You are on page 1of 23

Journal of the Korean Physical Society, Vol. 41, No. 5, November 2002, pp.

594616

Review Articles

Quantum Theory of the Harmonic Oscillator in Nonconservative Systems


Chung-In Um
Department of Physics, College of Science, Korea University, Seoul 136-701

Kyu-Hwang Yeon
Department of Physics, Chungbuk University, Cheongju, Chungbuk 306-763
(Received 7 September 2002)
Starting with the quantization of the Caldirola-Kanai Hamiltonian, we review in detail various
phenomenological methods to treat the damped harmonic oscillator as a dissipative system. We
show that the path integral method yields the exact quantum theory of the Caldirola-Kanai Hamiltonian without violating the Heisenbergs uncertainty principle. Through the dynamical invariant
together with the path integral, we also present systematically the exact quantum theories for various dissipative harmonic oscillators, as well as the relations between the canonical, and unitary
transformations for the classical, and quantum dissipative systems.
PACS numbers: 03.65.Ca, 03.65.Ge
Keywords: Harmonic oscillator, Exact theory

oscillator developed from early 1930s to the mid 1980s.


In Sec. III, we make use of the propagator method to
deal with the theories while we treat the theories in Secs.
IV, and V through the dynamical invariant and second
quantization methods. The relation between the canonical, and unitary transformations for the classical, and
the quantum dissipative systems are presented in Sec.
VI. Finally, we provide a summary in Sec. VII.

I. INTRODUCTION

Since Bateman proposed the time-dependent Hamiltonian in a classical context [1] for the description of
dissipative systems, there has been much attention paid
to quantum-mechanical treatments of nonconservative,
and nonlinear systems. In studying nonconservative systems, it is essential to introduce a time-dependent Hamiltonian which describes the damped oscillation, i.e., the
Caldirola-Kanai Hamiltonian. This was discovered first
by Caldirola [2], and rederived independently by Kanai
[3] via Batemans dual Hamiltonian, and afterward by
several others [46]. In the treatment of the Schr
odinger
equation, there are significant difficulties in obtaining the
quantum-mechanical solutions for the Caldirola-Kanai
Hamiltonian. It is confirmed clearly that, even though
the various solutions of the Schr
odinger equation can be
obtained, all these kinds of solutions always violate one
of the fundamental laws of quantum mechanics [7, 8].
However, in 1987, the present authors [9] evaluated the
exact quantum-mechanical solutions for the CaldirolaKanai Hamiltonian. The theory guarantee that all the
fundamental laws in quantum mechanics would be satisfied.
In this paper, we review the phenomenological approach to the exact quantum theory, which is based on
the present authors work (hereafter referred to as the
Um-Yeon solution). In Sec. II, we give a chronological
survey of the theories of the quantum damped harmonic
E-mail:

II. CHRONOLOGICAL SURVEY


In this section, we survey chronologically the various
approaches to linearly damped harmonic oscillators from
the year 1931 to the mid 1980s, following some parts of
Dekkers argument [10], and the historical approaches
of the theories through the Lagrangian, and Hamiltonian formalisms [11]. In 1931, Bateman [1] presented the
first investigation of the so-called dual or mirror-image
Hamiltonian. His Hamiltonian consisted of two different
time-dependent Hamiltonians: One represented the simple one-dimensional damped harmonic oscillator. The
energy dissipated by the oscillator was completely absorbed at the same time by the mirror-image oscillator,
and thus the energy of the total system was constant.
The other described a harmonic oscillator with an exponentially growing mass, that could be reproduced as the
so-called Caldirola-Kanai Hamiltonian. Batemans dual
Hamiltonian is given as
H = p
p (x
px
p) + w2 x
x,

cium@korea.ac.kr; Fax : +82-62-970-2204

-594-

(1)

Quantum Theory of the Harmonic Oscillator in Nonconservative Systems Chung-In Um and Kyu-Hwang Yeon

where x
is the mirror variable corresponding to the variable x. The Lagrangian corresponding to Eq. (1) is
2 x
x x),

L = x x
x + (xx

(2)

and from Eq. (2), two equations of motion could be


obtained: one for the variable x, and the other for x
:
x
+ x + 2 x = 0,

(3)

x
+ 2 x
x
= 0.

(4)

= dx/d(t).
Hence, x
Equation (4) clearly represents the time reversal process of Eq. (3), and is called the equation of motion for
the image-mirror oscillator of Eq. (3). The dynamical
variables x, p, and x
, p are the operators that should
satisfy the commutation relations [x, p] = i~, and [
x, p]
= i~.
Since the Hamiltonian of Eq. (1) is Hermitian, we may
introduce the creation, and the annihilation operators;
their inverse transformations, and conjugates, which are
made up of the variables x, p, and x
, p, can be easily obtained. With the use of these operators, and the
Baker-Hausdorff relation [12,13], we may reduce Eq. (1)
to a simpler form, and then find the eigenstates, and
eigenvalues for the Hamiltonian of Eq. (1). However,
the time-dependent uncertainty product obtained in this
method,
"
#
 2
  2  2
~
4
2
4t
(px x) =
e
1+4
sin t ,
2

(5)
tends obviously to zero. On the other hand, Eq. (5) violates Heisenbergs principle for 6=0, regardless of how
small the frictional coefficient is. Therefore, Batemans
dual Hamiltonian describes classical mechanics correctly,
but does not follow the fundamental principles of quantum physics. This problem is closely connected with the
treatment of eigenstates, and eigenvalues.
Caldirola [2] developed a generalized quantum theory
of a nonconservative system in 1941. Starting with the
Hamiltonian formalism, Caldirola built up the quantum
theory for a linear dissipative system. The equation of
motion of a single particle system subjected to a generalized nonconservative force Q can be written as


d T
T
V

=
+ Q(q),
(6)
dt q
q
q
where the potential V is only a function of q. Let us perform a nonlinear transformation on time as a canonical
variable,
t = (t),

Rt

dt = (t)dt ,

(t) = e

(t)dt

(7)

and assume the rth component of Q to have the form


s

Qr = (t)

X
T
= (t)
ark qk ,
qr
k=1

(8)

-595-

where (t) is an arbitrary function, and ark are constants. We may construct the Schr
odinger equation for
a nonconservative system from the classical Hamiltonian
through Eq. (7):


~2 2

+ V .
(9)
H = i~ =
t
2m
Transforming again the nonlinear time t into ordinary
time t in Eq. (9), we obtain a single-particle Schr
odinger
equation:



~2 0 t 2
i~
=
e
+ e0 t V ,
(10)
t
2m
where (t) has been taken as a time-independent constant 0 . Equation (10) is known as the Caldirola-Kanai
Hamiltonian, and yields the linear dissipative equation
of motion, Eq. (3). The commutation relation, and the
uncertainty product for the coordinate variable x, and
the corresponding kinetic momentum p become
[x, p ] = i~e0 t ,

(11)

~ 0 t
e .
(12)
2
As time goes to infinity, the uncertainty relation vanishes; thus, this formalism violates a fundamental principle in quantum physics.
In 1948, Kanai [3] derived the Caldirola-Kanai Hamiltonian from Batemans dual Hamiltonian by applying to
canonical transformations through several canonical generators. The transformation leads to the Caldirola-Kanai
Hamiltonian
1
1
(13)
H = e2t 2 + e2t 2 y 2. .
2
2
xpx

This Hamiltonian is identical to Eq. (10), and leads to


the linear dissipative equation of motion, Eq. (3). The
eigenstates of the Schr
odinger equation constructed from
the stationary Caldirola-Kanai Hamiltonian may be expressed by using the pseudostationary eigenstates
 1/4  1 1/2
n (x, t) =
~
2n n!
 


1
1
x2
exp i n +
t + t ( + i)e2t
2
2
2~
r 

w
Hn et x
.
(14)
~
Regarding the Hermite polynomial in Eq. (14), the inclusion of the et term in the argument makes this quite
different from the case of a simple harmonic oscillator.
It is interesting to compare these eigenstates with the
S
ussen-Hass-Albrecht results [14,15] when excluding et
in the pseudostationary states; both are identical to each
other.
The pseudostationary states may be used to evaluate the expectation values. By evaluating the quantum

-596-

Journal of the Korean Physical Society, Vol. 41, No. 5, November 2002

fluctuations for the proper mechanical momentum, and


canonical momentum, one finds the uncertainty relation
to be

2
2 2
1
2
4t ~
(PX X) = e
n+
.
(15)
2
2
The uncertainty product, Eq. (15) for the canonical momentum tends to zero as time goes to infinity. On the
other hand, the Hamiltonian in Eq. (13), as a generator
of motion, does not obey the fundamental principle in
quantum mechanics.
The Caldirola-Kanai Hamiltonian can be treated in an
electrical model. This model was proposed first by Pryce,
and Stevens [16] adopted the Pryce model, and showed
that it was possible to introduce damping into the quantum description of a harmonic oscillator that is made
of an LC circuit coupled to a semi-infinite transmission
line. These kinds of approaches have been investigated
by many other authors [1720], who have chosen a chain
of coupled harmonic oscillators as a heat reservoir. The
coupled oscillators must be in the limit of infinitely many
oscillators.
On the other hand, the damped equation of motion
for a LC circuit,
2

dQ Q
d Q
+R
+
= 0,
dt2
dt
C

(16)

can be derived from the Hamiltonian


H = eRt/L

P2
AQ2
+ eRt/L
,
2AL
2C

(17)

where the momentum P , and the coordinate Q correspond to the current, and the charge, respectively, in
the LC circuit. Eqs. (16), and (17) have the same form
as Eq. (3). The quantum mechanical treatment of the
Hamiltonian of Eq. (17) by Svinin [6] preserves the uncertainty relation artificially as time goes to infinity:
 2


~
~
~2
[px ()X()]2 =
coth2

.
2
2kB T
4
(18)
Within the framework of geometric quantization, Dedene [21] proposed a complex symplectic formulation of
damped harmonic oscillators. This theory was based on
the complex dynamical variables given by Dekker [22
24]. Dedene introduced the canonical transformation
1
z = (
p ix),
2

1
z = (
p + i x
),
2

(19)

where z means a formal complex mirror conjugation.


The dynamical variables z, and i
z correspond to a canonically paired coordinate, and momentum [
z = z = z ].
Substitution of Eq. (19) in the Bateman dual Hamiltonian yields Dedenes Hamiltonian
H = h + h

(20)

h = (w i)z z.

