You are on page 1of 41

STUDIES ON

PHASE AND COHERENCE IN QUANTUM


SYSTEMS

A THESIS
Submitted to the University of Madras
in Partial fulfilment of the requirement
of the Degree of

DOCTOR OF PHILOSOPHY

By

KUMAR ESWARAN
1

Department of Theoretical Physics


University of Madras
Madras 6000025
INDIA

January 1973

1
The support of the University Grants Commission is gratefully acknowledged
PREFACE

This thesis embodies the results of the investigations carried out by the au-
thor during the period 1969-1972 in the Department of Theoretical Physics of
the University of Madras, on the phase and coherence aspects of quantum sys-
tems.
The thesis consists of four chapters and an appendix. Chapter II embodies the
results of our studies on generalized phase operators for the quantum harmonic
oscillator. In Chapter III we present an approximate solution for the energy
levels of the anharmonic oscillator, which we obtain through the use of action
and phase variables. Chapter IV concerns itself with the behaviour of small
wavepackets in spherically symmetric potentials and, in particular, the question
whether there exist any wave packets which remain coherent for all time.

The work reported in this thesis has been done in collaboration with and
under the guidance of Professor P.M. Mathews. It forms the subject matter of
five papers, two of which have already been published:

1. On the Generalized Phase operators for the Quantum Harmonic


Oscillator, Il Nuovo Cimento 70 B, pp 1 -11 (1970)
2. On the Energy Levels of the Anharmonic Oscillator,
(with P.M. Mathews), Lett. al Nuovo Cimento 5, pp 15 - 18, (1972)

and three others (see footnote)1 which are in the course of publication.

To the knowledge of the author the results of the thesis have not formed any
part of any earlier thesis.

ACKNOWLEDGEMENTS

The author is very deeply indebted to Professor P.M. Mathews without


whose inspiring guidance and continuous encouragement throughout the entire
period of research, this work could never have been completed.

He is grateful to Dr. M. Seetharaman and Mr. J. Jayaraman for fruiful


discussions. He is also grateful to his colleagues Mr. Lakshmanan, Mr. Sekhar
Raghavan and Mr. J. Prabhakaran for their cooperation.

He wishes to acknowledge his gratitude to the University Grants Commis-


sion for financial assistance in the form of a research fellowship.

Finally thanks are due to Mr. Mohamed Jawad for his excellent typing of
the thesis.
1 NOTE ADDED: (by Author) These three papers have subsequently been published

and their references are as under:


3. Semi-Coherent States of the Quantum Harmonic Oscillator, (with P.M. Mathews),
Il Nuovo Cimento 17 B, pp 332 - 335 (1973)
4. Simultaneous Uncertainties of the Cosine and Sine Operators, (with P.M. Mathews),
Il Nuovo Cimento 19 B, pp 99 - 104 (1974)
5. On the Evolution of Wave Packets for Particles in Central Potentials,
(with P.M. Mathews), Journal of Physics A, 7, pp 1547 - 1556 (1974)

1
CONTENTS

Chapter I : Page 3

Introduction

Chapter II : Page 9

On the Generalized Phase Operators for the Quantum


Harmonic Oscillator

Chapter III : Page 21

The Action & Phase (Angle) Variables & the Energy Levels
of the Anharmonic Oscillator

Chapter IV : Page 26

Coherence in Quantum Mechanical Systems

Chapter V : Page 36
Appendix and References

2
Chapter 1

Introduction

This thesis deals with some of the problems of the quantum theory which de-
spite the varied successes of the theory since it first originated [1-4], have yet
remained unresolved. Broadly speaking the problems we consider have to do
with the role of the phase variable in relation to specific quantum systems (the
harmonic and the anharmonic oscillators) and the question of coherence in three
dimensional systems.

1. The Phase Operators for the Quantum Harmonic Oscillator

The problem of the phase variable for the quantum harmonic oscillator arises
when one tries to find a well behaved quantum mechanical opeator correspond-
ing to the phase or angle variable φ (canonically conjugate to the action variable
J) of the classical oscillator. The existence of a operator which is conjugate to
the number operator was postulated many years ago by Dirac [5]. Dirac pro-
posed to introduce a phase operator φop through the definition
1
a = exp(iφop )N 2 , (1.1)
where N is the number operator:
1
N = ~(N + ), N = a† a (1.2)
2
a and a† being the familiar annhilation and creation operators obeying the
commutation rule
[a, a† ] = 1 (1.3)
The equation (1.1) was meant to be the analogue of the classical expression
mω 1 1
( ) 2 (x + ip/mω) = eiφ J 2 , (1.4)
2
which follows from the definition of the classical phase through
2J 1 1
x=( ) 2 cosφ, p = (2Jmω) 2 sinφ, (1.5)

If exp(iφop ) of (1.1) is taken to be an unitary operator, as one would expect,
then the sustitution of (1.1) in (1.3) would immediately yield
exp(iφop ) N exp(−iφop ) = N + 1, (1.6)

3
But this cannot be: a similarity transformation of N cannot change it to N+1
since the spectrum of eigenvalues of the two operators cannot coincide. In fact
the impossibility of defining an unitary operator exp(iφop ) through (1.1) was
first pointed out by Susskind and Glogower [6]. Indeed if one assumed the
existence of such an operator satisfying the commutation relation

[N, φop ] = i (1.7)

suggested by the Poisson bracket relation {φ, J} = 1 , inconsistencies would


immediately arise. 1 For on sandwiching (1.7) between the number states |mi
and |ni one obtains
(m − n)hm| φop |ni = i δmn , (1.8)
which would require that expectation values of the supposedly hermetian oper-
ator φop should be infinite and imaginary!
Though an unique phase operator cannot be had for a quantum harmonic
oscillator, one can introduce the notion of phase as is shown in the pioneering
work of Louiselle[10 ] and the subsequent paper of Susskind and Glogower [6],
by the use of a pair of hermetian operators called the cosine and sine operators.
These operators 2 are intended to be the quantum equivalent of the cosine and
sine of the classical phase. The commutation relations of C and S with N are
suggested by the Poisson brackets of J with cosφ and sinφ :

[C, N ] = i S and [S, N ] = −i C (1.9)

or equivalently, with
U = C + iS , (1.10)
[U, N ] = U ; (1.11)
from (1.11) it is easy to see that U is a step down operator for the number
eigenstates.
Taking,
U |ni = (1 − δn0 )|n − 1i (1.12)
Susskind and Glogower [6], have solved the eigenvalue problem for the C and S
operators, and shown that their eigenvalue spectra are continuous and fill the
range (−1, 1) and that the eigenstates form a complete set.

1 Varga and Aks, Ref. [7], claim to have obtained a representation of equation (1.7) in a

basis related to the positive momentum eigenfunctions of a particle in a one dimensional box
(−1/2, 1/2), see also Refs. [8] and [9]. Their argument rests on the identity of the eigenvalue
spectra of VP (P,Q being the momentum and position operators of the particle in the box
and V the projection operator to the states of non-negative momentum) and of the number
operator N, from which they infer that these positive momentum states and the number
eigenstates are related by an unitary transformation. They propose then that φop be defined
through the unitary transformation V QV → φop . However, their arguments seem untenable:
while the zero eigenvalue of N is nondegenerate, that of VP is infinitely degenerate(since all
negative momentum eigenstates of P become eigenstates belonging to the eigenvalue zero of
VP). Thus the assumed unitary transformation does not exist.
2 A similar approach as has been employed by Louisell [10 ] is useful in dealing with the

problem of the azimuthal angle conjugate to Lz in angular momentum theory, eg. see Ref.
[11], for the discussion of the phase problem in angular momentum. Carruthers and Nieto
Ref. [12], provide an excellent review on these subjects.

4
The cosine and sine operators defined through (1.10) to (1.12) are not the
only possible operators that can be obtained. It was pointed out by Lerner [13]
that equation (1.12) is not really mandatory, one could have more generally

U |ni = f (n)|n − 1i ; with f (0) = 0 , (1.13)

where f (n) is arbitrary except for certain restrictions that are needed to ensure
that the C and S have eigenvalue spectra extending from -1 to +1. Lerner has
given examples for various forms for f (n) , following from the use of various
kinds of symmetrizations in going from the classical to the quantum variables.
The eigenstates for the C and S operators for a particular choice of f (n) have
been worked out by J. Zak [11]. Since U is not an Hermetian operator the eigen-
values of U will not be necessarily real. Lerner, Huang and Walters [14] have
proved the interesting fact that all complex numbers z with |z| < 1 are eigen-
values of U. The corresponding eigenstates form an overcomplete set. Recently,
some of the spectral properties of the phase operators have been obtained also
by Ifantis [15],[16] using the observation that C and S belong to a special class
of tridiagonal operators.

We have made a systematic study of the eigenvalue problem of the gener-


alized phase operators C and S defined by Lerner through (1.13), (1.10) and
(1.11) above. The particulars of this investigation are presented in Chapter II.
We prove that the coefficients of the transformation from the number eigenstates
to the eigenstates of the C and S are quite generally orthogonal polynomials in
the eigenvalue parameter. This enables us to show that the eigenstates of the
generalized C and S operators form a complete set. It is interesting that such
properties are demonstrable even when f (n) is arbitrary. In the special case of
f (n) = (1 − δn0 ) the coefficients of the expansion of the C and S eigenstates in
terms of the number eigenstates reduce to the familiar Gegenbauer polynomials.
We also treat the problem of the minimum uncertainty state for the general-
ized phase operators, following the approach of Jackiw [17] and prove that the
states which minimize [(4C)(4S)/hi[C, S]i]2 form an overcomplete set and are
labelled by a parameter λ taking values within an ellipse in the complex λ plane.
The coherent state representation of quantities of special interest involving the
C and S operators is then presented. We show that it is always possible to con-
struct phase operators C and S which commute within a class of coherent states
|αi characterized by |α| = constant. An example of such a pair of operators is
given. This is the closest we have been able to get to the quantum mechani-
cal counterpart of the Poisson bracket relation {Cosφ, Sinφ} = 0 reflecting the
uniqueness of phase in classical mechanics.

