =
i2 = h = 1/2, coupled according to the intermediate coupling scheme [cfJl [.".12 )..ia j J23
Evaluate the recoupling matrix in Eq. (15.37) for this case.
1:.
282 Chapter 15
Using Tables 15.6-7:

[¢1/2 [.,,1/2,.\1/2rf/2 -1/2 -/3/2 0 [ [¢1/2"l1/2] 0 ).1/2f/2

[¢1/2 [.,,1/2).1/2rf/ 2 = -/3/2 1/2 0 [[¢1/2.,,1/2 r ).1/2 r/ 2

[<jJ1/2 [.,,1/2 ).1/2]1 f/2 0 0 1 [ [<jJ1/2.,,1/2] 1 ).1/2f/2

One notes that the recoupling matrix is real and symmetric.

TABLE 15.6
Non-vanishing U-Coefficients with one argument zero

2J12 + 1
(2j1 + 1)(2h + 1) ,

(2jl + 1) (2h + 1)
Angular Momentum and Coupled States 283
TABLE 15.7
U-Coefficients with one argument half

J12 = J23 = h + 1/2 J 23 = i2 -1/2

J + 1/2 {b+h- J+1l2Hil-,i2+J+1i 2l {it +h+ J+3l 2H-it +h+ J +1i 2l


(2J+1)(2h+1) (2J+1)(2i2+1)

((il + h+J+3/2)(- il +i2+J+:U2) - ~it +i2- J+1i 2l(it -h+J+ 1l2l


J -1/2 \2J+iT(2i2+iT (2J+1)(2i2+1)

In some simple cases one can obtain the eigenfunctions of three angular momentum
operators by using symmetry arguments.

EXAMPLE 15.11
Consider for simplicity three eigenfunctions ¢i.:..n (1), ¢i~.n (2) and ¢i.:.D (3) each
having angular momentum quantum number il = i2 = h = 1/2 and two possible
projections mip mh, mh = ±1/2. Assume that they are eigenfunctions of three op-
erators J(l), J(2), and J(3), which satisfy Eq. (15.36). Construct the eigenfunctions
of the total angular momentum J = J(l) + J(2) + J(3) by symmetry arguments.
In this case since there are two possible eigenfunctions for each of these three
operators, namely <p~~~(i) and <p1};/2(i), (i = 1, 2, 3), one can thus construct eight
product states:
,1.1/2 ( ),1.1/2 ( ) 1/2 ( )
'I'±1/2 1 'I'±1/2 2 <P±1/2 3 .

Two of these product states, namely:


1/2( ),l.l/2( ),l.l/2() 1/2 ( ),1.1/2 ( ),1.1/2 ( )
<Pl/2 1 '1'1/2 2 '1'1/2 3, <P-l/2 1 '1'-1/2 2 '1'-1/2 3 ,
are symmetric under any interchange of particles 1, 2, and 3, whereas the rest are
neither symmetric nor anti symmetric under such an interchange. We can however
immediately classify the remaining six states according to their eigenvalue when they
are operated on by Jz = Jz (l) + J z(2) + Jz(3). Three have eigenvalue li/2, and three
eigenvalue -li/2. Consider the three states with eigenvalue li/2:
1/2( ) 1/2( )-1..1/2 () -I..1/2( )-1..1/2 ( ) 1/2() -1..1/2 ( )-I..l/2( )-I..l/2( )
<Pl/2 1 <Pl/2 2 '1'-1/2 3 , '1'1/2 1 '1'-1/2 2 <Pl/2 3 , '1'-1/2 1 '1'1/2 2 '1'1/2 3 .
284 Chapter 15
It is simple to get a normalized completely symmetric combination of these three
states, namely:

If one operates on this state (or on the two other symmetric states above) with

J2 = J2(1) + J2(2) + J 2(3) + 2 J(1)·J(2) + 2 J(1)·J(3) + 2 J(2)·J(3),


one obtains 15n2 /4 times this state which implies 1fJ1 is an eigenfunction of J2 with
eigenvalue j = 3n/2. For a second function with projection n/2 one can take the
product of the anti symmetric wave function Eq. (15.23) times an eigenfunction of
the third angular momentum which results in an overall projection n/2:

.1,2 _
'I" -
-.!...
J2 [",1/2(1)",1/2
'1"1/2 '1"-1/2
(2) _ ",1/2 (1)",1/2(2)] ",1/2(3)
'I" -1/2 '1"1/2 '1"1/2 .

Finally for the third function with projection n/2 one can take a linear combination
of symmetric two-angular momentum functions Eq. (15.22) times eigenfunctions of
the third angular momentum operator which again result in an overall projection of
n/2:
1 [1/2() 1/2 ()
1fJ = a J2 <P1/2 1 <P -1/2 2
3
+ <P1/2
-1/2 1 <P1/2 2 <P1/2 3 + f3<P1/2 1 <P1/2 2 <P -1/2 3 .
( ) 1/2()] 1/2() 1/2( ) 1/2( ) 1/2 ( )

The constants a, f3 can be determined from the requirement the functions 1fJ1, 1fJ2,
and 1fJ3 are orthonormal. This implies:

and /V2 + f32 = 1,


..... l' e
.. /V
..... = ± v'3'
1 f3 = TVf2"3 .
::r:

One thus obtains three states with projection mj = n/2:

Similarly one may obtain the three functions which have overall projection -n/2.
Thus there are four completely symmetric functions:
Angular Momentum and Coupled States 285

?jJ3/2
3/2 = </>~~~(1 )</>~~~(2)</>~~~(3)
?jJ3/2
1/2 = ~ [</>!~~(1 )</>!~~(2)</>~;/2(3) + </>!~~(1 )</>~12/2(2)</>!~~(3) + </>~12/2(1 )</>!~~(2)</>!~~(3)]
?jJ3/2
</>-1/2 1 '1'-1/2 2 '1'1/2 3 + '1'-1/2 1 '1'1/2 2 '1'-1/2 3 +
1 [1/2 ( )-1.1/2 ( )-I.1/2() -1.1/2 ( )-I.1/2( )-1.1/2 ( )
-1/2 v'3
.1. 1/ 2(1).1.1/ 2 (2).1.1/2 (3)]
'1'1/2 '1'-1/2 'I' -1/2
./,3/2
'1'-3/2 = </>~;/2(1 )</>~;/2(2)</>~;/2(3),
two wave functions antisymmetric under interchange of coordinates 1, and 2, but
with no definite symmetry under interchange of coordinates 1 or 2 with coordinate
3:

./,1/2
'1'1/2
= _1
v'2 [-1.1/2(1)-1.1/2
'1'1/2 '1'-1/2
(2) _ .1.1/ 2 (1)-1.1/2(2)] -1.1/ 2(3)
'I' -1/2 '1'1/2 '1'1/2

?jJ1}1~2 = ~ [</>~~!(1)</>1}12/2(2) - </>1};/2(1)</>~~!(2)] </>1};/2(3),


and finally two wave functions symmetric under interchange of coordinates 1, and 2,
but with no definite symmetry under interchange of coordinates 1 or 2 with coordinate
3:

./,1/2 =
'1'1/2
1 [1/2() 1/2 ()
- .j6 1/2 ( ) 1/2()] 1/2() f2 1/2( ) 1/2( ) 1/2 ( )
</>1/2 1 </>-1/2 2 + </>-1/2 1 </>1/2 2 </>1/2 3 + V3</>1/2 1 </>1/2 2 </>-1/2 3

./,1/2
'1'-1/2
= .j6 1/2 ( )-I.1/2( )] -1.1/2 ( )
</>1/2 1 </>-1/2 2 + </>-1/2 1 '1'1/2 2 '1'-1/2 3 -
1 [1/2() 1/2 ()

!a</>1}12/2(1 ) </>1}12/2 (2) </>~~~(3).


Thus in all one still has eight independent functions. More precisely the first four
correspond to: --

[[</>1/2</>1/2] 1]
</>1/2
3/2 [ 0 ]1/2 , and the last two to
,while the next two to: [</>1/2</>1/2] </>1/2

[[</>1/2</>1/2] 1]
</>1/2
112
. Here one has the equivalence 20202 = + 2 + 2. 4

Of the more complicated recoupling coefficents: 9-j's, 12-j's, 15-j's, etc. in the lit-
erature, the most frequently encountered are the 9-j's. They arise when one considers
286 Chapter 15
four angular momentum operators J{I), J(2), J(3), J(4), which operate on different
spaces such that
J = J(I) + J(2) + J(3) + J(4),
is also an angular momentum operator (i.e. J x J = J).
An important case in point is if one has two particles with orbital and spin angular
momenta. Two convenient couplings are the j-j (intermediate) coupling scheme:

(15.40)

and the L-8 (intermediate) coupling scheme:

(15.41)

where as the names imply in the former case one couples the total individual angular
momenta to form the total angular momentum, and in the latter the total orbital
and spin angular momenta to form the total angular momentum.
To change the intermediate coupling one uses 9-j's. Thus

[[l112]L[8182]S]J =
M
L:
h, i2
A (~~L :~S ~~)
J
[[l18di1 [1282]hr
M
(15.42)

The A function in Eq. (15.42) is related to the 9-j (which as the name implies
involves nine angular momenta) as follows:

(15.43)

and in Table 15.8 the values of the A's are given for the simple case 81 = 82 = 1/2
(electrons, protons, etc.). The 9-j is a very symmetric quantity. Thus it is invariant
under reflection about either diagonal. Also, if one or more entries of a 9-j are zero
it simplifies. Thus:

{H c }
f
o
= Dei Dgh
{-1 )b+c+d+g
y'(2c+l){2g+1)
{a be}
e d 9
,

{ oad 0be}
e f
0
= Da.d Dbe Dc!
1
yi{2a+1)(2b+1){2c+l)
.
Angular Momentum and Coupled States 287
TABLE 15.8

A(
It 1/2
12
L S
1/2 h)
j coefficients for spin 1/2 particles

S=o S=l
L=J L = J-1

(11 +12-J)(11 +12+J+11 (11 +12-J+1)(l1 +12-JHh -12+ J )(l2- ll +Jl


j1 = 11 - 1/2 2(211 +1)(212+1)
- 2J(2J+1)(2h +1)(212+1)
)2=12 -1/2

j1 = h -1/2 - (11-12+J)(l2-ll +J+1)