(21)

In agreement with Dedene [21], z, and z are the annihilation, and the creation operators in a generalized Hermitian form. If the usual Weyl symmetrization [25] is
allowed, the Hamiltonian of Eq. (20) can be separated
into two uncorrelated parts, i.e., h+ , and h , which are
mixtures of the physical oscillator, and its mirror image.
Incorporating a nonzero separation constant, ,
H = h + h+ ,




1
1
h = zz + ~ i zz + ~ ,
2
2




1
1
h+ = z z + ~ + i z z + ~ ,
2
2

(22)

(23)

where =1, and =0 exhibit the Weyl, and the normal orderings [12], respectively. The eigenvalues of the Hamiltonian are given by




1
1

~

i
n
+

~,
h()
=
n
+
n
2
2
n = 0, 1, 2, .
(24)
Taking =0, and =2, we can reduce Eq. (24) to the
eigenvalue expression. The choice of =1, and =0 yields
Bopps spectrum [26].
The equation of motion can be expressed in the conventional commutator form, and thus the general functions F+ (z , z ), and F (z, z) can be written as sums
of the factorizing terms F+ F given by F (z, z , z, z ).
Then, one can obtain the mean value of the correct
equation of motion. Through this procedure, one can
obtain the position, and the momentum spreads of the
damped oscillator. The uncertainty product obtained in
the complex symplectic Hamiltonian is identical to Eq.
(5). Therefore, Heisenbergs principle is apparently violated again. The main flaw in the complex Hamiltonian
comes from the incorrect fundamental commutation relation. On the other hand, the physical oscillator, and its
mirror mathematical adjoint do not commute with each
other.
Bopp introduced the pseudo-density operator W as
the projection operator w = | >< |, and derived the
general density matrix in terms of the coherent states
| > as
Z
2
n,m = P (0 )(0)
(25)
nm d 0 ,
(0)

where P () is a quasi-probability density, and nm is


the density operator. Equation (25) satisfies the master
equation. Using Eq. (25), one obtains the expectation
value for the uncertainty relation:


~ 2t

2
(X)2 =
e
1 + sin 2t + 2 2 sin2 t
2

~
+ (1 + e2t ),
(26)
2

Quantum Theory of the Harmonic Oscillator in Nonconservative Systems Chung-In Um and Kyu-Hwang Yeon
2

(PX )



~ 2t

2
2
=
e
1 sin 2t + 2 2 sin t
2

2
~
+
(1 + e2t ).
(27)
2

The first parts in Eqs. (26), and (27) are exactly the
same as those of Eq. (5), but there are extra terms that
guarantee the correct uncertainty product.
The evaluation of the propagator can be a convenient
method for finding the quantum-mechanical solution for
a given system. Especially, the propagator for a given
quadratic Hamiltonian can be expressed exactly as a
path integral [27, 28]. Tikochinsky [7] has shown how
to solve this problem as an initial value problem. We
consider the quadratic Hamiltonian given by
H(x, p, t) = 0 + 1 x + 2 x2 + 3 p + 4 p2
+ 5 xp + 5 px,

(28)

where i = i(t) . The Hamiltonian of Eq. (28) satisfies


the Schrodinger equation
i~

K
= HK.
t

(29)

If the Hamiltonian is quadratic, then the propagator [29]


can be written as


i
K(x, t) = F (t) exp Sc (x, t) ,
(30)
~
where F (t) is a factor depending only on the time interval, and Sc (x, t) is the classical action. Substituting Eq.
(30) into Eq. (29) under the assumption


S
S
+ H x,
, t = 0,
(31)
t
t
together with
S(x, t) = a(t) + b(t)x + c(t)x2 ,

(32)

one obtains coupled first-order differential equations for


a(t), b(t), c(t), and i (t) from the Hamilton-Jacobi equation. When the i is not time-dependent, the timedependent solutions for a(t), b(t), and c(t) can be obtained. For the simple harmonic oscillator, we get 0 =
1 = 3 = 5 = 0, 2 = mw2 /2, and 4 = 1/2m.

For convenience, let us write the Caldirola-Kanai


Hamiltonian, Eq. (10), as
h(t) = e2t

p2
1
+ e2t m 2 x2 ,
2m
2

(33)

which yields the linearly damped equation, Eq. (3). In


this case, we have 0 = 1 = 3 = 5 = 0, 2 =
e2t m 2 /2 and 4 = e2t /2m. Then, we obtain
1
1
mx02 cos t/ sin t + x02 ,
2
2
b(t) = mx0 exp(t)/ sin t,
1
c(t) = me2t ( cos t sin t)/(sin t),
2
F (t) = et/2 [m/(2ih sin )]1/2 ,
a(t) =

(34)
(35)
(36)

= ( 2 2 )1/2 .
(37)
Setting =0, Eqs. (34)-(37) are reduced to those of the
simple harmonic oscillator. We will determine the validity of such expressions in the next section. It is easy to
show that as tends to zero, the propagator is reduced
to the product of the two propagators for simple harmonic oscillators. In this theory it is instructive to note
that one may bypass the equations of motion, and obtain
the classical action, and propagator directly as solutions
of an initial value problem.
Cheng [8] evaluated the propagator for the CaldirolaKanai Hamiltonian by using a modified Feynman path
integral in configuration phase space via Montrolls
method [30]. Furthermore he showed that the propagator
for the damped harmonic oscillator could be evaluated
at, and beyond caustics with the help of the Horv
athyFeynman formula [25].
The propagator can be expressed as a path integral in
phase space as


Z
1
K(q 00 , q 0 , T ) = exp (pq H(p, q)dt) DpDq (38)
~
where H(p, q) is the Hamiltonian of the system considered, and DpDq is the two-dimensional path differential measure in phase space. One obtains the propagator for the damped harmonic oscillator described by the
Caldirola-Kanai Hamiltonian as

!1/2
0
00
h m 
i
0
00
me(t +t )/2)
K[q , q ; T ] =
exp
(et q 02 et q 002 )
2i~ sin t
4i~



0
00
0
00
im
exp
[(et q 02 + et q 002 ) cos T 2e(t +t )/2 q 0 q 00 ] .
2~ sin t
00

-597-

In the case of a bound system, the propagator is ex-

(39)

pressed in terms of the time-dependent wavefunction

-598-

Journal of the Korean Physical Society, Vol. 41, No. 5, November 2002

n (x, t) [29,30] as
K(q 00 , t00 ; q 0 , t0 ) =

n (q 0 , t0 )n (q 00 , t00 ).

(40)

n=0

Comparison of Eq. (40), and Eq. (39) yields


n(q,t) = Nn exp(iEn t/~)
exp[(m/2~)(1 + i/2)et q 2 ]
Hn [(m/~)1/2 ]et/2 q]
with the energy eigenvalues



1
1
En =
n+
+ ~,
2
4

(41)

n = 0, 1, 2, , (42)

where the normalization factor is Nn


=
(m/~)1/4 (2n n!)1/2 and Hn (X) is the nth-order
Hermite polynomial. Equations (41), and (42) are
equivalent to those of Kerner [31], and Hasse [32]. The
propagator, Eq. (30), was also evaluated by Khandekar,
and Lawande, and Jannussis et al. [33]. It should be
noted that the derivation of this propagator is carried
out by using the method generalized by Cheng [34].
However, Cheng did not show whether or not this
derivation guaranteed the uncertainty relation. The
correct derivation, i.e., the Um-Yeon solution, will be
shown in the next section.
In this section, we have reviewed various previous theories for the dissipation of a linear damped harmonic
oscillator in quantum mechanics. Besides these theories, there are many other theories, such as the modified Bopp-Dekker master equation for the reduced density operator in the treatment of the quantum optics
oscillator [3537], the nonlinear frictional quantum theory in a Gaussian approximation due to Hasse [38], the
Hamilton-Jacobi formalism [30,3840], the quantization
of the novel Hamiltonian in the Schr
odinger-Razavy variation procedure [4,41], and so on. We should also point
out that none of the theories for this linear damped harmonic oscillator are perfect. Although they may satisfy
one of the fundamental principles in quantum mechanics,
they do not guarantee the others. In the next section,
we will present what we believe to be a more elegant
quantum theory for the damped harmonic oscillator.

III. DAMPED QUANTUM HARMONIC


OSCILLATION
1. Path Integral Methods

Although the Feynman path integral formulation


[29] offers a general approach for treating quantummechanical systems, only a few time-dependent
Schr
odinger equations can be solved exactly. Most of the
previous theories for the Caldirola-Kanai Hamiltonian violate the uncertainty relation. This difficulty is critically

reviewed by Dodonov-Manko [41], and others [42]. In


this section, we discuss the quantum-mechanical solution (Um-Yeon solution) of the Caldirola-Kanai Hamiltonian for this dissipative system by using the propagator method developed by the present authors [9].
We introduce the Caldirola-Kanai Hamiltonian with a
time-dependent external driving force f (t), defined as
the time-dependent damped driven harmonic oscillator
(DDHO), as


p2
mw02 x2
H = et
+ et
xf (t) .
(43)
2m
2
Hamiltons equations
Lagrangian

1
L = et
mx 2
2

of motion for Eq. (43) yields the



1
mw02 x2 + xf (t) ,
2

(44)

and the corresponding equation of motion is


x
+ x + w02 = f (t)/m.

(45)

The mechanical energy can be expressed as


1
E = e2t p2/2m + mw02 x2 .
2

(46)

Here, the energy expression in Eq. (46) is not equal to


the Hamiltonian itself.
In the path integral formulation, the solution of the
Schr
odinger equation is given by the path-dependent integral equation with the propagator K:
Z
(x, t) = K(x, t; xo , 0)(x0 , 0)dx0
(47)
which provides the wavefunction (x, t) at time t in
terms of the wavefunction (x0 , 0) at time t=0. The
Hamiltonian of Eq. (1), and the propagator K in Eq.
(47) should satisfy the Schr
odinger equation [Eq. (29)].
If the Hamiltonian is quadratic, the propagator [29] can
be written as Eq. (30). Making use of the Hamiltonian
of a free particle, we can express the propagator for the
damped quadratic Hamiltonian can be expressed as
K(x, t; x0 , 0) =

mwet/2
2i~ sin wt

1/2

exp


i
Sc(x, x0 , t) .
~
(48)

The classical action of the DDHO Hamiltonian is




Z
0
1
1
Sc = et
mx 2 mw02 x2 + xf (t0 ) dt0 .
2
2

(49)

In Eq. (49) for small w0 , the kinetic energy is dominant,


with the Lagrangian acting like a damped free particle,
so that one may take the propagator for DDHO as having
the form

Quantum Theory of the Harmonic Oscillator in Nonconservative Systems Chung-In Um and Kyu-Hwang Yeon







2

2
K(y, p; y0 , 0) = exp a(p) exp i p y + b(p) exp i p y + c(p) ,
w0
w0

where we have changed the variables as


r
w0
mw0
p=
t,
y=
x.
2i
~

2 (t) =

b =

c =


1
+ w cot wt ,
2

F (t) =

mwwt/2
2i~ sin wt

(53)

(54)

w t/2
e
,
sin wt

(55)
1/2

(56)

When one takes =0, Eq. (52) becomes the familiar


propagator of the simple harmonic oscillator.
The Hamiltonian of DDHO, Eq. (43), reduces to a
quadratic form for t=0, and f (0)=0, i.e. a simple harmonic oscillator. Then, the corresponding wave function
n (x, 0), and the energy eigenvalues are given by


1
n (x, 0) = N0 Hn (0 x) exp 0 x2 ,
(57)
2
En =

1
n+
2

~w0 .

(58)

From Eq. (47) together with Eqs. (43), (52), and (57),
we obtain the wavefunction at time t in the form
Z
n (x, t) =
K(x, t; x0 , 0)n (x0 )dx0





1
1
= N n 1/2 exp i n +
cot1
+ cot wt
2
2w
(2 n!)
exp[Ax2 ]Hn [D(x)].

(50)

(61)

(51)

Substituting Eqs. (43), and (51) into Eq. (29), we can


determine the time-dependent coefficients in Eq. (50).
To simplify the expression, we take f (t)=0 in Eq. (43);
then, the propagator can be expressed as


im 2
2
K(y, p; y0 , 0) = F (0) exp
a
x + ebx0 2 + 2
cx0 x ,
2
(52)
where the new time-dependent coefficients are


1
a
= + w cot wt et ,
2

2
+ cos t + cos ec2 t,
4 2

-599-

(59)

Here, the time-dependent coefficients are given by


 mw 1/4 exp 1 t
4
N=
,
(60)
~
(t)(sin wt)1/2


m t
1
A(t) =
e
2
2h
(t) sin2 t


/2 + cot t

cot t +
,
+i
2
(t)2 sin2 t

D(t) =

0 et/2
.
(t) sin t

(62)

(63)

The mechanical energy, Eq. (47), can be expressed


as the energy operator E, whose expectation values take
the form


~2 2t 2
1
< E >mn =
e
+ m02 < x2 >mn .
2
2m
x mn 2
(64)
The evaluations of the expectation value for x2 , and /
are straightforward. Here, we write only the diagonal
element of < E >mn as follows:




1 ~ t
1
.
< E >nn = n +
e
(t)2 sin2 t +
2 2
(t)2 t
(65)
To evaluate the uncertainty relation, in a similar way
to that used to obtain < x2 >, and < 2 /x2 >, we
can calculate the expectation values, i.e., < x >mn , and
< /x >mn .
Then, the uncertainty relation corresponding to the
diagonal is given by


1
2
2 1/2
[(x) (p) ]n,n = n +
~
2
n
h

1+
cot t (t)2 sin2 t
2

i2 1/2
+
+ cot t
.
2
(66)
When f (t) 6==0, the progagator [Eq. (52)], and the
wavefunction [Eq. (59)] are of a new form, but we have
not expressed these here because of their complicated
tructure. Equation (52) has the same structure as that
obtained by Cheng [8], Khandekar, and Lawande [33],
and Janussis et al. [33]. The propagator [Eq. (52)] requires that the Hamiltonian be identical to the energy of
the system. In this sense, the mechanical energy operator
[Eq. (64)] is not identical to the Hamiltonian operator

-600-

Journal of the Korean Physical Society, Vol. 41, No. 5, November 2002

Fig. 1. Energy eigenvalue En,n (t) =< E >nn [(65)] for


/2 = 0.1.