2. Action and Phase (Angle) Variables and the Energy Levels of


the Anharmonic Oscillator

The problem of the anharmonic oscillator has been the subject of a great
deal, [18] -[20], of recent work which has been motivated largely by the hope
of exploiting the analogy of the anharmonic oscillator and the nonlinear fields
to gain some insight into the latter. The results are however of considerable
interest. Bender and Wu [19] showed that if the energy eigenvalues En of the

5
hamiltonian
p2
H= + αx2 + βx4 (1.14)
2m
are calculated by a perturbation method in powers of β the series fails to
converge for any β . They have traced the reason for this divergence to the
peculiar analytical properties of En (β) in the complex β plane. Despite the
singularities of En (β), it turns out that the values of En for physical β can be
clearly reproduced by P adé approximants [21]-[22]. Other methods of comput-
ing E (eg. in terms of the roots of a Hill determinant) have also been explored
recently[23]-[25]. However throughout this work one does not find a simple for-
mula for quick and practical computation of the energy levels. We show in
Chapter III of this thesis that by a semiclassical method based on the use of
the action and angle variables an approximate formula which serves this pur-
pose can be obtained. It not only reduces to the harmonic oscillator spectrum
as β → 0 but also reproduces the known asymptotic (large n) expression for
En (β) . We also derive a scaling relation (for the exact eigenvalues) which is
more general than any previously known in the literature [20] and show that
our approximate formula for En (β) does obey this scaling relation.

3. Coherence of Wave Packets in Three Dimensions

It is well known that attempts to reconcile wave-mechanics with classical


mechanics by interpreting particles as wavepackets were not completely sucess-
ful because quantum mechanical wavepackets do not keep a constant size for
all time [26],[27] . The size of a free wave packet increases indefinitely as time
passes. Though the rate of spreading in the case of macroscopic objects is so ex-
tremely small as to be totally unnoticable, the very possibility of spreading has
appeared to some as too disturbing to be acceptable. Louis de Broglie notably
has suggested [28] that the Schrodinger equation should be replaced by some
nonlinear equation which would preserve the main quantum mechanical struc-
ture but inhibit the spread of such wavepackets3 . As an alternative possibility,
the use of singular solutions4 of the Schrodinger equation has been advocated,
for describing point particles [32]-[34],[28] . There has also been an attempt
[35] to link the problem of the spread of wavepackets with the curvature of
space-time in the general theory of relativity and show that a theory that takes
this into accout will permit the motion of wavepackets along definite trajecto-
ries without unlimited spreading. None of these attempts can be said to have
really made much headway, and the standard form of quantum mechanics has
the allegiance of practically all physicists. It is surprising however that hardly
any attempt has been made within the ambit of orthodox theory, to study the
motion of wavepackets even in simple realistic systems. Besides the free particle,
the harmonic oscillator (in one or more dimensions) seems to be the only other
system to have been extensively investigated. It has been long recognised that
the latter, unlike the free particle has wavepacket states such that the size of
the packet remains time independent. These are the coherent states [36]- [39]
3 For a discussion of the spreading of a ‘classical wavepacket’ (i.e. the group of samples of

a statistical ensemble around some point in phase space) see M.Born and D.J. Hooton, and
M.Born Refs. [29]-[31].
4 Such solutions have been called ‘The Double solutions’, for a discussion of this, see Ref.

[28].

6
already referred to in Sec.1, and they are characterised by a minimum value for
the product of the extensions in coordinate and momentum spaces. But even in
the case of such an important system as a particle in a coulomb-type potential,
nothing seems to be known about the behaviour of a small three dimensional
wavepacket as time progresses, and in particular, as to whether there exist co-
herent (in the sense of non-spreading) wave packets in such a potential.

In view of this situation it seemed interesting to investigate the behaviour of


wavepackets bound in spherically symmetric potentials of three dimensions. We
present in Chapter IV the details of such a study. Our aim is to check whether
there exist wave packet states in which the uncertainties in the various compo-
nents of position, momentum etc. are time independent. We will refer to such
states as coherent states, for brevity, but it must be noted that the usage of this
term in the context of the harmonic oscillator implies much more than the mere
constancy of the uncertainties. To emphasise this point as well as to introduce
the method of attack in the more complicated three-dimensional case, we first
obtain the equations which govern the time variation of position and momentum
uncertainties associated with wavepackes in the harmonic oscillator potential.
We show that there exist a large class of states representing wavepackets whose
mean positions move simple harmonically and whose sizes remain time indepen-
dent but which are not necessarily coherent states in the standard terminology,
we term such states as semi-coherent states. An interesting special class of such
states is exhibited.

We then turn to study the coherence aspect of a wavepacket in an arbitrary


central potential in three dimensions. The interlinking of different degrees of
freedom through the potential makes the problem quite complex, but when the
trajectory of the mean centre of the packet is circular the symmetry of the
system can be taken advantage of to make the problem tractable. It turns out
that the uncertainty in the position and momentum components perpendicular
to the plane of the orbit have a simple behaviour. As far as the uncertainties
within the plane of the orbit are concerned, it is an advantage to define them
through the following quantities:
χR = h(~x − h~xi)2R i, (1.15a)
χT = h(~x − h~xi)2T i, (1.15b)
πR = h(~p − h~pi)2R i, (1.15c)
πT = h(~ pi)2T i,
p − h~ (1.15d)
where the subscripts R,T refer to component in the radial and the tangential
direction respectively:
h~xi
(~x − h~xi)R = (~x − h~xi). ,
R
h~
pi
(~x − h~xi)T = (~x − h~xi). , (1.16)
P
etc., where R, P are the radius of the circular orbit and the mean momentum
in the orbit respectively.
The time derivative of the quantities (1.15) evaluated using the quantum
mechanical equation of motion(in the approximation of small size of the wave

7
packet) involve various cross correlations eg. h(~x − h~xi)R (~
p − h~
pi)T i. It turns
out that one can get a closed system of first order differential equations with
constant coefficients for a set of six such correlated functions together with the
four quantities (1.15). Taking these ten quantities as the elements of a ten
component column vector u, we can write the equations in the matrix form:
du
= ωM u, (1.17)
dt
where M is a 10×10 matrix in which the particular central potential enters, only
through a single parameter, and ω is the frequency of the orbit. The question
of coherence of the wave packet then boils down to whether (1.17) has solutions
such that dudt = 0. We find that formally such solutions do infact exist, since
M happens to have zero eigenvalues (four of them); but it turns out that none
of these solutions is compatible with the essential positivity of the quantities
(1.15) . This difficulty manifests itself through the non-diagonalizability of the
matrix M . Positivity can be ensured only by allowing some of the uncertainties
defined through (1.15) to increase indefinitely. This is the case in particular for
the uncertainty in the tangential components of position: There is no choice of
initial conditions by which spreading of the wavepacket in this direction can be
prevented. The case of πR (which measures the uncertainty in the radial com-
ponent of momentum) is similar. On the other hand the tangential component
of momentum and the radial component of position as well the components per-
pendicular to the plane of motion remain well defined, their uncertainties being
constant or varying harmonically at worst. The details of these calculations are
carried out for the case of particular interest, namely the coulomb potential.
But we observe that the results hold quite generally, for central potentials, the
only exception (where true coherent states exist) being the three dimensional
harmonic oscillator. In this case the M matrix becomes diagonalizable.

8
Chapter 2

On the Generalized Phase


Operators for the Quantum
Harmonic Oscillator

1. Introduction

We have noted in the last chapter, though an unique well-defined phase op-
erator conjugate to the number operator N of the quantum harmonic oscillator
does not exist, the notion of phase can be introduced through cosine and sine
operators C and S defined through the commutation relations:
[C, N ] = i S and [S, N ] = −i C (2.1)
or equivalently, with
[U, N ] = U where U = C + iS . (2.2)
Equation (2.2) implies that
U |ni = f (n) |n − 1i , with f (0) = 0 , (2.3)
where the sequence of numbers f (n) has to be so chosen as to ensure that the
hermetian operators C and S have the right kind of eigenvalue spectrum, namely
a continuous one extending from -1 to +1 , consistent with their interpretation
as cosine and sine operators. It was Lerner’s observation that the necessary and
sufficient conditions for this purpose5 are that
f (n) → 1 , as n→∞ (2.4)
and that it be possible to express the f (n) in terms of another sequence
gn (0 ≤ gn ≤ 1) (2.5)

1 2
f (n) = gn (1 − gn−1 ) . (2.6)
4
5 See Lerner,Huang and Walters Ref. [14] for a detailed proof of the ‘chain sequence condi-

tion’ (2.6) , and for the spectral properties of U .

9
Subject to these conditions one can make any of a vast variety of choices for f (n)
and hence for C and S. This freedom of choice raises interesting possibilities.
One may ask, for instance, whether one can define C, S in such a way that

hα|[C, S]|αi = 0 (2.7)

within some class of coherent states. Such a question is prompted by the fact
that coherent states are wave packets with a classical motion and classically,
the cosine and sine functions have a vanishing Poisson bracket, the phase angle
being uniquely defined. We will consider this in section 5, but first we solve the
the eigenvalue problem for C and S in all generality. We show in section 2 that
if
C |λi = λ |λi , (2.8)
and |λi is expanded in terms of the number eigenstates |ni as
n=∞
X
|λi = Cn (λ) |ni , (2.9)
n=0

then the Cn (λ) form a set of orthogonal polynomials in λ . We also verify that
the states |λi form a complete set, with the eigenvalue parameter λ ranging
from -1 to +1 (provided the chain sequence condition holds). It is remarkable
that there are such elegant properties associated with the phase operators and
that they are demonstrable even when the f (n) is quite arbitrary. For the famil-
iar special case [6],[12] when f (n) = (1 − δn0 ) , the coefficients of transformation
Cn (λ) reduce to the Gegenbauer polynomials. Similar properties can be de-
duced with regard to eigenstates of the sine operator.

In section 3 we give a formal extension of Jackiw’s result regarding the


number-phase minimum uncertainty states to the case of the generalized phase
operators. The next section is devoted to the interesting problem of states which
minimize the uncertainty product (4C)2 (4S)2 /hAi2 with iA = [C, S]. Jackiw
[17] has attempted the minimization of (4C)2 (4S)2 and concluded that the
extrememum condition cannot be satisfied by any normalizable state 6 . He has
explained this in terms of the fact that (4C)2 (4S)2 can in fact approach zero,
as in the case of coherent states with average occupation numbers tending to
infinity [12],[41]. When the quantity (4C)2 (4S)2 /hAi2 is considered, however
the minimization procedure leads to a well defined class of solutions. In fact the
states for which this quantity is a minumum appear as solutions of an eigen-
value problem which includes as a special case the eigenvalue problem for U .
The condition of normalizability of these states constrains the eigenvalue pa-
rameter λ to lie within an ellipse in the complex λ-plane. The set of states form
an overcomplete set.

