• No results found

Thermal state of the general time-dependent harmonic oscillator

N/A
N/A
Protected

Academic year: 2022

Share "Thermal state of the general time-dependent harmonic oscillator"

Copied!
14
0
0

Loading.... (view fulltext now)

Full text

(1)

P

RAMANA c Indian Academy of Sciences Vol. 61, No. 1

—journal of July 2003

physics pp. 7–20

Thermal state of the general time-dependent harmonic oscillator

JEONG-RYEOL CHOI

Department of Electronic Physics, College of Natural Science, Hankuk University of Foreign Studies, Yongin 449-791, Korea

Email: choiardor@hanmail.net

MS received 25 July 2002; revised 15 January 2003; accepted 21 February 2003

Abstract. Taking advantage of dynamical invariant operator, we derived quantum mechanical solu- tion of general time-dependent harmonic oscillator. The uncertainty relation of the system is always larger than~=2 not only in number but also in the thermal state as expected. We used the diagonal elements of density operator satisfying Leouville–von Neumann equation to calculate various ex- pectation values in the thermal state. We applied our theory to a special case which is the forced Caldirola–Kanai oscillator.

Keywords. Time-dependent harmonic oscillator; thermal state; density operator.

PACS Nos 03.65.-w; 05.30.-d; 05.70.-a

1. Introduction

Vibration may be one of the most dominant physical aspect that we come upon in everyday life [1]. For small oscillation, it can be approximated to the motion of harmonic oscillator.

Harmonic oscillator that has time-dependent mass or frequency may be a good example of time-dependent Hamiltonian systems. Although a large number of dynamical systems have been investigated using approximation and perturbation method in the literature [2,3], we confine our concern to the exact quantum solution of the time-dependent system. There are about three kinds of methods to solve the quantum solution of time-dependent har- monic oscillator. These are propagator method [4–6], unitary transformation method [7–9]

and invariant operator method [10–16]. We will use invariant operator method and uni- tary transformation method together to evolve the quantum theory and investigate thermal state of the general time-dependent harmonic oscillator. The time-dependent harmonic oscillator has several applications such as electrical behavior of LC-circuit that has time- dependent parameter [17] and/or RLC-circuit [18], path-integral formulation of real-time finite-temperature field theory [19–21], dissipative quantum tunnelling effect in macro- scopic system [22–25] and quantum motion of an ion in a Paul trap [7,26,27].

When a system interacts with environment, its coupling parameters may explicitly de- pend on time. Even if the system is closed so that it is conserved as a whole, its subsystem

(2)

may implicitly depend on time through interaction with the remnant of the system. The main purpose of this paper is to evolve the thermal state of the general time-dependent harmonic oscillator. The Liouville–von Neumann equation [28,29] for non-equilibrium dynamics can be applicable to both time-dependent harmonic and unharmonic oscillator.

The density operator of the system can be obtained using the wave function satisfying Schr¨odinger equation and can be used to derive various expectation values of variables in the thermal state.

Inx2, we investigate quantum mechanical solution of the general time-dependent har- monic oscillator. The thermal state of the system is discussed in x3 on the basis of Liouville–von Neumann approach. Inx4, we will apply our theory for a special case which is the forced Caldirola–Kanai oscillator. Finally,x5 summarizes this paper and concludes about some physical results of the system.

2. Hamiltonian, invariant operator and wave function

The Hamiltonian of general time-dependent harmonic oscillator can be written as

Hˆ(xˆ;pˆ;t)=A(t)pˆ2+B(t)(x ˆˆp+p ˆˆx)+C(t)pˆ+D2(t)xˆ2+D1(t)xˆ+D0(t); (1) where A(t) C(t)and Di(t)(i=0;1;2)are time-dependent coefficients. These coefficients are real and differentiable with respect to t and note that A(t)6=0. The corresponding equation of motion can be derived from Hamilton’s equation of motion as

¨ˆx A˙ A˙ˆx+

2 ˙AB

A 2 ˙B 4B2+4AD2

ˆ x+AC˙

A C˙ 2BC+2AD1=0: (2) The introduction of invariant operator may save the labor of finding quantum mechanical solution of the system. We let the trial invariant operator as the form