Fig. 2. Uncertainty relation as the (n, n) state oscillates


with period [Eq. (66)].

[Eq. (43)]. Therefore, one assumes that this Hamiltonian


represents a quantum-mechanical dissipative system.
The energy expectation values given in Eq. (65)
contain the term et , and thus decay exponentially.
The second off-diagonal elements, En+2,n , not shown
here, depend only on the exponential decay constant
. Figure 1 illustrates the decay of the energy eigenvalue, Enn (t). Note that the results of Dodonov, and
Manko [43] can be obtained by taking the driving force
as f (t) = f0 sin(t + ).
The uncertainty for the (n, n) states with period [Eq.

(66)] is reduced to that of the harmonic oscillator at 180


 ,
1

and 0 . The uncertainty for the (n, n) states is n + 2 ~.


Figure 2 illustrates the uncertainty for the (n, n) states.
It does not decay exponentially, but oscillates with the
period ; thus, the uncertainty relation is satisfied. One
should recognize that every fundamental law in quantum
theory is guaranteed in this Um-Yeon solution.

The coherent states can be defined by the eigenstates


of the non-Hermitian operator a i.e., a| > | >. Using the completeness relation for the number representations, we can expand | > as

2. Coherent States

Coherent states for the harmonic oscillator have been


studied, and are widely used to describe many fields of
physics [4448]. In the case of a quantum-mechanical
model of a damped harmonic oscillator, Dodonov, and
Manko [43] introduced the Caldirola-Kanai Hamiltonian
with an external force, and constructed the coherent
states, and Um et al. [4951] constructed the correct
coherent states for the damped harmonic oscillator described by the Caldirola-Kanai Hamiltonian.
Here, we define the creation, and the annihilation operator a , and a, and using these operators, we construct
coherent states with the following properties: (i) They
are eigenstates of the annihilation operator, (ii) they are
created from the vacuum or the ground states by a unitary oprerator, (iii) they represent the minimum uncertainty states, and (iv) they are not orthogonal but complete, and normalized. To obtain these operators, we will
make use of the propagator in Eq. (59).

| e(1/2)||

X
2

n
|n >= e(1/2)|| ea |0 >, (67)
n!
n=0

where |0 > is the vacuum or ground state, and is independent of n. The calculation of < | > in Eq. (67)
gives
< | e1/2(||

+||2 )

+ .

(68)

Since Eq. (68) has nonzero values for = , the states


are not orthogonal. The eigenvalues of the coherent
states are complex numbers u+iv; thus, the completeness
relation of coherent states can be written as
Z
d2
| >< |
= 1.
(69)

To define a , and a for the damped oscillator, we make


use of the wavefunction in Eq. (59) for the eigenvalues
of < x >mn , and < p >mn :

1/2
1
< x >mn = n +
(t)m,n+1 + n1/2 (t)m,n1 ,
2
(70)

< p >mn =

n+

1
2

1/2

(t)m,n+1 + n1/2 (t)m,n1 ,


(71)

where
(t) =

h
h
ii
1
(Re A)1/2 exp i cot1
+ cot(t) ,
2
2
(72)

(t) =

2i~

h
h
ii
A
exp i cot1
+ cot(t) ,
D
2

(73)

Quantum Theory of the Harmonic Oscillator in Nonconservative Systems Chung-In Um and Kyu-Hwang Yeon

with the relation = i~. Then, the creation,


and the annihilation operators for the damped harminic
oscillator can be defined as
a=

1
(x p),
i~

a =

(74)

1
( p x).
i~

(75)

We can easily confirm that a, and a are not Hermitian operators, but the following relations are preserved:
[x, p] = i~, and [a, a ]=1. We can evaluate the transformation function < x| > from the coherent states to the
coordinate representation |x > from Eqs. (74) as

1 2
1/4
x
< x| (2 )
exp
2i~


1
1 2
+ x ||2
.
(76)

2
2
Finally, we can show that a coherent state represents
a minimum uncertainty state. It is straightforward to
evaluate the expectation values of x, p, x2 , and p2 in
state | >. We obtain
(x)2 =< x2 > < x >2 = ,
(p)2 =< p2 > < p >2 = .

(77)

Therefore, the uncertainty relation becomes


(x)(p) = ||2 ||1/22 =

(t) =
+

(78)

2 !1/2
1 h i2
sin (t) +
sin(2t)
.
8
2

IV. HARMONIC OSCILLATOR WITH


TIME-DEPENDENT FREQUENCY AND
EXTERNAL FORCE
In general, the solutions of the Schrodinger equation
for explicit time-dependent systems have been investigated by many authors [48,55,56]. Camiz et al. [57], and
Khandekar, and Lawande [33] have obtained the wavefunctions of a time-dependent harmonic oscillator with or
without an inverse quadratic potentials. In this section,
we study the propagator, and the quantum mechanical
solutions for a forced harmonic oscillator with a timedependent frequency through the path integral method
[58,59].
Consider a forced harmonic oscillator with a timedependent frequency (t), whose Hamiltonian is of the
form
H=

p2
m
+ w2 (t) f (t)x
2m
2

(79)

Equation (78) is obviously the minimum uncertainty corresponding to Eqs. (66) in the (0,0) state. The uncertainty for the (n, n) state [Eq. (78)] oscillates with the
period , which corresponds to the half period of a simple harmonic oscillator. From all of the above, we conclude that the coherent states for the damped harmonic
oscillator with the Caldirola-Kanai Hamiltonian satisfy
properties (i)-(iv).
We notice that the Um-Yeon solution guarantees that
the fundamental laws in quantum mechanics, and especially, Heisenbergs uncertainty principle, are preserved.
This theory has been successfully applied to a molecular
system absorbed on a dielectric surface [52], a charged
particle in an infinite square-well potential with a constant electric field [53], a driven coupled harmonic oscillator, and a coupled damped driven harmonic oscillator
[54].

(80)

where f (t) is an external driving force. The correasponding Lagrangian is


1
1
mx 2 m 2 (t)x2 + f (t)x,
2
2
with the classical equation of motion
L=

(81)

1
dx2
+ 2 (t)x = f (t).
dt2
m

(82)

For the case (t) = 0 , the solution of Eq. (82) represents harmonic motion; otherwise, it is not easy to evaluate the exact solution.
Since the Lagrangian is quadratic, we can assume the
propagator to have the form [9,54,60]
K(x, t; x0 , t0 ) = exp[a(t, t0 )x2 + b(t, t0 )xx0
+c(t, t0 )x02 + g(t, t0 )x
+h(t, t0 )x0 + d(t, t0 )].

 h i
1 3
1+
8

h ii

~
(t),
2

-601-

(83)

Substituting Eqs. (80), and (83) into Eq. (29), we can


determine the time- dependent coefficients in the propagator. The propagator for a forced time-dependent harmonic oscillator, Eq. (83), be expressed as


m( )
1/2
0 0
K(x, t; x , t ) =
2i~ sin( 0 )



im 2 0 02
exp
x 0x
2~


im
exp
[(x
2 + 0 x02 )
2~ sin( 0 )

p
2
0
0
0
cos( ) 2 xx +
x
m
Z t
f (s)

ds p
sin[(s) ]
(s)

t0

Z
2 0 t
f (s)
+
x
ds p
sin[ (s)]
m
(s)

t0

-602-

Journal of the Korean Physical Society, Vol. 41, No. 5, November 2002

Z t
1
f (s)
2
ds p
sin[ (s)]
m t0
(s)

#)
Z t
f (p)

dp p
sin[(p) ]
.
(p)

t0

Hn

(85)

and the function q(t) satisfies the differential equation


d2 q(t)
+ 2 (t)q(t) = 0.
dt2

(86)

Here, (t), and (t) are real quantities. From Eqs. (85)(86), the real, and the imaginary parts of the differential
equation are given as
(t) (t)(t)
2 + 2 (t)(t)
= 0, 2(t)
(t)

+ (t)
(t) = 0,
(t)2 (t)

= ,

(87)

where the constant is a time-invariant quantity.


In Eq. (84), the unprimed, and the primed variables
denote quantities which are functions of t0 , and t, respectively. In the case of (t) = 0 (real constant), we have
(t)=1, and (t) = 0 t. Then the propagator, Eq. (84),
reduces to the usual expression for a forced harmonic oscillator. We can obtain the wave function directly from
the definition of the propagator
K(q 00 , t00 ; q 0 , t0 ) =

n (q 0 , t0 )n (q 00 , t00 )

(88)

n=0

and from Mehlers formula [61]


exp[(X 2 + Y 2 2XY )/(1 Z 2 )]

1 Z2

n
2
2 X Z
= exp(X +Y )
Hn(X) Hn(Y ) .
n
2 n!
n=0

(89)

Comparison of the result of Eq. (84) with Eq. (88)


through Eq. (89) gives the wavefunction as
 



1
n (x, t) = exp i (t) n +
(t) n (x, t),
2
(90)
where
 1/2 #1/2
1

n (x, t) = n
2 n! ~
(
"
#)
Z t
im
2p
f (s)
exp
x
x

ds p
cos[ (s)]
2~
m
(s)

"
#2
Z

m p
1 t
f (s)
cos[ (s)]
exp
x

ds p

2
m
(s)

"

m
~

(
)#
Z
p
1 t
f (s)
x

ds p
.
sin[ (s)]
m
(s)

(91)

(84)

Here, the relation between (t), and (t) in Eq. (84) is


given by
q(t) = (t)ei(t) ,

"r

The wave function, Eq. (90), is simply a unitary transformation of n (x, t), where n (x, t) satisfies all the properties associated with n (x, t).
The Hamiltonian, Eq. (80), and Lagrangian, Eq. (81),
represent the time-dependent energy. Therefore, one
must derive a time-invariant energy operator. Let (t)
be a particular solution of Eq. (82). Making use of Eqs.
(86), we can express the energy as
~2 2 2
m
Eop =
+ ( 2 + 2 )x2
2
2m
x

2



2x
+ 1 + ( ) i~
+ mx

2
x
x
m
m
2.
+ m 2 x + 2 2 + ( )
(92)
2
2
Equation (91) can be simplified to the form
n (x, t)

1/2
2
2
1 2

eix +ix e 2 (x) Hn [(x )]


= n
2 n!

1/2
2

= n
eAx +Bx Hn [(x )],
(93)
2 n!
where the new coefficients are
 m 1/2
=
,
~
m
(t) =
,
2~
Z t
1p
f (s)
(t) =

ds p
cos[ (s)],
~
(s)

Z t
p
1

(t) =
dsf (s) (s)

sin[ (s)],
m
A = i 2 /2,
B = i + 2 .

(94)

With Eq. (91), it is straightforward to evaluate the


energy expectation values and the uncertainty relations.
The diagonal elements of the energy expectation value,
and uncertainty relation are given by


~ 2
1
En,n = En = ( )(2n

+ 1) = ~ n +
, (95)
2
2

(xp)n,n =

2
1+ 2 2

1/2 

1
n+
2

(96)

The energy expectation value is obviously a timeinvariant quantity. The minimum uncertainty of Eq.
(96) is larger than ~/2; thus, the minimum uncertainty
states are needed.

