Sunteți pe pagina 1din 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 Hamil-
tonian without violating the Heisenbergs uncertainty principle. Through the dynamical invariant
together with the path integral, we also present systematically the exact quantum theories for var-
ious 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
I. INTRODUCTION
Since Bateman proposed the time-dependent Hamil-
tonian 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 sys-
tems, it is essential to introduce a time-dependent Hamil-
tonian which describes the damped oscillation, i.e., the
Caldirola-Kanai Hamiltonian. This was discovered rst
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 signicant diculties in obtaining the
quantum-mechanical solutions for the Caldirola-Kanai
Hamiltonian. It is conrmed 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 Caldirola-
Kanai Hamiltonian. The theory guarantee that all the
fundamental laws in quantum mechanics would be satis-
ed.
In this paper, we review the phenomenological ap-
proach 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: cium@korea.ac.kr; Fax : +82-62-970-2204


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 canon-
ical, and unitary transformations for the classical, and
the quantum dissipative systems are presented in Sec.
VI. Finally, we provide a summary in Sec. VII.
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 Hamilto-
nian formalisms [11]. In 1931, Bateman [1] presented the
rst investigation of the so-called dual or mirror-image
Hamiltonian. His Hamiltonian consisted of two dierent
time-dependent Hamiltonians: One represented the sim-
ple one-dimensional damped harmonic oscillator. The
energy dissipated by the oscillator was completely ab-
sorbed 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 expo-
nentially growing mass, that could be reproduced as the
so-called Caldirola-Kanai Hamiltonian. Batemans dual
Hamiltonian is given as
H = p p (x p xp) +w
2
x x, (1)
-594-
Quantum Theory of the Harmonic Oscillator in Nonconservative Systems Chung-In Um and Kyu-Hwang Yeon -595-
where x is the mirror variable corresponding to the vari-
able x. The Lagrangian corresponding to Eq. (1) is
L = x

x
2
x x +(x

x x

x), (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 = 0. (4)
Hence,

x = dx/d(t).
Equation (4) clearly represents the time reversal pro-
cess 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 eas-
ily obtained. With the use of these operators, and the
Baker-Hausdor relation [12,13], we may reduce Eq. (1)
to a simpler form, and then nd the eigenstates, and
eigenvalues for the Hamiltonian of Eq. (1). However,
the time-dependent uncertainty product obtained in this
method,
(p
x
x)
2
=
_

2
_
2
e
4t
_
1 + 4
_

_
2
_

_
2
sin
4
t
_
,
(5)
tends obviously to zero. On the other hand, Eq. (5) vi-
olates Heisenbergs principle for =0, regardless of how
small the frictional coecient is. Therefore, Batemans
dual Hamiltonian describes classical mechanics correctly,
but does not follow the fundamental principles of quan-
tum 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 gener-
alized nonconservative force Q can be written as
d
dt
_
T
q
_

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

= (t), dt = (t)dt

, (t) = e

t
0
(t)dt
, (7)
and assume the rth component of Q to have the form
Q
r
= (t)
T
q
r
= (t)
s

k=1
a
rk
q
k
, (8)
where (t) is an arbitrary function, and a
rk
are con-
stants. We may construct the Schr odinger equation for
a nonconservative system from the classical Hamiltonian
through Eq. (7):
H

= i

=
_


2
2m

2
+V

_
. (9)
Transforming again the nonlinear time t

into ordinary
time t in Eq. (9), we obtain a single-particle Schr odinger
equation:
i

t
=
_


2
2m
e

0
t

2
+e

0
t
V
_
, (10)
where (t) has been taken as a time-independent con-
stant
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

] = ie

0
t
, (11)
xp
x


2
e

0
t
. (12)
As time goes to innity, the uncertainty relation van-
ishes; thus, this formalism violates a fundamental prin-
ciple in quantum physics.
In 1948, Kanai [3] derived the Caldirola-Kanai Hamil-
tonian from Batemans dual Hamiltonian by applying to
canonical transformations through several canonical gen-
erators. The transformation leads to the Caldirola-Kanai
Hamiltonian
H =
1
2
e
2t

2
+
1
2
e
2t

2
y
2.
. (13)
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 ex-
pressed by using the pseudostationary eigenstates

n
(x, t) =
_

_
1/4
_
1
2
n
n!
_
1/2
exp
_
i
_
n +
1
2
_
t +
1
2
t ( +i)e
2t
x
2
2
_
H
n
_
e
t
x
_
w

_
. (14)
Regarding the Hermite polynomial in Eq. (14), the in-
clusion of the e
t
term in the argument makes this quite
dierent 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 e
t
in the pseudostationary states; both are identical to each
other.
The pseudostationary states may be used to evalu-
ate the expectation values. By evaluating the quantum
-596- Journal of the Korean Physical Society, Vol. 41, No. 5, November 2002
uctuations for the proper mechanical momentum, and
canonical momentum, one nds the uncertainty relation
to be
(P
X
X)
2
= e
4t

2
_
n +
1
2
_
2
. (15)
The uncertainty product, Eq. (15) for the canonical mo-
mentum tends to zero as time goes to innity. 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 rst by Pryce,
and Stevens [16] adopted the Pryce model, and showed
that it was possible to introduce damping into the quan-
tum description of a harmonic oscillator that is made
of an LC circuit coupled to a semi-innite 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 innitely many
oscillators.
On the other hand, the damped equation of motion
for a LC circuit,
L
d
2
Q
dt
2
+R
dQ
dt
+
Q
C
= 0, (16)
can be derived from the Hamiltonian
H = e
Rt/L
P
2
2AL
+e
Rt/L
AQ
2
2C
, (17)
where the momentum P, and the coordinate Q corre-
spond 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 un-
certainty relation articially as time goes to innity:
[p
x
()X()]
2
=
_

2
_
2
coth
2
_

2k
B
T
_


2
4
.
(18)
Within the framework of geometric quantization, De-
dene [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
z =
1

2
( p ix), z =
1

2
( p +i x), (19)
where z means a formal complex mirror conjugation.
The dynamical variables z, and i z correspond to a canon-
ically paired coordinate, and momentum [ z = z

= z

].
Substitution of Eq. (19) in the Bateman dual Hamilto-
nian yields Dedenes Hamiltonian
H = h +h

(20)
h = (w i)z z. (21)
In agreement with Dedene [21], z, and z are the annihi-
lation, and the creation operators in a generalized Her-
mitian 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
+
, (22)
h

=
_
zz +
1
2

_
i
_
zz +
1
2

_
,
h
+
=
_
z

+
1
2

_
+i
_
z

+
1
2

_
, (23)
where =1, and =0 exhibit the Weyl, and the normal or-
derings [12], respectively. The eigenvalues of the Hamil-
tonian are given by
h
()
n
=
_
n +
1
2

_
i
_
n +
1
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 con-
ventional commutator form, and thus the general func-
tions 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 vio-
lated again. The main aw in the complex Hamiltonian
comes from the incorrect fundamental commutation re-
lation. 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

n,m
=
_
P(
0
)
(0)
nm
d
2

0
, (25)
where P() is a quasi-probability density, and
(0)
nm
is
the density operator. Equation (25) satises the master
equation. Using Eq. (25), one obtains the expectation
value for the uncertainty relation:
(X)
2
=

2
e
2t
_
1 +

sin2t + 2

2
sin
2
t
_
+

2
(1 +e
2t
), (26)
Quantum Theory of the Harmonic Oscillator in Nonconservative Systems Chung-In Um and Kyu-Hwang Yeon -597-
(P
X
)
2
=

2
e
2t
_
1

sin2t + 2

2
sin
2
t
_
+

2
2
(1 +e
2t
). (27)
The rst 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 nding 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
x
2
+
3
p +
4
p
2
+
5
xp +

5
px, (28)
where
i
=
i(t)
. The Hamiltonian of Eq. (28) satises
the Schrodinger equation
i
K
t
= HK. (29)
If the Hamiltonian is quadratic, then the propagator [29]
can be written as
K(x, t) = F(t) exp
_
i