Iˆ(t)=α1(t)[pˆ pp(t)]2+α2(t)f[xˆ xp(t)][pˆ pp(t)]

+[pˆ pp(t)][xˆ xp(t)]g+α3(t)[xˆ xp(t)]2; (3) whereα1(t) α3(t)are time-variable functions which should be determined afterwards and xp(t)is a particular solution of the equation of motion in ˆx space, (eq. (2)) and pp(t)is the corresponding particular solution of the equation of motion in ˆp space. We can choose the dimension of ˆI(t)the same as that of the Hamiltonian. By virtue of its definition, the invariant operator must satisfy the following relation:

d ˆI(t) dt =

Iˆ(t)

t + 1

i~[Iˆ(t);Hˆ]=0: (4)

Substituting eqs (1) and (3) in the above equation gives

α1(t)=c1ρ12(t)+c2ρ1(t)ρ2(t)+c3ρ22(t); (5) α2(t)= 1

4Af4[c1ρ12(t)+c2ρ1(t)ρ2(t)+c3ρ22(t)]B

[2c1ρ1(t)ρ˙1(t)+c2ρ˙1(t)ρ2(t)+c2ρ˙2(t)ρ1(t)+2c3ρ2(t)ρ˙2(t)]g; (6)

(3)

α3(t)= 1 2A2

f

1

2[c1ρ˙12(t)+c2ρ˙1(t)ρ˙2(t)+c3ρ˙22(t)]

B[2c1ρ1(t)ρ˙1(t)+c2ρ˙1(t)ρ2(t)+c2ρ˙2(t)ρ1(t)+2c3ρ2(t)ρ˙2(t)]

+2B2[c1ρ12(t)+c2ρ1(t)ρ2(t)+c3ρ22(t)]g; (7) where c1–c3are constants and ¨ρ1;2(t)are two independent solutions of the following dif- ferential equation

ρ¨1;2(t) A˙

Aρ˙1;2(t)+

2 ˙AB

A 2 ˙B 4B2+4AD2

ρ1;2(t)=0: (8) To simplify the invariant operator, we introduce the unitary operator defined as

Uˆt=Uˆ00Uˆ0Uˆ; (9)

Uˆ =exp

i

~

xppˆ

exp

i

~

ppxˆ

; (10)

Uˆ0=exp

i α21~xˆ2

; (11)

Uˆ00=exp

i 4~

(x ˆˆp+p ˆˆx)ln(1)

: (12)

We can transform the invariant operator using the above operator as

Iˆ0=UˆtI ˆˆUt: (13)

Then, ˆI0reduces to the following simple form Iˆ0=1

2pˆ2+1

2xˆ2; (14)

where

ω2=4(α1α3 α22)=

1

4A2(ρ1ρ˙2 ρ˙1ρ2)2(4c1c3 c22)=constant: (15) For convenience, we only discuss the system forω2>0. Since eq. (14) is the same as that of the Hamiltonian of the simple harmonic oscillator with unit mass, we can introduce the ladder operators defined as

ˆb=

rω

2~xˆ+ i

p

~pˆ; (16)

ˆb=

rω

2~xˆ i

p

~pˆ: (17)

These satisfy the boson commutation relation[ˆb;ˆb]=1. In terms of these operators, eq.

(14) can be simplified to

(4)

Iˆ0=~ω

ˆbˆb+1 2

: (18)

The eigenvalue equation for ˆI0can be written as

Iˆ0jn0(t)i=λnjn0(t)i: (19)

We can easily identify the eigenstate in ˆx space as

hxˆjn0(t)i=

ω

~π

1=4 1

p

2nn!Hn

rω

~

ˆ x

exp

ω

2~xˆ2

: (20)

The eigenstate of the untransformed invariant operator can be obtained from

hxˆjn(t)i=Uˆthxˆjn0(t)i: (21) Substituting eqs (9) and (20) into the above equation gives

hxˆjn(t)i=

ω

1~π

1=4

1

p

2nn!Hn

r ω

1~(xˆ xp)

exp

i

~

ppxˆ

exp

1 2α1~

ω 2 +iα2

(xˆ xp)2

: (22)