Quantum Theory of the Harmonic Oscillator in Nonconservative Systems Chung-In Um and Kyu-Hwang Yeon

To obtain the coherent states, we construct a creation


operator a , and annihilation operator a for the timedependent harmonic oscillator.

1/2 


m

a =
1+i
xi
,
(97)
2~

M

a=

m
2~

1/2 



p
1+i
x+i
.

M

(98)

It is easy to obtain the representation (x, p) in terms


of (a , a). The commutation relations, [x, p] = i~, and
[a, a ]=1, are preserved. Through a similar process to
obtain Eq. (76), the eigenvector of the operator a in the
coordinate representation |x > can be calculated, and
we can evaluate the expectation values in the state | >.
Then, the uncertainty relation becomes
"

 #1/2
2
~
eta
xp =
1+
,
(99)
2

which is the minimum value allowed by Eq. (96).
Here, the Hamiltonian, Lagrangian, and mechanical
energy have the units of energy, but are not time invariant. For this reason, we have used the time-invariant
operators, and derived the energy operator from the classical equation of motion. We then used the energy operator to calculate the energy expectation values. This
energy operator is very similar to the Ermakov-Lewis invariant operator [55]. The uncertainty relation is time
dependent, but consistent with Heisenbergs principle.

V. TIME-DEPENDENT BOUND, AND


UNBOUND QUADRATIC HAMILTONIAN
SYSTEM: DYNAMICAL INVARIANT
METHOD
In this section, we investigate the exact quantum theory of a general time-dependent bound, and unbound
quadratic Hamiltonian system [62] through the dynamical invariant, and path-integral methods. We find the
relation between the quantum mechanical solution, and
the dynamical invariant [6366].

1. Time-dependent Bound Quadratic Hamiltonian

One may consider a system with a time-dependent


quadratic Hamiltonian of the type
1
H = [A(t)p2 + B(t)(xp + px) + C(t)x2 ],
2

(100)

where A(t) is a nonzero time-dependent function, and


B(t), and C(t) are time- dependent functions of an arbitrary form, which are continuously differentiable with

-603-

respect to time. From Hamiltons equations of motion,


the classical equation of motion becomes
"

A(t)B(t)
A(t)
x + A(t)C(t) +
x

A(t)
A(t)
i

B 2 (t) B(t)
x = 0.
(101)
However, the solution with a general form for arbitrary
time-dependent coefficients is not known. For simplicity,
let us write Eq. (101) as
x
+ (t)x + (t)x = 0.

(102)

A general solution of Eq. (102) can be expressed in the


form
x = (t)ei(t) ,

(103)

where the functions (t), and (t) must be determined


from Eq. (102). Substitution of Eq. (103) in Eq. (102)
yields the real, and the imaginary parts of Eq. (102) as
2 + (t) + (t) = 0,

(104)

+ 2 + (t) = 0.

(105)

The first invariant quantity (t) can be found from Eq.


(105) in the form
=

2
.
A(t)

(106)

Equation (106) is a time-invariant quantity with an auxiliary condition given by Eq. (102). If is not equal
to zero, then is not constant, and the position has the
form of a complex function of time. Since the particle
of the system will pass through more than two points on
the trajectory, the motion of the system will be found in
some restricted region. If is equal to zero, the motion
of the system will be unbound.
To find another classical invariant quantity with the
auxiliary condition given by Eq. (102), let us assume
that this invariant quantity I(t) is given by
I(t) =


1
(t)p2 + 2(t)xp + (t)x2 ,
2

(107)

where (t), (t), and (t) are all real time-dependent


functions. From Hamiltons equations of motion, the
time derivative of I(t) becomes
dI
I
=
+ [I, H].
dt
t

(108)

Combining Eqs. (100), and (107) with (108), we can determine the time-dependent coefficients, (t), (t), and
(t). Finally, one obtain the invariant quantity
(
2 

2 )

B(t)
1
I(t) =
x +

x + p
. (109)

A(t)
A(t)

-604-

Journal of the Korean Physical Society, Vol. 41, No. 5, November 2002

Note that I(t) is always positive in an unbound system,


and negative in a bound system.
As we mentioned in the previous section, the propagator for a quadratic Hamiltonian has the following form
of Eq. (103):
K(x, t; x0 , t0 )


i
= exp
[a(t, t0 )x2 + b(t, t0 )xx0 + c(t, t0 )x02 d(t, t0 )] .
~
(110)
Instead of Eq. (110), we can introduce the definition of
the propagator for the bound system given by Eq. (88).
For the unbound system, the propagator is given as
Z
K(x, t; x0 , t0 ) = dkk (x, t)k (x0 , t0 ).
(111)
The above two propagators, Eqs. (88), and (111), must
satisfy the Schr
odinger equation, Eq. (29), and its complex conjugate. Substituting Eqs. (100), and (110) into
Eq. (29), we can determine the time-dependent coefficients, a(t), b(t), and c(t). From the auxiliary condition,
i.e., the classical solution, if (t)ei(t) is that solution,
then (t)ei(t) is also a solution of that system. The
classical solution can be written as
q(t, t0 ) = 0 sin( 0 ).

(112)

With the use of Eq. (112) together with the coefficients


a(t), b(t), and c(t), we can write the propagator, Eq.
(110), as
1/2

1/2 01/2
K(x, t; x0 , t0 ) =
2i~ sin( 0 )A1/2 A01/2



i

exp
+ cot( 0 ) B x2
2~A


i

0
0
0
+

cot(

)
+
B
x02
2~A0

)

1/2
i 0
xx0
+
.
~ AA0
sin( 0 )


(113)

Comparing Eq. (113) with Mehlers formula, Eq. (89)


through Eq. (88), we find the wavefunction of the system:

1/4 
1/2
1
1
n (x, t) =
ei[(1/2)+n]
A
2n n!
(
1/2 )
1
Hn
x
~A



 
1

exp
i
B x2 .
(114)
2~A

Equation (114) is the wavefunction of the bound system with the auxiliary condition of the classical solution.

The uncertainty relation becomes


 "

2 #1/2

1
1
~ 1+ 2
B(t)
.
(xp)n,n = n +
2

(115)
Note that the diagonal element of the uncertainty relation in the ground state is larger than the minimum
uncertainty value, ~/2.
The quantum invariant operator corresponding to Eq.
(109), i.e., the classical quantity, can be defined as
 2

1
+1
2 2
I=
[B(t) ] x2
2
2 A2(t)


2 2
+
[B(t) ](xp

+ px) + p .
(116)
A(t)
I(t) should satisfy the quantum condition that corresponds to the classical condition, Eq. (88),
dI
I
1
=
+ [I, H] = 0,
dt
t
i~

(117)

with the Hamiltonian given by Eq. (100). The expectation values of the quantum invariant operator, Eq. (116),
are given by

 2
1

h
m,n
< m|I|n > = n +
2
A


1
= n+
hm,n .
(118)
2
Here, the propagator, and the wavefunction have been
obtained, and the wavefunction has been expressed in
terms of the classical solution. In the evaluation of the
uncertainty relation, and the expectation values, we have
used the wavefunction, together with the invariant operator, which is inferred from its classical counterpart. The
expectation values of the quantum mechanical invariant
operator I also satisfy the uncertainty relation.

2. Time-dependent Unbound Quadratic Hamiltonian

For convenience, we express the Hamiltonian of Eq.


(100) as
H=

1
[A(t)p2 + B(t)(pq + qp) + C(t)q 2 ],
2

(119)

where q, and p are the canonical coordinate, and its conjugate momentum, respectively. The classical equation
of motion is given as
q + (t)q + (t)q = 0.

(120)

Let the general solution of Eq. (120) be in the form


q = C1 (t) exp[(t)] + C2 (t) exp[(t)],

(121)

Quantum Theory of the Harmonic Oscillator in Nonconservative Systems Chung-In Um and Kyu-Hwang Yeon

where C1 , and C2 are integral constants. Substituting


Eq. (121) into Eq. (120), we obtain two differential
equations, one for e(t), and the other for (t). For e(t),
and (t) to satisfy these two equations simultaneously,
the following differential equations must be satisfied:
+ (t) = 0,
+ 2 + (t)rho

(122)

+ 2 + (t)
= 0.

(123)

Equation (123) offers the invariant quantity quantity


(t):
(t) =

2
.
A(t)

(124)

One can find another classical time-invariant quantity


from Eq. (120):
I(q, p, t) =

1
[(t)p2 + 2(t)qp + (t)q 2 ],
2

(125)

where , , and are real time-dependent functions.


Combining Eqs. (119), and (125) with Eq. (108), we
can determine these time-dependent functions, and the
invariant quantity can be expressed as
 2 

2

B(t)
1
I(q, p, t) =
+

q + p . (126)

A(t)
A(t)
The eigenvalue of the coefficient matrix of the
quadratic variable in Eq. (126) is given as
(
 2 2 
2 )
1

B(t)
1
2
=
+
+

. (127)
2

A(t)
A(t)
When is real, one of the eigenvalues is positive, and the
other is negative, so that the invariant quantity I(q, p, t)
is a hyperbola in phase space, which represents unbound
motion.
Consider a quantum unbound system with a real .
Taking the propagator in the form of Eq. (110), and
using Eqs. (29), we obtain the coupled differential equations for the coefficients a(t), b(t), and c(t). Since both
r,
and are real in an unbound system, according to
Eq. (121), the solution to Eq. (120) can be expressed as
q = 0 sinh(r r0 ).

(128)

Solving the coupled differential equations for the coefficients, together with Eq. (128), we obtain the timedependent coefficients; thus, the propagator for a system
with an unbound time-dependent quadratic Hamiltonian
is given by
K(q, t; q 0 , t0 )

 12
1/2 01/2
=
2i~A1/2 A01/2 sinh(r r0 )



i

exp
+ ctg~(r r0 )
2~A(t)

-605-

 

 
i

i
qq 0
B(t) q 2 exp
2~A(t)
~ 0 sinh(r r0 )

 0

 
i

i
exp
0 + 0 ctg~(r r0 ) B(t0 ) q 02 .
2~A(t)

2
(129)

It is well known that the Hamiltonians for a free particle,


a damped free particle, and an overdamped, an underdamped, and a negative harmonic oscillator belong to
the unbound quadratic Hamiltonian system. Here, two
time-invariant quantities with an auxiliary condition are
obtained. One of these invariants can be used to determine whether or not the system is unbound. The
propagator, Eq. (129), for the unbound system with a
quadratic Hamiltonian enables us to find the propagators
for several unbound systems.

VI. QUANTUM DAMPED HARMONIC


OSCILLATOR-DYNAMICAL INVARIANT
In the treatment of a time-dependent quantum system, Lewis, and Riesenbeld [63, 64] introduced the dynamical invariant method to obtain the exact quantummechanical solutions. The invariants received primary
concern because of their value in discussing physical
problems [6769], and because of the possibility to apply
them to classical, and quantum physics [70,71]. In this
section, we make use of the dynamical invariant quantity to evaluate the creation, and annihilation operators,
and then the exact quantum mechanical solutions for the
quantum damped driven harmonic oscillator.