S
c
(x, t)
_
, (30)
where F(t) is a factor depending only on the time inter-
val, and S
c
(x, t) is the classical action. Substituting Eq.
(30) into Eq. (29) under the assumption
S
t
+H
_
x,
S
t
, t
_
= 0, (31)
together with
S(x, t) = a(t) +b(t)x +c(t)x
2
, (32)
one obtains coupled rst-order dierential equations for
a(t), b(t), c(t), and
i
(t) from the Hamilton-Jacobi equa-
tion. When the
i
is not time-dependent, the time-
dependent solutions for a(t), b(t), and c(t) can be ob-
tained. For the simple harmonic oscillator, we get
0
=

1
=
3
=
5
= 0,
2
= mw
2
/2, and
4
= 1/2m.
For convenience, let us write the Caldirola-Kanai
Hamiltonian, Eq. (10), as
h(t) = e
2t
p
2
2m
+e
2t
1
2
m
2
x
2
, (33)
which yields the linearly damped equation, Eq. (3). In
this case, we have
0
=
1
=
3
=
5
= 0,
2
=
e
2t
m
2
/2 and
4
= e
2t
/2m. Then, we obtain
a(t) =
1
2
mx
2
cos t/ sint +
1
2
x
2
, (34)
b(t) = mx

exp(t)/ sint, (35)


c(t) =
1
2
me
2t
(cos t sint)/(sint), (36)
F(t) = e
t/2
[m/(2ihsin)]
1/2
,
= (
2

2
)
1/2
. (37)
Setting =0, Eqs. (34)-(37) are reduced to those of the
simple harmonic oscillator. We will determine the valid-
ity 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 har-
monic 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 Caldirola-
Kanai Hamiltonian by using a modied Feynman path
integral in conguration 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 athy-
Feynman formula [25].
The propagator can be expressed as a path integral in
phase space as
K(q

, q

, T) =
_
exp
_
1

(p q H(p, q)dt)
_
DpDq (38)
where H(p, q) is the Hamiltonian of the system consid-
ered, and DpDq is the two-dimensional path dieren-
tial measure in phase space. One obtains the propaga-
tor for the damped harmonic oscillator described by the
Caldirola-Kanai Hamiltonian as
K[q

, q

; T] =
_
me
(t

+t

)/2)
2i sint
_
1/2
exp
__

m
4i
_
(e
t

q
2
e
t

q
2
)
_
exp
__
im
2 sint
_
[(e
t

q
2
+e
t

q
2
) cos T 2e
(t

+t

)/2
q

]
_
. (39)
In the case of a bound system, the propagator is ex- 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

, t

; q

, t

) =

n=0

n
(q

, t

n
(q

, t

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

n(q,t)
= N
n
exp(iE
n
t/)
exp[(m/2)(1 +i/2)e
t
q
2
]
H
n
[(m/)
1/2
]e
t/2
q] (41)
with the energy eigenvalues
E
n
=
__
n +
1
2
_
+
1
4

_
, n = 0, 1, 2, , (42)
where the normalization factor is N
n
=
(m/)
1/4
(2
n
n!)
1/2
and H
n
(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 the-
ories for the dissipation of a linear damped harmonic
oscillator in quantum mechanics. Besides these theo-
ries, there are many other theories, such as the modi-
ed Bopp-Dekker master equation for the reduced den-
sity operator in the treatment of the quantum optics
oscillator [3537], the nonlinear frictional quantum the-
ory 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 vari-
ation procedure [4,41], and so on. We should also point
out that none of the theories for this linear damped har-
monic 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] oers a general approach for treating quantum-
mechanical systems, only a few time-dependent
Schr odinger equations can be solved exactly. Most of the
previous theories for the Caldirola-Kanai Hamiltonian vi-
olate the uncertainty relation. This diculty is critically
reviewed by Dodonov-Manko [41], and others [42]. In
this section, we discuss the quantum-mechanical solu-
tion (Um-Yeon solution) of the Caldirola-Kanai Hamil-
tonian for this dissipative system by using the prop-
agator method developed by the present authors [9].
We introduce the Caldirola-Kanai Hamiltonian with a
time-dependent external driving force f(t), dened as
the time-dependent damped driven harmonic oscillator
(DDHO), as
H = e
t
p
2
2m
+e
t
_
mw
2
0
x
2
2
xf(t)
_
. (43)
Hamiltons equations of motion for Eq. (43) yields the
Lagrangian
L = e
t
_
1
2
m x
2

1
2
mw
2
0
x
2
+xf(t)
_
, (44)
and the corresponding equation of motion is
x + x +w
2
0
= f(t)/m. (45)
The mechanical energy can be expressed as
E = e
2t
p
2/2m
+
1
2
mw
2
0
x
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 in-
tegral equation with the propagator K:
(x, t) =
_
K(x, t; x
o
, 0)(x
0
, 0)dx
0
(47)
which provides the wavefunction (x, t) at time t in
terms of the wavefunction (x
0
, 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; x
0
, 0) =
_
mwe
t/2
2i sinwt
_1/2
exp
_
i

Sc(x, x
0
, t)
_
.
(48)
The classical action of the DDHO Hamiltonian is
Sc =
_
e
t

_
1
2
m x
2

1
2
mw
2
0
x
2
+xf(t

)
_
dt

. (49)
In Eq. (49) for small w
0
, 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 -599-
K(y, p; y
0
, 0) = exp
_
a(p) exp
_
i
2
w
0
p
_
y
2
+b(p) exp
_
i

w
0
p
_
y +c(p)
_
, (50)
where we have changed the variables as
p =
w
0
2i
t, y =
_
mw
0

x. (51)
Substituting Eqs. (43), and (51) into Eq. (29), we can
determine the time-dependent coecients in Eq. (50).
To simplify the expression, we take f(t)=0 in Eq. (43);
then, the propagator can be expressed as
K(y, p; y
0
, 0) = F(0) exp
_
im
2
ax
2
+ ebx
2
0
2 + 2 cx
0
x
_
,
(52)
where the new time-dependent coecients are
a =
_

1
2
+wcot wt
_
e
t
, (53)

b =
_
1
2
+wcot wt
_
, (54)
c =
w
sinwt
e
t/2
, (55)
F(t) =
_
mww
t/2
2i sinwt
_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 har-
monic oscillator. Then, the corresponding wave function

n
(x, 0), and the energy eigenvalues are given by

n
(x, 0) = N
0
H
n
(
0
x) exp
_

1
2

0
x
2
_
, (57)
E
n
=
_
n +
1
2
_
w
0
. (58)
From Eq. (47) together with Eqs. (43), (52), and (57),
we obtain the wavefunction at time t in the form

n
(x, t) =
_

K(x, t; x
0
, 0)
n
(x
0
)dx
0
= N
1
(2
n
n!)
1/2
expi
__
n +
1
2
_
cot
1
_

2w
+ cot wt
_
_
exp[Ax
2
]H
n
[D(x)]. (59)
Here, the time-dependent coecients are given by
N =
_
mw

_
1/4
exp
_
1
4
t
_
(t)(sinwt)
1/2
, (60)

2
(t) =

2
4
2
+

cos t + cos ec
2
t, (61)
A(t) =
m
2h
e
t
_
1
(t)
2
sin
2
t
+i
_

2
cot t +
/2 + cot t
(t)
2
sin
2
t
__
, (62)
D(t) =

0
e
t/2
(t) sint
. (63)
The mechanical energy, Eq. (47), can be expressed
as the energy operator E, whose expectation values take
the form
< E >
mn
=

2
2m
e
2t
_

2
x
2
_
mn
+
1
2
m
2
0
< x
2
>
mn
.
(64)
The evaluations of the expectation value for x
2
, and /
are straightforward. Here, we write only the diagonal
element of < E >
mn
as follows:
< E >
nn
=
_
n +
1
2
_