2(21t +1)(212+1)
(11 +12+J+1)(11 +12-J+1)(l2-ll +J +1)(12-11 +J)
2J(2J+1)(2lt +1)(212+1)
j2 = h + 1/2

j = 11 + 1/2 (12 -It +J){l1- 12+J+1)


2(211 +1)(212+1)
- {It +b+J+1HIt +b-J+1)(1t -12+J+1){1t -12+J)
2J(2J+1)(21t +1)(212+1)
j2 = 12 - 1/2

)1 = it + 1/2 {It +12+J+2){l1 +12-J+1)


2(2!t +1)(212+1)
{ll+ldJ+2)(lt +12+J+1){lt -12+J){l2-!t +J)
2J(2J+1)(21t +1)(212+1)
j2 = 12 + 1/2

S=l S=l
L=J L= J+1

it = 11 - 1/2 (12-11) x {It +12+J+2){h +12+J+1){1t -12+J+1){f2-1t +J+1)


y'J(J+1) (2J+1)(2J+2)(2/t +1)(212+1)
j2 = 12 - 1/2 (Entry for S = 0, L = J)

_ (/1 +12+1) X {It +12+J+2){l1 +12-J){l1-12+J+1){1t -12+J)


)1 = 11 - 1/2 (2J+1)(2J+2)(211 +1)(212+1)
V J (J+1)
)2 = 12 + 1/2 (Entry for S = 0, L = J)

) = it + 1/2 {It +12+1) X


y'J(J+1)
- {It +l2+J+2){1t +l2-J){l2-1t +J+1){l2-lt +J)
(2J+1)(2J+2)(2h +1)(212+1)
)2 = 12 - 1/2 (Entry for S = 0, L = J)

{It +12-J + 1Hit +12- J )(l1 -12+J +1)(l2-1t +J+1)


j1 = It + 1/2 _(1 1-12)
_VJ(J+1)
X -
(2J+1)(2J+2)(211 +1)(212+1)
j2 = 12 + 1/2 (Entry for S = 0, L = J)
288 Chapter 15
EXAMPLE 15.12
Obtain the orthogonality conditions satisfied by the 9-j's.
Consider

r
Hence,

j52 (2i1 + l)(2h + l)(2L + 1)(2S + }{ ~,


i1 i1
~hv6ss'.
81 81

1) ~ 82 i2 82 i2 }
S J S' J
(15.44)
Similarly one obtains:

L:{2i1+1)(2h+1)(2L+1)(2S+1){~:L
L,8
;:
S J
~:}{ ~~L ;~S

EXAMPLE 15.13
Consider two p (l = 1) spin 1/2 Fermions. Write the possible states inj-j coupling
and using Table 15.8 express these states in L-S coupling.

One uses for simplicity notation pj( i) here which stands for [1 1/2]; and 28+1 LJ
where L = 0 => S, L = 1 => P, L = 2 => D. In the j-j intermediate coupling scheme
one has:

where, in all, there are 3 ® 2 ® 3 ® 2 = 36 = 1 + 3 + 1 + 3 + 5 + 7 + 3 + 5 + 3 + 5 states.


Using Table 15.8
Similarly one obtains:

[P3/2(1 )P3/2(2)] 0

[P3/2(1 )P3/2(2) r

[P3/2(1 )P3/2(2) r

[P3/2(1 )P3/2(2) r =

[P3/2(1)Pl/2(2)r =

[P3/2(1 )Pl/2(2)] 2

[Pl/2(1 )P3/2(2) r

[Pl/2(1 )P3/2(2) r
290 Chapter 15
EXAMPLE 15.14
Pick out the anti symmetric states (hence only acceptable states if the two particles
are identical) in Example 15.13.

For the states


[Pl/2(1 )Pl/2(2)]: ' [P3/2(1 )113/2(2)]: '
using the first of the properties of Wigner coefficients given in Eq. (15.27) it follows
that:
[Pl/2(1)Pl/2(2)r, and [p3/2(1)113/2(2)t 2
are antisymmetric since

j being half integer.


Finally one has to take the anti symmetric combinations:
1 {[P3/2(1 )Pt/2(2) ] I, 2 - [P3/2(2)Pt/2(1) ] 1, 2}
,fi

I.e. ~ {[P3/2(1)Pl/2(2)f + [Pl/2(1)113/2(2)f} , ~ {[P3/2(1)Pl/2(2)r - [Pl/2(1)p3/2(2:

In all one thus has 1+1+5+3+5=15 (acceptable) anti symmetric functions. The re-
maining 21 are (unacceptable) symmetric functions.
Schematically:

1
73 -ft,0 0 0
[Pl/2(1 )pl/2(2)] 0 ISO

r
[113/2(1 )P3/2(2)
-ft,
1
-73 0 0 0 3po

~ {[P3/2(1)Pl/2(2)f + [PI/2(1)113/2(2)r} = 0 0 1 0 0 3P1

[P3/2(1)P3/2(2)] 2 0 0 0 A 1
73
3P2

~ {[P3/2(1)Pl/2(2)r - [Pl/2(1)P3/2(2)]2}
0 0 0 1 ID2
-73 -ft,
CHAPTER 16

Tensor Operators, and Evaluation of Matrix Elements

Given an 'irreducible, tensor' operator Tqk. This loosely speaking is an operator


which behaves like a wave function 'IjJ: as concerns angular momentum coupling. Thus
for a given k this operator must, like 'IjJ:, have 2k + 1 components q.
If one has such an operator in an integral

one can then write

Extracting the Wigner coefficient and summation from the integral:

l(jl' k,j2, mIl q, m2) = ~ (kj2q m 21hq + m2) 'IjJ*~1


33
J [Tk'IjJh ]~~m2 dr.

But for the integral on the r.h.s. not to vanish, jl and h must be able to couple to
angular momentum zero, (with projection zero), i.e. form a scalar integrand. This
implies only the term j3 = jl survives and also that q + m2 = ml (since analogously
with Y,;:I(O, ¢) = (-l)mY~m(O, ¢), the complex conjugate of 'ljJt:. 1 , 'IjJ*!:.1 '" 'IjJ::ml as
concerns angular momentum coupling) . Thus

l(jll k,j2, ml, q, m2) = (kj2q m 21jlmd J 'IjJ*!:.1 [Tk'IjJh 1:1 dr.
Moreover the integral on the r.h.s. being a scalar, it should not depend on ml which
can therefore be dropped in writing down the result:

l(jll k,j2, mIl q, m2) = (kj2q m 2ljlml) J 'IjJ*it [Tk'IjJi2 t dr. (16.1 )

This is the famous Wigner-Eckart theorem. An almost obvious result, once one has
seen the derivation, it is basic in the evaluation of numerous matrix elements which
arise in Atomic, Molecular and Nuclear Physics. It is useful because it removes the
"geometry", i.e. the projections mIl q, m2 from the matrix element evaluation by
concentrating them in a Wigner coefficient which one can look up in tables (cf. Tables

t
15.2-5), and additionally one has merely to evaluate a single projection-independent
"reduced matrix element" f 'IjJ*it [Tk'IjJi2 dr in order to obtain the value of the in
292 Chapter 16
principle (2jl + 1)(2k + 1)(2h + 1) matrix elements I(h,k,h,ml,q,m2) (many of
which are zero however since I(jl,k,h,mhq,m2) = 0 unless q + m2 = ml) which
arise for a particular jl, k and h.
If the irreducible operator is for example the multipole operator:

(16.2)

this automatically behaves like a wave function. For a three-component vector oper-
ator, e.g. 5, Lor 1, one must choose combinations of the three components which
behave like an (irreducible) tensor of rank 1. This can be done by using the spherical
harmonics (d. Table 15.1) as prototypes, i.e. take

Jz since Yo(I)«()) fI cos () = [~2rY;fI]


= ~2Y; z ,

Jx+iJ
V2
y •
sInce y;(I}«(),A..)
1 '/' = - [If!] 1 . ()
-2 -1r -V2 -2r -1r V2 Y:
z,/, = - [If!](X+i
sm exp (.,A..)

'1/;:1 /"oJ JxV2


- iJy since y(~)- «() 4» = [~ fI]_l sin () exp ( -i4» = [~ fI]
2Y; V2 2rY;
(x -
V2
iy) ,
where the components must have the same relative normalizations as the V's, but
overall constant factors (in this case {J3/1r }/ 2r) may be dropped.

EXAMPLE 16.1
Obtain the irreducible tensors which can be formed from the two vectors 1(1), and
1(2).
There are nine components here, and since the angular momentum operators
operate on different spaces they commute and one need not worry about the order in
which these combinations appear. In detail one has

A single scalar can be formed from these components, by analogy with rl·r2.

From the discussion immediately preceding Example 16.1 one sees that a three-
component irreducible tensor of rank one can be formed as follows:
Tensor Operators, and Evaluation of Matrix Elements 293
If a == i(l)xi(2),

then the three-component tensor of rank 1 is:

,p~ C"

,p}
Cx + iCy
..;2
Cx -iCy
,p:l ..;2
Finally one forms an irreducible tensor of rank 2 by analogy with the five compoent
spherical harmonics of rank 2, Y,! (d. Table 15.1) or by recalling that

T/J!(1,2) = L (llmlm212m) Y;;l (I)Y;;2(2),


mI(m2)

and using Table 15.3. The result is:

'" 2- {[Jx(l) ..;2


T..;2
± iJy (l)] J (2) [Jx(2) ± iJy (2)] J (I)}
,,+ ..;2 %

[Jx(l) ± iJy (I)] [Jx(2) ± iJy (2)]


..;2 ..;2
Thus the reducible nine ( = 3 X 3) components above have been expressed as the
(same number but) more convenient nine irreducible tensor components: 1 (from the
scalar, ,p8) + 3 (from the tensor of rank 1, ,pI) + 5 (from the irreducible tensor of
rank 2, ,p2).
Similarly one can build up tensors of higher rank if desired, using the general
expression
T; L
= (k 1k 2qlq2Ikq) U~l V: '
ql(q2)
(16.3)

where U~l, V~2 are irreducible tensor operators.


294 Chapter 16

EXAMPLE 16.2
Evaluate the angular part of the matrix element

(11V12) = (21V11) ,

in Example 12.12 without using the Wigner-Eckart theorem.