In section 5 we present the coherent state representation of various quantities


of physical interest constructed out of the phase (C and S ) operators, We then
show that it is possible to construct phase operators such that hα|[C, S]|αi = 0
within a class of coherent states.

6 Actually there is one normalizable state satisfying his extrememum condition, namely | 0i,

the ground state . However this is a state in which (4C)2 (4S)2 has its maximum value.

10
2. The Eigenstates of the Generalized Cosine and Sine Operators

Consider now the eigenvalue problem for C, eq. (2.8), with C given in terms
of U defined by (2.3), as
1
C = ( U + U† ) (2.10)
2
On sustituting (2.9) in (2.8) and using (2.10) and (2.3) we obtain the following
recursion relations for the coefficients Cn (λ) :
2 λ C0 (λ) = C1 (λ) f (1),
2 λ C1 (λ) = C2 (λ) f (2) + C0 (λ) f (1),
− − − − − − − − − − − − − − −,
2 λ Cn (λ) = Cn+1 (λ) f (n + 1) + Cn−1 (λ) f (n), (2.11)
To solve eqs. (2.11) it is necessary to rewrite them in the form:
f 2 (n)
Fn+1 (λ) = 2 λ − , (2.12)
Fn (λ)
where
Cn (λ)
Fn (λ) = f (n), (2.13)
Cn−1 (λ)
By virtue of ( 2.13) we can write
Qn
Fr (λ)
Cn (λ) = Qr=1 n C0 (λ), (2.14)
r=1 f (r)
Qn
Further it is a consequence of (2.12) thet the product r=1 Fr (λ) is a polynomial
of degree n in λ. It follows from (2.14) that Ĉn (λ) ≡ Cn (λ)/C0 (λ) is also such
a polynomial. It may be verified that its explicit form is
Ĉn (λ) ≡ Cn (λ)/C0 (λ)
" n #−1 { n }
Y X2

Ĉn (λ) = f (r) (2λ)n−2r (−1)r Dr(n−1) (f 2 ), (2.15)


r=1 r=0
where { n2 } stands for (n − 1)/2 or n/2, according as n is odd or even, and
(n)
Dr (f 2 ) is defined by
(n)
D0 (f 2 ) = 1 (f or all n)
X
Dr(n) (f 2 ) = f 2 (a1 )f 2 (a2 )....f 2 (ar ) (2.16)
Here the arguments a1 , a2 , ...., ar are integers from the set 1, 2, 3, ....n taken in
such away that no two of them differ by less than two and the summation in
(2.16) is over all possible distinct sets of r integers which can be chosen in this
manner7 . It may be noted that the definition (2.16) is only possible for those
7 For example we write:
(5)
D3 (f 2 ) = f 2 (1)f 2 (3)f 2 (5)
(6)
D3 (f 2 ) = f 2 (1)f 2 (3)f 2 (5) + f 2 (1)f 2 (3)f 2 (6) + f 2 (1)f 2 (4)f 2 (6) + f 2 (2)f 2 (4)f 2 (6)
and
(3)
D1 (f 2 ) = f 2 (1) + f 2 (2) + f 2 (3)

11
(n)
values of r such that r ≤ { n+1 2
2 }. For other values of r , Dr (f ) could be
defined to be zero; in any case they do not occur within the range of summation
in the expansion (2.15). We observe further that the definitions (2.16) lead to
the identity
(n−2)
Dr(n−1) (f 2 ) + Dr−1 (f 2 ) f 2 (n) = Dr(n) (f 2 ) , (2.17)

which is very useful in verifying explicitly that (2.15) for Cn (λ) is indeed the
solution of the recursion relations (2.11).
We will now show that the polynomials Ĉn (λ) form an orthogonal set, that
is, a weight function ρ(λ) can be found such that
Z +1
ρ(λ) Ĉm (λ) Ĉn (λ) dλ = δmn (2.18)
−1

To prove this we need the following theorem [40] :


If any set of polyomials Pn (x) satisfies a recursion relation of the type

Pn (x) = (An x + Bn )Pn−1 (x) − Ln Pn−2 (x), (2.19)

(where An , Bn and Ln are constants, with An > 0 and Ln > 0 ) and if the
coefficient of xn in Pn (x) is kn then the necessary and sufficient condition for
Pn (x) to be orthonormal is that

kn An kn kn−2
An = ; Ln = = 2 (2.20)
kn−1 An−1 kn−1

We can write (2.11), using (2.15), in the form (2.19) with

2 f (n − 1) 2n
An = ; Bn = 0 ; Ln = and kn = Qn (2.21)
f (n) f (n) r=1 f (r)

It is evident that they satisfy (2.20). Hence the polynomials Ĉn (λ) form an
orthonormal set, as asserted earlier. Since the spectrum of C is known [13],[14]
to extend from -1 to +1, this is the domain over which the orthogonality relation
holds, and this fact has been used in writing the limits of integration in (2.18).
We have not been able to determine the weight function ρ(λ), however. If
we choose thep coefficient C0 (λ) (which is not determined by the recurrence
relations) as ρ(λ), we can write (2.18) in view of (2.15) as
Z +1
Cm (λ) Cn (λ) dλ = δmn (2.22)
−1

It is expected that the infinite set of orthonormal functions Cm (λ) form a com-
plete set for normalizable functions of λ over the interval (−1, 1). In fact this can
be inferred from the orthonormality of the states |λi which are the eigenstates
of the hermitian operator C :
X
Cm (λ) Cm (λ0 ) = hλ|λ0 i = δ(λ − λ0 ) (2.23)
m

12
whence for arbitrary f (λ),
Z
f (λ0 ) δ(λ − λ0 ) dλ0
f (λ) =
X Z 
= Cm (λ) f (λ0 )Cm (λ0 ) dλ0 (2.24)
m

Thus any f (λ) can be expanded in terms of the set Cm (λ) which is therefore
complete.
While the polynomials Ĉn (λ) cannot be identified with any known set when
the f (n) are arbitrary, in the special case when f (n) = (1 − δn0 ) they reduce to
the Gegenbaur polynomials. This can be verifiedp by inspection of (2.15). The
weight function ρ(λ) in this case is ρ(λ) = (1 − λ2 ). Similar properties can
be proved for the coefficients of expansion (in terms of number eigenstates) of
the eigenstates of the S-operator. In fact it turns out that the coefficients are
just in Cn (λ).

3. The Number Phase Minimum Uncertainty States

Since the commutator of the number operator N and the cosine operator
C is the operator i S (and not a c-number) the equation which determines the
state for which (∆N )(∆C) is a minimum is quite complicated as pointed out
by Jackiw. But if the uncetainty product is defined to be
(∆N )2 (∆C)2
,
hSi2
the condition for this to be a minimum is relatively simple,8 and following
Jackiw, we will adopt this definition for our purposes. The relevant equation is
then
(N + i γ C) |ψi = λ |ψi, (2.25)
where
1
λ = hN i + i γ hCi and γ = [(∆N )2 /(∆C)2 ] 2 , (2.250 )

Jackiw has obtained the solution of this equation for the special case f (n) =
(1 − δn0 ), and we will generalize it here for arbitrary f (n). For a state |ψi
n=∞
X
|ψi = an (λ) |ni , (2.26)
n=0

which satisfies the minimum condition (2.25), we obtain the following recurrence
relations with the aid of equations (2.2) and (2.3):

a1 (λ) f (1) = λ a0 (λ) (2.27)
2
− − − − − − − − − − − − − − −,
2(λ − n)
an (λ) = an+1 (λ) f (n + 1) + an−1 (λ) f (n), (2.28)

8 See Jackiw, Ref [17 ]; and Carruthers and Nieto, Ref.[12], for a further discussion of this

question.

13
Observing that φn (λ) , defined by

an (λ)
φn (λ) ≡ f (n),
an−1 (λ)
satisfies
2 (λ − n) f 2 (n)
φn+1 (λ) = − , (2.29)
iγ φn (λ)
one finds, using the same procedure as in the solution of (2.11), that the ratio
Qn
an (λ) φr (λ)
= Qr=1n , (2.30)
a0 (λ) r=1 f (r)

th
is a poly nomial of
 the 2n degree
 in λ. Eq. (2.26) with (2.30) gives the state
(∆N ) (∆C)2
which minimises hSi2 .

4. Simultaneous Uncertainties in C and S

We now turn to the determination of the states which minimise :


(∆C)2 (∆S)2
 
Q ≡ .
hi[C, S]i2

Jackiw has pointed out the relevance of this kind of uncertainty product when
dealing with a pair of operators whose commutator is not a c-number. From his
treatment it follows that the states which minimise Q must satisfy the equation:
! !
Ĉ Ŝ Ĉ Ŝ
−i +i |ψi = 0 ; (2.31)
∆C ∆S ∆C ∆S

where
Ĉ = C − hCi and Ŝ = S − hSi. (2.32)
This would be ensured of course if
!
Ĉ Ŝ
+i |ψi = 0 , (2.33)
∆C ∆S

and we now proceed to solve this more restrictive equation. We first rewrite it
as
(1 + γ)U + (1 − γ)U † |ψi = 2 λ |ψi ,
 
(2.34)
where
 12
(∆C)2

γ = , (2.35)
(∆S)2
and
λ = hCi + iγ hSi. (2.36)
If we now assume an expansion
n=∞
X
|ψi = an |ni , (2.37)
n=0

14
for |ψi in terms of the number eigenstates, we obtain from (2.34) the following
recurrence relations for the expansion coefficients:

(1 + γ)f (n + 1) an+1 + (1 − γ)f (n) an−1 = 2 λ an . (2.38)

With the substitution


1−γ
an = q n/2 bn ; q= , (2.39)
1+γ

eq.(2.38) simplifies to

f (n + 1) bn+1 + f (n) bn−1 = 2 x bn . (2.40)

where
λ
x= p . (2.41)
1 − γ2
Eq. (2.40) is identical to (2.11), so that we can write down the solution imme-
diately:
#−1 {n
2}
" n
Y X
bn (x) = b0 (x) f (r) (2x)n−2r (−1)r Dr(n−1) (f 2 ), (2.42)
r=1 r=0

Hence
#−1 {n
2}
" n
Y X
n/2
an (x) = a0 (x) q f (r) (2x)n−2r (−1)r Dr(n−1) (f 2 ), (2.43)
r=1 r=0