1.
Quantum-mechanical Treatment for the
Damped Harmonic Oscillator

Consider an external driving force with a tail, g(t),


with the form [72]
Z t
g(t) =
dt0 (t t0 )f (t0 ),
(130)

where f (t) is an instantaneous force at a given time t,


and (t t0 ) satisfies the properties of the -function.
Then, the Hamiltonian of a damped harmonic oscillator
with the above force has the form


p2
1
H(
p, q, t) = et
+ et m02 q2g (t)
q . (131)
2m
2
Equation (131) is exactly the same as Eq. (43), i.e.,
the so-called Caldirola-Kanai Hamiltonian. The classical
Lagrangian, and the corresponding equation of motion
are given by


1
1
L(q,
q, t) = et mq2 m02 q2 + g(t)q ,
(132)
2
2

-606-

Journal of the Korean Physical Society, Vol. 41, No. 5, November 2002

g(t)
.
m

(133)

where

Through the same procedure as in Sec. V, we can obtain an invariant operator, I(p, q, t), that satisfies Hamiltons equation, Eq. (117), within the assumption of a
quadratic form for p, and q:

 
1
2
I = et m2 02
(
q q0 )2
2
4


1
et (
p + p0 )2 + m(
q q0 )2 ,
(134)
2

(t) =

q + q + 02 q =

where p0 (t) = mert q0 (t), and q0 (t) is the solution to the


differential equation
q0 + q0 +

02 q0 (t)

g(t)
=
.
m

(135)

From Eq. (135), q0 (t) may be regarded as a particular


solution of Eq. (133):
Z t
Z t0
0
0
00

0
q0 (t) =
dt
dt00 e(tt )/2 e(t t )
md t0
t0
sin[d (t t0 )]f (t00 ),

(136)

where d = (02 2 /4) 2 . Note that the past affects


influences in the time- dependent function q0 (t):
The dynamical invariant operator can be written in
terms of the annihilation, and the creation operators as
follows:

1/2
1
a
(t) =
2~md e t
n

o

met d + i
[
q (t)] + i
p ,
(137)
2
1/2
1
a
(t) =
2~md e t
n

o

met d i
[
q (t)] i
p ,
2

(138)

q0 (t) +



1/2

2
q0 (t)
q

(t)
+
i
1

.
0
202
402
0
(139)

These operators satisfy the commutation relation


[a, a ]=1. The dynamical invariant operator, I(q, p, t),
can be represented in terms of a, and a as


1
I = ~d a
a
+
.
2

Using the eigenstates of the invariant operators, we


can easily obtain the exact wavefunctions satisfying
the Schr
odinger equation. Since the solution to the
Schr
odinger equation differs only by a time-dependent
phase factor from the eigenstate of the invariant operator [64], we can write directly the wavefunction, n (x, t)
as
n (q, t) = n (q, t)e(i/h)R(t) .

n (q, t)n (q 0 , t0 )

(141)

Then, the exact form of the wavefunction is given by


1/4



met d
1
md et
2
exp
(q q0 )
n (q, t) =
~
2~
n!2n
"
#
1/2
md et
i(n+1/2)wd (tt0 )
e
Hn
(q q0 )
~



i met
q2
exp
(q q02 ) 2q0 (q q0 ) + 02
~ 2
2
20
 Z t

i
dt0 [L0 (t0 ) (t0 )] .
(142)
exp
~ t0
An exact closed form of the propagator can be obtained
readily from the expansion formula, Eq. (88):

#1/2
0
md (et et )1/2
K(q, t, q , t ) =
=
2i~ sin d (t t0 )
n=0


i
im h t 


0
2
t0
0
2
exp
e
d cot d (t t )
(q q0 ) + e
d cot(t t ) +
(q q0 )
2
2
2
(
)


0
imd (et et )1/2
im t
0
0
t0 0
0
0
exp
(q q0 )(q q0 ) exp
[e q0 (q q0 ) e q0 (q q0 )]
~ sin d (t t0 )
~



Z t
i m t 2
t0 2
00
00
00
exp
(e q0 e q0 )
dt [L0 (t ) (t )] .
~ 402
t0
0

(140)

"

With the use of the wavefunction, Eq. (142), we can

(143)

obtain the uncertainty relations for various states, and

Quantum Theory of the Harmonic Oscillator in Nonconservative Systems Chung-In Um and Kyu-Hwang Yeon

the uncertainty relation corresponding to the diagonal


element is given by
(qp)n,n =

~
(2n + 1).
2

(144)

Note that the off-diagonal elements of the uncertainty


relations, (x, P )n1,n , are governed by the past
in terms of q0 (t) and p0 (t), and that the relations
(x, P )n2,n , and (x, P )n2,n have the same forms
as they do for the simple harmonic oscillator.

However, if is imaginary, Eq. (151) is a hyperbola in


phase space, and q can take any value in space, making
the system unbound.
In order to obtain the eigenfunctions, and the eigenvalues of the invariant operator, we define the annihilation,
and the creation operators a, and a by the relations





1
m 1
q + ip ,
(152)
a=

2m~
a =

2. Harmonic Oscillator with a Time-dependent


Frequency

The generalization of the relation between the dynamical invariant, and the solution of the Schr
odinger equation for the time-dependent oscillator offers wide applications to various fields [7377]. Let us investigate the
quantum solutions of the harmonic oscillator with a timedependent frequency via the dynamical invariant, and
second quantization methods [78,79].
The Hamiltonian for the harmonic oscillator with a
time-dependent frequency is given by
2

H=

p
1
+ m 2 (t)q 2 ,
2m 2

(145)

where 2 (t) is a real positive function, and the classical


equation of motion is
q + (t)2 q = 0.

(146)

The solution to Eq. (146) can be written as


q = (t)ei(t) ,

(147)

where (t), and (t) are determinable from Eq. (146).


Substitution of Eq. (147) into Eq. (146) yields the real,
and the imaginary differential equations:
2 + (t)2 = 0,

(148)

+ 2 = 0.

(149)

From Eq. (149), an invariant quantity can be found:


= m2 ,

(150)

with an auxiliary condition being given by the classical


solution. Another time-invariant quantity can be obtained from Eq. (108). Through a similar calculation,
we obtain this classical invariant quantity as


1 2 2
2
2
I=
q + (p m q)

.
(151)
2 2
The invariant quantities , and I are measures of the
bound system. If is real, Eq. (151) is an elliptic equation in phase space. Thus, as the values of q, and p in the
system are limited in same region, it is a bound system.

-607-





1

m 1 +
q ip ,

2m~

(153)

with the auxiliary conditions of Eqs. (148), and (149).


These operators satisfy the commutation relation. Then,
the invariant quantity of Eq. (151) can be expressed by
a, and a as


1

,
(154)
I = ~ aa +
2
and the eigenstates, and the eigenvalue spectrum of the
invariant operator are given as
aa |n n|n >, n = 0, 1, 2,


1
= n+
, n = 0, 1, 2,
2

(155)

Applying the annihilation operator to the ground state,


and then solving this equation, we obtain the normalized
ground state as

1/4


m
m
2
u0 =
(1 i
)q

.
(156)
exp
~
2~
We may apply the creation operator a n times to the
eigenfunction of the ground state, Eq. (156). we, thus,
obtain the nth excited state. In consideration of Eq.
(141), the exact wavefunction of the nth state of the
system corresponding to the Schr
odinger equation can
be expressed as

1/2 
1/4
1
m
(q, t) = n
e(1/2+n)
2 n!
~
"


 
1/2 #
m

m
2
exp
1i
q Hn
q
. (157)
2~

~
With the help of Mehlers formula, Eq. (89), the propagator of the system is given as


1/2
m 0
K(q, t; q 0 , t0 ) =
2i~ sin( 0 )



im 2
1
exp
q 0 q 02 +
2~

sin( 0 )
io
p
[(q
2 + 0 q 02 ) cos( 0 ) 2 0 qq 0 ] .
(158)
Note that Eqs. (146), and (158) are the same as the
previous results, i.e., Eqs. (82), and (84) for f (t)=0,
which were derived by using the propagator method [58].

-608-

Journal of the Korean Physical Society, Vol. 41, No. 5, November 2002

capacitances, and C the coupling capacitance. These


equations of motion can be formulated from the following
Hamiltonian [16]:
p21
q2
p2
+ et 1 + et 2
2L1
2C1
2L2
2
2
q
(q

q
)
1
2
+et 2 + et
,
2C2
2C
Hq = et

Fig. 3. Diagram of mesoscipic caacitance coupled circuit.

3. Mesoscopic Capacitance Coupled Circuit

The advent of mesoscopic physics [80], and nanotechnology [81] was stimulated, and encouraged by
a host of intriguing theories, such as localization in
lower dimension. The rapid development of nanotechnology, and nanoelectronics [82] has made it possible to
minimize integrated circuits, and components towards
atomic-scale dimensions [83]. In this section, we investigate the quantum-mechanical quantities of a mesoscopic
capacitance-coupled circuit in a vacuum state [84, 85].
We consider the RLC circuit in Fig. 3, which couples
two loops via a capacitance in the presence of a power
(t) in one part of the circuit. The classical equations of
motion are
L1

d2 q1
q1
q1 q2
+ R1 q 1 +
+
= (t),
2
dt
C1
C

(159)

L2

q2
q1 q2
d2 q2
+ R2 q 2 +
+
= 0,
2
dt
C2
C

(160)

where qi (i = 1, 2) stand for the electric charges in the


circuits, Li the inductances, Ri the resistances, Ci the

b1

b
1

b2

where we have assumed (t) = 0, and = R1 /L1 =


R2 /L2 for simplicity. Here, the electric charges qi substitute for the conventional coordinates while their conjugation variables pi represent the electric current, instead
of the momentum.
Through a procedure similar to the one used in Sec. V,
we may obtain the dynamical invariant quantity, which
has quadratic form, as
( 2 



1
1
1 X t p2i
t 2
I=
e
+ qi pi +
+
e qi
2 i=1  Li
C
Ci
2
et q1 q2 .
(162)
C
To treat the system quantum mechanically, we replace
the canonical variables of the classical system with the
corresponding quantum operators; then, we obtain the
quantum invariant operators
( 2 
X
1
p2

I =
et i + (
qi pi + pi qi )
2 i=1
Li
2




1
1
2
+
et qi2 et q1 q2 .
(163)
+
C
Ci
C
Here, we define the time-dependent canonical creation,
and annihilation operators, bi , and bi , to obtain the
eigenfunctions, and eigenvalues of the invariant operator:

!
r
L

1
2
4
= p
1 L1 L2 e
cos q1 4
sin q2
L2
2
L1
2
2~1 L1 L2
"
!
!#)
r
r
r
r

L2

L1

p
L1

L2

2 t 4
4
t
4
4
+i e
cos p1
sin p2 +
L1 L2 e 2
cos q1
sin q2
,
L1
2
L2
2
2
L2
2
L1
2
!
(
r
r
p

1
L

1
2
= p
1 L1 L2 e 2 t 4
cos q1 4
sin q2

L2
2
L1
2
2~1 L1 L2
!
!#)
"
r
r
r
r

L2

L1

p
L1

L2

4
4
2 t 4
t
4
i e
cos p1
sin p2 +
L1 L2 e 2
cos q1
sin q2
,
L1
2
L2
2
2
L2
2
L1
2
!
(
r
r
p

1
L

1
2
= p
2 L1 L2 e 2 t 4
sin q1 +4
cos q2

L2
2
L1
2
2~2 L1 L2
1

2t

(161)

(164)

(165)

Quantum Theory of the Harmonic Oscillator in Nonconservative Systems Chung-In Um and Kyu-Hwang Yeon

"

b
2

!
!#)
r
r
r

L2

L1

L2

L1
4
t
4
4
+i e
sin p1 +
cos p2 +
L1 L2 e 2
sin q1 +
cos q2
,
L1
2
L2
2
2
L2
2
L1
2
(
!
r
r
p

1
L

1
2
= p
2 L1 L2 e 2 t 4
sin q1 +4
cos q2

L2
2
L1
2
2~2 L1 L2
"
!
!#)
r
r
r
r

L2

L1

p
L1

L2

2 t 4
4
t
4
4
i e
sin p1 +
cos p2 +
L1 L2 e 2
sin q1 +
cos q2
,
L1
2
L2
2
2
L2
2
L1
2
2 t

where
s

i
,
4
L1 L2
r


L2 1
1

1 =
+
cos2
L1 C
C1
2
r


L1 1
1
sin
+
+
sin2 +
,
L2 C
C2
2
C
r


L2 1
1

2 =
+
sin2
L1 C
C1
2
r


L1 1
1
sin
+
+
cos2
,
L2 C
C2
2
C

2 L1 L2
tan =
.
(168)
L2 (1 + C/C1 ) L1 (1 + C/C2 )
There operators satisfy the commutation relation, and
obey the usual properties of creation, and annihilation
operators. The invariant quantity of Eq. (163) can be
expressed as


2
X
1

,
(169)
I=
~i bi bi +
2
i=1
i =

and the eigenstates, and the eigenvalue spectrum of Eq.