The matrix element involved is

where

28 1/2 R20 (r ) [Y °X1/2] m1/2, = R20 (r ) Yoo( fJ, </1 )Xmj'


1/2

One notes that z = (z/r)r = r cos fJ = 2rV'lr/3 Yo1(fJ, </1) (d. Table 15.1), behaves
like an irreducible tensor operator of rank 1 and projection 0, i.e. can be treated like
a wave function 'l/JJ as concerns angular momentum coupling. Thus:

(11V12) = (R20(r)YoO(fJ, </1)X~; I~c2r v'i"Yo1(fJ, </1)1 R21(r) X

{(l~Omjl ~mj) Yo1(fJ, </1)X;,c; + (1~lmj -11 ~mj) Ytl(fJ, </1)Xt,!.:-l

~C2v'i" (R2o(r) Irl R21 (r)) (1 ~omj I~mj) (Yo 0 ( fJ, </1) IYo1 (0, </1)1 Yo1(fJ, </1))
= ;;(_I)m j +1/2 (R20(r)lrIR21(r)).

EXAMPLE 16.3
Solve Example 16.2 using the Wigner-Eckart theorem.
Tensor Operators, and Evaluation of Matrix Elements 295
Using Eqs. (16.1), (15.37) and Table 15.6,

(1\V12) = ~c2Ji (R 2o (r) [yOXI/2]~: Iryol(O)1 R21(r) [yIXI/2]~:) = ~c 2 Ji


x (l~Omjl ~mj) (R20(r)lrIR21(r» ([yOXI/2f/21 [yl [ylxl/2f/2r/ 2).
But

([YOi/ 2f/21 [Yl[YIXI/2f/2f/) = ~U(l1~~j J~) ([yOX I/2f/ 21[[ylylt XI/2f/)

= U(l1~~jO~)(yO[ylylt)·
Using the results

( yO [ylyl]O) __1_ ([ylyl]O) __1_ (VaIVaI) __1_ 1


- .;;r; - .;;r; (1100100) - .;;r; (-f173) ,
one has:

(1IVI2) = ~c (2Ji) (_1j;+1/2 (R20(r)\r\R21(r») [-~l (~)


= ;;(_1)mj +1/2 (R20(r)\r\R21(r» ,

i.e. one obtains (as one must) the same result as in Example 16.2.

EXAMPLE 16.4
Evaluate the matrix element

Using the complete set (cf. Eq. 15.6):

L: Iy.!::,) (y.!::,1 = 5(cos () - cos 0') 5(1) - 1>'),


~~ .

one can write


296 Chapter 16
Applying the Wigner-Eckart theorem (cf. Eq. (16.1) ):

I~ky~,) = L: (kl'qm'lllm") IY~;') (yill [ykyi'] ill) .


I", mil

Premultiplying the above expression by (kl'qm'llm) and summing over q one


obtains with the help of Eq. (15.34)

L: (kl'qm'llm) I~ky~,) = L: (kl'qm'llm) (kl'qm'll"m") IY~:') (yill [ykyi'().


q lI',m",q

i'~11 DIII,I Dmll,m IY~:') (ylll [ykyi'() .

= IY.!.'+q) (yl [ykyl'f) ,


i.e. for any value of 0, </>,

L.J (kl 'qm, I1m) ~k(0, </> )Ym'


'"' I' ( 0, </> ) = Ym'I +q(O, </» \/ yI [ yky1']1) . (16.4)
q

It is convenient here to use the following property of the spherical harmonics, namely:

(16.5)

Substituting the value ° = 0 in Eq. (16.4) one thus obtains:

whence one obtains:

(21' + 1)(2k + 1)
(16.6)
(21 + 1)47r
Finally then:

(21' + 1)(2k + 1)
(Y.!.I~kl y~,) = (kl'qm'llm) (kl'OO Ito)
(21 + 1)47r

EXAMPLE 16.5
Evaluate the matrix element
Tensor Operators, and Evaluation of Matrix Elements 297
As in Example 16.4 one writes:

L [yilt Xl/2]~It) ([yilt Xl/2]~1t I~kl [yll Xl/2]~,)

)(
I", j"

~ "~" (kj' qm'li"m") [I" ~r [l' ~r IH'~rT") .


Premultiplying both sides by (kj'qm'ljm) and summing over q one obtains:

(16.7)

For a particular j there are two possible values of [" with opposite parities, namely
[" = j + 1/2, or [" = j - 1/2 .

However for the reduced matrix element ([[" ~r [k [I' ~r'r) not to vanish, ["

must have the same parity as [' if k is even parity or opposite parity than l' if
k is odd parity. Hence in this case (since one is coupling [' to spin 1/2 ) only one
I" which, to simplify the notation at this point, one can calli, survives. This is not
the case however if one for instance couples [' to a spin 2 1. Thus in the spin =1 case
there are generally three possible values for I (cf. Example 16.6 for general spin S).
In the present case therefore

Again setting () = 0 and using Eq. (16.5) this reduces to:

l.e.

/[I~]j [k[l'~]jl]j) = (kj'Omljm) (1'!Omlj'm) (2k + 1)(2[' + 1) (16.8)


\ 2 2 (qOm Ijm) + 1)47l'
(21
298 Chapter 16
This reduced matrix element should not depend on m, hence one may choose m = 1/2
in Eq. (16.8). One obtains finally:

/ [y, 1/2]i Iykl [Y" 1/2]i') = (kj'qm/ljm) (kj'O~ Ij~) (t/~0~ Ij'~) (2k + 1)(21' + 1)
\ X m q X m' (ItOt Ijt) (21 + 1 )411"

EXAMPLE 16.6
Evaluate (for arbitrary S) the matrix element:

Using the Wigner-Eckart Theorem, Eq. (16.1) and Eq. (15.37) one obtains:

M.E. = I
(kj'qm/ljm) ([IS]j [k [I'S]ir)

= (kj'qm/ljm) 11 U (kl'jSj Ij') ([IS]i I[[kl']I S]i)


= (kj'qm/ljm) U (kl'jSj Ii') (/ 1[kl']')

(21' + 1)(2k + 1)
= (kj'qm/!jm) U (kl'jSj Ij') (kl'OO I/O)
(21 + 1)411"
where one has also used Eq. (16.6).
If S = 1/2, comparison of this result with Example 16.5 implies

(kJ"011'1) (1/10 11"1)


U (kll . ~. 1") (kl'OO 110) = '2 J'2 '2 '2 J '2 .
J 2,] (ItOt Ijt)
EXAMPLE 16.7
For a vector operator A evaluate the diagonal (in j) matrix elements :

where

with analogous expressions for JJ Jf J: u and hence obtain the generalized Lande
formula:
Tensor Operators, and Evaluation of Matrix Elements 299
Applying the Wigner-Eckart theorem to the matrix element (W!..IA!I w!,..), one ob-
tains:
(16.9)

and similarly for (w!n IJ~ Iw~,) ,


(16.10)

Taking the ratio of Eqs. (16.9) and (16.10) one has:

(w!n IA!I w~,) _ (j [A1j]i)


(16.11)
(w!"IJ:1 w!,.,) - (j [pj]i)
Consider next

since
J . A -- J+A_1 +2 L1A+ + Jz Az -- - J11 A1-1 - J1-1 A11 + J.tOJ-'-O·
Ll1

If one inserts a complete set of states L:jll, mil Ij"m"} (i"m"l between the J's and A's
in Eq. (16.12) one obtains:

J(J. A) = -J: L:
j",m"
Ij"m"} (i"m"l A: 1 - J~l L:
j",m"
Ij"m"} (j"m"l A~

+JJ L: Ij"m"} (i"m"l A~


jllt m"

However any component of J1, the angular momentum operator cannot change the
angular momentum of a state. Thus only the term j" = j contributes to these sums.
Also matching the projections one has:

+JJ Ijm} (jml A~ Iw!,.,)


= - (w!..IJ: Ijm -1} (im -11 A: 1Iw!n)
- (w!..1 J~lljm + 1} (im + 11 Allw!n) + (w!..1 JJ Ijm} (iml A~I w!..).
300 Chapter 16
One may now apply the Wigner-Eckart theorem to the r.h.s. of the above expression:

{(lj1m -lljm)(lj -lmljm -1) + (lj -1m + 1Ijm)(lj1mljm + 1)

+(ljOmljm)(ljOmljm)} . (16.13)

A formally identical derivation yields

{(lj1m - 1Ijm)(lj - 1mljm - 1) + (lj - 1m + 1Ijm)(lj1mljm + 1)

+(ljOmljm)(ljOmljm)} . (16.14)

Taking the ratio of Eqs. (16.13) and (16.14) one obtains:

(w!..IJ . AI w!..) (j [Alj]i) (16.15)


(w!n IJ· JI w!n) = (j [pj]i) .

Equating Eqs. (16.11) and (16.15) and recalling that (w!..IJ . JI w!..) = (w!..1 J 2 Iw!..) =
j(j + 1)1i2 one finally obtains the generalized Lande formula.

EXAMPLE 16.8
With the help of the generalized Lande formula (d. Example 16.7) show that:

If

A = J(l), q = 0, J = J(l) + J(2), i.e. J(2) = J - J(l),


the generalized Lande formula becomes:
Tensor Operators, and Evaluation of Matrix Elements 301
But

('11t,.;\ Jz(I) \'11t,.) = E timjl (j1i2mjl m j21 jmj)2, ('11t,.;\ Jz \'11t,.) = timj ,
m;l(mh)

Hence

('11!,.;\J.J(I)\'11!"j) = ('11!,.j\J 2 +J2 (i- J2 (2) \W!,.)

=
ti 2 {j(j + 1) +h(j1 + 1) - h(h + I)}
2
Substituting one obtains finally (where this expression may be compared to Eq.
(15.34) ):

" _ ( ... _I· _)2_ mj{j(j+1)+h(j1+ 1)-j2(h+ 1)}


L.J m'l JtJ2 m 31m]2 }m, - 2j(j + 1) .
mh(mj2)

EXAMPLE 16.9
Evaluate the (possibly) off-diagonal matrix element (cf. Example 16.7 for the
diagonal case) in I and j :

One applies the Wigner-Eckart theorem to the r.h.s of this expression:

M.E. = (1j'qm'ljm) ([l~r [AI [/'~r])


= (1j'qm'!im) t,: U (ll'j~j I"i') ([/~r I[[All']'" ~r) .
If A only operates on the spatial coordinates (e.g. iffor instance A l = £1)

(16.16)

If A only operates on the spin coordinates (e.g. if for instance Al = SI), with the
help of Eq. (15.37) one may write:

M.E. = (1j'qm'ljm)([~/rl[AI [~l'r]) (_1)I/2+/-j+1/2+I'-j'


302 Chapter 16

= (l/qm'ljm)~U (l~jl';Sj') ([~z]' [[Al~r 1']) (_I)1+I+I'-i-l


= (l/qm'ljm) U (l~jl'; ~/) (_I)I+I+I'-i-i' (~ [Al~] 1/) 611' ,

i.e.