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

_
1 +
__

2
cot t
_
(t)
2
sin
2
t
+
_

2
+ cot t
__
2
_
1/2
.
(66)
When f(t) ==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)] re-
quires 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 E
n,n
(t) =< E >
nn
[(65)] for
/2 = 0.1.
[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 e
t
, and thus decay exponentially.
The second o-diagonal elements, E
n+2,n
, not shown
here, depend only on the exponential decay constant
. Figure 1 illustrates the decay of the energy eigen-
value, E
nn
(t). Note that the results of Dodonov, and
Manko [43] can be obtained by taking the driving force
as f(t) = f
0
sin(t +).
The uncertainty for the (n, n) states with period [Eq.
(66)] is reduced to that of the harmonic oscillator at 180

,
and 0

. The uncertainty for the (n, n) states is


_
n +
1
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 satised. One
should recognize that every fundamental law in quantum
theory is guaranteed in this Um-Yeon solution.
2. Coherent States
Coherent states for the harmonic oscillator have been
studied, and are widely used to describe many elds 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 de-
scribed by the Caldirola-Kanai Hamiltonian.
Here, we dene the creation, and the annihilation op-
erator 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 uni-
tary oprerator, (iii) they represent the minimum uncer-
tainty states, and (iv) they are not orthogonal but com-
plete, and normalized. To obtain these operators, we will
make use of the propagator in Eq. (59).
Fig. 2. Uncertainty relation as the (n, n) state oscillates
with period [Eq. (66)].
The coherent states can be dened by the eigenstates
of the non-Hermitian operator a i.e., a| > | >. Us-
ing the completeness relation for the number representa-
tions, we can expand | > as
| e
(1/2)||
2

n=0

n!
|n >= e
(1/2)||
2
e
a

|0 >, (67)
where |0 > is the vacuum or ground state, and is inde-
pendent of n. The calculation of < | > in Eq. (67)
gives
< | e
1/2(||
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
_
| >< |
d
2

= 1. (69)
To dene 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
:
< x >
mn
=
_
n +
1
2
_
1/2
(t)
m,n+1
+n
1/2
(t)
m,n1
,
(70)
< p >
mn
=
_
n +
1
2
_
1/2
(t)
m,n+1
+n
1/2

(t)
m,n1
,
(71)
where
(t) =
1
2
(Re A)
1/2
exp
_
i cot
1
_

2
+ cot(t)
__
,
(72)
(t) =

2i
A
D
exp
_
i cot
1
_

2
+ cot(t)
__
, (73)
Quantum Theory of the Harmonic Oscillator in Nonconservative Systems Chung-In Um and Kyu-Hwang Yeon -601-
with the relation

= i. Then, the creation,


and the annihilation operators for the damped harminic
oscillator can be dened as
a =
1
i
(x p), (74)
a

=
1
i
(

x). (75)
We can easily conrm that a, and a

are not Hermi-


tian operators, but the following relations are preserved:
[x, p] = i, and [a, a

]=1. We can evaluate the transfor-


mation function < x| > from the coherent states to the
coordinate representation |x > from Eqs. (74) as
< x| (2

)
1/4
exp
_
1
2i

x
2
+

x
1
2
||
2
1
2


2
_
. (76)
Finally, we can show that a coherent state represents
a minimum uncertainty state. It is straightforward to
evaluate the expectation values of x, p, x
2
, and p
2
in
state | >. We obtain
(x)
2
=< x
2
> < x >
2
=

,
(p)
2
=< p
2
> < p >
2
=

. (77)
Therefore, the uncertainty relation becomes
(x)(p) = ||
2
||
1/22
=

2
(t), (78)
(t) =
_
1 +
__
1
8
_

_
3
+
_

__
sin
2
(t) +
1
8
_

_
2
sin(2t)
_
2
_
1/2
. (79)
Equation (78) is obviously the minimum uncertainty cor-
responding to Eqs. (66) in the (0,0) state. The uncer-
tainty for the (n, n) state [Eq. (78)] oscillates with the
period , which corresponds to the half period of a sim-
ple harmonic oscillator. From all of the above, we con-
clude 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 espe-
cially, 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 innite square-well potential with a con-
stant electric eld [53], a driven coupled harmonic oscil-
lator, and a coupled damped driven harmonic oscillator
[54].
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 investi-
gated by many authors [48,55,56]. Camiz et al. [57], and
Khandekar, and Lawande [33] have obtained the wave-
functions 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 time-
dependent frequency through the path integral method
[58,59].
Consider a forced harmonic oscillator with a time-
dependent frequency (t), whose Hamiltonian is of the
form
H =
p
2
2m
+
m
2
w
2
(t) f(t)x (80)
where f(t) is an external driving force. The correaspond-
ing Lagrangian is
L =
1
2
m x
2

1
2
m
2
(t)x
2
+f(t)x, (81)
with the classical equation of motion
dx
2
dt
2
+
2
(t)x =
1
m
f(t). (82)
For the case (t) =
0
, the solution of Eq. (82) repre-
sents harmonic motion; otherwise, it is not easy to eval-
uate the exact solution.
Since the Lagrangian is quadratic, we can assume the
propagator to have the form [9,54,60]
K(x, t; x

, t

) = exp[a(t, t

)x
2
+b(t, t

)xx

+c(t, t

)x
2
+g(t, t

)x
+h(t, t

)x

+d(t, t

)]. (83)
Substituting Eqs. (80), and (83) into Eq. (29), we can
determine the time- dependent coecients in the propa-
gator. The propagator for a forced time-dependent har-
monic oscillator, Eq. (83), be expressed as
K(x, t; x

, t

) =
_
m( )
1/2
2i sin(

)
_
exp
_
im
2
_

x
2

x
2
__
exp
_
im
2 sin(

)
[( x
2
+

x
2
)
cos(

) 2
_

xx

+
2


m
x

_
t
t

ds
f(s)
_
(s)
sin[(s) ]
+
2


m
x

_
t
t

ds
f(s)
_
(s)
sin[ (s)]
-602- Journal of the Korean Physical Society, Vol. 41, No. 5, November 2002

1
m
2
_
t
t

ds
f(s)
_
(s)
sin[ (s)]

_
t
t

dp
f(p)
_
(p)
sin[(p) ]
__
. (84)
Here, the relation between (t), and (t) in Eq. (84) is
given by
q(t) = (t)e
i(t)
, (85)
and the function q(t) satises the dierential equation
d
2
q(t)
dt
2
+
2
(t)q(t) = 0. (86)
Here, (t), and (t) are real quantities. From Eqs. (85)-
(86), the real, and the imaginary parts of the dierential
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 t

, and t, respec-
tively. 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 os-
cillator. We can obtain the wave function directly from
the denition of the propagator
K(q

, t

; q

, t

) =

n=0

n
(q

, t

n
(q

, t

) (88)
and from Mehlers formula [61]
exp[(X
2
+Y
2
2XY )/(1 Z
2
)]

1 Z
2
= exp
(X
2
+Y
2
)

n=0
Z
n
2
n
n!
H
n(X)
H
n(Y )
. (89)
Comparison of the result of Eq. (84) with Eq. (88)
through Eq. (89) gives the wavefunction as

n
(x, t) = exp
_
i
_
(t)
_
n +
1
2
_
(t)
__

n
(x, t),
(90)
where

n
(x, t) =
_
1
2
n
n!
_

_
1/2
_
1/2
exp
_
im
2
_

x
2
m
_
x
_
t
ds
f(s)
_
(s)
cos[ (s)]
__
exp
_
_
_
m
2
_
_
x
1
m
_
t
ds
f(s)
_
(s)
cos[ (s)]
_
2
_
_
_
H
n
_
_
m

_
_
x
1
m
_
t
ds
f(s)
_
(s)
sin[ (s)]
__
.
(91)
The wave function, Eq. (90), is simply a unitary trans-
formation of
n
(x, t), where
n
(x, t) satises all the prop-
erties 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
E
op
=

2
2m

2
x
2
+
m
2
(
2
+
2
)x
2


2
_
2x

x
+ 1
_
+ (

)
_
i

x
+m x
_
+m
2
x +
m
2

2

2
+
m
2
(

)
2
. (92)
Equation (91) can be simplied to the form

n
(x, t)
=
_

2
n
n!