(169) are the same as those of Eq. (155).
We introduce a linear transformation of the canonical
i , Pi ):
operators (
qi , pi ), and new variables (Q
q
q

 



L1
L2
4
4
cos

sin
1
t
q1
Q
L
2
L
2
2
1

2
q
q
,
2 = e
4

L1
L2
q2
4
Q
L2 sin 2
L1 cos 2
(170)


P1
P2

cos 2
q

L2
4
L1 sin 2

t
=e2

L2
L1

t
p
+
L1 L2 e 2
2

L1
L2

L1
L2

sin
cos

2
cos 2 4 L
sin 2

q
q L1
4

L1
L2
4
sin
cos
L2
2
L1
2

L1
L2

p1
p2


p
iPi ],
L1 L2 Q

(167)

(173)

(174)
The above wavefunction should satisfy the Schrodinger
equation for the Hamiltonian, Eq. (161):

q |n ,n (t) > .
|n1 ,n2 (t) H
(175)
1
2
t
The wavefunction in the q representation, |n1 ,n2 (t) >,
representation by
may be connected with that in the Q

some unitary transformation U :


i~

|n1 ,n2 (t) >= hatU |n1 ,n2 > .

(176)

Substitution of Eq. (176) in Eq. (175) gives a Hamiltonian which represent the simple harmonic oscillator
having frequencies w1 , and w2 :
2

1 H
qU
i~U
1 U = p1
H
q = U
t
2 L1 L2
2
p
1
p

1p
+
L1 L2 12 q12 + 2
+
L1 L2 22 q22 ,
2
2
2 L1 L2
(177)

(171)
1
U
3
U

"

(178)
#

!
L1

ln
+ t (
p1 q1 + q1 p1 )
L2
2
!
#
r
i
L2

4
exp
ln
+ t (
p2 q2 + q2 p2 ) ,
2~
L1
2


i
= exp
(
p1 q2 p2 q1 ) ,
~2



i L1 L2 2
= exp
(
q1 + q22 ) .
(179)
4~

1 = exp
U

Then, we may express Eqs. (164)-(167) by using Eqs.


(56)-(57), as
p
1
bi = p
+ iPi ],
[i L1 L2 Q
(172)

2~i L1 L2

(166)

From Eqs. (172), and (173) we confirm that bi , and bi


are the annihilation, and the creation operators, respec repretively, for a simple harmonic oscillator in the Q
sentation respectively. The wavefunction |n1 ,n2 (t) >
differs by only a time-dependent phase factor from the
wavefunction, |n1 , n2 , t > of the Schr
odinger equation:
"
#


2
X
1
|n1 ,n2 ,t (t) exp i
ni +
i t |n1 , n2 , t > .
2
i=2

where
=U
1 U
2 U
3 ,
U


q1
q2

1
b = p
[i

i
2~i L1 L2

-609-

i
2~
"

-610-

Journal of the Korean Physical Society, Vol. 41, No. 5, November 2002

Using Eqs. (176), together with Eq. (77), we can


obtain the uncertainty relations for the charge, and the
current of the system for n1 = n2 = 0 as
(q1 )2 (p1 )2




2

2
4
=
1+
cos
+
1
+
sin4
2
2
41
2
42
2



2
1

varphi ~2
2
+
+
+
sin2 cos2
.(180)
1
2
21 2
2
2
4
(q2 )2 (p2 )2




2

2
4
sin
+
1
+
cos4
=
1+
2
2
41
2
42
2


 2
2
1
2

~
+
+
+
sin2 cos2
.
1
2
21 2
2
2 4

P = e(t) p (t)e(t) q.

(185)

Here, (t), and (t) are real, and differentiable functions


of t, and the relations between the new, and the old
variables are simply time-dependent linear relations. If
(Q, P ) are to be canonical coordinates, there should exist
a new Hamiltonian where (Q, P ) is determined by the old
Hamiltonian, and the transformation, Eqs. (184), and
(185). For all trajectories in phase space, these variables
must satisfy the relation
P Q HQ (Q, P, t) = pq H(q, p, t) +

dF (Q, P, t)
,
dt
(186)

(181)

From Eq. (168), we know that one frequency, w1 , cannot


be equal to the other frequency, w2 . The uncertainty relations do not satisfy the minimum uncertainty relation,
(xi , pi )min =~/2, eventhough is zero. The uncertainty relation approaches the minimum relation when
r = w1 /w2 1 or n.

where the generating function, F (Q, P, t), is a timedependent function in phase space. The coefficients
(P, p) of Q, P are given by
P p

F (Q, P, t)
q
=
,
Q
Q

q
F (Q, P, t)
=
.
P
P

(187)

(188)

Combining Eqs. (184), and (185) with Eqs. (187), and


(188), we obtain the generating function
VII. LINEAR CANONICAL, AND UNITARY
TRANSFORAMTIONS ON GENERAL
HAMILTONIAN SYSTEMS
It is well known that a simple unitary transformation
of the time variable transforms the Schr
odinger equation for a damped harmonic oscillator with the CaldirolaKanai Hamiltonian [Eq. (43)] into the Schr
odinger equation for an undamped harmonic oscillator [86,87]. In this
section, we review the relations for the canonical transformations in classical mechanics, and the unitary transformations in quantum mechanics [8890], and provide
the applications of this theory to the time-dependent harmonic oscillators [91,92].

1. Linear Canonical Transformations for Classical, and Quantum Systems

Consider a system with the Hamiltonian, and the Lagrangian given by


2

H(q, p) =

p
+ V (q),
2m

(182)

L(q, q)
=

1 2
mq V (q).
2

(183)

Let us introduce a linear transformation of the canonical


variables (q, p) to new variables (Q, P ):
Q = e(t) q,

(184)

F (Q, P, t) = Q2 .
2

(189)

The new Hamiltonian, and Lagrangian are given by



p2

HQ (Q, P, t) = e2
+ + e2
PQ
2m
m


1 2 2
+
e
+ 2 + Q2+V (Q) ,
(190)
2
m
2 + )QQ
t) = e2 Q 2 (me
LQ (Q, Q,
1
+ (m 2 e2 )Q
2 V (Q).
2

(191)

The relation between the canonical momentum, and the


kinetic momentum can be obtained from the Lagrangian:
2 + )Q.
P = e2 Pk (me

(192)

The differentiation of the inverse canonical transformation, Eqs. (184), and (185), gives the relation between
the kinetic momenta of the new, and old systems as
+ e Pk .
pk = mq

(193)

The mechanical energy of the new coordinate system becomes



P2 

E = e4
+ + e2 e2 P Q
2m
m
2
1 
2
+
m + e
Q2 + V (Q),
(194)
2m
where Eq. (194) includes , but does not really depend
on .

Quantum Theory of the Harmonic Oscillator in Nonconservative Systems Chung-In Um and Kyu-Hwang Yeon

-611-

One can make use of Feynmans path integral to investigate the quantum-mechanical relation between two
systems connected canonically [29,93]. From the definitions of the propagator, and the wavefunction, Eqs. (47),
(29), and (88), together with Eq. (192), we obtain the
Schr
odinger equation as

We can prove that U (x)(x) is the solution of Eq. (195)


in some x-space if (x) is a solution of Eq. (197):

(Q, t)
Q (Q, t),
=H
(195)
t
where


2
2
i~ 

2 ~
2

HQ = e

+e
2Q
+1
2m Q2
2
m
Q


1 2 2
+
e
+ 2 + Q2 + V (Q).
(196)
2
m

(Q,
P , t) < Q| U
(
(Q, t) = U
q , p, t)(e q),(204)

| > .
| U

The operation of < q| or < Q| on Eq. (203) yields the


relation between the wavefunctions in each space as

i~

This Hamiltonian operator is the same as the classical


Hamiltonian of Eq. (190). Using Eqs. (183), and (47)
together with the integral form of the propagator, we
obtain the Schrodinger equation obtained as
i~

(q, t)
~2 2 (q, t)
=
+ V (q)(q, t).
t
2m q 2

Consider the unitary operator connecting two quantum systems given by






i
i
2

U (
q , p, t) = exp
q exp (
q p + pq)(199)
.
2~
2
Using Eq. (199), we obtain the quantum operator relations corresponding to the classical relations, Eqs.(184),
and (185):
=U
qU
= e q,
Q
pU
= e p e q.
P = U

q (
(Q(
q , p, t), P (
U
q , p, t) = U
q , p, t), t)




i
i
q p + pq) 2q 2 .
= exp e2 q2 exp (
2~
2~
(205)
Utilization of Eq. (203) yields the relation between the
quantum average of the position, and the momentum
operators in the old, and the new spaces as

(201)

(202)

qU
+ | >
< |
q | < |U
= e < |
q | >

>,
= e < |Q|

(206)

< |
p| e < |(
p +
q )| >

= e < |(P + Q)|


>
6= q < |P | >,

(207)

e < |
< |Q|
q | >,

(208)

< |P | e < |
p| > e e < |
q | > .(209)
Equations (205)-(208) have the same forms as Eqs.
(184)-(185), and the inverses of Eqs. (184)-(185), respectively. From these relations, we can confirm that the two
systems are related by a canonical transformation, and
form distinct quantum spaces.
From Eqs. (205)-(208), the relation between the quantum uncertainty for the old and the new systems can be
evaluated:
(qp)2 = (P Q)2 + 2 (Q)4
P + P Q)|

+(Q)2 [< |(Q


>

(200)

The inverse forms of Eqs. (200), and (201) can be easily


obtained by using the unitary operator:
Q (Q,
P , t) = U
(
P , t), p(Q,
P , t), t)
U
q (Q,


i
2
= exp e2 Q
2~


i
P + P Q)
+ 2Q
2] .
exp [(Q
2~

q (
where the unitary transformation operator, U
q , p, t),
corresponding to that of the new coordinate, Eq. (202),
is given by

(197)

Here, one can confirm that the quantum Hamiltonian


has the same form as the classical Hamiltonian, Eq.
(182), whose variables are replaced by their corresponding quantum operators. With the use of Eqs. (182), and
(191), the integral factor, and the integral definition of
the propagator, we can obtain the relation between the
propagators of the new, and the old systems [94]:


i
22 2
K(Q2 , t2 ; Q1 , t1 ) = exp 2 e q2
2


i
1 e21 q12
exp
2~
K(q2 , t2 ; q1 , t1 ).
(198)

(203)

>< |P | >]
2 < |Q|

(210)

(P Q)2 = (qp)2 + 2 E 4 (q)4


e2 (q)2 [< |(
pq + qp)| >
< |
q | >< |
p| >].

(211)

or

Note that for =0, the quantum uncertainties for the


two systems are the same.
The kinetic momentum is equal to the canonical momentum in the old system, but it is not equal to it in

-612-

Journal of the Korean Physical Society, Vol. 41, No. 5, November 2002

the new system. From Eq. (190), the kinetic momentum


operator in the new system is

Pk = e2 P + (m + e2 )Q.