The condition for orthonormality of |ψi now becomes


1
|a0 (x)|2 = P∞ n 2
. (2.44)
1+ n=1 q |bn (x)|

P∞ , a0 (x) 6= 0 (otherwise an (x) = 0 for all n,


For a non-trivial solution of eq.(2.38)
see eq.(2.43)). Thus the series n=1 q n |bn (x)|2 should converge. This requires
in turn (by the ratio test) that
2
q n+1  b
n+1 (x)
lim . <1 (2.45)

n→∞ q n bn (x)

or
2
1
|α| < (2.46)
q
where
bn+1 (x)
α = lim (2.47)
n→∞ bn (x)
The value of α may be found from eq.(2.40) reexpressed as

bn+1 bn
f (n + 1) + f (n) = 2 x . (2.48)
bn−1 bn−1

15
On letting n → ∞, remembering that f (n) → 1, we get

α2 − 2 x α + 1 = 0. (2.49)

The relevant9 root is p


α=x+ x2 − 1 (2.50)
Equations (2.50) and (2.46) lead to the condition

p
2
1
2
|x + x − 1 | < (2.51)
q

for normalizability. On using definitions (2.41) and (2.39) of x and q, this


becomes p
|λ + λ2 + γ 2 − 1 |2 < (1 + γ)2 (2.52)
Note that γ is, by definition, positive. The boundary of the allowed domain of
λ is thus characterized by the equality
p
|λ + λ2 + γ 2 − 1 |2 = (1 + γ)2 . (2.53)

This equation describes an ellipse in the λ-plane. we now proceed to prove this
assertion. There are two cases to be considered : γ 2 > 1 and γ 2 < 1. If γ 2 > 1,
we write p
λ = γ 2 − 1 sinh θ, (2.54)
where θ is in general complex,

θ = u + iv.

Then eq. (2.53) reduces to


1+γ
eu = p . (2.55)
γ2 − 1
Feeding this back into eq. (2.54) we have

1 + γ iv γ 2 − 1 −iv
λ= e − e ≡ µ + iν (say) (2.56)
2 2(1 + γ)

It is easily verified now that


ν2
µ2 + = 1. (2.57)
γ2
p
When γ 2 < 1 , a similar calculation with λ written as (1 − γ 2 ) sin θ enables
us to show once again that the allowed region of values of λ is once again the
region of an ellipse (2.57). It is proved therefore that the domain of convergence
of the series in the denominator of (2.44) is an ellipse and hence the states |ψi
defined by (2.33) are normalizable for all λ within this ellipse. It may be noted
that for the special case when (∆C)2 = (∆S)2 , i.e. γ = 1, the states |ψi are
eigenstates of the operator U , see (2.34). In this case the ellipse (2.57) becomes
a circle and the minimum uncertainty states with ∆C = ∆S coincide with the
9 Sincebn+1 (x)/bn (x) is a ratio of polynomials of degree n + 1 and n respectively, it
√should
behave like x as x → ∞. Equation (2.50) is consistent with this; the choice α = x − x2 − 1
would violate this requirement.

16
eigenstates of U found by Lerner at. al. [14]. It has also been shown by Lerner
et. al. that the eigenstates of U form a complete (or rather an over- complete)
set. By using a similar argument as that used by Cahill [42], we can assert that
for general values of γ , the set of all states |ψi characterized by λ lying within
the ellipse of convergence form an over complete set.

5. The C and S operators and the Coherent state


It is well known that if any oscillator is at some time a coherent state |αi then
the subsequent time variation of the state is given completely by a change in
argument in arg α : α = ρeiφ → α(t) = ρei(φ−ωt) . Further the expectation
value of position varies simple harmonically with phase (φ − ωt). Therefore it
is natural that one should think of the coherent states in connection with the
phase operators. Though the coherent states are eigenstates of the annihilation
operator a rather than of C and S , it may nevertheless be expected that the
expection values of of C and S in coherent states should show an intimate
relation to cos φ and sin φ ( φ = arg α). Actually Carruthers and Nieto
[12],[41] have considered these and various related quantities in the special case
f (n) = (1 − δn0 ), and found that apart from factors depending on ρ = |α|, the
expected correspondence does exist. Here we cary out the simple extension to
the case of general f (n), and verify that the same type of relations continue to
hold. Following this we raise the interesting question whether it is possible to
choose f (n) in such a way that the commutator [C, S] has vanishing expectation
values within the coherent states. We show by an example that this is indeed
possible for all the coherent states with a fixed value of |α|, though the example
itself is rather artificial.
Let us define the following relations10

X Nn
ψ1 (N ) = √ f (n + 1) ,
n=0
n! n + 1

X N n f (n + 1)f (n + 2)
ψ2 (N ) = ,
n=0
n! {(n + 1)(n + 2)}1/2

X Nn 2
ξ1 (N ) = {f (n + 1) + f 2 (n)} , (2.58)
n=1
n!

X Nn 2
ξ2 (N ) = {f (n + 1) − f 2 (n)} .
n=1
n!

By using the representation [37]


 ∞
1 2 X αn

|αi = exp − |α| |ni (2.59)
2 n=0
{n!}1/2

we define
1
α = N 2 eiφ
10 We have defined these functions using the same notation as in Ref. [12], except that here

f (n) is quite general.

17
for coherent states the following expressions can be easily obtained :
1
hα|C|αi = N 2 exp(−N ) cos φ ψ1 (N ), (2.60a)
1
hα|S|αi = N exp(−N ) sin φ ψ1 (N ),
2 (2.60b)
1 1 N
hα|C 2 |αi = exp(−N ) [ f 2 (1) + ξ1 (N ) + cos 2φ ψ2 (N )], (2.60c)
4 4 2
1 1 N
hα|S 2 |αi = exp(−N ) [ f 2 (1) + ξ1 (N ) − cos 2φ ψ2 (N )], (2.60d)
4 4 2
1 1
hα|C 2 + S 2 |αi = exp(−N ) [ f 2 (1) + ξ1 (N )], (2.60e)
2 2
hα|C 2 − S 2 |αi = N exp(−N ) cos 2φ ψ2 (N ), (2.60f)
hα|CS + SC|αi = N exp(−N ) sin 2φ ψ2 (N ), (2.60g)
i
hα|CS − SC|αi = exp(−N ) [ξ2 (N ) + f 2 (1)]. (2.60h)
2
The quantities:

P (N, φ) = (∆N )2 (∆C)2 /hSi2 , (2.61a)


R(N, φ) = (∆N )2 (∆S)2 /hCi2 , (2.61b)

which measure the degree of incompatibility of the number phase operators can
be evaluated with the aid of eqs (2.60). Note that for the coherent states

(∆N )2 = hN 2 i − hN i2 = N

We will now consider the following question:


It is well known that the cosine and sine operators cannot commute, this
being a reflection of the impossibility of defining an unique phase operator. But
in view of the fact that a phase φ can be associated with a coherent state in
the sense indicated above, one may ask whether the commutator of C and S
cannot be made to vanish atleast weakly, in the sense of the vanishing of its
expectation value with respect to the coherent states. A quick check with the
aid of eq (2.60h) shows that the familiar C and S operators (ie. with f (n) = 1
for all n ≥ 1) as well as others considered explicitly in Ref. [13] do not have
this property. However with the freedom of choice of the f (n), available to us
in turns out that we can define C and S in such away that within a class of
coherent states hα|[C, S]|αi = 0. In fact if we choose

N
f 2 (0) = 0, f 2 (1) =
N −1
f 2 (r) = 1 (f or r > 1), (2.62)

then we see that equation (2.60h) is satisfied because

N [f 2 (1) − f 2 (2)] = f 2 (1) (2.63)

We must verify that f (n) defined in (2.62) do follow the chain sequence condi-

18
11
tion:
1 2
f (n) = gn (1 − gn−1 ), (2.64a)
4
with g0 = 0 and 0 ≤ gn ≤ 1. (2.64b)

Now since
1 2
f (1) = g1 (1 − g0 ), (2.65)
4
from eq.(2.62) we have  
1 N
g1 = (2.66)
4 N −1
Since
1 2
g2 (1 − g1 ) = f (2)
4
1
=
4
we have
1
g2 = /(1 − g1 ).
4
1/4
gn = .
1 − gn−1
Hence,

1/4
gn = (2.67)
1/4
1−
1/4
..... −
1 − g1

Now using the method of continuous fractions [44] and eq(2.66) it turns out
that one can sum up eq. (2.67) , to obtain

1 nN − 2(n − 1)
gn = (2.68)
2 (n + 1)N − 2n

We see that for all n > 1,


1
gn (1 − gn−1 ) =
. (2.69)
4
Hence the f (n) given in (2.62) obey eq (2.64a). Now all the gn ’s must obey the
inequality (2.64b) i.e.
0 ≤ gn ≤ 1 . (2.70)
Now if
gn+1 ≤ 1 . (2.71)
11 The choice of f (n) given in our paper [43] , is incorrect for, though they satisfy the chain

sequence conditions (2.64) they do not have the property Limit: f (n) = 1 as n → ∞. It
is for this reason, the spectra of C and S will not span the entire region (-1,1). We are
grateful to Professor Lerner for pointing this out (Private Communication). The choice of
(2.62) overcomes this difficulty.

19
i.e.
1 N (n + 1) − 2n
≤1 .
2 N (n + 2) − 2(n + 1)
i.e.
N n − N − 2n ≤ 2nN + 4N − 4n − 4,
whence
2(n + 2)
≤ N. (2.72)
n+3
The maximum value of the l.h.s. of (2.72) is attained when n → ∞, which is 2.
Therefore if
N ≥ 2, (2.73)
equation (2.71) is satisfied. Since we must also have

gn+1 ≥ 0, (2.74)

i.e.,  
n
N ≥2 (2.75)
n+1
The maximum value of r.h.s of (2.75) is 2. Therefore we see that (2.73) ensures
that (2.70) holds. We have shown that there exist phase operators such that
hα|[C, S]|αi = 0 for all coherent states having |α| = N with N ≥ 2. Since
the chain sequence conditions with the requirement limn→∞ f (n) = 1 are sat-
isfied the spectra of these operators will span the entire region (-1,1). It is
apparant from eq. (2.60h) that there could be other phase operators which are
more complicated than the simple choice given in (2.62) which will satisfy these
conditions.