(16.17)

x(-I)
1 .., /1
-3-3 \2 [A 11]1/2)
2 611' .

EXAMPLE 16.10
Evaluate the matrix element in Example 16.9 if A! = L!.
From Eq. (16.16):

([I~1:1 L! I[l'~]:,) = (lj'qm'ljm) U (ll'j~; lj') (I [LIZ'f)·


It remains to evaluate the reduced matrix element (1 [LIZ']I) .
The operator L only connects diagonal matrix elements in I. Thus:

(}[I IL~I }[I') = (ll'OZlll) (Z [LIZ'f) 611' = 1MII' .


Using Table 15.3

(Z [LIZf) = ( 1~~lll) = Iii = -liJZ(l + 1) ,


1 -I/y/(l + 1)
and
([/~]: IL! I[I' ~]:,) = - (lj'qm'ljm) U (llj~; 1/) liJI(l + 1) 6/1' .
One distinguishes four possibilities and uses Tables 15.3, 15.7. There are two
diagonal cases in j, namely:
., Z 1 . 1 1
J = +2' J = +2'

for which values:

(ljOjljj) = -J ! j 1 U(llj~; Ij) = (21 + 3)1


(1+1)(2/+1) ,
Tensor Operators, and Evaluation of Matrix Elements 303
·, I 1 . I 1
an d J= -'2' J= -'2'
for which values:

(ljOjljj) = -j j ~ 1' U(llj~; Ij) =


(21- 1)(1 + 1)
1(21+1)

For both these diagonal cases the result for ([I~J:I L~ I[I~]:) using the above

expressions is in agreement with the generalized Lande formula of Example 16.7,


namely:

h2 [j(j + 1) + 1(1 + 1) - 3/4] jh


2h2j(j + 1)

lh, if j=j'=1+1/2,

(21-1)(l+1)h if j=j'=1-1/2.
21 + 1 '
Finally for the two off-diagonal cases

Jl(21~1) if J., = 1 +'2'


U(ll·~.l·') 1 j = 1- ~ and
J 2' J , 2 '

1 . 1
U(llj~;lj') , if J., = 1 - -1 , J =
1
+-.
(l + 1)(21 + 1) 2 2

Thus

J·=I-~2'

and

(lj'Qm'ljm)hj 21 ~1' if j' = 1 - ~, j = 1+ ~.


EXAMPLE 16.11
Write the tensor force:

S12 (1', 0-(1), 0-(2)) == 3(0-(1). f)(0-(2) . f) - 0-(1) . 0-(2), (16.18)


304 Chapter 16

r= r~ - d Ar
where r"2 an r = Ir'I '
and the 0"(1) and 0"(2) operate on the spins of two spin 1/2 particles, in terms of
irreducible tensors.
One notes:

(111 - 1100)A~B:l + (1100100)A~B~ + (11 -11100)A:'IB;


111 111111
y'3A1B_1 - y'3 A oBo + y'3A_1B1

Hence

SIZ (f, 0"(1), 0"(2))

To change intermediate couplings one uses Eq. (15.42):

Thus one can write:

where one notes that e = 0, 1, or 2, and that e = j, since they couple to J = O.


But

A
(
ad
: f 0)
~ =6 ad 6be 6ef
J +2c + 1+
. (2a 1 )(2b 1) .

Hence:
Since, using Eq. (16.19)

the above expression further simplifies:

But (d. Example 15.14)

So finally one has:

If one writes Eq. (16.20) out in detail with the help of the Wigner coefficients:

(22 ± 2 =F 2100) = If, (22 ± 1 =F 1100) = -If, (2200100) = If '


one has:

S12 (f, u(l), u(2)) = 3 { [u 1(1)u 1 (2)]: [flfl]~2 - [u 1 (1)u 1 (2)]: [flfl]~1
+ [flfl]: - [ul(1)Ul(2)]~1 [flfl]:
[u 1 (1)u 1 (2)]: (16.21)

+ [ul(1)Ul(2)]~2 [flfl]: } .
306 Chapter 16
In order to operate with Eq. (16.21) one must write out the various terms in this
equation in a simpler form. Thus.

Yo
2(0.1..)
,'fJ = V[5
'i'(i;
3z -2
r2'
r2 Y,12(O,.I..) = _
'fJ V{15 x + iy ~,
8" r r
y;2(O, .1..) =
2 'fJ
J 15
3211"
(X + iy'\
r)
Similarly one obtains:
[f1f1]: = (1110121)f~f~ + (1101121)f~f~
1 AlAI 1 AlAI
..j2 rlrO + ..j2 rOrl

= 2 ~ (- Xr: ; ) ~ = - ~ iy ~ =
X fJ ~2(O, ¢» ,
and

[r1rl]~ = (1111122)r~r} = (- Xr: ; ) (- Xr: ; ) = fJ y.}(O, ¢».

Summarizing (where [rlfl]~ = [flrl]~M' a result which may be obtained using the
second property of Wigner coefficients noted in Eq. (15.27) ):

[rAlrAI] 2±2 = Vrs:;


15 Y±2(O,
2
¢» ,

[flfl]:l fJ Yil(O, ¢» , (16.22)

[flfl]: = fJ Yo 2 (O, ¢» .
The spin operators must also be simplified:
[a 1(1)a l (2)]:2 = (11 ± 1 ± 112 ± 2)a(1)i1a(2)i1
= ( ax (l) ± ia y (l)) ( a x (2) ± ia y (2))
1=..j2 1=..j2
1 2
= 2a±(1)a±(2) = 1i,2 S±(1)S±(2) ,
Tensor Operators, and Evaluation of Matrix Elements 307
where (cf. Eq. (15.25) )
hu ~-----------
S±=T' and S±X~.=hV(sTm8)(s±m8+1)X~.±1·
Similarly one obtains:

and

Summarizing
2 1 2
[
u1(1)u 1(2) ] ±2 = '2 u±(1)u±(2) = h2S±(1)S±(2)
[u 1 (1)u 1(2)]:1 = T~ {u±(1)uz(2) + uz(1)u±(2)} ,
2
= T h2 {S±(1)Sz(2) + Sz(1)S±(2)} , (16.23)

[u 1(1)u1(2)]: = ~ {3uz(1)uz(2) - u(l) . u(2)}


4
= h2V6 {3Sz(1)Sz(2) - S(l) . S(2)} .

Combining terms one has finally

S12 (r, u(l), u(2)) = h12 J961r


5 { S+(1)S+(2) y_ 2(0,
2

+ {S+(1)Sz(2) + Sz(1)S+(2)} Y~I(O, 4»

+ ~ {3Sz(1)Sz(2) - S(l). S(2)} Yi(O, 4» (16.24)


- {S_(1)SzC2) + Sz(1)S_(2)} 1';.2(0, 4»
+ S_(1)S_(2) Y}(O, 4» } •

The 0, and the 4> in Eq. (16.24) refer to relative angles.


The term in Eq. (16.24) which usually contributes to matrix elements is
308 Chapter 16
since for spin 1/2 particles when

S(l). S(2) = {S(l) + S(2)}2 - S(1)2 - S(2)2


2
operates on triplet spin states it yields

ti21(1 + 1) - 3/4 - 3/4 = ti 2 .


2 4
The other terms in Eq. (16.24) involve S+, S_ etc. and these operators sometimes
make the contributions of these terms zero.

EXAMPLE 16.12
Given the tensor operator (cf. Example 16.11), evaluate the matrix element:

Using the results of Examples 16.4 and 16.11, one has:

I = (~lx~1 SI21~IXO
= ~
ti 2
/967r5 V~3' ti2 (y} 1-v.21 Y,I)
2
I 0 I

1 {9&; f2 f5ti 2
= ti 2 VT V3' "2 V4; (2101111)(2100110)
~2 /9:7r ~ ~2 (f Jio (-fs) = -~.
EXAMPLE 16.13
Given the tensor operator (cf. Example 16.11), evaluate the matrix element:

Using the results of Examples 16.4 and 16.11 one has:

I (~ [~IX~ -l';;txO ISI21 ~ [~IX~ - Yolxn)


= /9~7r ~ [vi (-1) (~1IYo21 ~l) vi (~) (Yo11Yo21Yo1)
+ + v'2 (~11~21 Yl) 1
Tensor Operators, and Evaluation of Matrix Elements 309

~ l:' ~ ff (2100110) [-y1 (2101111) +y1 m +v'2(2100110) (2110111) 1


~ ff-~ff (-/0 [-y1~+~H-~)+v'2(-v'4)l
226
= 5+5+5=2.