_
1/2
e
ix
2
+ix
e

1
2

2
(x)
2
H
n
[(x )]
=
_

2
n
n!

_
1/2
e
Ax
2
+Bx
H
n
[(x )], (93)
where the new coecients are
=
_
m

_
1/2
,
(t) =
m
2

,
(t) =
1

_

_
t
ds
f(s)
_
(s)
cos[ (s)],
(t) =
1
m


_
t
dsf(s)
_
(s) sin[ (s)],
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
E
n,n
= E
n
=

2
(
2
)(2n + 1) =
_
n +
1
2
_
, (95)
(xp)
n,n
=
_
1 +

2

2

2
_
1/2
_
n +
1
2
_
(96)
The energy expectation value is obviously a time-
invariant 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 -603-
To obtain the coherent states, we construct a creation
operator a

, and annihilation operator a for the time-


dependent harmonic oscillator.
a

=
_
m
2
_
1/2
__
1 +i


_
x i
p
M
_
, (97)
a =
_
m
2
_
1/2
__
1 +i


_
x +i
p
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
xp =

2
_
1 +
_

eta

_
2
_
1/2
, (99)
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 in-
variant. For this reason, we have used the time-invariant
operators, and derived the energy operator from the clas-
sical equation of motion. We then used the energy op-
erator to calculate the energy expectation values. This
energy operator is very similar to the Ermakov-Lewis in-
variant 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 the-
ory of a general time-dependent bound, and unbound
quadratic Hamiltonian system [62] through the dynam-
ical invariant, and path-integral methods. We nd the
relation between the quantum mechanical solution, and
the dynamical invariant [6366].
1. Time-dependent Bound Quadratic Hamilto-
nian
One may consider a system with a time-dependent
quadratic Hamiltonian of the type
H =
1
2
[A(t)p
2
+B(t)(xp +px) +C(t)x
2
], (100)
where A(t) is a nonzero time-dependent function, and
B(t), and C(t) are time- dependent functions of an ar-
bitrary form, which are continuously dierentiable with
respect to time. From Hamiltons equations of motion,
the classical equation of motion becomes
x

A(t)
A(t)
x +
_
A(t)C(t) +

A(t)B(t)
A(t)
B
2
(t)

B(t)
_
x = 0. (101)
However, the solution with a general form for arbitrary
time-dependent coecients 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)e
i(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 rst 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 aux-
iliary 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 nd 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
2
_
(t)p
2
+ 2(t)xp +(t)x
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
dt
=
I
t
+ [I, H]. (108)
Combining Eqs. (100), and (107) with (108), we can de-
termine the time-dependent coecients, (t), (t), and
(t). Finally, one obtain the invariant quantity
I(t) =
_
_

x
_
2
+
__
B(t)
A(t)

1
A(t)

_
x +p
_
2
_
. (109)
-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 propaga-
tor for a quadratic Hamiltonian has the following form
of Eq. (103):
K(x, t; x

, t

)
= exp
_
i

[a(t, t

)x
2
+b(t, t

)xx

+c(t, t

)x
2
d(t, t

)]
_
.
(110)
Instead of Eq. (110), we can introduce the denition of
the propagator for the bound system given by Eq. (88).
For the unbound system, the propagator is given as
K(x, t; x

, t

) =
_
dk
k
(x, t)

k
(x

, t

). (111)
The above two propagators, Eqs. (88), and (111), must
satisfy the Schr odinger equation, Eq. (29), and its com-
plex conjugate. Substituting Eqs. (100), and (110) into
Eq. (29), we can determine the time-dependent coe-
cients, a(t), b(t), and c(t). From the auxiliary condition,
i.e., the classical solution, if (t)e
i(t)
is that solution,
then (t)e
i(t)
is also a solution of that system. The
classical solution can be written as
q(t, t

) =

sin(

). (112)
With the use of Eq. (112) together with the coecients
a(t), b(t), and c(t), we can write the propagator, Eq.
(110), as
K(x, t; x

, t

) =
_

1/2

1/2
2i sin(

)A
1/2
A
1/2
_1/2
exp
_
i
2A
_

+ cot(

) B
_
x
2
+
i
2A

cot(

) +B

_
x
2
+
i

AA

_
1/2
xx

sin(

)
_
. (113)
Comparing Eq. (113) with Mehlers formula, Eq. (89)
through Eq. (88), we nd the wavefunction of the sys-
tem:

n
(x, t) =
_
1


A
_
1/4
_
1
2
n
n!
_
1/2
e
i[(1/2)+n]
H
n
_
_
1


A
_
1/2
x
_
exp
_

1
2A
_
i
_

B
__
x
2
_
. (114)
Equation (114) is the wavefunction of the bound sys-
tem with the auxiliary condition of the classical solution.
The uncertainty relation becomes
(xp)
n,n
=
_
n +
1
2
_

_
1 +
1

2
_

B(t)
_
2
_
1/2
.
(115)
Note that the diagonal element of the uncertainty re-
lation 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 dened as
I =
1
2
__

2
+1
A
2(t)
[B(t)
2
]
2
_
x
2
+

A(t)
[B(t) ](xp +px) +
2
p
2
_
. (116)
I(t) should satisfy the quantum condition that corre-
sponds to the classical condition, Eq. (88),
dI
dt
=
I
t
+
1
i
[I, H] = 0, (117)
with the Hamiltonian given by Eq. (100). The expecta-
tion values of the quantum invariant operator, Eq. (116),
are given by
< m|I|n > =
_
n +
1
2
_
h

2

A

m,n
=
_
n +
1
2
_
h
m,n
. (118)
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 oper-
ator, 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 Hamil-
tonian
For convenience, we express the Hamiltonian of Eq.
(100) as
H =
1
2
[A(t)p
2
+B(t)(pq +qp) +C(t)q
2
], (119)
where q, and p are the canonical coordinate, and its con-
jugate 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 = C
1
(t) exp[(t)] +C
2
(t) exp[(t)], (121)
Quantum Theory of the Harmonic Oscillator in Nonconservative Systems Chung-In Um and Kyu-Hwang Yeon -605-
where C
1
, and C
2
are integral constants. Substituting
Eq. (121) into Eq. (120), we obtain two dierential
equations, one for e(t), and the other for (t). For e(t),
and (t) to satisfy these two equations simultaneously,
the following dierential equations must be satised:
+
2
+(t)

rho +(t) = 0, (122)
+ 2 +(t) = 0. (123)
Equation (123) oers the invariant quantity quantity
(t):
(t) =

2

A(t)
. (124)
One can nd another classical time-invariant quantity
from Eq. (120):
I(q, p, t) =
1
2
[(t)p
2
+ 2(t)qp +(t)q
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
I(q, p, t) =
_

_
2
+
__
B(t)
A(t)

1
A(t)

_
q +p
_
2
. (126)
The eigenvalue of the coecient matrix of the
quadratic variable in Eq. (126) is given as
=
1
2
_

2
+
_

_
2
+
_
B(t)
A(t)

1
A(t)

_
2
_
. (127)
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 dierential equa-
tions for the coecients 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 =

sinh(r r

). (128)
Solving the coupled dierential equations for the co-
ecients, together with Eq. (128), we obtain the time-
dependent coecients; thus, the propagator for a system
with an unbound time-dependent quadratic Hamiltonian
is given by
K(q, t; q

, t

)
=
_

1/2

1/2
2iA
1/2
A
1/2
sinh(r r

)
_
1
2
exp
__
i
2A(t)
_

+ctg(r r

)
_

i
2A(t)
B(t)
_
q
2
_
exp
__
i

sinh(r r

)
_
qq

_
exp
__
i
2A(t)
_

ctg(r r

)
_

i
2
B(t

)
_
q
2
_
.
(129)
It is well known that the Hamiltonians for a free particle,
a damped free particle, and an overdamped, an under-
damped, 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 de-
termine whether or not the system is unbound. The
propagator, Eq. (129), for the unbound system with a
quadratic Hamiltonian enables us to nd the propagators
for several unbound systems.
VI. QUANTUM DAMPED HARMONIC
OSCILLATOR-DYNAMICAL INVARIANT
In the treatment of a time-dependent quantum sys-
tem, Lewis, and Riesenbeld [63, 64] introduced the dy-
namical invariant method to obtain the exact quantum-
mechanical 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 quan-
tity 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]
g(t) =
_
t

dt

(t t

)f(t

), (130)
where f(t) is an instantaneous force at a given time t,
and (t t

) satises the properties of the -function.