20
Chapter 3

The Action and Phase


(Angle) Variables and the
Energy Levels of the
Anharmonic Oscillator

1. Introduction

In this chapter we treat the problem of the anharmonic oscillator in the


action and phase variables and show how an approximate formula for the energy
levels can be obtained. We demonstrate that this formula satisfies a general
scaling law which must be obeyed by the exact eigenvalues.
In Sec. 2 of this Chapter we derive this general scaling law. In Sec 3 we
introduce the action and angle variables through the use of Jacobi’s elliptic
functions. The Hamiltonian is dependent only on the action variable J, and is
independent of the phase (angle) variable. By a semiclassical treatment based
on the Bohr-Sommerfeld quantum rules as modified in the W.K.B. approach,
we obtain in Sec 4, an expression for the energy levles En of the anharmonic
oscillator. This formula exhibits the correct limits for small anharmonicity and
gives the correct behaviour (corresponding to that of a pure quartic potential)
for large n or alternatively for large anharmonicities.

2. A general Scaling Law for the Anharmonic Oscillator

The hamiltonian for the anharmonic oscillator with a quartic term is

p2
H= + αx2 + βx4 (3.1)
2m
If the nth energy level of (3.1) is denoted by En (m, α, β) , then a general scaling
law can easily be established by writing m = µm0 and x = ξ y (ξ = constant)
in the eigenvalue equation

~2 ∂ 2 ψ
− + (αx2 + βx4 ) ψ = En (m, α, β) ψ,
2m ∂x2

21
to obtain
~2 ∂ 2 ψ
− + (α µ ξ 4 y 2 + β µ ξ 6 y 4 ) ψ = ξ 2 µ En (m, α, β) ψ .
2m0 ∂y 2
Comparing these two equations, we discover that

En (m, α, β) = f En (m0 , α0 , β0 )
(3.2a)
where f is given by
 1/2  1/3
a b a2
f= = = ; (3.2b)
µ µ2 b
in terms of the ratios
m α β
µ= , a= and b = ; (3.2c)
m0 α0 β0
which have to be so chosen that

µ a3 = b2 . (3.2d)

Our scaling law is a generalization of that found in Symanzik [20]. The latter
corresponds to setting µ = 1 in our equation (3.2).
A familiar special case, ( Simon Ref. [20]) is obtained when µ = 1,
b = β and α = 1,
2
En (m0 , 1, β) = β 1/3 En (m0 , β − 3 , 1) (3.3)

which exhibits the β 1/3 type of behaviour of the energy levels and the cube root
type of singularity at β = 0.

3. Action and Phase (Angle) Variables

The identification of the action angle variables is facilated by the knowledge


that in the solution of the classical anharmonic oscillator problem one encounters
elliptic functions in place of the trignometric functions of the harmonic oscillator
case.12 Thus we write

x = A cn(φ, k) , (3.4a)
1
2βA2 2

1
p = mẋ = − (2mα) A 1 +
2
sn(φ, k) dn(φ, k) , (3.4b)
α
d
where the fact that du cn(φ, k) = −sn(u, k) dn(u, k) and φ̇ has been written as a
constant since φ is an angle variable. The value of the constant and the relation
between A and the modulus k of the elliptic functions are so chosen that H is
free of φ. We have
βA2 αk 2
k2 = or A2 = (3.5)
α + 2βA2 β(1 − 2k 2 )
12 For definitions of the elliptic functions used inthis Chapter and some of their elementary

properties, see Ref. [45].

22
On using (3.4) and (3.5) H reduces to

H = A2 (α + βA2 ), (3.6)

where the standard properties of the elliptic functions such as

cn2 (φ, k) + sn2 (φ, k) = 1

dn2 (φ, k) + k 2 sn2 (φ, k) = 1, (3.7)


are taken into account,
The action variable J conjugate to φ may now be determined as a function
of A by requiring that
{x, p}J,φ = 1. (3.8)
We have
∂A
{x, p}J,φ = {x, p}A,φ . (3.9)
∂J
Because of the dependence of the modulus k on A, {x, p}A,φ cannot be evaluated
through a straight-forward use of eqs. (3.4). However by differentiating the
identity
p2
+ αx2 + βx4 = A2 α + βA2

(3.10)
2m
with respect to A and φ respectively:

2βA2
 
p ∂p 3 ∂x

+ 2αx + 4βx = 2αA 1 + , (3.11a)
m ∂A ∂A α

p ∂p  ∂x
+ 2αx + 4βx3 = 0, (3.11b)
m ∂φ ∂φ
∂x ∂x
and multiplying (3.11a) by ∂φ and (3.11b) by ∂A , subtracting and using eqs.
(3.4) we have
 12
2βA2

1
{x, p}A,φ = (2mα) A 1 +
2
. (3.12)
α
Hence from (3.8) and (3.9) we obtain

∂J 1 1
= (2m) 2 A α + 2βA2 2
∂A
Z A
1 1
J = (2m) 2 A α + 2βA2 2 dA
0
 12 "  32 #
2mα3 2βA2
= 1+ −1 . (3.13)
6β α
Thus: (
 12 ) 43 
2

α  2
H = 1+ 3βJ − 1 . (3.14)
4β mα3

23
4. The Energy Levels

The quantized energy levels will now be obtained by requiring that


I
1
J dφ = (n + ) h, n = 0, 1, 2, ... (3.15)
2
This is of course a heuristic procedure which is based on the Bohr-Sommerfeld
approach of the old quantum theory, but with the important modification (re-
placement of n by n + 12 on the right hand side) suggested by the W.K.B.
treatment. Since the elliptic functions introduced in eqs. (3.4) are periodic in
φ with period 4K(k) , where K(k) is the complete elliptic function of the first
kind Z 1
dt
K(k) = p (3.16)
0 (1 − t ) (1 − k 2 t2 )
2

eq. (3.15) requires that


1
4 K(k) J = (n + ) h, n = 0, 1, 2, ... (3.17)
2
On substituting the solutions Jn of this equation for J in (3.14) we get the
energy levels of the quantum anharmonic oscillator.
Eq(3.17) is a transcendental equation , since the modulus of the elliptic func-
tion, is a function of J, c.f. (3.13) and (3.5). However K(k) is quite insensitive
to the parameters involved. The extreme limits are:
 2
βA π
→ 0, k → 0, K(0) = (3.18a)
α 2
  2
βA2
   
1 1 1 1
→ ∞, k→√ , K √
= √ Γ ≈ 1.18 K(0)
α 2 4 π
2 4
(3.18b)
Thus as the anharmonicity increases from a vanishing value to infinity, K(k)
changes no more than 18 % . In the limit (3.18a) the energy levels are expected
to tend to those of the harmonic oscillator and indeed they do. For (3.18a) and
(3.17) give Jn = (n+ 12 )~, and when this is introduced in (3.14) the leading term
 12
in the binomial expansion is taken, we get En = (n + 12 )~ ω0 with ω0 = 2α m .
At the opposite extreme (large anharmonicity or high energy) (3.18b) applies
and then one gets from (3.17)
3
2π 2 1
Jn =  2 (n + )~. (3.19)
Γ 14 2


On substituting this in eq. (3.14) and noting that the second term in curly
brackets now dominates we get:
 4
 18π 3  12 n + 1   3  β~4  13
2
En ≈  1 2 (3.20)
 Γ  4m2
4

24
This expression coincides 13 with the asymptotic formula given in Titch-
marsh [46] for the eigenvalues in the case of the pure x4 potential. It may be
noted that the energy increases as n4/3 β 1/3 for large n. Finally it is a trivial
matter to verify that our approximate formula obeys the scaling law eqs. (3.2)
satisfied by the exact energy levels. Actual numerical computation shows that
even for low lying energy levels the formula reproduces the energy values within
a few percent, the error being of the same order as the W.K.B. evaluation [47]
which is more complicated.

13 The apparent missing of a factor of 2 within the curly brackets is because the range of x

in our case is −∞ to +∞ and not from 0 to ∞ as is the case in Ref. [46].

25
Chapter 4

Coherence in Quantum
Mechanical Systems

1. Introduction

In this Chapter we study the behaviour of small wavepackets in spherically


symmetric potentials in three dimensions. In particular we investigate whether
any coherent states exist in such potentials, i.e. whether it is possible to have
wavepackets which move along classical orbits without spreading. Before pro-
ceeding to this question we consider, in Sec. 2, the familiar problem of the
harmonic oscillator in one dimension, and obtain a simple criterion for the con-
stancy for the position and momentum uncertianties. Among the variety of
states which satisfy this criterion (besides the coherent states) we identify an
interesting class of semicoherent states. It is conceivable that such states might
exist in more general cases and might show up if one demands coherence only in
the loose sense of constancy in position and momentum uncertainties. Actually
our investigations in Secs. 3-5, show that no small wavepackets which are co-
herent even in this loose sense can exist14 in any central poential other than the
three dimensional harmonic oscillator potential. We define a set of quantities
which measure the uncertainties in position and momentum in various direc-
tions, and solve the set of coupled equations governing their time variation. We
show that the initial conditions required to ensure the constancy of the var-
ious uncertainties are unphysical, so that the wavepackets necessarily spread
indefinitely. Particulars of derivation of the equations employed in Sec. 3, are
indicated in the Appendix.

2. The Wavepacket in the Harmonic Oscillator Portential


It is a simple matter to obtain the class of harmonic oscillator states char-
acterised by constant values of ∆x and ∆p. Defining:

χ = (∆x)2 = hx2 i − hxi2 ,

π = (∆p)2 = hp2 i − hpi2 , (4.1)


14 Strictly speaking our proof is only for wavepackets moving in circular orbits, but it is

most unlikely that the case of other types of orbits would be very different.

26
and calculating the time derivatives using the quantum equation of motion:
dA 1 ∂A
= [A, H] + , (4.2)
dt i~ ∂t
with
p2 1
H = + mω 2 x2 , (4.3)
2m 2
one finds from the expressions χ̇ and π̇ involve also hxp + p xi, but

m2 ω 2 χ̇(t) + π̇(t) = 0 , (4.4)

as can be seen from an application of (4.2). A dot denotes differentiation with


respect to time. The second time derivatives yield the equations15
2
χ̈(t) = −2ω 2 χ(t) + π(t) , (4.5a)
m2
π̈(t) = 2m2 ω 4 χ(t) − 2ω 2 π(t) , (4.5b)
which may be solved immediately by rewriting them as

π̈(t) + m2 ω 2 χ̈(t) = 0 , (4.6a)

π̈(t) − m2 ω 2 χ̈(t) + 4ω 2 π(t) − m2 ω 2 χ(t) = 0 ,


 
(4.6b)
We obtain
−1  2 2
χ(t) = 2m2 ω 2 m ω χ(0) + π(0) + m2 ω 2 χ(0) − π(0) cos2ωt
 
 
−1 1  2 2
+ 2m2 ω 2

m ω χ̇(0) − π̇(0) sin2ωt (4.7a)

π(t) = π(0) + m2 ω 2 χ(0) − m2 ω 2 χ(t) .