EXAMPLE 16.14
Given the tensor operator (cf. Example 16.11), evaluate the matrix element:

Using the results of Examples 16.4 and 16.11 one has:

I = (~ [YlX:l - YoIX~ + Y2 xt] 1S121 ~ [~IX:l - YOIX~ + Y2 xt])


1 1

=
fg6;{ V/23" 2"n2 (~11Yo21 ~1) + V/2(3" -n2) (Yo11 Yo21 Yo1) + V/23" 2"n2 (1 IYo2IY-l1)
3"1 n12 V5 Y- 1

- 2 (~) (Yl11~2IYol) + 2 (n2) (~111';2IY21) + 2 (~) (Yoll~2IY21)}


= ~ /9!7r If; (2100110) {~ ~ (2101111) - ~ (2100110)
+ ~ ~ (210 -111 -1) - ~ (2110111) + 2 (212 -1111) + ~ (211 -1 11O)}
= ! 54 (_ ~) [ ~ ! _1 _ ~ (_ ~) + ~!_1
3 V"5 V3" 2 v'IO V3" V"5 V3" 2 v'IO

- ~ (-V;;) +2 /r+ ~ V;;]


=
1 (/2)
3" 54 - V5
[1 1 2 1 1 3 6
2" v'i5 + v'i5 + 2" v'i5 + v'i5 + v'i5 + v'i5
3]

= ~ 54 (-Is) [~] = -4 .
Comparing the results of Examples 16.12-14 one notes that the relative strength
of the tensor force in the three triplet P states 3 P2 , : 3 PI , : 2 Po , is as -2/5 : 2 :
-4 , where (3 R0,1,2 = [y1X1jO,l,2) .
310 Chapter 16

EXAMPLE 16.15
Given the tensor operator (d. Example 16.11), evaluate the matrix element:

Using the results of Examples 16.4 and 16.11 one has:


I (2101111) (Yixi ISI21 y;,°xO + (2110111) (Y?X~ ISI21 y;,°xO
+ (212 - 1111) (Y;2X:I ISui y;,°Xn
= _1 2- J967r . f!.. fi,2 (y::21y::21 y::O) _ f3 2- J967r (_~) (Y? ly?1 y::O)
v'IO fi,2 5 V'3 2 ° 0 0 V10 fi,2 5 .../2 1 1 0
+ ViF5'. .!...J967r
fi,2 5 (y:21
2 y:21Y::°)
2 0

= 1 roo; /21 1 f3 roo;( 1) 1 f3 roo; 1


v'IO Vs V'3"2 0fi - V10 VS-5- -.../2 0fi + V5' Vs 0fi
.../2 + 3.../2 + 6.../2 = 10.../2 = ..2V2 = v'8 .
=
5 5 5 5
EXAMPLE 16.16
Given the tensor operator (cf. Example 16.11), evaluate the matrix element:

One first notes that:

Using the results of Examples 16.4 and 16.11 one has:


I = 110 (Yixi - V3 ~2X~ + v'6 y22X:II S121y;,2 Xi - V3 ~2X~ + V6 YiX:I)
= 110 (;2 J9~7r ) [(YilYilYo2) fa (~2) + 3(~21y;,21~2) fa (_fi,2)
+ 6(y;2 1y;,2IYi) ~ (~2) _ 2V3 (y;,2IY!II~2) (~) + 2V6 (YilY!21Y;2) (fi,2)
Tensor Operators, and Evaluation of Matrix Elements 311

- 6J2 (l;2I Y!1I Y22) ( - ~)]


= ~
10
(J967r)
5 V-:t; . Pi. ~ - 3(2201121) ~
f5 (2200\20) [(2200\20) Va2 Va
+ 6 (2202122) ~ ~2 -
Va 2V3 (22 - 11120) ~ + 2v'6 (22 -
v2
22120)

- 6J2 (22 -12121) ( - ~)]


= ~ v'24 ( __2 ) [_ _
2 ~ ~ _ 3 ( __
1 ) ~ + 6 (_2) ( ~) 1
10 y'I4 y'I4 V"3 2 y'I4 V"3 y'I4 V"3 2

- 2V3 (_1 ) _1 + 2v'6


y'I4 V2 V"7
~1
~ + 6V"7
1 (2) [1 3 6 3 12 18]
-~+~+~-~+~+~
=1Ov'24-y'I4

= 110 v'24( - ~) [~] = -2.


The matrix elements evaluated in Examples 16.15, 16.16 are needed to solve the
deuteron problem (in nuclear physics).
CHAPTER 17

Applications of Quantum Mechanics

Since its inception, the applications of Quantum Mechanics have been legion.
Starting with the one-electron hydrogen atom, the Schrodinger equation was success-
fully applied to many-electron atoms, molecules, solids, and nuclei, and has been able
to explain scattering phenomena, effects induced by electric and magnetic fields, spin
polarization etc. Moreover when Dirac's equation is used instead of Schrodinger's,
Quantum Mechanics successfully explains the spin and gyromagnetic ratio of the elec-
tron, the existence of particle-antiparticle pairs etc. These applications and related
problems are scattered through the literature. Given below are some applications of
Quantum Mechanics to Mathematics and Optics.

It has recentlyl,2,7 been shown that the quantum-mechanical formalism can be


used to obtain new and interesting series expressions for the mathematical constant
1r. The following ten examples are simple illustrations of this technique.

EXAMPLE 17.1
Consider the complete set for a particle in the one-dimensional oscillator well:

x 2 /2, x ~ 0,
v= {
00, x < o.
The complete discrete set for this potential is :

~n(X) = 2(n~1)/2 (n!~) 1/2 exp ( - ~2)Hn(x) ,


where n is odd (see also Eq. (2.12) ).
It is clear that n must be odd in this case since for this potential ~n(O) = 0 and
only for odd n is Hn(O) = O.
Using this basis evaluate J~i(X)X2~I(X)dx, and with the help of closure obtain
an infinite series expansion for 1r •
One quickly obtains J cpi(x )X2~1(X )dx = 3/2.
But
Applications of Quantum Mechanics 313

= f:
n=1,3 0
1 (>~(x)x(>n(x)dx 1 (>~(XI)XI(>l(x')dx'
00

0
00

= nt 3 [1 00
(>n(X)X(>l(x) dx f
(using Eq. (2.6), interchanging integration and summation, and dropping the * since
these (>(x)'s are all real).
Thus:

(17.1)

But
10 00
iPt(x)xiPt(x)dx =~, 1000
iP t (x)xiP 3 (x)dx = f[; ,
10'>0 iP 1(x)x(>s(x)dx = -J3~11" •

Thus
3 4 2 1
'2 = ;; + 311" + 3011" + ... ,
or

(17.2)
From Eq. (17.1) it is clear that this is a monotonic series, binding 11" from below.

EXAMPLE 17.2
Noting that I;' iPt(x)iPt(x)dx = Iooo iPt (x)(l/x) x iPt(x)dx = 1, use closure and
the complete set in Example 17.1 to get a second infinite-series expansion for 11".

(17.3)

where integrations and summations have been interchanged and complex conjugates
have been dropped since these iP's are real.
Evaluating
314 Chapter 17

10
(00 <I>1(X)~<I>5(X)dx
x
= J 3
0
1 1r
,

and also using the integrals listed in Example 17.1 one obtains:

1 =

=
J;J; -
4 2
;: - 31r - -10-7r - ••• ,
aa -J1~1r J3~1r
1

or
(17.4)
A careful examination of the nth term of this series shows that each term is
negative for n > 1, i.e. this series bounds 1r from above.

EXAMPLE 17.3
Use the complete basis for a two-dimensional harmonic oscillator

1 2
V ="2 P , P ~ 0,

<I> ( ,/..) _ Wnm (p) exp (±im¢»


nm p,'I' - Vii v'2i '
n = 0, 1, ... , m = 0, 1, 2, ... ,
where

(See Eq. (8.14) ), and a procedure similar to that used in Example 17.1 to get an
infinite monotonic series for 1/1r.
The simplest case one can examine involves n, m = o.
r'lr (00 <I>~(p, ¢»p2<I>oo(p, ¢»pd¢>d</> = 10{OO woo(p)lwoo(p)dp = 1.
10 10
Thus, using closure:

However, since p is a scalar, integrating over </>, </>' one finds that m = 0, i.e. that

(17.5)
Applications of Quantum Mechanics 315
Evaluating

rOO ..ft rOO ..ft


io woo(p)pwoo(p)dp = ""2 ' io woo(p)p wlO(p)dp = -4 '
rOO ..ft
io woo(p)pw2o(p)dp = -16" '

and substituting in Eq. (17.5) one obtains:


1 = 7r/4 + 7r/16 + 7r/256 + ... ,or
1 1 1 1
;- = 4 + 16 + 256 + .... (17.6)

and since Eq. (17.5) is a monotonic series this expansion bounds 1/7r from below.

EXAMPLE 17.4
With the complete set of Example 17.3 and the methods of Example 17.2 obtain
a second infinite series which bounds 1/7r from above.
Using the techniques of the previous three examples,

rOO woo(p)woo(p)dp =
io
1= f n
roo woo(p)pwno(p)dp roo Wno(P')!...woo(p')dp'.
io io p'
(17.7)

Evaluating

l o
co 1
woo(p)-woo(p)dp =
p
..ft , 1 woo(p)-wlO(p)dp
o
00 1
p
= -
..ft
2
,

1 00

o
1
woo(p)-w2o(p)dp = -
p
3..ft
8
.

Using in addition the integrals listed in Example 17.3 and substituting into Eq.
(17.7) one thus obtains:

1= (..ft/2) (..ft) - (ft/4) (..ft/2) - (..ft/16) (3ft/8) - ... ,i.e.

1 = 7r/2 - 7r/8 - 37r/128 - ... , or


1 1 1 3
------ (17.8)
7r 2 8 128
A careful examination of this expression shows it bounds 1/7r from above since
all terms, other than the first, are negative.
316 Chapter 17
EXAMPLE 17.5
Show that a different, more quickly converging infinite series than Eq. (17.6)
results if one starts with n = 0, m = 1 (rather than n = 0, m = 0) in Example 17.3.

One first evaluates 10'' ' WOl (p )P2WOl (p )dp = 2.

By the same arguments as in Example 17.3, one then obtains:

(17.9)

Evaluating

roo WOl(P)pWOl(P)dp = -4-


10
3ft
,
roo wOl(p)p wl1(p)dp =
10
3y127r
-~ ,

roo WOl(P)PW21(P)dp = -3'2


10
J31r
'

and substituting in Eq. (17.9) one obtains: 2 = 911"/16+911"/128+3'11"/1024+ ... , or


1 9 9 3
-; = 32 + 256 + 2048 + .... (17.10)

Since, like Eq. (17.5), Eq. (17.9) involves a monotonic series, if truncated, Eq.
(17.10) yields a lower bound for 1/11". The series Eq. (17.10) however converges faster
than that of Eq. (17.6) as can be seen by inspection.

EXAMPLE 17.6
Show that a different, more quickly converging infinite series than Eq. (17.8)
results if one starts with n=O, m=l (rather than n=O, m=O) in Example 17.4.
Using the same techniques as the previous Examples one may write:

1= 00 100 WOl(P)pWnl(P)dp100 Wnl(P')-WOl(P')dP'


E 1 . (17.11)
n 0 0 I
Evaluating

roo WOl(p)-pWOl(p)d
10
1 ft
p = """2 '
Applications of Quantum Mechanics 317
Using additionally the integrals listed in Example 17.5 and substituting into Eq.
(17.11) one thus obtains:

1 = (3..(i/4)(..(i/2) - (3V21f/16) (V21f/8) - (v'37r/32) (v'37r/16) - ... , I.e.