Then, the Hamiltonian of a damped harmonic oscillator
with the above force has the form
H( p, q, t) = e
t
p
2
2m
+e
t
_
1
2
m
2
0
q
2g
(t) q
_
. (131)
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
L( q, q, t) = e
t
_
1
2
m q
2
1
2
m
2
0
q
2
+g(t)q
_
, (132)
-606- Journal of the Korean Physical Society, Vol. 41, No. 5, November 2002
q + q +
2
0
q =
g(t)
m
. (133)
Through the same procedure as in Sec. V, we can ob-
tain an invariant operator, I(p, q, t), that satises Hamil-
tons equation, Eq. (117), within the assumption of a
quadratic form for p, and q:

I =
1
2
e
t
_
m
2
_

2
0


2
4
_
( q q
0
)
2

_
e
t
( p +p
0
)
2
+
1
2
m( q q
0
)
2
__
, (134)
where p
0
(t) = me
rt
q
0
(t), and q
0
(t) is the solution to the
dierential equation
q
0
+ q
0
+
2
0
q
0
(t) =
g(t)
m
. (135)
From Eq. (135), q
0
(t) may be regarded as a particular
solution of Eq. (133):
q
0
(t) =

m
d
_
t
t
0
dt

_
t

t
0
dt

e
(tt

)/2
e
(t

)
sin[
d
(t t

)]f(t

), (136)
where
d
= (
2
0

2
/4)
1
2
. Note that the past aects
inuences in the time- dependent function q
0
(t):
The dynamical invariant operator can be written in
terms of the annihilation, and the creation operators as
follows:
a(t) =
_
1
2m
d
e

t
_
1/2

_
me
t
_

d
+i

2
_
[ q (t)] +i p
_
, (137)
a(t)

=
_
1
2m
d
e

t
_
1/2

_
me
t
_

d
i

2
_
[ q (t)] i p
_
, (138)
where
(t) =
_
q
0
(t) +

2
2
0
q
0
(t)
_
+i
_
1

2
4
2
0
_
1/2
q
0
(t)

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

I =
d
_
a

a +
1
2
_
. (140)
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 diers only by a time-dependent
phase factor from the eigenstate of the invariant opera-
tor [64], we can write directly the wavefunction,
n
(x, t)
as

n
(q, t) =
n
(q, t)e
(i/h)R(t)
. (141)
Then, the exact form of the wavefunction is given by

n
(q, t)=
1

n!2
n
_
m
d
e
t

_
1/4
exp
_

me
t

d
2
(q q
0
)
2
_
e
i(n+1/2)w
d
(tt

)
H
n
_
_
m
d
e
t

_
1/2
(q q
0
)
_
exp
_

me
t
2
_

2
(q q
2
0
) 2 q
0
(q q
0
) +
q
2
0
2
2
0
__
exp
_
i

_
t
t
0
dt

[L
0
(t

) (t

)]
_
. (142)
An exact closed form of the propagator can be obtained
readily from the expansion formula, Eq. (88):
K(q, t, q

, t

) =

n=0

n
(q, t)

n
(q

, t

) =
_
m
d
(e
t
e
t

)
1/2
2i sin
d
(t t

)
_
1/2
exp
_
im
2
_
e
t
_

d
cot
d
(t t

)

2
_
(q q
0
)
2
+e
t

d
cot(t t

) +

2
_
(q q
0
)
2
_
_
exp
_

im
d
(e
t
e
t

)
1/2
sin
d
(t t

)
(q q
0
)(q

0
)
_
exp
_
im

[e
t
q
0
(q q
0
) e
t

0
(q

0
)]
_
exp
_

_
m
4
2
0
(e
t
q
2
0
e
t

q
2
0
)
_
t
t
0
dt

[L
0
(t

) (t

)]
__
. (143)
With the use of the wavefunction, Eq. (142), we can obtain the uncertainty relations for various states, and
Quantum Theory of the Harmonic Oscillator in Nonconservative Systems Chung-In Um and Kyu-Hwang Yeon -607-
the uncertainty relation corresponding to the diagonal
element is given by
(qp)
n,n
=

2
(2n + 1). (144)
Note that the o-diagonal elements of the uncertainty
relations, (x, P)
n1,n
, are governed by the past
in terms of q
0
(t) and p
0
(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.
2. Harmonic Oscillator with a Time-dependent
Frequency
The generalization of the relation between the dynam-
ical invariant, and the solution of the Schr odinger equa-
tion for the time-dependent oscillator oers wide appli-
cations to various elds [7377]. Let us investigate the
quantum solutions of the harmonic oscillator with a time-
dependent 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
H =
p
2
2m
+
1
2
m
2
(t)q
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)e
i(t)
, (147)
where (t), and (t) are determinable from Eq. (146).
Substitution of Eq. (147) into Eq. (146) yields the real,
and the imaginary dierential equations:

2
+(t)
2
= 0, (148)
+ 2 = 0. (149)
From Eq. (149), an invariant quantity can be found:
= m
2
, (150)
with an auxiliary condition being given by the classical
solution. Another time-invariant quantity can be ob-
tained from Eq. (108). Through a similar calculation,
we obtain this classical invariant quantity as
I =
1
2
_

2
q
2
+ (p m
2
q)
2
_
. (151)
The invariant quantities , and I are measures of the
bound system. If is real, Eq. (151) is an elliptic equa-
tion in phase space. Thus, as the values of q, and p in the
system are limited in same region, it is a bound system.
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 eigenval-
ues of the invariant operator, we dene the annihilation,
and the creation operators a, and a

by the relations
a =
1

2m
_
m
_
1


_
q +ip
_
, (152)
a

=
1

2m
_
m
_
1 +


_
q ip
_
, (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
I =
_
aa

+
1
2
_
, (154)
and the eigenstates, and the eigenvalue spectrum of the
invariant operator are given as
aa

|n n|n >, n = 0, 1, 2,
=
_
n +
1
2
_
, n = 0, 1, 2, (155)
Applying the annihilation operator to the ground state,
and then solving this equation, we obtain the normalized
ground state as
u
0
=
_
m

_
1/4
exp
_

m
2
(1 i )q
2
_
. (156)
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
(q, t) =
_
1
2
n
n!
_
1/2
_
m

_
1/4
e
(1/2+n)
exp
_

m
2
_
1 i


_
q
2
_
H
n
_
_
m

q
_
1/2
_
. (157)
With the help of Mehlers formula, Eq. (89), the propa-
gator of the system is given as
K(q, t; q