(212)

With Eqs. (207)-(208), the relation between the quantum averages of the kinetic momentum operators for
both systems becomes
< |
< |P | e < |
pk | > +me
q | > . (213)
From Eqs. (207)-(208), and (212), the relation between
the quantum uncertainties of the position, and the kinetic momentum operators for both systems can be obtained as
n
(QPk )2 = e4 (pk q)2 + m2 2 (q)4
2

+ m(q)
[< |(
pk q + qpk )| >
< |
q | >< |
pk | >]}

(214)

or

2. Quantum Treatment for Special Types of Linear Canonical Transformations

Consider the special case where (t) is set to zero, such


that Eqs. (184)-(185) take on much simpler forms. From
Eqs. (195), and (196), the Schr
odinger equation of this
new system becomes
i~

(Q, t)
~2 2 (Q, t) i~
= e2

t
2m Q2
2


(Q, t)
2Q
+ 1 + V (Q)(Q, t).
Q
(218)

If we know the solution to Eq. (195) together with


Eq. (203), then we can find the solution to the new
Schr
odinger equation. For example, let us give the relation between the general solutions of the classical equation of motion for the harmonic, and the damped harmonic oscillators with the same frequency as
Q = e0 t q,

(219)

(qpk ) = e
{(QPk )
2 2
2 )(Q)4
(m + 2me
2 (Q)2 [< |(Q
P + P Q)|

me
>

2 < |Q| >< |P | >]}.

and the unitary operator connecting them as


h
i
P , t) = exp iover2~0 t(Q
P + P Q)
.
U (Q,
(215)

Though Eqs. (213), and (214) contain (t), the quantum uncertainty does not depends on , but only on .
Therefore, although the old system satisfies the uncertainty principle, the new system may not satisfy it in
some cases if < 0. Furthermore, if , the uncertainty of the new system goes to zero.
The mechanical energy operator for the new system
can be defined from Eq. (194) as
2

p
1

+Q
P )
+ ( + e2 )e2 (P Q
2m 2
m
1
2 + V (Q).

+
(m + e2 )2 Q
(216)
2m

Eq = e4

If the potential energyis assumed to have a quadratic


form in the position, the relation between the quantum
averages of the mechanical energy operators for both systems becomes
h
Q | e2 < |E
Q | > + m 2 < |q 2 | >
< |E
2
#

+ < (hatq p + p
q )| >
2

2 ) < |E
Q | > m(2
Q | >
= e2 < |E
+ me
P + P Q)|

2m < |(Q
>,
(217)
is the mechanical energy of the old system. Alwhere E
though appears in Eq. (216), this equation does not
depend on , but only on .

(220)

From Eq. (151), the Schr


odinger solution for the damped
harmonic oscillator can be evaluated as
P , t) = U (Q,
P , t)HO (Q,
P , t)
DHO (Q,
(m/~)1/4 0 t/2 20 t(m/2h)Q2

e
e
1/4 2n n!
r


m
Q .
Hn e0 t
~

(221)

The damped harmonic oscillator can be represented by


another Hamiltonian as
P , t) = e0 t
HDHO (Q,

P 2
m 2 2
+ e0 t
Q .
2m
2

(222)

Note that equation (221) is the damped harmonic oscillator Hamiltonian corresponding to the classical case of
the harmonic oscillator.
For the scale transformation, since =0, Eqs. (209),
and (210) are equal to each other, i.e., qp) =
(DeltaP Q). For the damped harmonic oscillator in
this case, we can obtain the uncertainties of the system
by using the canonical, and kinetic momentum operators, respectively:


1
(P Q)DHO = n + ~ ,
(223)
2

1 

2 2
1
(QPk )DHO = e20 t 1 + 02
n+
~.
4w
2
(224)

Quantum Theory of the Harmonic Oscillator in Nonconservative Systems Chung-In Um and Kyu-Hwang Yeon

The uncertainty in Eq. (223) vanishes as time goes to infinity. However, the uncertainty in Eq. (222) is constant.
This means that the uncertainty exists, even though the
oscillations have stopped, and is in agreement with Eq.
(66), except for the oscillations.
From Eq. (216), the quantum average of the mechanical energy operator for the damped harmonic oscillator
is




02
1
0 t

(E)DHO = e
1+
~ n +
.
(225)
8w2
2
The mechanical energy, Eq. (224), vanishes as times goes
to infinity, which agrees with Eq. (65).
From the above general results, we can treat the
canonical transformation where the equation of motion
is unique. If (t) is zero, the canonical transformation
equations, Eqs. (184)-(185), become
Q = q,

P = p (t)q.

(226)

We see that one position has innumerable counterpart


canonical momenta of the system due to the arbitrariness of (t). In this case, the Hamiltonian of Eq. (190)
becomes


p2

1 2
HQ (Q, P, t) =
+ PQ +
+ Q2 + V (Q).
2m m
2 m
(227)
In Eq. (226), depending on (t), there are innumerable
Hamiltonians, and Lagrangians that give rise to one dynamical equation for the system. Since the gauge invariance is independent of the choice of (t) in Eq. (226),
Eq. (226) is the gauge transformation. The kinetic momentum for the transformation system becomes
pk (t) = mq = (t)q + P.

(228)

If (t)=0, Eqs. (200), and (201) are reduced to Eq.


(225). In this case, the Hamiltonian of Eq. (196) becomes


2

i~

Q = ~
H

2Q
+
1
2m Q2
2 m
Q


1 2
+
+ Q2 + V (Q).
(229)
2 m
From Eqs. (207), and (208) with (t)=0, the quantum
average of q is invariant for any space, but the momentum operator is not. From Eq. (227), we can define the
kinetic momentum as
Pk(t) = (t)
q + P ,
(230)
which also is invariant. If gauge invariance holds in
classical mechanical treatments, then it also holds for
quantum-mechanical treatments.
From Eq. (210), the relation of the canonical momentums uncertainty for the gauge transformation becomes
(QP )2 = (qp)2 + 2 (q)4
(q)2 [< |
pq + qp)| > 2 < |
q >< |
p| >].
(231)

-613-

The uncertainty, Eq. (230), does not violate Heisenbergs


principle, but depends on the gauge chosen, so Eqs.
(213), and (214) are equal to each other, (QPk ) =
(qpk ). This does not depend on the gauge chosen.
Therefore, the uncertainties of the canonical, and kinetic momentum operators for the harmonic oscillator
become, after the gauge transformation, respectively,
 12 


2
1
n+
,
(232)
(QP )HO = ~ 1 + 2 2
m
2


1
(QPk )HO = ~ n +
.
2

(233)

Under the gauge transformations of Eq. (225), the


Hamiltonian for the damped harmonic oscillator of Eq.
(221) becomes
2
P , t) = e0 t P + e0 t (P Q
+Q
P )
HGDHO (Q,
2m 
2m

m 2
2 2
+ e0 t
+ e0 t
Q .
(234)
2
2m
The relation of the canonical momentum operators uncertainty between the two systems for this gauge transformation can be obtained as
(QP )2GDHO = (QP )2DHO
 2
 
2
~2

0
1
+ 2
+
n+
.
( 02 /4) m2
m
2

(235)

Substitution of Eq. (222) into Eq. (234) yields the


canonical momentum operators uncertainty corresponding to the Hamiltonian of Eq. (233):
(QP )GDHO

1 
2
1
~2
2
0 2
n+
= 1+ 2
+
. (236)
( 02 /4) m2
m
2
The uncertainties of the position, and the momentum
for the system vary with the choice of gauge, but do not
violate Heisenbengs principle; on the other hand, the
uncertainties for the position, and the kinetic momentum
for the system might not satisfy Heisenbergs principle,
but are invariant with respect to the choice of gauge.
Generally, it is not easy to calculate the exact
quantum-mechanical solution for the simple harmonic
oscillator with various potentials, such as the asymmetric quantum septic harmonic oscillator [95], the timedependent harmonic plus inverse harmonic potential [96],
and various other cases [9799]. One can apply the
canonical, and the unitary transformations to evaluate
the exact wavefunction for a harmonic plus an inverse
harmonic potential with a time-dependent mass, and
frequency [100,101].

VIII. SUMMARY

-614-

Journal of the Korean Physical Society, Vol. 41, No. 5, November 2002

This paper consists of two parts: the first presented


quantum-mechanical treatments of various damped harmonic oscillators by using the path-integral method, and
the second part gave the exact quantum theory for the
damped harmonic oscillator through the dynamical invariant method together with the path-integral method.
The Um-Yeon solution for the Caldirola-Kanai Hamiltonian, obtained through the path-integral method, was
presented in Sec. III, where the propagator, wavefunctions, energy eigenvalues, uncertainty relations, and coherent states were discussed explicitly. This theory guarantees that Heisenbergs uncertainty principle and the
other fundamental laws in quantum mechanics are satisfied. Some examples for application of this theory were
given.
In Sec. IV, the procedure used in Sec. III was introduced to evaluate exactly the propagator, wavefunctions, energy expectation values, uncertainty relations,
and coherent states for a harmonic oscillator with a
time-dependent frequency, and a time-dependent external driving force. Section V was devoted to the phenomenological theory of time-dependent bound, and unbound quadratic Hamiltonian systems, which was based
on the path-integral, and dynamical invariant methods;
the propagator, wavefunctions, and expectation values
were evaluated explicitly. We showed the relation between the wavefunction, and the dynamical invariants,
which determined whether or not the system was bound.
The exact quantum-mechanical solutions to the
damped harmonic oscillator with the Caldirola-Kanai
Hamiltonian, to the harmonic oscillator with a timedependent frequency, and to a mesoscopic capacitance
coupled circuit were rederived through the dynamical
invariant in Sec. VI. For the harmonic oscillator with
a time-dependent frequency, two quantum invariant operators were found together with an auxiliary condition.
This theory was applied to several problems.
Section VII described the quantum correspondence
for linear canonical transformations, which are combinations of the scale, and the gauge transformations of
general Hamiltonian systems. We also showed that a single system had innumerable Schr
odinger equations when
a gauge transformation was used; however the quantum
averages of functions of the position, and the kinetic momentum operators were invariant for all solution as for
classical cases.
Although we did not discuss all the theories developed
so far, we adequately covered the necessary studies of
dissipative systems for the quantum damped harmonic
oscillator from the mid 1980s to the present day. The
phenomenological theory is widely applicable to fields
such as the fission of heavy nuclei [102,103], electric conductivity [104], optical resonant cavities [105], quantum
Hall effects [106], tunneling problems through potential
barriers [107], Josephson currents [108, 109], and quantum chaos with dissipation [110,111], and is still an open
problem in physics.

ACKNOWLEDGMENTS
This work was supported by the Korea Research Center for Theoretical Physics and Chemistry (2002), and
by Korea University.