1 3 3 3
;; = 8- 64 - 512 - .... (17.12)

A careful examination of Eq. (17.12) shows that like Eq. (17.8) this series bounds
1/7r from above since all terms, other than the first, are negative. However the series
Eq. (17.12) converges to 1/7r faster than the series Eq. (17.8).

EXAMPLE 17.7
Use the complete basis for the three-dimensional harmonic oscillator
VCr) = r 2 /2 , r ~ 0, namely:
Unl(r) 1
Wnlm(r,lJ, ¢) = --Ym(O,
r
¢), n = 0,1,2 ... , 1 = 0, 1,2 ... , Iml = 0,1, 2... ~ 1,

( ) _{2r(n+l+3/2)}1/2 r'+1exp(-r2/2) F (
Unl r - n! r(1+3/2)
1 3
1 1 -n, + 2,r
2)
(see Eq. (8.13)), to obtain a similar series to that in Example 17.1 but which
converges more quickly than Eq. (17.2).
By the same arguments as were used in Examples 17.1, 17.3 one obtains:

(17.13)

°
Consider the case n = 0, 1 = 1. (The case n = 0, 1= gives identical expressions as
Example 17.1 since in this case there is no angular momentum term 1i, 21(l + 1)/2mr2
in the Hamiltonian and the potentials in these two examples become identical.)
Evaluating

['>Co
Jo uOl(r) ru n(r)dr = -v[32
45; ;
one may then substitute in Eq. (17.13) to obtain:
5 64 32 8
2 = 97r + 457r + 3157r··· ,
or
128 64 16
11" = 45 + 225 + 1575··· . (17.14)
318 Chapter 17
This series converges to 1r more quickly than Eq. (17.2) .

EXAMPLE 17.8
Use the same basis as in Example 17.7 to get a series similar to that of Example
17.2 but which converges more quickly than Eq. (17.4).
By the same type of arguments as are used in Examples 17.2, 17.4 one obtains:

(17.15)

Evaluating

1o
00 1
uOl(r)-uOl(r)dr
r
4
= 3y'-;
1r
1o
00 1
uOl(r)-un(r)dr
r
=
{f
-;
451r

Jor;,o uOl(r)~u21(r)dr
51r
r
= 13 2 ,

and using additionally the integrals given in Example 17.7, for the case n = 0, 1 = 1,
Eq. (17.15) yields:

8 4 {32{8 i8/2
1 = 3";; 3";; - V45;V 45; - V31[;; V35; ... '
or
32 16 4
1r = "'9 - 45 - 105···' (17.16)
a series which converges more quickly than Eq. (17.4).

One may note here that the leading terms of the series obtained in Examples
17.1-8 are all odd or even truncations of Wallis's3 expression for 1r
1r 2 x 2 x 4 x 4 x 6 x 6 x 8.. .
2' = 1 x 3 x 3 x 5 x 5 x 7 x 7... .
Other monotonic series for 1r, 1/1r whose leading terms are not a truncation of
the Wallis's expression may be obtained by using higher odd values of n (i.e. 3,5 ...
rather than 1) in Examples 1,2, or higher values of n, (i.e. 1,2,3 ... rather than 0 )
in Examples 3-8.

EXAMPLE 17.9
Consider the complete infinite square-well set over the range -1 < x < 1, namely:

cI>;(x)= cos en:


1) 1rX (even) = 0, 1, 2, ... n
cI>;;;(x) = sin m1rX (odd) m = 1, 2, ....
Applications of Quantum Mechanics 319
By evaluating
00

(cPtllxI2IcPt) = E(cPtllxllcPp)(cPpllxllcPt) (17.17)


p=o

obtain an expression for 71"4.


From parity considerations p will involve only even-parity states since cPo has even
parity and Ixl is an even-parity function.
Now

while 1 cPt(x)lxlcPt(x)dx = -2 -
1

-1
1 2
2" '
71"
and

Therefore substituting in Eq. (17.17):

1 71"22
3"- =
(1'2- 71"22 )2 + (-71"22)2 + ( -(371")2
2)2 + ( -(371")2
2)2 + ( -(571")2
2)2 + ...

I.e.

or
(17.18)

a standard result4 •

EXAMPLE 17.10
Consider the complete infinite square-well set over the range -1 < x < 1, namely:

cP~(x) = cos en: 1) 7I"X (even) n= 0, 1, 2, ...


cP;;;,(x) = sin m7l"x (odd) m = 1, 2, ....
By evaluating
00

(cPtl x2 lcPt) = E(cPtl x IcP p) (cPpl x IcPt) (17.19)


p=o
obtain an expression for 71"4 (different from the one in Example 17.9 ).
320 Chapter 17
From parity considerations p in Eq. (17.19) will only involve odd-parity states
since <Po has even parity and x is an odd-parity function.
Now

1 <Pt(x)x cl>t(x)dx
-1
1
2
1
= - - '2 '
3
2
7r

while

m = 1,2, 3 ....
Thus
1 1

-1
<Pt(x) x <Pi(x)dx = '2
7r 1
1 25
23 2 ' etc.

Therefore substituting in Eq. (17.19):

7r
2 {~2 _ 2} = 210 C1 ~2 3)4 + (3 ~25)4 + (5 !27)4 .. -} . (17.20)

Combining Eq. (17.20) with Eq. (14.28) namely

one obtains
210 { 7 X 12 19 X 22 39 X 32 }
7r
2
= ""3 (1 X 3)5 + (3 X 5)5 + (5 X 7)5 ... , (17.21)

10 { 23 X 12 83 X 22 183 X 3 2 }
7r
4 = 2 (1 X 3)5 + (3 X 5)5 + (5 X 7)5 ... . (17.22)

In Quantum Optics it is of interest to explore the product:

of the Nth order fluctuations (where the operators chosen are such that (wlxlw) =
(wlplw) = 0). It turns out that 5 the state W which minimizes this product in the
case (WlxNIW) = (WlpNlw) can be found by solving the eigenvalue equation:

~ (pN + xN) W = AN W, (17.23)


Applications of Quantum Mechanics 321
where Afv = UN.
< <
If one relaxes the condition "if! Ix N I"if! ) = "if! IpN I"if! ) this leads to so called "squeezed"
states where for example there can be less uncertainty in x than p. Squeezed light
fields have recently been generated by a number of groups working in this field which
has undergone an explosive growth in the past few years 6 .

EXAMPLE 17.11
In the case N = 2 find the state "if! which minimizes U2 and the corresponding
second-order fluctuation.
In the case N = 2 Eq. (17.23) reduces to the well-known oscillator problem with
solutions "if! = <pn(x) (c.f. Eq. (2.12) with b = 1) in the coordinate representation,
while

EXAMPLE 17.12
For the general case N = 4, 6... find (UN )min. if one assumes the solutions of
Eq. (17.23) are the same states as those which arise for N = 2, i.e. the functions in
Example 17.11.
Premultiplying both sides of Eq. (17.23) by <Po and integrating over x, one obtains

But
1 ~
00
<Pox N <Podx = 100

~
<PoPN <podx ;;;;r (N
.1
= VW -2+-1) .
Thus
A = _1 r
N ft (N 2+ 1) = (U N
)l/~
mm. ' (17.24)

under this assumption for the wave function. (Example 17.11 is a special case of this
expression with N = 2, in which case A2 = (11ft) r (3/2) = 112).
If
N=4, A4 = _l_r (~) = _1_~~y7r = ~, etc.
ft 2 ..fo22 4
Using the duplication formula

r (~) r(2N) = 22N- 1 r(N) r(N +~),


and the identity

N! = (N _1)!!2N/2 (~)! (N even),

one can also write Eq. (17.24) as follows:


322 Chapter 17

(N -I)!!
AN = 2N/ 2 •

EXAMPLE 17.13
Show that the operator

only connects oscillator states which are diagonal, or differ by 4 or more quanta.
For diagonal states there are non-zero matrix elements. Thus in Example 17.12
one obtains (q>oIOIq>o) = (N -1)!!/2 N / 2 • Similarly one obtains

( n.. lOIn.. ) = (N - I)!! {N4 + 4N3 + 20N2 + 32N + 24}


'J.'4 'J.'4 2N / 2 24 ' etc ..

Obviously the operator 0 does not connect states which differ by an odd number
(1, 3, 5... ) of quanta since, like q>n, 0 has definite (even) parity.
Consider next states which differ by 2 quanta, in particular:

if N = 2.
But

=- L: q>2(X) (X2 -1) q>o(x)dx = -(q>2Ix 21q>o).


Similarly, if N = 4

If N = 6,
Applications of Quantum Mechanics 323
and so on for higher even powers of N. Thus generally

Finally consider states which differ by 4 quanta, for instance

A straightforward calculation yields the result:

{q; IOIq; } = (N -I)!! {N(N - 2)}.


4 0 2N/2 2~

Thus the operator 0 connects diagonal elements and off-diagonal elements which
differ by 4 quanta.

EXAMPLE 17.14
Solve the eigenvalue problem, Eq. (17.23), for

o = ~ (pN + x N) N = 2, 4, 6 ... ,

making the more realistic assumption that:

(rather than just q;o(x) ) and find the resulting AN.


The operator equation one has in this case is :

Premultiplying this expression by q;o(x), q;4(X) respectively and integrating over


x one obtains two equations expressible by the matrix relationship:

which in turn implies that the determinant:

{q;oIOIq;o} - AN (q;oIOIq;4)

{q;4101q;o} {q;4101q;4} - AN
equals zero. Otherwise the two linear equations above for 0:, f3 will be singular. This
determinant condition yields a quadratic equation for AN.
324 Chapter 17
Substituting into this equation the values of

obtained in Example 17.13 and noting that all these quantities are real (hence
(cJ>oIOIcJ>4) = (cJ>4101cJ>o) ), the equation can be solved for A, and after some algebra
one obtains:

_ (N-l)!! {f(N)-Jf(N)2_ 768 9(N)}


Aground - 2N/2 48 ' (17.25)

where
f(N) == (N 4 + 4N3 + 20N 2 + 32N + 48) ,
and
g(N) == (N 3 + 2N2 + 4N + 3).
This complicated expression simplifies as N ---t 00 in which case:

(N -I)!! {8 9(N)}
Aground ---t 2N/2 f(N) ,

where the term in the braces and hence also Aground goes to zero as N ---t 00.