, t

) =
_
m

2i sin(

)
_
1/2
exp
_
im
2
__

q
2

q
2
_
+
1
sin(

)
[( q
2
+

q
2
) cos(

) 2
_

qq

]
__
. (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
Fig. 3. Diagram of mesoscipic caacitance coupled circuit.
3. Mesoscopic Capacitance Coupled Circuit
The advent of mesoscopic physics [80], and nan-
otechnology [81] was stimulated, and encouraged by
a host of intriguing theories, such as localization in
lower dimension. The rapid development of nanotech-
nology, and nanoelectronics [82] has made it possible to
minimize integrated circuits, and components towards
atomic-scale dimensions [83]. In this section, we investi-
gate 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
L
1
d
2
q
1
dt
2
+R
1
q
1
+
q
1
C
1
+
q
1
q
2
C
= (t), (159)
L
2
d
2
q
2
dt
2
+R
2
q
2
+
q
2
C
2
+
q
1
q
2
C
= 0, (160)
where q
i
(i = 1, 2) stand for the electric charges in the
circuits, L
i
the inductances, R
i
the resistances, C
i
the
capacitances, and C the coupling capacitance. These
equations of motion can be formulated from the following
Hamiltonian [16]:
H
q
= e
t
p
2
1
2L
1
+e
t
q
2
1
2C
1
+e
t
p
2
2
2L
2
+e
t
q
2
2
2C
2
+e
t
(q
1
q
2
)
2
2C
, (161)
where we have assumed (t) = 0, and = R
1
/L
1
=
R
2
/L
2
for simplicity. Here, the electric charges q
i
substi-
tute for the conventional coordinates while their conju-
gation variables p
i
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
I =
1
2
_
2

i=1
_
e
t
p
2
i
L
i
+q
i
p
i
+
_
1
C
+
1
C
i
_
e
t
q
2
i
_

2
C
e
t
q
1
q
2
_
. (162)
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

I =
1
2
_
2

i=1
_
e
t
p
2
i
L
i
+

2
( q
i
p
i
+ p
i
q
i
)
+
_
1
C
+
1
C
i
_
e
t
q
2
i
_

2
C
e
t
q
1
q
2
_
. (163)
Here, we dene the time-dependent canonical creation,
and annihilation operators,

b
i
, and

b

i
, to obtain the
eigenfunctions, and eigenvalues of the invariant opera-
tor:

b
1
=
1
_
2
1

L
1
L
2
_

1
_
L
1
L
2
e

2
t
_
4
_
L
1
L
2
cos

2
q
1

4
_
L
2
L
1
sin

2
q
2
_
+i
_
e

2
t
_
4
_
L
2
L
1
cos

2
p
1

4
_
L
1
L
2
sin

2
p
2
_
+

2
_
L
1
L
2
e

2
t
_
4
_
L
1
L
2
cos

2
q
1

4
_
L
2
L
1
sin

2
q
2
___
, (164)

1
=
1
_
2
1

L
1
L
2
_

1
_
L
1
L
2
e

2
t
_
4
_
L
1
L
2
cos

2
q
1

4
_
L
2
L
1
sin

2
q
2
_
i
_
e

2
t
_
4
_
L
2
L
1
cos

2
p
1

4
_
L
1
L
2
sin

2
p
2
_
+

2
_
L
1
L
2
e

2
t
_
4
_
L
1
L
2
cos

2
q
1

4
_
L
2
L
1
sin

2
q
2
___
, (165)

b
2
=
1
_
2
2

L
1
L
2
_

2
_
L
1
L
2
e

2
t
_
4
_
L
1
L
2
sin

2
q
1
+
4
_
L
2
L
1
cos

2
q
2
_
Quantum Theory of the Harmonic Oscillator in Nonconservative Systems Chung-In Um and Kyu-Hwang Yeon -609-
+i
_
e

2
t
_
4
_
L
2
L
1
sin

2
p
1
+
4
_
L
1
L
2
cos

2
p
2
_
+

2
_
L
1
L
2
e

2
t
_
4
_
L
1
L
2
sin

2
q
1
+
4
_
L
2
L
1
cos

2
q
2
___
, (166)

2
=
1
_
2
2

L
1
L
2
_

2
_
L
1
L
2
e

2
t
_
4
_
L
1
L
2
sin

2
q
1
+
4
_
L
2
L
1
cos

2
q
2
_
i
_
e

2
t
_
4
_
L
2
L
1
sin

2
p
1
+
4
_
L
1
L
2
cos

2
p
2
_
+

2
_
L
1
L
2
e

2
t
_
4
_
L
1
L
2
sin

2
q
1
+
4
_
L
2
L
1
cos

2
q
2
___
, (167)
where

i
=

L
1
L
2


2
4
,

1
=
_
L
2
L
1
_
1
C
+
1
C
1
_
cos
2

2
+
_
L
1
L
2
_
1
C
+
1
C
2
_
sin
2

2
+
sin
C
,

2
=
_
L
2
L
1
_
1
C
+
1
C
1
_
sin
2

2
+
_
L
1
L
2
_
1
C
+
1
C
2
_
cos
2

2

sin
C
,
tan =
2

L
1
L
2
L
2
(1 +C/C
1
) L
1
(1 +C/C
2
)
. (168)
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

I =
2

i=1

i
_

b
i
+
1
2
_
, (169)
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
operators ( q
i
, p
i
), and new variables (

Q
i
,

P
i
):
_

Q
1

Q
2
_
= e
t
2
_
_
4
_
L
1
L
2
cos

2

4
_
L
2
L
1
sin

2
4
_
L
1
L
2
sin

2
4
_
L
2
L
1
cos

2
_
_
_
q
1
q
2
_
,
(170)
_

P
1

P
2
_
= e
t
2
_
_
4
_
L
2
L
1
cos

2

4
_
L
1
L
2
sin

2
4
_
L
2
L
1
sin

2

4
_
L
1
L
2
cos

2
_
_
_
p
1
p
2
_
+

2
_
L
1
L
2
e
t
2
_
_
4
_
L
1
L
2
cos

2

4
_
L
2
L
1
sin

2
4
_
L
1
L
2
sin

2
4
_
L
2
L
1
cos

2
_
_
_
q
1
q
2
_
.
(171)
Then, we may express Eqs. (164)-(167) by using Eqs.
(56)-(57), as

b
i
=
1
_
2
i

L
1
L
2
[
i
_
L
1
L
2

Q+i

P
i
], (172)

i
=
1
_
2
i

L
1
L
2
[
i
_
L
1
L
2

Qi

P
i
], (173)
From Eqs. (172), and (173) we conrm that

b
i
, and

b

i
are the annihilation, and the creation operators, respec-
tively, for a simple harmonic oscillator in the

Q repre-
sentation respectively. The wavefunction |
n
1
,n
2
(t) >
diers by only a time-dependent phase factor from the
wavefunction, |n
1
, n
2
, t > of the Schr odinger equation:
|
n
1
,n
2
,t
(t) exp
_
i
2

i=2
_
n
i
+
1
2
_

i
t
_
|n
1
, n
2
, t > .
(174)
The above wavefunction should satisfy the Schrodinger
equation for the Hamiltonian, Eq. (161):
i

t
|
n
1
,n
2
(t)

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

Q representation by
some unitary transformation

U:
|
n
1
,n
2
(t) >= hatU|
n
1
,n
2
> . (176)
Substitution of Eq. (176) in Eq. (175) gives a Hamil-
tonian which represent the simple harmonic oscillator
having frequencies w
1
, and w
2
:

H
q
=

U
1

H
q

U i

U
1


U
t
=
p
2
1
2

L
1
L
2
+
1
2
_
L
1
L
2

2
1
q
2
1
+
p
2
2
2

L
1
L
2
+
1
2
_
L
1
L
2

2
2
q
2
2
,
(177)
where

U =

U
1

U
2

U
3
, (178)

U
1
= exp
_
i
2
_
ln
4
_
L
1
L
2
+

2
t
_
( p
1
q
1
+ q
1
p
1
)
_
exp
_
i
2
_
ln
4
_
L
2
L
1
+

2
t
_
( p
2
q
2
+ q
2
p
2
)
_
,

U
1
= exp
_

2
( p
1
q
2
p
2
q
1
)
_
,

U
3
= exp
_

L
1
L
2
4
( q
2
1
+ q
2
2
)
_
. (179)
-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 n
1
= n
2
= 0 as
(q
1
)
2
(p
1
)
2
=
__
1 +