REFERENCES
[1] H. Bateman, Phys. Rev. 38, 815 (1931).
[2] P. Caldirola, Nuovo Cimento 18, 393 (1941); ibid 77,
241 (1983).
[3] E. Kanai, Prog. Theor. Phys. 3, 440 (1948).
[4] P. Havas, Nuovo Cimento, Suppl. 5, 363 (1957).
[5] H. H. Denman, Am. J. Phys. 34, 1147 (1966).
[6] I. R. Svinin, Teor. Mat. Fiz. 22, 107 (1975).
[7] Y. Tikochinsky, J. Math. Phys. 19, 888 (1978).
[8] B. K. Cheng, J. Phys. A: Math. Gen. 17, 2475 (1984).
[9] C. I. Um, K. H. Yeon and W. H. Kahng, J. Phys. A:
Math. Gen. 20, 611 (1987); J. Korean Phys. Soc. 19,
888 (1978); C. I. Um, K. H. Yeon and T. F. George,
Phys. Rep. 362, 63 (2002).
[10] H. Dekker, Phys. Rep. 80, 1 (1981).
[11] H. Haken, Rev. Mod. Phys. 47, 67 (1975).
[12] W. H. Louisell, Quantum Statistical Properties of Radiation (Wiley, NewYork, 1973).
[13] O. H. Weiss and A. A. Maradudim, J. Math. Phys. 3,
771 (1962).
[14] R. W. Hasse, Repl. Prog. Phys. 41, 1027 (1978).
[15] K. Albecht, Phys. Lett. B 56, 127 (1975).
[16] K. W. H. Stevens, Proc. Phys. Soc. London 77, 515
(1961).
[17] I. R. Senitzkey, Phys. Rev. 119, 670 (1960).
[18] E. Braun and S. V. Godoy, Physica 86A, 337 (1977).
[19] H. Dekker, Opt. Commun. 10, 114 (1974).
[20] P. L. Torres, J. Math. Phys. 18, 301 (1977).
[21] G. Dedene, Physica A 103, 371 (1980).
[22] H. Dekker, Rep. Prog. Phys. 42, 1937 (1979).
[23] H. Dekker, Phys. Rev. A 16, 2126 (1977); Physica A
95, 311 (1979).
[24] H. Dekker, Phys. Lett. A 80, 369 (1980).
[25] P. A. Horv
athy, Int. J. Theoret. Phys. 18, 245 (1979).
[26] F. Bopp and Sitz-Ber. Bauer, Akad. Wiss. MathNatusw. K 1, 67 (1973).
[27] C. Garrod, Rev. Mod. Phys. 38, 483 (1966).
[28] R. P. Feynman, Phys. Rev. 84, 108 (1051).
[29] R. P Feynman and A. R. Hibbs, Quantum Mechanics
and Path Integrals (McGraw-Hill, New York, 1965).
[30] E. W. Montroll, Commun. Pure Appl. Math. 5, 415
(1952).
[31] K. H. Kerner, Can. J. Phys. 36, 371 (1958).
[32] R. W. Hasse, J. Math. Phys. 16, 2005 (1975).
[33] A. D. Jannussis, G. N. Brodimas and A. Streclas,
Phys. Lett. 74A, 6 (1979); D. C. Kandekar and S. R.
Lawande, J. Math.Phys. 16, 384 (1975).
[34] B. K. Cheng. Rev. Bras. Fis. 13, 360 (1983).
[35] R. W. Zwanzig, Physica 30, 1109 (1064).
[36] R. Benquria and M. Kac, Phys. Rev. Lett. 46, 1 (1981).
[37] H. Dekker, Physica 83, C183 (1976).
[38] R. W. Hasse, Phys. Lett. B 85, 197 (1979).
[39] K. K. Kan and J. J. Griffin, Phys. Lett. B 50, 241
(1974).

Quantum Theory of the Harmonic Oscillator in Nonconservative Systems Chung-In Um and Kyu-Hwang Yeon
[40] W. Stocker and K. Albrecht, Ann. Phys. 117, 436
(1976).
[41] V. V. Dodonov and V. I. Manko, Phys. Rev. A 20, 550
(1979).
[42] D. M. Greenberger, J. Math. Phys. 20, 762 (1979); J.
J. Cervero and J. Villaroel, J. Phys. A: Math. Gen. 17,
2963 (1984).
[43] V. V. Dodonov and V. I. Manko, Nuovo Cimento B
44, 265 (1978).
[44] E. Schr
odinger, Naturwissenschaften 14, 166 (1926).
[45] Z. E. Zimmerman and A. H. Silver, Phys. Rev. 167,
418 (1968).
[46] J. C. Botke, D. J. Scalapino and R. L. Sugar, Phys.
Rev. D 9, 813 (1974).
[47] M. M. Nieto and L. M. Simons, Jr., Phys. Rev. D 20,
1321 (1979); ibid 20, 1342 (1979).
[48] J. G. Hartley and J. R. Ray, Phys. Rev. D 25, 382
(1982).
[49] K. H. Yeon, C. I. Um and T. F. George, Phys. Rev. A
36, 5287 (1987).
[50] H. G. Oh, H. R. Lee, T. F. George and C. I. Um, Phys.
Rev. A 39, 5515 (1989).
[51] H. G. Oh, C. I. Um, W. H. Kahng and S. T. Choh. J.
Korean Phys. Soc. 22, 14 (1989).
[52] H. G. Oh, H. R. Lee, T. F. George and C. I. Um, Phys.
Rev. A 40, 45 (1989).
[53] K. H. Yeon, C. I. Um, T. F. George and L. N. Pandey,
Can. J. Phys. 72, 591 (1994); K. H. Yeon and C. I. Um,
J. Korean Phys. Soc. 25, 398 (1993).
[54] O. V. Manko, V. V. Dodonov, T. F. George, C. I. Um
and K. H. Yeon, J. Sov. Laser. Rev. 12, 385 (1991); K.
H. Yeon, C. I. Um, W. H. Kahng and T. F. George,
Phys. Rev. A 38, 6224 (1998); K. H. Yeon, C. I. Um
and W. H. Kahng, J. Korean Phys. Soc. 23, 82 (1990).
[55] H. R. Lewis, Jr., Phys. Rev Lett. 18, 510, 639 (1967).
[56] J. R. Burgan, M. R. Feix, E. Fijalkov and A. Munier,
Phys. Lett. 74A, 11 (1979).
[57] P. Camiz, A. Gerardi, C. Marchioro, E. Presutti and E.
Scacciatelli, J. Math. Phys. 12, 2040 (1971).
[58] K. H. Yeon, T. F. George and C. I. Um, Workshop on
Squeezed State and Uncertainty Relations, editted. by
D. Han, Y. S. Kim and W. W. Zachary, NASA Conference Publication 3135, 347 (1992).
[59] K. H. Yeon and C. I. Um, J. Korean Phys. Soc. 24, 369
(1991).
[60] G. J. Papadopoulous, Phys. Rev. D 11, 2870 (1975).
[61] A. Erdelyi, Higher Transcendental Functions (McGrawHill, New York, 1953), Vol. 2, p. 194.
[62] K. H. Yeon, K. K. Lee, C. I. Um, T. F. George and L. N.
Pandey, Phys. Rev. A 48, 2716 (1993); Nuovo Cimento
111, 963 (1996).
[63] H. R. Lewis, Jr., J. Math. Phys. 9, 1976 (1968).
[64] H. R. Lewis, Jr. and W. B. Riesenfeld, J. Math. Phys.
10, 1458 (1969).
[65] D. C. Khandekar, S. V. Lawande and K. V. Bhagwat,
Path-Integral Methods and Their Applications (World
Scientific, Singapore, 1993).
[66] G. Junker, A. Inomata and P. Wang, Phys. Lett. 110A,
195 (1985).
[67] H. J. Korsch, Phys. Lett. 74A, 294 (1979).
[68] H. Kohl and R. M. Dreizler, Phys. Lett. 98A, 95 (1983).
[69] D. C. Khandekar and S. V. Lawande, J. Math. Phys.
20, 1870 (1979).

-615-

[70] H. R. Lewis, Jr., Phys. Rev. 172, 1313 (1968).


[71] C. C. Gerry, Phys. Lett. 109A, 149 (1985).
[72] C. I. Um, K. H. Yeon, T. F. George and L. N. Pandey,
Phys. Rev. A 54, 2707 (1996).
[73] R. J. Glauber, Phys. Rev. 131, 2766 (1963).
[74] R. B. Levien, M. J. Collett and D. F. Walls, Phys. Rev.
A 47, 2324 (1993).
[75] K. H. Gheri, C. Saavendra and D. F. Walls, Phys. Rev.
A 48, 1532 (1993).
[76] P. D. Drummond, J. Phys. A 13, 2353 (1980); Phys.
Rev. A 43, 6194 (1991).
[77] Z. Ficek and P. D. Drummond, Phys. Rev. A 43, 6247
(1991).
[78] K. H. Yeon, H. J. Kim, C. I. Um, T. F. George and L.
N. Pandey, Phys. Rev. A 50, 1035 (1994).
[79] C. I. Um, I. H. Kim, S. K. Hong, K. H. Yeon and D. H.
Kim, J. Korean Phys. Soc. 30, 1 (1997).
[80] M. B tticker, H. Thomas and A. Prtre, Phys. Lett. A
180, 364 (1993).
[81] G. Mahler and V. A. Weberrub, Quantum Networks
(Springer-Verlag, 1995).
[82] F. A. Buot, Phys. Rep. 234, 73 (1993).
[83] R. G. Garcia, Appl. Phys. Lett. 60, 1960 (1992).
[84] S. Zhang, J. R. Choi, C. I. Um and K. H. Yeon, Phys.
Lett. A 289, 257 (2001); S. Zhang, C. I. Um and K. H.
Yeon, Chin. Phys Lett. 19, 985 (2002).
[85] S. Zhang, J. R. Choi, C. I. Um and K. H. Yeon, Phys.
Lett. A 294, 319 (2002); S. Zhang, J. R. Choi, C. I. Um
and K. H. Yeon, J. Korean Phys. Soc. 40, 325 (2002).
[86] G. Crespo, A. N. Proto, A. Plastino and D. Otero, Phys.
Rev. A 42, 3608 (1990).
[87] K. Husimi, Prog. Theor. Phys. (Kyoto) 9, 381 (1953).
[88] K. H. Yeon, D. F. Walls, C. I. Um, T. F. George and
L. N. Pandey, Phys. Rev. A 58, 1765 (1998).
[89] R. Shankar, Principles of Quantum Mechanics
(Plenum, New York, 1994).
[90] V. V. Dodonov, I. A. Malkin and V. I. Manko. Teor.
Mat. Fiz. 24, 164 (1975).
[91] K. H. Yeon, D. H. Kim, C. I. Um, T. F. George and L.
N. Pandey, Phys. Rev. A 55, 4023 (1997).
[92] C. I. Um, S. M. Shin, K. H. Yeon and T. F. George,
Phys. Rev. A 58, 1574 (1998).
[93] W. Dittrich and M. Reuter, Classical and Quantum Dynamics (Springer-Verlag, Berlin, 1993).
[94] L. Chetouani, L. Guechi and T. F. Hamman, Phys. Rev.
A 40, 1157 (1989).
[95] J. Y. Lee, K. L. Liu and C. F. Lo, Phys. Rev. A 58,
3433 (1998).
[96] R. S. Kaushal and D. Parashar, Phys. Rev. A 55, 2610
(1997).
[97] M. Maanache, K. Bencheikh and H. Hachemi, Phys.
Rev. A 59, 3124 (1999).
[98] T. Toyoda and S. Wakayama, Phys. Rev. A 59, 1021
(1999).
[99] Bo Gao, Phys. Rev. A 58, 1728 (1998).
[100] I. A. Pedrosa, Phys. Rev. A 55, 3129 (1997).
[101] B. D. Simons, P. A. Lee and B. L. Altshula, Phys. Rev.
Lett. 72, 64 (1994).
[102] T. Rentzsch, W. Schmidt, J. A. Maruhn, H. Stoecker
and W. Greiner, in Proceedings of Gross Properties of
Nuclei and Nuclear Excitations (Hischegg, 1988), p.
149.
[103] M. Bleicher, M. Reiter, A. Dumitru, J. Brachmann, C.

-616-

[104]
[105]
[106]
[107]

Journal of the Korean Physical Society, Vol. 41, No. 5, November 2002
Spieles, S. A. Bass, H. Stocker and W. Greiner, Phys.
Rev. C 59, 1844 (1999).
H. N. Nagashima, R. N. Onody and R. M. Faria, Phys.
Rev. B 59, 905 (1999).
C. W. Gardner, Quantum Noise (Springer-Verlag, New
York, 1991), Chap. 3.
S. L. Sondhi, S. M. Girvin, J. P. Carini and D. Shahar,
Rev. Mod. Phys. 69, 315 (1997).
A. Wolter, P. Rannou and J. P. Travers, Phys. Rev. B
58, 7637 (1998).

[108] A. Gold, Z. Phys. B 83, 499 (1991); ibid B 81, 155


(1990).
[109] A. P. Betenev and V. V. Kurin, Phys. Rev. B 56, 7855
(1997).
[110] E. Ott, Chaos in Dynamical Systems (Cambridge University Press, Cambridge, 1993).
[111] J. R. Ackerhalt, P. W. Milonni and M.-L. Shih, Phys.
Rep. 128, 205 (1985).

You might also like