One may generalize Eq. (17.23), to include odd positive integer as well as non-
integer values of N (where N > -1), by considering the eigenvalue equation:

(17.26)

which results from finding the \]i which minimizes:

(17.27)

where as above AN = (UN )1/2. These expressions and the corresponding results
reduce to those obtained previously, if N = 2, 4, ....

EXAMPLE 17.15
Obtain AN = (UN )1/2 for general positive N if one assumes \]i = cJ>o(x), the
simple-harmonic oscillator ground-state wave function (d. Example 17.11).

One must evaluate the two integrals:


Applications of Quantum Mechanics 325
and

Firstly

2 10
00
{)/4 exp ( - x;) } x {1r~/4 exp ( - ~2) }dx
N = J,r 10 x (-x )dx
00
N exp 2

= J,rr ( 1) .
N;

The evaluation of I~oo W"'p,N wdp, is easiest done in momentum space (d. Chapter
3) which is an acceptable procedure since the values of matrix elements are indepen-
dent of the basis representation used to evaluate them. Solving for c}o(x) using Eq.
(3.1) one obtains
c}o(p) = 1r1/4 (p2)
1
exp -"2 .
Hence by analogy with the above result:

1 w", INwd
00
-00
p p-
__
1
y1r (N 2+ 1) '
t:;
r and 1 (N- 2+ -1) .
>'N = .,fir
Thus for general N > -1 one merely gets an analytic continuation of the result given
in Eq. (17.24) which was derived for positive, even integer values of N.

References
1. H. A. Mavromatis, J. Approx. Theory 60 1990, pp. 1-10.

2. R. Lynch and H. A. Mavromatis, J. Compo & App. Math. 30 1990, pp.


127-137.
3. H. Eves, An Introduction to the History of Mathematics, Holt, Rinehart & Win-
ston, New York, Fourth Edition (1976), pp. 96-102.
4. 1. S. Gradshteyn and I. M. Ryzhik, Tables of Integrals, Series and Products,
Fourth (English) Edition, prepared by A. Jeffrey Academic Press, New York
(1980), p.7.
5. R. Lynch and H. A. Mavromatis, J. Math. Phys. 31 1990, pp. 1947-1950.
6. Squeezed States of the Electromagnetic Field, J. Opt. Soc. Am. B. 4, 10
(1987).
7. H. A. Mavromatis, Int. J. of Compo Math., 38, 1991, pp. 91-100.
INDEX