2
4
2
1
_
cos
4

2
+
_
1 +

2
4
2
2
_
sin
4

2
+
_

1
+

1

2
+

2
2
1

2
_
sin
2

2
cos
2
varphi
2
_

2
4
. (180)
(q
2
)
2
(p
2
)
2
=
__
1 +

2
4
2
1
_
sin
4

2
+
_
1 +

2
4
2
2
_
cos
4

2
+
_

1
+

1

2
+

2
2
1

2
_
sin
2

2
cos
2

2
_

2
4
. (181)
From Eq. (168), we know that one frequency, w
1
, cannot
be equal to the other frequency, w
2
. The uncertainty re-
lations do not satisfy the minimum uncertainty relation,
(x
i
, p
i
)
min
=/2, eventhough is zero. The uncer-
tainty relation approaches the minimum relation when
r = w
1
/w
2
1 or n.
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 equa-
tion for a damped harmonic oscillator with the Caldirola-
Kanai Hamiltonian [Eq. (43)] into the Schr odinger equa-
tion for an undamped harmonic oscillator [86,87]. In this
section, we review the relations for the canonical trans-
formations in classical mechanics, and the unitary trans-
formations in quantum mechanics [8890], and provide
the applications of this theory to the time-dependent har-
monic oscillators [91,92].
1. Linear Canonical Transformations for Classi-
cal, and Quantum Systems
Consider a system with the Hamiltonian, and the La-
grangian given by
H(q, p) =
p
2
2m
+V (q), (182)
L(q, q) =
1
2
m q
2
V (q). (183)
Let us introduce a linear transformation of the canonical
variables (q, p) to new variables (Q, P):
Q = e
(t)
q, (184)
P = e
(t)
p (t)e
(t)
q. (185)
Here, (t), and (t) are real, and dierentiable 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

QH
Q
(Q, P, t) = p q H(q, p, t) +
dF(Q, P, t)
dt
,
(186)
where the generating function, F(Q, P, t), is a time-
dependent function in phase space. The coecients
(P, p) of Q, P are given by
P p
q
Q
=
F(Q, P, t)
Q
, (187)
p
q
P
=
F(Q, P, t)
P
. (188)
Combining Eqs. (184), and (185) with Eqs. (187), and
(188), we obtain the generating function
F(Q, P, t) =

2
Q
2
. (189)
The new Hamiltonian, and Lagrangian are given by
H
Q
(Q, P, t) = e
2
p
2
2m
+
_

+e
2

m
_
PQ
+
1
2
_
e
2

2
m
+ 2

+
_
Q
2+V (Q)
, (190)
L
Q
(Q,

Q, t) = e
2

Q
2
(m

e
2
+)Q

Q
+
1
2
(m

2
e
2
)Q
2
V (Q). (191)
The relation between the canonical momentum, and the
kinetic momentum can be obtained from the Lagrangian:
P = e
2
P
k
(m

e
2
+)Q. (192)
The dierentiation of the inverse canonical transforma-
tion, Eqs. (184), and (185), gives the relation between
the kinetic momenta of the new, and old systems as
p
k
= m

q +e

P
k
. (193)
The mechanical energy of the new coordinate system be-
comes
E = e
4
P
2
2m
+
_

+

m
e
2
_
e
2
PQ
+
1
2m
_
m

+e
2
_
2
Q
2
+V (Q), (194)
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 in-
vestigate the quantum-mechanical relation between two
systems connected canonically [29,93]. From the deni-
tions of the propagator, and the wavefunction, Eqs. (47),
(29), and (88), together with Eq. (192), we obtain the
Schr odinger equation as
i
(Q, t)
t
=

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

H
Q
= e
2

2
2m

2
Q
2

i
2
_

+e
2

m
_
_
2Q

Q
+ 1
_
+
1
2
_
e
2

2
m
+ 2

+
_
Q
2
+V (Q). (196)
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)
t
=

2
2m

2
(q, t)
q
2
+V (q)(q, t). (197)
Here, one can conrm that the quantum Hamiltonian
has the same form as the classical Hamiltonian, Eq.
(182), whose variables are replaced by their correspond-
ing quantum operators. With the use of Eqs. (182), and
(191), the integral factor, and the integral denition of
the propagator, we can obtain the relation between the
propagators of the new, and the old systems [94]:
K(Q
2
, t
2
; Q
1
, t
1
) = exp
_

i
2

2
e
2
2
q
2
2
_
exp
_
i
2

1
e
2
1
q
2
1
_
K(q
2
, t
2
; q
1
, t
1
). (198)
Consider the unitary operator connecting two quan-
tum systems given by

U( q, p, t) = exp
_

i
2
q
2
_
exp
_

i
2
( q p + p q)
_
. (199)
Using Eq. (199), we obtain the quantum operator rela-
tions corresponding to the classical relations, Eqs.(184),
and (185):

Q =

U

U = e

q, (200)

P =

U

U = e

p e

q. (201)
The inverse forms of Eqs. (200), and (201) can be easily
obtained by using the unitary operator:

U
Q
(

Q,

P, t) =

U( q(

Q,

P, t), p(

Q,

P, t), t)
= exp
_

i
2
e
2

Q
2
_
exp
_

i
2
[(

Q

P +

P

Q) + 2

Q
2
]
_
. (202)
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):
|

U| > . (203)
The operation of < q| or < Q| on Eq. (203) yields the
relation between the wavefunctions in each space as
(Q, t) =

U(

Q,

P, t) < Q|

U( q, p, t)(e

q), (204)
where the unitary transformation operator,

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

U
q
( q, p, t) =

U(

Q( q, p, t),

P( q, p, t), t)
= exp
_

i
2
e
2
q
2
_
exp
_

i
2
( q p + p q) 2q
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
< | q| < |

U q

U
+
| >
= e

< | q| >
= e

< |

Q| >, (206)
< | p| e

< |( p + q)| >


= e

< |(

P +

Q)| >
= q < |

P| >, (207)
< |

Q| e

< | q| >, (208)


< |

P| e

< | p| > e

< | q| > .(209)


Equations (205)-(208) have the same forms as Eqs.
(184)-(185), and the inverses of Eqs. (184)-(185), respec-
tively. From these relations, we can conrm that the two
systems are related by a canonical transformation, and
form distinct quantum spaces.
From Eqs. (205)-(208), the relation between the quan-
tum uncertainty for the old and the new systems can be
evaluated:
(qp)
2
= (PQ)
2
+
2
(Q)
4
+(Q)
2
[< |(

Q

P +

P

Q)| >
2 < |

Q| >< |

P| >] (210)
or
(PQ)
2
= (qp)
2
+
2
E
4
(q)
4
e
2
(q)
2
[< |( p q + q p)| >
< | q| >< | p| >]. (211)
Note that for =0, the quantum uncertainties for the
two systems are the same.
The kinetic momentum is equal to the canonical mo-
mentum 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

P
k
= e
2

P + (m

+e
2
)

Q. (212)
With Eqs. (207)-(208), the relation between the quan-
tum averages of the kinetic momentum operators for
both systems becomes
< |

P| e

< | p
k
| > +m

e

< | q| > . (213)


From Eqs. (207)-(208), and (212), the relation between
the quantum uncertainties of the position, and the ki-
netic momentum operators for both systems can be ob-
tained as
(QP
k
)
2
= e
4
_
(p
k
q)
2
+m
2

2
(q)
4
+m

(q)
2
[< |( p
k
q + q p
k
)| >
< | q| >< | p
k
| >]} (214)
or
(qp
k
)
2
= e
4
{(QP
k
)
2
(m
2

2
+ 2m

e
2
)(Q)
4
m

e
2
(Q)
2
[< |(

Q

P +

P

Q)| >
2 < |

Q| >< |

P| >]}. (215)
Though Eqs. (213), and (214) contain (t), the quan-
tum uncertainty does not depends on , but only on .
Therefore, although the old system satises the uncer-
tainty principle, the new system may not satisfy it in
some cases if < 0. Furthermore, if , the uncer-
tainty of the new system goes to zero.
The mechanical energy operator for the new system
can be dened from Eq. (194) as
E
q
= e
4
p
2
2m
+
1
2
(

+

m
e
2
)e
2
(

P

Q+

Q

P)
+
1
2m
(m

+e
2
)
2

Q
2
+V (

Q). (216)
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 sys-
tems becomes
< |