(4.7b)
 2 2
Actually the general solution contains another term m ω χ̇(0) + π̇(0) t,which
however vanishes identically for any state according to eq.(4.4). Thus both χ
and π are harmonically varying in general with an angular frequency 2ω. If they
are time independent we must have

m2 ω 2 χ(0) − π(0) = 0 , (4.8a)

m2 ω 2 χ̇(0) − π̇(0) = 0 , (4.8b)


These can be expressed more simply in terms of the creation and annihilation
operators. On setting
  12   12
~ † m~ ω
a − a†
 
x= a+a p= (4.9)
2mω 2

and simplifying, the conditions (4.8) become

ha2 (0)i = ha(0)i2 , (4.10a)


15 For slightly different forms of equations (4.5) see Yu E. Murakhver, Ref. [48] and A.

Messaih Ref. [49].

27
2
h a† (0)i = ha† (0)i2 , (4.10b)
These are necessary and sufficient conditions for a harmonic oscillator wavepacket
to maintain a constant size or extension. They are of course trivially satisfied
by the coherent states |αi which are eigenstates of a. At the other extreme we
have the energy eigenstates which being stationary states have constant values
of χ and π. In between there is a whole class of semicoherent states for which
eqs (4.10) are valid. An interesting example is provided by states of the type

|ψi = N [ |αi − |βi hβ|αi ] , (4.11)

where |αi and |βi are coherent states and N is a normalization factor, N =
 − 1
1 − |hβ|αi|2 2 . It may be easily verified that the equations (4.10) do hold good
in these states. Note that (4.11) defines |ψi as the projection of a coherent state
|αi orthogonal to another coherent state |βi. Therefore the ordinary coherent
states |αi are not obtainable as the limit of |ψi when the two superposed states
|αi and |βi are made identical : in the limit α → β, |ψi does not exist. We will
refer to the states (4.11) as semicoherent states. it is a curious fact that in the
semicoherent state (4.11), the mean position or (momentum) of the oscillator
is exactly the coherent state |αi i.e. hψ|x(t)|ψi = x̄α (t) ≡ hα|x(t)|αi. The
position probability density |ψ(x, t)|2 in the state (4.11) is describable in terms
of two gaussian packets of constant amplitude and width (oscillating simple
harmonically with phases characterised by arg α and arg β ) together with an
interference term which ensures that χ and π remain time independent. Their
values are
2 |hα|βi|2 |α − β|2
  
~
χ = 1+ , (4.12a)
2mω (1 − |hα|βi|2 )
2 |hα|βi|2 |α − β|2
  
mω~
π = 1+ . (4.12b)
2 (1 − |hα|βi|2 )
Clearly the product of these exceed the minimum value ~ 2 /4 unlike the case of
true coherent states.
This example illustrates the wide variety of states which can satisfy the con-
dition of constant values for the uncertainties ∆x and ∆p. Of these the true
coherent states have the property of being small wavepackets around the mean
position. In the search for coherent states in three dimensions, in the remaining
sections of this chapter, we use the constancy of uncertainties of various quanti-
ties as a criterion. But we also demand small size, so that semi-coherent states
(if at all they exist) are excluded from consideration.

3. Behaviour of a Three Dimensional Wavepacket in a


Spherically Symmetric Potential

We now consider a wavepacket representing a particle in a central potential


in three dimensons and determine how various parameters representing the ex-
tension of the wavepacket (in the configuration and momentum spaces) change
with time. To avoid nonessential complications we will assume that the mean
position moves in a circular orbit of radius R. Taking advantage of the geome-
try of the system, we consider the uncertainties in the components of position

28
and momentum in the radial and tangential directions as well as the directions
perpendicular to the plane of the orbit. We define the quantities
* 2 +
h~xi
χR = · (~x − h~xi) , (4.13a)
R
* 2 +
1 h~pi
χT = 2 2 · (~x − h~xi) , (4.13b)
m ω R
D E
2
χz = {z − hzi} , (4.13c)
* 2 +
1 h~xi
πR = 2 2 · (~p − h~pi) , (4.13d)
m ω R
* 2 +
1 h~
pi
πT = 4 4 · (~p − h~pi) , (4.13e)
m ω R
1 D 2
E
πz = 2 2 {pz − hpz i} , (4.13f)
m ω
The subscript R indicates the radial component, T the component tangential
to the orbit (and hence parallel to h~ pi ) and z the component normal to the
plane of the orbit which is taken to be the x − y plane. The factors of mω
are included to reduce the dimensions of all the quantities to the dimension of
(length2 ). Here m is the mass of the particle and ω the angular frequency of
its motion in the circular orbit. Of course by the radius of the orbit we mean
|h~xi|. It is evident that
 
dV
mω 2 R = ≡ V 0 (R) . (4.14)
dr r=R

On calculating the rates of change of these quantities as dictated by the Hamil-


tonian
p2
H = + V (r) , (4.15)
2m
with the assumption that the wavepacket is very small (χR , χT , χz << R2 )
one obtains a set of coupled differential equations which involve not only the
quantities defined above in (4.13), but also various cross correlations between
position and momentum components. However, the equations for χz and πz
decouple from those for uncertainties in the plane of the orbit and can be solved
easily. We will now obtain the relevant equations (relegating the proof to the
Appendix) and solve them. In particular the question whether time independent
solutions (corresponding to the coherent wavepacket) exist will be considered.

3.1 Spread of a wavepacket normal to the plane of the orbit

The equations for χz and πz can be shown to be reducible to the following


closed system:
2
χ̈z = 2ω 2 πz − V33 χz , (4.16a)
m

29
1
π̇z = − V33 χ̇z , (4.16b)
m2 ω 2
where V33 is defined by
V0
 2     2

∂ V 2
00 z 0z
V33 = = V −V 3 +
∂z 2 0 r r r 0
V 0 (R)
= = m ω2 . (4.17)
R
The subscript 0 indicates that the quantity concerned is to be evaluated at the
position of the orbit; the fact that the orbit has been chosen to be in the x-y
plane (hzi = 0) has been used in the above reduction. Equations of exactly
the same form16 as (4.16), with V33 = m ω 2 , hold also for a one dimensional
oscillator.On solving these equations we obtain
1 1
χz (t) = (χz (0) + πz (0)) + (χz (0) − πz (0)) cos2ωt
2 2
1
+ χ̇z (0) sin2ωt (4.18a)

πz (t) = (χz (0) + πz (0)) − χz (t) . (4.18b)
Clearly the extension of the wavepacket perpendicular to the orbital plane re-
mains bounded; infact, initial conditions may be so chosen as to make χz and
πz time independent.

3.2 Spread of wavepacket in the orbital plane

Expressions for the first time derivative of χR , χT , πR and πT involve also


the following quantities:
 
1 1
χRT = {h~xi · (~x − h~xi)} {h~pi · (~x − h~xi)} , (4.19a)
mω R2
 
1 1
πRT = 3 3 {h~xi · (~
p − h~pi)} {h~pi · (~p − h~pi)} , (4.19b)
m ω R2
 
1 1
µRR = {h~xi · (~x − h~xi)} {h~xi · (~p − h~
pi)}
mω R2
 
1 1
+ {h~xi · (~p − h~
pi)} {h~xi · (~x − h~xi)} (4.19c)
mω R2
 
1 1
µRT = 2 2 {h~xi · (~x − h~xi)} {h~pi · (~p − h~pi)} , (4.19d)
m ω R2
 
1 1
µT R = 2 2 {h~pi · (~x − h~xi)} {h~xi · (~p − h~pi)} , (4.19e)
m ω R2
 
1 1
µT T = 3 3 {h~pi · (~x − h~xi)} {h~pi · (~p − h~
pi)}
m ω R2
 
1 1
+ 3 3 {h~pi · (~p − h~pi)} {h~
pi · (~x − h~xi)} (4.19f)
m ω R2
16 Infact Yu. E. Murakhver Ref.[48], has obtained the same set of equations for an arbitrary

potential in one dimension. While the equations are exact for the harmonic oscillator, the
derivation of these equations for other potentials involves the neglect of third and higher order
derivatives of V, in a Taylor’s expansion around the point h~ xi, the ‘centre’ of the wavepacket,
a procedure which is valid for ‘small’ wavepackets (i.e. χz << R2 ).

30
These are the correlations 17 along various compnents of (~x − h~xi) and
p − h~
(~ pi). The µ’s contain one component of (~x − h~xi) and one of (~p − h~pi) and
the convention regarding the subscripts in this case is that the first subscript
indicates which component of (~x − h~xi) is involved while the second subscript
p −h~
refers to the component of (~ pi). The equations for the ten coupled quantities
written in matrix form are as follows:

χ̇R 0 0 2 0 0 0 1 0 0 0 χR

χ̇T

0
0 −2 0 0 0 0 0 0 1 χT
χ̇RT −1 1 0 0 0 0 0 1 1 0 χRT

π̇R 0 0 0 0 0 2 η 0 0 0 πR

π̇T
= ω
0 0 0 0 0 −2 0 0 0 −1 πT

π̇RT

0
0 0 −1 1 0 0 η −1 0 πRT
µ̇RR 2η 0 0 2 0 0 0 2 2 0 µRR

µ̇RT

0
0 −1 0 0 1 −1/2 0 0 1/2 µRT

µ̇T R

0
0 η 0 0 1 −1/2 0 0 1/2 µT R

µ̇T T 0 −2 0 0 2 0 0 −2 −2 0 µT T
(4.20)
or
du
= ω M u, (4.21)
dt
Where M and u stand for the square matrix and the column vector which
appear in the r.h.s. of eq.(4.20). The parameter η is defined as

r ∂ω 2
 
η = − 1+ 2
ω ∂r r=R

which by (4.14) can also be written as


 00 
rV
η = − . (4.22)
V 0 r=R

One is familiar with solutions of the form


X
u= u(i) eλi ωt (4.23)
i

for equations of the above type, where λi and u(i) are respectively the eigenvalues
and column eigenvectors of M . In the present case however, the solutions are
not of this kind, because it so happens that M is not a diagonalizable matrix (it
does not have ten independent eigenvectors) except for the special case η = −1.
The characteristic equation for M can be verified to be

λ4 (λ2 + 3 − η)2 (λ2 + 12 − 4η) = 0 , (4.24)


1 1
so that the eigenvalues are 0 (4 times) , ±(η − 3) 2 (twice each ) and ± 2(η − 3) 2 .
If M were diagonalizable it would satisfy the equation
M (M 2 + 3 − η)(M 2 + 12 − 4η) = 0. Actual evaluation shows that
M (M 2 + 3 − η)(M 2 + 12 − 4η) becomes (η + 1) times a non- zero matrix, so
17 It may be mentioned that the quantities in eqs. (4.19) which do not appear to symmetrized

are symmetric for circular orbits (see Appendix).