Accidental degeneracy(ies) 42, 193, 212 270, 271, 272, 273


Airy function 39, 69, 71, 138, 164 Closed-form expression 251
Analytic continuation 325 Closure vii, 16,24, 63, 172,201,229,263,
Angular momentum vii, 79, 80, 81, 196, 312, 313, 314
220, 262, 299 Commutator(s) 228, 262, 263
coupling(s) xi, 274, 291, 294 relation 256
operator(s) 194,268,269,280,283,284, relationship 252
286, 292, 299 Complete basis 276, 314, 317
quantum number 144, 273, 274, 277, 283 Complete infinte square-well set 318, 319
Antisymmetric combinations 194, 196, Complete set 16, 17, 18, 27, 28, 31, 32,
197, 273, 275, 290 133, 157, 201, 263, 265, 267, 279, 280,
eigenfunctions 196, 275 281,299,312,313,315
state(s) 197,274,277,290 Completeness vii, 16, 27, 263
wave function(s) 193, 197,284 Complex integration 14
Approximate inversion technique 223 Conjugate variables 1
Atomic physics 230, 291 Continuity 4, 6, 84, 97, 122, 176
Attractive potential 71 considerations 74
Average value 110 equation 62
Axis orientation 206, 208 Continuous functions 16, 17
representation 22
Basis (states) 200,201,318 Continuum states 24, 229, 245
representation 325 Converging series 180, 243
Bessel function(s) 79, 121 Coordinate representation 321
Bohr('s) 1 Coordinate space xiii, 36, 45
result 42 Coulomb force 1
Born approximation vii, 74, 84, 87, 88, potential 217, 220
90, 91, 92, 223, 224, 225 system 193
Bose particle 277 Coupled states vii, 262, 274
Bound states 74, 79,99 wave functions 273
Boundary condition(s) 38, 95, 113, 114, Coupling coefficient(s) 269,273,276,277,
115, 124, 126, 133, 138, 186 279
Box with infinite walls 17 Cross normalized 175
function(s) 159, 160, 171,229
Cartesian basis 200, 203 Cross section 80, 82, 85, 89
coordinate(s) 89, 188, 189, 190, 192 Current density 62
coordinate solutions 188 Cylindrical Bessel function( s) 117, 121
Central potential(s) 24, 80, 126,211 coordinates 188, 189
Centripetal term 144
Clebsch-Gordan (coupling) coefficients Dalgarno-Lewis approach 247
Index 327
expressions 230, 261 Eigenvectors 159
function 230, 233, 236 Electric dipole field 208, 209
technique(s) vii, xiii, 227, 232, 249 Electric Field 252
De Broglie 1 Energy eigenvalue(s) 39, 41, 187
Degeneracy(ies) vii, 186, 187, 188, 189, level(s) xiii, 1,4, 7, 8, 9, 10, 11, 12, 13,
190, 191, 192, 193, 194, 196, 197, 198, 42, 113, 114, 119, 189, 190, 192, 193, 197,
199,210,211,212,213 198, 202, 214
Degenerate basis 207, 210, 211 sequence 6
eigenfunctions 205 spectrum 1
functions 199 to first order 166
perturbation theory 204, 208 Even eigenfunctions 176
result 202, 209 functions 184
states 191, 194, 198, 199, 200, 201, 207 integral 38
Delta function(s) vii, 16, 17, 18, 19,20, parity bound states 100
22,23,24,32,47,53,56,57,59,60, 140, parity function 319
179,216,264 parity solutions 43, 44, 45, 139, 177
potential{s) 98, 100, 179, 217 parity state energy 139
property 34 parity trial wave function(s) 139, 140
Determinant 158 solutions 6
Deuteron problem 311 Exact answer(s) 149,227
Diagonal (matrix) elements 199, 323 eigenfunctions 176
Differential (scattering) cross section 80, energy(ies) 144,158,159,173,182,211,
83, 84, 87, 88, 91, 92 231, 238, 241
Dirac's equation 312 first excited-state energy 236, 237
Discontinuity(ies) 78,81, 176 ground state 142, 232, 237, 239, 240
Discrete functions 16 ground-state energy 147, 150,235,247
set 16, 17, 18, 312 ground-state wave function 146, 165,
Doublet 278 166
Duplication formula 321 result 164, 170, 180
second-order contribution 243
Ehrenfest's expression(s) 106, 108, 110 second-order result 245
Eigenfunction(s) 17, 18, 19, 32, 34, 112, solution 39, 144, 257
117,126,139,143,157,158,161,172,175, wave function( s) 160, 170, 175, 178, 204
176,181,187,194,196,199,201,214,227, Exchange degeneracy 196
262,269,273,274,275,276,277,280,282, operator 196, 273
283, 284 Excited Energy Estimates 73
Eigenstates 208, 246, 274, 277 Expectation value(s) 35, 102, 107, 110,
Eigenvalue(s) 117, 139, 143, 157, 158, 161, 126, 229
181,196,199,205,208,210,211,262,269, External wave functions 86
275, 276, 283, 284
equation(s) 139, 207, 320,324 Fermi particle 277
problem 139, 323 Fifth order 249
328 Index
-order expression 249 function 45
Fine structure constant 230 wave function 45, 112, 134, 142, 167,
Finite square well (system) 4, 118,121 232,252324
First Born Approximation 80, 83 Group velocity 54
excited state 31, 192,203,210,211,212,
213, 227, 241, 243 Hamiltonian(s) 45, 103, 110, 133, 139,
excited-state wave functions 237 144,148,149,150,156,157,160,161,162,
order 159, 164, 167, 169, 176, 179, 182, 166,168,169,172,176,178,179,182,183,
183, 201, 203, 230, 231, 232 193,196,197,198,199,200,202,204,205,
order correction 170, 231 207,210,211,227,230,237,239,240,241,
order energy(ies) 162, 168, 180, 236, 253 243,246,247,251,317
order (energy) contribution(s) 184,238, Harmonic oscillator potential 18, 43
239 Heisenberg formulation 6
order energy term 227 representation(s) vii, 102, 103, 104, 108
order expression 208 Hermiticity 227
order perturbation theory 168, 181 High-lying states xiii
order results 227 Hille- Hardy formula 25
Five-dimensional harmonic oscillator 212 Hydrogen 42, 208
Four-dimensional harmonic oscillator 212 atom 1, 42, 208
Fourier expansion 18 problem 72
sine transform 222
transform 34 Infinite barrier result 84
Fourth order 249 discontinuities 80
contribution 251 repulsive barrier result 86
Free parameter 140, 141, 142 series (expansion) 255, 264, 312, 313,
particle(s) vii, 36, 54, 74, 102 315,316
particle Schrodinger equation 52 square-well potential 17
particle wave function(s) 22, 23, 47, 53 sum( s) 176, 227, 267
walls 1
Gel'fand-Levitan-Marchenko equation we1143,44
221 Intermediate coupling(s) 280, 282, 286,
Generalized Lande formula 298, 300, 303 304
Graphical solution 202 j-j coupling 286, 288
techniques 201 orderings 280
Green's function 53, 62 Internal wave functions 86
Ground state xiii, 27, 29, 30, 31, 72, 107, Inverse problem vii, 214
108,109,133,180,184,188,192,197,205, Irreducible (tensor) operator(s) 291, 292,
210, 211, 212, 213, 22~ 237, 251, 254 293,294,304
eigenvalue 146
energy(ies) vii, 68, 69, 70, 71, 73, 134, Kinetic energy 5, 107
142,146,148,161,168,169,170,179,181, operator 108
182,237,239,240 Kramers' formula 127
Index 329
relation 131 Nine-j 286, 287, 288
techniques 132 Non-diagonal matrix elements 107,227
type expressions vii, 124 Normalization(s) 18, 21, 23, 38, 40, 112,
177,193,197,292
Lamb shift 208, 209 condition 121, 265
Legendre's duplication formula 25 constant 265, 267
polynomials 265 Nth-order fluctuations 320
Linear approximation 61, 62 Nuclear physics 291, 311
equations 323
term(s) 187 Odd eigenfunctions 173
Lower bound 316 functions 184, 265
L-S coupling 286, 288 parity 242, 265, 297
-parity bound states 100
Matrix element(s) vii, 132, 166,227,229, -parity function 320
291,292,294,295,296,298,299,301,302, -parity solutions 43, 44, 45, 111, 112,
307, 308, 309, 310, 311, 322, 325 139
Mehler's formula 19 -parity state energy 139
Minimization condition 151 -parity trial functions 139, 140
Minimum energy 149 -parity wave functions 43
values 150 solution 6
Mirror reflection 43 One-dimensional case 126, 128, 256
Molecular Physics 291 Coulomb potential 217
Moments 126 eigenvalue equation 214
Momentum-dependent potential 36 expression 24
distribution 64, 65 Hamiltonian 228
representation 36, 37, 39, 42 (infinite) harmonic oscillator 25, 126,
space vii, xiii, 34, 35, 36, 39, 47,52,325 142, 161
-space function 35 oscillator well 312
-space wave function 45 potential 74
Monotonic series 313, 314, 315, 316 problem(s) vii, 111, 112, 126, 127,220
Multipole operator 292 Schrodinger equation 24, 111
system(s) 113, 186,228,241,258,259
N-dimensional Coulomb potential (well) well 114
119, 120 One-electron hydrogen atom 312
harmonic oscillator 120, 213 Optics 312
potential 120 Optimization requirement 149
problem(s) vii, 111, 118 Orbital angular momentum (operator)
Schrodinger equation 120 262, 268, 286
Negative energies 79, 80, 115 quantum number 24
Neumann functions 79 Original energy 31
New basis 199 Orthogonality 265, 280, 281
Newton's second law 104 Orthonormal 21, 280, 284
330 Index
eigenfunctions 275 Projection(s) 196,273,274,278,283,284,
set 16 291
states 278
symmetric states 273 Quadratic equation 323
Orthonormality 63, 279 Quanta 322
Oscillator problem 321 Quantization condition 15, ll8, 121, 186
Overall phase 178, 205 Quantized energies 1, 15
Overlap 31 Quantum Mechanics iii, vii, xi, 6, 15, 214,
262, 312
Parity vii, 129, 208, 228, 242, 265, 297, Quantum-mechanical formalism ix, 312
322 number 274
considerations 46, 109, 133, 141, 183, operators 104
184,209,229,234,235,236,243,247,249, paradigm ix
253,319,320 scattering theory 221, 222
operator 138, 227, 228 state 125
Partial-wave Born approximation 80, 82, system( s) vii, xiii, 68, 71, 133, 227
89, 91, 93, 222 wave functions 24
results 83 Quantum optics 320
Particle in box with infinite walls 1, 7 Quartet 278
interchange 196 Quintuplet 276
Permutation operator 196, 197
Perturbation(s) 157, 166, 168, 175, 176, Radial Schrodinger (eigenvalue) equation
179, 180, 181, 198, 210, 2ll, 236, 237, 24, 187
252, 256, 258, 259, 260 wave function 79
expansion 198 Rate of convergence 245
result 175 Rayleigh-Schrodinger expansion 159
theory vii, xiii, 157, 160, 161, 165, 166, expression 160
171,182,183,184,201,209,227,230,232 Recoupling coefficient(s) 281, 286
theory decomposition 241, 244 matrix 282
Phase 265, 278 Recursion relations 127
Phase shift(s) vii, 74, 79,80,82,84,214, Reduced matrix element 292, 297, 298,
222, 223, 224, 225, 236 302
Plane wave 56, 57, 64 Reflection amplitude 74, 79
expression 56 probability 74
Pole(s) xiii, 74, 76, 77, 80, 82, 98, 99 time(s) 53, 77, 94
Potential Energy 4, 6 Representation(s) 17, 18, 19, 20, 32, 56,
Probability 267, 268 57, 59, 60, 179
density 62
distributions 15 Scalar 292
Product functions 276, 277 Scattering amplitude xiii, 80, 82, 84, 87,
solution 193 90,98,99
states 273, 274, 283 cross section 82
Index 331
data 218 operators 306
Schrodinger (eigenvalue) equation xiii, 2, Spinless particle 219
4,5,6,8,9,10,15,34,35,36,37,39,53, Square normalizable 40
76, 139,157, 172, 181, 186, 191,201,210, well 4, 86
220, 312 -well scattering 86
formulation 6 Squeezed states 321
representation 102, 103 Stationary phase arguments 95
result 8, 13 Step barrier 94
Second energy eigenvalue 112 potential 219
excited state 109, 200 Sum rules 125
order 160, 164, 167, 170, 171, 175, 182, S-wave phase shift 89, 91, 93
183, 184, 185, 232, 250 scattering 83
-order contribution(s) 170, 183, 184, Symmetric combinations 194, 196, 197,
239, 240, 241 273,275
-order correction 164 eigenfunctions 196, 275
-order differential equation 186 potential( s) 43, 44
-order energy 163, 180, 243 state(s) 196, 197,274,277
-order energy contribution 207 wave functions 193, 197
-order energy expansion 253, 255 well 44
-order (energy) term(s) 228, 245 Symmmetry arguments 283
-order fluctuation 321
-order result 163 Taylor series 178
-order series 180 Tensor algebra xi
-order term(s) 163, 185 force 303, 310
Series expansion 180 operator(s) vii, 291, 308,309,310
Shielded Coulomb potential 94, 220 Third-order 159, 160, 164, 175, 230, 231,
Short range central three-dimensional po- 239, 240
tential219 contribution(s) 239, 240
Simultaneous eigenfunctions 263 energy 164, 229
Simple harmonic oscillator 107, 109, 324 result 164
potential 106, 108, 109 term 164
Singlet 274, 276 Three angular momenta 281
Spatial coordinates 301 Three-dimensional analogue 80
localization 47 Coulomb potential 192
Spherical Bessel function(s) 26, 117, 121, Coulomb system 193
222 delta function 23
coordinates 23, 192 harmonic oscillator 213, 317
harmonic(s) (functions) 117, 263, 265, harmonic oscillator potential 24
266, 267, 292, 293, 296 inverse problems 219
Neumann functions 79 potential 80, 83, 124, 219
Spin angular momentum 286 problem(s) vii, 111, 127, 193
coordinates 301 radial function 112
332 Index
radial potential 111
simple harmonic oscillator (system) 144, U-coefficients 281, 282, 283
180, 192, 193 -functions 281
system(s) 12,24,79,191,212,251,260 Uncertainty principle vii, xiii, 47, 68, 151
well 112, 116 relation 43
Time delay vii, 74, 75 Unperturbed energy(ies) 175, 179, 210,
Time-independent potential 102 246, 254
-dependent Schrodinger equation 62 Hamiltonian 203, 236, 253, 254, 255
Total angular momentum (operator) 269, wave functions 176
274,277,280,283,286 Upper bound(s) vii, 133, 139, 144, 148,
(differential, scattering) cross section 149, 230, 232, 238, 240
80, 84, 87, 88, 91, 92 -bound results 144
energy 4, 6
Transcendental equation 177 Variance 42, 47, 229
Transit time 100 Variational approach 164,227
Transition rates 15 calculation 69, 71
Transmission 96, 98 results 164
amplitude 74, 76, 79, 80 technique xiii
probability 74 Vector (operator) 292, 298
time 53,77 Virial theorem vii, 124, 125, 156
Trial function 151
wave function(s) 133, 134, 136, 137, 138, Wallis's expression 318
140,142,143,144,145,146,147,148,149, Wave function(s) 27, 29, 31, 32, 34, 38,
150, 154, 155, 181, 232 42, 46, 47, 51, 52, 55, 56, 74, 75, 79, 80,
Triplet 274, 276 81, 84, 97, 98, 103, 113, 115, 116, 118,
(spin) states 308, 310 119, 121, 124, 132, 138, 165, 170, 171,
Truncation(s) 318 175, 176, 186, 187, 188, 189, 193, 196,
Two angular momenta 268, 269 197, 205, 214, 215, 216, 217, 218, 219,
-dimensional Coulomb potential (well) 220, 227, 230, 231, 232, 250, 274, 276,
115, 191, 193 277,285,291,292,294,321
-dimensional (harmonic) oscillator po- to first order 176, 229, 230, 232
tential188, 193 Wavepacket(s) vii, 47, 49, 53, 54,62, 74,
-dimensional oscillator 200, 203, 314 77,95, 100
-dimensional oscillator well 116 Weak potentials 80
-dimensional problems vii, 111 Wigner (coupling) coefficient(s) 270, 271,
-dimensional Schrodinger equation 115 273,276,277,280,290,291,305,306
-dimensional square box 181,210 -Eckart theorem 291,294,296,298,299,
-dimensional system(s) 12, 187, 213 300,301
-dimensional well 117 Wilson-Sommedeld procedure 15
-fold degenerate 191 Quantization condition vii, xiii, 1, 3, 4,
-particle system(s) 193, 197 5,6,7,8, 13
-step barrier 96 result 39, 45, 70, 71
Index 333
W. K. B. approximation 15
approximation result 15

Yukawa potential 94

Zero-order 159
-order eigenfunctions 175
-order energy 211
Kluwer Texts in the Mathematical Sciences

1. A.A. Harms and D.R. Wyman: Mathematics and Physics of Neutron Radiography.
1986 ISBN 90-277-2191-2
2. H.A. Mavromatis: Exercises in Quantum Mechanics. A Collection of Illustrative
Problems and Their Solutions. 1987 ISBN 90-277-2288-9
3. V.I. Kukulin, V.M. Krasnopol'sky and J. Horacek: Theory of Resonances. Principles
and Applications. 1989 ISBN 90-277-2364-8
4. M. Anderson and Todd Fell: Lattice-Ordered Groups. An Introduction. 1988
ISBN 90-277-2643-4
5. J. Avery: Hyperspherical Harmonics. Applications in Quantum Theory. 1989
ISBN 0-7923-0165-X
6. H.A. Mavromatis: Exercises in Quantum Mechanics. A Collection of Illustrative
Problems and Their Solutions. Second Revised Edition. 1992 ISBN 0-7923-1557-X

KLUWER ACADEMIC PUBLISHERS - DORDRBCHT / BOSTON / LONDON

You might also like