E
Q
| e
2
_
< |

E
Q
| > +
m
2

2
< |q
2
| >
+

2
< (hatq p + p q)| >
_
= e
2
< |

E
Q
| > m

(2 +m

e
2
) < |

E
Q
| >
2m

< |(

Q

P +

P

Q)| >, (217)
where

E is the mechanical energy of the old system. Al-
though appears in Eq. (216), this equation does not
depend on , but only on .
2. Quantum Treatment for Special Types of Lin-
ear 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)
t
= e
2

2
2m

2
(Q, t)
Q
2

i
2

_
2Q
(Q, t)
Q
+ 1
_
+V (Q)(Q, t).
(218)
If we know the solution to Eq. (195) together with
Eq. (203), then we can nd the solution to the new
Schr odinger equation. For example, let us give the rela-
tion between the general solutions of the classical equa-
tion of motion for the harmonic, and the damped har-
monic oscillators with the same frequency as
Q = e

0
t
q, (219)
and the unitary operator connecting them as
U(

Q,

P, t) = exp
_
iover2
0
t(

Q

P +

P

Q)
_
. (220)
From Eq. (151), the Schr odinger solution for the damped
harmonic oscillator can be evaluated as

DHO
(

Q,

P, t) = U(

Q,

P, t)
HO
(

Q,

P, t)
=
(m/)
1/4

1/4

2
n
n!
e

0
t/2
e
2
0
t(m/2h)Q
2
H
n
_
e

0
t
_
m

Q
_
. (221)
The damped harmonic oscillator can be represented by
another Hamiltonian as
H
DHO
(

Q,

P, t) = e

0
t

P
2
2m
+e

0
t
m
2
2

Q
2
. (222)
Note that equation (221) is the damped harmonic oscil-
lator 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) =
(DeltaPQ). For the damped harmonic oscillator in
this case, we can obtain the uncertainties of the system
by using the canonical, and kinetic momentum opera-
tors, respectively:
(PQ)
DHO
=
_
n +
1
2

_
, (223)
(QP
k
)
DHO
= e
2
0
t
_
1 +

2
0
4w
2
_
1
2
_
n +
1
2
_
.
(224)
Quantum Theory of the Harmonic Oscillator in Nonconservative Systems Chung-In Um and Kyu-Hwang Yeon -613-
The uncertainty in Eq. (223) vanishes as time goes to in-
nity. 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 mechan-
ical energy operator for the damped harmonic oscillator
is
(

E)
DHO
= e

0
t
_
1 +

2
0
8w
2
_

_
n +
1
2
_
. (225)
The mechanical energy, Eq. (224), vanishes as times goes
to innity, 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 arbitrari-
ness of (t). In this case, the Hamiltonian of Eq. (190)
becomes
H
Q
(Q, P, t) =
p
2
2m
+

m
PQ+
1
2
_

2
m
+
_
Q
2
+V (Q).
(227)
In Eq. (226), depending on (t), there are innumerable
Hamiltonians, and Lagrangians that give rise to one dy-
namical equation for the system. Since the gauge invari-
ance is independent of the choice of (t) in Eq. (226),
Eq. (226) is the gauge transformation. The kinetic mo-
mentum for the transformation system becomes
p
k
(t) = m q = (t)q +P. (228)
If (t)=0, Eqs. (200), and (201) are reduced to Eq.
(225). In this case, the Hamiltonian of Eq. (196) be-
comes

H
Q
=

2
2m

Q
2

i
2

m
_
2Q

Q
+ 1
_
+
1
2
_

2
m
+
_
Q
2
+V (Q). (229)
From Eqs. (207), and (208) with (t)=0, the quantum
average of q is invariant for any space, but the momen-
tum operator is not. From Eq. (227), we can dene the
kinetic momentum as

P
k(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 momen-
tums uncertainty for the gauge transformation becomes
(QP)
2
= (qp)
2
+
2
(q)
4
(q)
2
[< | p q + q p)| > 2 < | q >< | p| >].
(231)
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, (QP
k
) =
(qp
k
). This does not depend on the gauge chosen.
Therefore, the uncertainties of the canonical, and ki-
netic momentum operators for the harmonic oscillator
become, after the gauge transformation, respectively,
(QP)
HO
=
_
1 +

2
m
2

2
_
1
2
_
n +
1
2
_
, (232)
(QP
k
)
HO
=
_
n +
1
2
_
. (233)
Under the gauge transformations of Eq. (225), the
Hamiltonian for the damped harmonic oscillator of Eq.
(221) becomes
H
GDHO
(

Q,

P, t) = e

0
t

P
2
2m
+e

0
t

2m
(

P

Q+

Q

P)
+
_
e

0
t
m
2
2
+e

0
t

2
2m
_

Q
2
. (234)
The relation of the canonical momentum operators un-
certainty between the two systems for this gauge trans-
formation can be obtained as
(QP)
2
GDHO
= (QP)
2
DHO
+

2
(
2

2
0
/4)
_

2
m
2
+

0
m
_

_
n +
1
2
_
2
. (235)
Substitution of Eq. (222) into Eq. (234) yields the
canonical momentum operators uncertainty correspond-
ing to the Hamiltonian of Eq. (233):
(QP)
GDHO
=
_
1 +

2
(
2

2
0
/4)

2
m
2
+

0
m
_
1
2
_
n +
1
2
_
2
. (236)
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 asymmet-
ric quantum septic harmonic oscillator [95], the time-
dependent 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 rst presented
quantum-mechanical treatments of various damped har-
monic oscillators by using the path-integral method, and
the second part gave the exact quantum theory for the
damped harmonic oscillator through the dynamical in-
variant method together with the path-integral method.
The Um-Yeon solution for the Caldirola-Kanai Hamil-
tonian, obtained through the path-integral method, was
presented in Sec. III, where the propagator, wavefunc-
tions, energy eigenvalues, uncertainty relations, and co-
herent states were discussed explicitly. This theory guar-
antees that Heisenbergs uncertainty principle and the
other fundamental laws in quantum mechanics are satis-
ed. Some examples for application of this theory were
given.
In Sec. IV, the procedure used in Sec. III was in-
troduced to evaluate exactly the propagator, wavefunc-
tions, energy expectation values, uncertainty relations,
and coherent states for a harmonic oscillator with a
time-dependent frequency, and a time-dependent exter-
nal driving force. Section V was devoted to the phe-
nomenological theory of time-dependent bound, and un-
bound 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 be-
tween 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 time-
dependent 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 op-
erators 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 combi-
nations of the scale, and the gauge transformations of
general Hamiltonian systems. We also showed that a sin-
gle system had innumerable Schrodinger equations when
a gauge transformation was used; however the quantum
averages of functions of the position, and the kinetic mo-
mentum 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 elds
such as the ssion of heavy nuclei [102,103], electric con-
ductivity [104], optical resonant cavities [105], quantum
Hall eects [106], tunneling problems through potential
barriers [107], Josephson currents [108, 109], and quan-
tum chaos with dissipation [110,111], and is still an open
problem in physics.
ACKNOWLEDGMENTS
This work was supported by the Korea Research Cen-
ter 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 Ra-
diation (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. Math-
Natusw. 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. Grin, Phys. Lett. B 50, 241
(1974).
Quantum Theory of the Harmonic Oscillator in Nonconservative Systems Chung-In Um and Kyu-Hwang Yeon -615-
[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. Schrodinger, 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 Confer-
ence 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 (McGraw-
Hill, 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
Scientic, 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).
[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 Dy-
namics (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- 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).
[104] H. N. Nagashima, R. N. Onody and R. M. Faria, Phys.
Rev. B 59, 905 (1999).
[105] C. W. Gardner, Quantum Noise (Springer-Verlag, New
York, 1991), Chap. 3.
[106] S. L. Sondhi, S. M. Girvin, J. P. Carini and D. Shahar,
Rev. Mod. Phys. 69, 315 (1997).
[107] 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 Uni-
versity Press, Cambridge, 1993).
[111] J. R. Ackerhalt, P. W. Milonni and M.-L. Shih, Phys.
Rep. 128, 205 (1985).

S-ar putea să vă placă și