31
that M is diagonalizable if η = −1 but not for any other value of η. However
one has for any η,

M 3 (M 2 + 3 − η)2 (M 2 + 12 − 4η) = 0 , (4.25)

which shows that in general M has only two (instead of four) independent eigen-
vectors belonging to the eigenvalue 0 and only one eigenvector each (instead of
1
two) corresponding to the eigenvalues ±(η − 3) 2 . This means that the Jordan
canonical form [50] of M is

0 1 . . . . . . . .

. 0 1 . . . . . . .

. . 0 . . . . . . .

. . . 0 . . . . . .

. . . . β 1 . . . .
Mc = SM S −1 =

(4.26)
. . . . . β . . . .
. . . . . . −β 1 . .

. . . . . .
. −β . .
. . . . . . . . 2β .

. . . . . . . . . −2β

Where we have defined, for convenience,


1
β ≡ (η − 3) 2

The consequences of the nondiagonalizability of M will be explored in the par-


ticular case of the coulomb potential in the next section. Here we remark that
the general solution of eq (4.20) has sinusoidal or exponential parts according as
η < 3 or η > 3 for the potential considered. But the basic question which we are
interested in is whether initial conditions can be so presented that the solutions
of the equations become time independent. We now proceed to analyse this
problem in the context of the coulomb potential, which is of course physically
the most important case.

4. Solution of Eq. (4.21) for the Coulomb Potential

When
α
V (r) = (4.27)
r
η = 2, (4.28)
the eigenvalues of M are
0, ±i , ±2i . (4.29)
Rewriting eq. (4.21) as

dv
= ω Mc v , v = Su ; (4.30)
dt
we can readily obtain the general solution for this set of coupled equations for

32
the components of

v = c1 + c2 ωt + c3 (ωt)2 v (1) + (c2 + 2 c3 (ωt)) v (2)


 

+ 2 c3 v (3) + c4 v (4) + (c5 + c6 (ωt)) eiωt v (5)


+ c6 eiωt v (6) + (c7 + c8 (ωt)) e−iωt v (7) + c8 e−iωt v (8)
+ c9 e2iωt v (9) + c10 e−2iωt v (10) (4.31)

where the v (i) are column vectors with the elements


(i)
vj = δij . (4.32)

With this definition, it is easy to see from the form of M , eq.(4.26), with η = 2,
that

Mc v (1) = 0 ; Mc v (2) = v (1) ; Mc v (3) = v (2) ; Mc v (4) = 0 ;


Mc v (5) = i v (5) ; Mc v (6) = i v (6) + v (5) ;
Mc v (7) = −i v (7) ; Mc v (8) = −i v (8) + v (7) ;
Mc v (9) = 2i v (9) ; Mc v (10) = −2i v (10) ; (4.33)

The general solution for u is now obtained from (4.31) as

u = S −1 v
= c1 + c2 ωt + c3 (ωt)2 u(1) + (c2 + 2 c3 (ωt)) u(2)
 

+ 2 c3 u(3) + c4 u(4) + (c5 + c6 (ωt)) eiωt u(5)


+ c6 eiωt u(6) + (c7 + c8 (ωt)) e−iωt u(7) + c8 e−iωt u(8)
+ c9 e2iωt u(9) + c10 e−2iωt u(10) (4.34)

where u(i) = S −1 v (i) . The u(i) evidently satisfy equations (4.33) with Mc
replaced by M . It can easily be verified that since M is a real matrix

u(5)∗ = u(7) ; u(6)∗ = u(8) ; u(9)∗ = u(10) ; (4.35)

and hence for the reality of the solution (which is required in view of the nature
of the components of u ) we must choose

c∗5 = c7 ; c∗6 = c8 ; c∗9 = c10 . (4.36)

The vectors u(i) can be explicitly shown to be given by:18

18 The u(i) in (4.37) are not normalized and u(1) , u(4) , u(5) , and u(9) are arbitrary up to

a factor. However such a factor can be absorbed into the integation constants c1 , c2 , .... etc.,
and one can easily check by differentiation that (4.34) with (4.37) is the most general solution
of (4.21), with η = 2.

33
u(1) u(2) u(3) u(4) u(5) u(6) u(9)
             
0 0 2/9 1 0 2i/3 1
1  0   0  3  2   2i/3   −4 
             
0 −1/3  0  0  −i/2  −1/2   2i 
             
1  0   0  0  1   i/3   −1 
             
0  0   1/18  1  0   i/3   1 
 
0

−1/6
 
 0 
  
0
 
 −i/2 

 0 
 
 i 
 (4.37)
             
0  2/3   0  0  i   1/3  −2i
             
0  0  −1/9  −1  0   −i/2   −1 
             
−1  0   0  −1 −3/2   −i/2   2 
0 1/3 0 0 i 1/3 −4i

4.1 Spreading of wavepacket in a Coulomb Potential


The question of constancy of the size of the wavepacket now resolves into
the following: Is it possible to choose initial conditions on the packet in such a
way that all coefficients ci except c1 and c4 vanish? Even if this is not possible
can we atleast arrange to make c2 = c3 = c6 = c8 = 0 so that the size of the
wavepacket has at worst an oscillating behaviour but no indefinite increase? The
answer turns out to be in the negative. To see this let us consider the constant
c3 which appears as the coefficient of t2 . It can be expressed in terms of the
initial values of u as
1
c3 = χR (0) + πT (0) + 2 µRT (0) (4.38)
9
The vanishing of this quantity is a necessary condition for the constancy of the
size of the wave packet. However this condition cannot be met as we will now
show.
We note first in view of the definitions (4.13) and (4.19) of χR , πT and µRT ,
the r.h.s. of the above quantity is equal to
h~xi 1 h~
pi 1
h{ · (~x − h~xi)}2 it=0 + 4 4 h{ · (~
p − h~pi)}2 it=0 +
R m ω R m2 ω 2
h~xi h~
pi h~
pi h~xi
h{ · (~x − h~xi)}{ · (~
p − h~
pi)} + { · (~
p − h~
pi)}{ · (~x − h~xi)}it=0
R* R R R
2 +
h~xi h~
pi
= · (~x − h~xi) + 2 2 · (~ p − h~pi) .
R m ω R
t=0

Since the quantity in curly brackets is evidently a Hermetian operator the above
expectation value can vanish only if
 
h~xi h~
pi
· (~x − h~xi) + 2 2 · (~p − h~pi) |ψi = 0 (4.39)
R m ω R t=0

Now let us suppose that the x-axis is chosen so as to pass through the mean
position at t = 0, so that hxi = R, hyi = hzi = 0 at t = 0 and hpx i = hpy i = 0,
and hpz i = mωR. Then eq.(4.39) reduces to
  
~ ∂
(x − R) + −R ψ(x, y, z) = 0 , (4.40)
imω ∂y

34
ie.
1 ∂ψ mω
= i (2R − x) . (4.41)
ψ ∂y ~
Hence  h mω i
ψ = exp i (2R − x) y f (x, z). (4.42)
~
where f (x, z) is any arbitrary function of x and z. It is evident that the wave-
function merely oscillates and does not fall off in the y direction. Thus there
does not exist any compact wavepacket which does not spread indefinitely.
It is to be noted however that χR (the uncertainty in the radial component of
position) and πT ( the uncertainty in the tangential component of momentum)
either remain constant or vary utmost periodically. This is because, as may be
seen from the explicit expressions (4.37) for the u(i) , the columns u(1) , u(2) and
u(5) do not contribute to χR and πT and these are the columns which appear
with factors of t or t2 in (4.34). On the other hand u(1) and u(5) contribute to
χT and this quantity increases indefinitely i.e. the wavepacket (even if it is well
defined initially) spreads gradually in the tangential direction.

5. Wavepacket in an Arbitrary Spherically Symmetric Potential

As we have already noted M is diagonalizable only for η = −1. The only


potential for which η ≡ − (rV 00 /V 0 ) is identically equal to −1 is the parabolic
potential (i.e. the harmonic oscillator in three dimensions) as is seen by solving
the equation rV 00 = V 0 . In other potentials there may be some particular radius
R at which η = −1, and if this happens, a wavepacket moving in a circular orbit
at that particular radius can retain its size indefinitely.19 In all other cases
the wavepacket must necessarily spread, and the spreading is in the direction
tangential to the orbit. This can be verified by following the procedure adopted
in the coulomb case, or more directly by observing that the postulate du dt = 0,
i.e., M u = 0, leads to the same constraint eq. (4.39) as in the coulomb case,
which cannot be satisfied by a well defined wavepacket.

19 This statement is of course for ‘small’ wavepackets, subject to the approximaton of ne-

glecting third and higher derivatives of V in a Taylor’s series expansion around hxi. In the case
of the harmonic oscillator of course all such derivatives vanish identically and the nonspreading
is a rigorous property.

35
Chapter 5

Appendix

We will now derive the equations of spread (4.20), for a wavepacket moving in
a circular orbit in an arbitrary central potential V (r).
The Hamiltonian for such a system is given by
3
X pi2
H = + V (r) (5.1)
i=1
2m

For our derivation we will have occasion to use the equation of motion (4.2) and
the commutation relations
∂f (x)
[pi , f (x)] = −i~ (5.2)
∂xi
together with the Taylor’s expansion of V (~x), about the ‘centre’ h~xi of the
wavepacket:
3  
X ∂V
V (~x) = V (h~xi) + . (xi − hxi i)
i=1
∂xi ~x=h~xi
3 3 
∂2V

1 XX
+ . (xi − hxi i) (xj − hxj i)
2! i=1 j=1 ∂xi ∂xj ~x=h~xi
+ ... (5.3)

where for a central potential


   
∂V 1 ∂V
Vi ≡ = hxi i (5.4a)
∂xi ~x=h~xi r ∂r r=R

∂2V
    
1 ∂V h xi i h xj i ∂ 1 ∂V
Vij ≡ = δij + (5.4b)
∂xi ∂xj ~ xi
x=h~ r ∂r r ∂r r ∂r r=R

Equations (4.20) have been derived for a bound wavepacket moving in a circular
orbit. For such an orbit we have
2
h p~ i = m2 ω 2 R2 (5.5a)

h p~ i . h ~x i = 0 , (5.5b)

36
and
1 ∂V
m ω2 = . (5.5c)
R ∂R
When these results are used, some of the expressions (4.13) and (4.19) simplify
somewhat. For example
1
χR = hxi i hxj i hxi xj i − R2 (5.6a)
R2 
1 1
χRT = hp i i hxj i hx x
i j i , (5.6b)
mω R2
 
1 1
πR = 2 2 hx i i hx j i hp p
i j i , (5.6c)
m ω R2
 
1 1
πRT = 3 3 hxi i hpj i hpi pj i , (5.6d)
m ω R2
 
1 1
µRR = hxi i hxj i hpi xj + xi pj i . (5.6e)
mω R2
Here and in the sequel, we use the summation convention with respect to re-
peated indices, which range from 1 to 3.
To derive the equation corresponding to the first row in the matrix equation
(4.20), we note from (5.6a) that

dχR 1 1 d
i~ = 2 hxi i hxj i h[xi xj , H]i + i~ 2 hxi xj i (hxi i hxj i) , (5.7)
dt R R dt
d
where we have used the quantum equation of motion to evaluate dt hxi xj i . On
d
evaluating the commutator and substituting dt hxi i = hp i i/m, we get

dχR hxi i hxj i hxi xj i


= hxi pj + pi xj i + {hxi i hpj i + hpi i hxj i} ,
dt mR2 mR2
in view of (5.6b) and (5.6e) this reduces to

1 dχR
= µRR + 2 χRT , (5.8)
ω dt
which corrsponds to the first row in the matrix equation (4.20). This equation
is exact.
We will now derive the equation for πR corresponding to the 4th row in
(4.20), this however involves an approximation depending on the small size of
the wave packet. From (5.6c) we have
 
dπR 1 hxi i hxj i d hpi pj i d
= 2 2 hpi pj i + {hxi i hxj i}
dt m ω R2 dt R2 dt
    
1 hxi i hxj i ∂V ∂V
= 2 2 − pj + p i
m ω R2 ∂xi ∂xj
 
1 hpi pj i
+ 2 2 {hxi i hpj i + hpi i hxj i} , (5.9)
m ω mR2

37
d
where the quantum equation of motion has been used to evaluate dt hpi pj i. To
proceed further we expand ∂V /∂xi in a Taylor’s series about the position h~xi :
 
∂V
pj = Vi hpj i + Vit h( xt − hxt i) pj i
∂xi
1 ∂V
= hxi i hpj i + {hxt pj i − hxt i hpj i} Vit . (5.10)
R ∂R
Vi and Vit are given in eqs. (5.4). We have dropped the third and higher order
derivatives in the Taylor’s expansion. This is a good approximation for most
potentials of interest provided the size of the wave packet is much less than the
radius of the orbit. From (5.10) we have
 
hxi i hxj i ∂V hxi i hxj i
− p j = − {hxt pj i − hxt i hpj i} Vit . (5.11)
R2 ∂xi R2
Note that the contribution to (5.11) from the first term in (5.10) is zero because
of (5.5b). Similarly the second term in (5.11) is zero, hence
 
hxi i hxj i ∂V hxi i hxj i
− pj = − hxt pj i Vit .
R2 ∂xi R2
  
hxi i hxj i 1 ∂V hxi i hxt i ∂ 1 ∂V
= − hxt pj i δit +
R2 R ∂R R ∂R R ∂R
  
hxt i hxj i 1 ∂V ∂ 1 ∂V
= − hxt pj i +R
R2 R ∂R ∂R R ∂R
R ∂ ω2
 
hxt i hxj i 2
= − hxt pj i mω 1 + 2
R2 ω ∂R
mω 2
= hxt i hxj i hxt pj i η (5.12)
R2
where we define
R ∂ ω2
 
η = − 1+ 2
ω ∂R
r V 00
 
= − (5.13)
V 0 r=R
Putting (5.12) into (5.9) and using (5.6d) and (5.6e) we have
1 dπR
= 2 πRT + η µRR , (5.14)
ω dt
which is the desired equation. Similarly all the other equations in (4.20) can be
derived.
Before we close the appendix we make a final remark, namely that all the
quantities which are defined in (4.19) and which do not appear to be properly
symmetized are symmetric for circular orbits. Consider for example µT R , eq.
(4.19c) :
 
1 1
µT R = 2 2 {h~p i · (~
x − h~
x i)} {h~
x i · (~
p − h~
p i)} ,
m ω R2
1 hpi i hxj i
= 2 2 {hxi pj i − hxi i hpj i} (5.15)
m ω R2

38
In the case of circular orbits this reduces to
1 hpi i hxj i
µT R = hxi pj i
m2 ω 2 R2
1 hpi i hxj i
= 2 2 hpj xi + i ~ δij i
m ω R2
1 hpi i hxj i
= 2 2 hpj xi i (5.16)
m ω R2
confirming that the order of factors in the product whose expectation value is
taken is irrelevant.

*********************************************

REFERENCES

1. W. Heisenberg : Z. Physik 33, 879 (1925)


2. M. Born and P. Jordon : Z. Phyzik 34, 858 (1925)
3. E. Schrodinger : Ann. Physik 79, 361, 489, 734 (1926)
4. Van der Waerden : “ Sources of Quantum Mechanics”, North Holland Publ.
Amsterdam (1967)
5. P.A.M. Dirac : Proc. of the Royal Soc. A114, 243 (1927)
6. L. Susskind and J. Glogower : Physics 1, 49 (1964)
7. B.B. Varga and S. Aks : Phys. Lett. 31A, 40 (1970)
8. J. Garrison and J. Wong : Journ. of Math. Phys. 11, 2242 (1970)
9. S.G. Eckstein, B.B. Varga and S. Aks : Nuovo Cimento 8B, 451, (1972)
10. W.H. Louisell : Phys. Lett. 7, 60 (1963)
11. J. Zak : Phys. Rev. 187, 1803 1969
12. P. Carruthers and M.M. Nieto : Rev. of Mod. Phys. 40, 411 (1968)
13. E.C. Lerner : Nuovo Cimento 56B, 183 (1968); erratum 57B, 251(1968)
14. E.C. Lerner and H.W. Huang and G.E. Walters : Journ. of Math. Phys.
11, 1679 (1970)
15. E.K. Ifantis : Journ. of Math. Phys. 11, 3138 (1970)
16. E.K. Ifantis : Journ. of Math. Phys. 12, 1021, 2512 (1971)
17. R. Jackiw : Journ. of Math. Phys. 9, 339 (1968)
18. C.M. Bender and T.T. Wu : Phys. Rev. Lett. 21, 406 (1968)
19. C.M. Bender and T.T. Wu : Phys. Rev. 184, 1231 (1969)
20. B. Simon : Annals of Phys. 58, 76 (1970)
21. J.J. Loeffel, A. Martin, B. Simon : Phys Lett. 30B, 656 (1969)
22. S. Graffi, V. Grecchi and B. Simon : Phys. Lett. 32B, 631 (1970)
23. S.N. Biswas, K. Datta, R.P. Saxena, P.K. Srivastava and V.S. Varma :
Phys. Rev. D4, 3617 (1971)
24. S.N. Biswas, K. Datta, R.P. Saxena, P.K. Srivastava and V.S. Varma :
Pre-print Delhi Univ., (1971)
25. For a variation calculation for the anharmonic oscillator, see : N. Bazley
and D. Fox : Phys. Rev. 124, 483 (1961)
26. C.G. Darwin : Proc. of the Royal Soc. A117, 258 (1927)

39
27. E.H. Kennard : Z. Physik 44, 326 (1927)
28. Louis de Broglie : “Nonlinear Wave Mechanics”, Elsever Publ.,
New York (1960)
29. M. Born and D.J. Hooton : Z. Physik 142, 201 (1955)
30. M. Born and D.J. Hooton : Proc. Camb. Phil. Soc. 52, 287 (1956)
31. M. Born : Z. Physik 153, 372 (1958)
32. G. Petiau : Comptes Rendus 238, 998, 1568 (1954)
33. G. Petiau : Comptes Rendus 239, 344, 792, 1158 (1954)
34. G. Petiau : Comptes Rendus 240, 2491 (1955)
35. F. Karolhazy : Nuovo Cimento 42A, 390 (1966)
36. E. Schrodinger : Z. Physik 14, 664 (1926)
37. R.J. Glauber : Phys. Rev. 131, 2766 (1963)
38. P. Carruthers and M.M. Nieto : Amer. Journ. of Phys. 33, 537 (1965)
39. E.C.G. Sudarshan : Phys. Rev. Lett. 10, 277 (1963)
40. G. Szego : “ Orthonormal Polynomials”, American Mathematical Society
Colloquium Publ. Providence R.I. Vol. 23 (1965)
41. P. Carruthers and M.M. Nieto : Phys. Rev. Lett. 14, 387 (1965)
42. K.E. Cahill : Phys. Rev. 138, B 1566 (1965)
43. K. Eswaran : Nuovo Cimento 70B, 1 (1970)
44. H. Barnard and J.M. Child : “ Higher Algebra” , Macmillan and Co. (1959)
45. E.T. Whittaker and G.N. Watson : “ A Course of Modern Analysis”,
Cambridge Univ. Press, Cambridge (1952) Chapter XXXI.
46. E.C. Titchmarsh : “ Eigenfunction Expansion”, Oxford Univ. Press,
Oxford (1962) Chapter VII, p 144.
47. Pao Lu : Journ. of Chem. Phys. 47, 815 (1967)
48. Yu. E. Murakhver : Soviet Phys. Doklady 10, 1079 (1966).
(Translated from Doklady Akad. Nauk. SSSR 165, 5261 (1965))
49. A. Messiah : “Quantum Mechanics”, North Holland Publ.
Amsterdam (1966) Vol I, Chapter VI.
50. F.R. Gantmacher : “ The Theory of Matrices ”, Chelsea Publ. New York
(1960) Vol I, Chapter VII.

**********************************************

40

You might also like