The ˆx space Schr¨odinger solutionhxˆjψniof the Hamiltonian, eq. (1), is the same as the eigenstate of ˆI, except for some time-dependent phase factor,εn(t)[30]

hxˆjψn(t)i=exp[iεn(t)]hxˆjn(t)i: (23) Inserting the above equation into Schr¨odinger equation, we derive the relation

~ε˙n(t)=hn(t)j

i~

t Hˆ

jn(t)i: (24)

Using eqs (1), (22) and (24),εn(t)can be obtained as εn(t)= ω

n+1 2

Zt 0

A(t0) α1 dt

0

1

~ Zt

0

Hp(xp(t0);pp(t0);t0)dt0; (25) where Hp(xp(t);pp(t);t)is defined as

Hp(xp(t);pp(t);t)=A(t)p2p(t)+C(t)pp(t) D2(t)x2p(t)+D0(t): (26) Substituting eq. (25) into (23), we can obtain the exact wave function as

hxˆjψn(t)i=hxˆjn(t)iexp

iω

n+1 2

Zt 0

A(t0) α1 dt

0

i

~ Zt

0

Hp(xp(t0);pp(t0);t0)dt0

: (27)

(5)

The ˆp space wave function is related to the ˆx space by the Fourier transformation

hpˆjψn(t)i= 1

p

~

Z

hxˆjψn(t)iexp

ip ˆˆx

~

d ˆx: (28)

Using (28), the above equation can be calculated as

hpˆjψn(t)i=( i)n

2ωα1

~π

1=4 1

p

2nn!

(ω 2iα2)n

(ω+2iα2)n+1

1=2

Hn

"s

1ω

~(ω2+4α22)(pˆ pp)

#

exp

i

~

xp(pˆ pp) α1(pˆ pp)2

~(ω+2iα2)

exp

iω

n+1 2

Zt 0

A(t0) α1 dt

0

i

~ Zt

0

Hp(xp(t0);pp(t0);t0)dt0

: (29)

To express ˆI(t)in a simple form, we introduce another ladder operator as ˆ

a(t)= 1

pω~α1

hω 2 +iα2

(xˆ xp)+iα1(pˆ pp)

i

; (30)

ˆ

a(t)= 1

pω~α1

h ω 2 iα2

(xˆ xp) iα1(pˆ pp)

i

: (31)

These operators also satisfy[aˆ;aˆ]=1. In terms of eqs (30) and (31), (3) can be expressed as

Iˆ(t)=~ω

ˆ aaˆ+1

2

: (32)

From eqs (30) and (31), we can confirm that the coordinate and the momentum can be expressed as

ˆ x=

r

~α1

ω (aˆ+aˆ)+xp; (33) ˆ

p=i

s

~

α1ω

hω 2 +iα2

ˆ a

ω 2 iα2

ˆ a

i

+pp: (34)

Using eqs (27), (33) and (34), we can calculate the following expectation values

hψnjxˆjψni=xp; (35)

hψnjpˆjψni=pp; (36)

hψnjxˆ2jψni=

~α1

ω (2n+1)+x2p; (37)

(6)

hψnjpˆ2jψni=

~

α1ω

ω2 4

+α22

(2n+1)+p2p; (38)

hψnj(x ˆˆp+p ˆˆx)jψni=

2~

ω (2n+1)+2xppp: (39) Then, we can easily identify the uncertainty relation as

x∆ˆ pˆ=[hψnjxˆ2jψni (hψnjxˆjψni)2

]

1=2

[hψnjpˆ2jψni (hψnjpˆjψni)2

]

1=2

=

~

ω

q

ω2+4α22

n+1 2

: (40)

This is always larger than~=2 as expected. The uncertainty relation of q-deformed har- monic oscillator also differs from the uncertainty relation for the simple harmonic oscillator [31].

By performing a similar procedure, we obtain the expectation value of Hamiltonian as

hψnjHˆjψni=

~

ω[α1D2 2α2B+α3A](2n+1)+Hp(xp(t);pp(t);t): (41)

3. Thermal state

We consider an ensemble of particles that satisfies the given general time-dependent har- monic oscillator motion. Let us assume that these particles conform to the Bose–Einstein distribution function.

Density operator of the system may satisfy Liouville–von Neumann equation as

∂ρˆ(t)

t + 1

i~[ρˆ(t);Hˆ]=0: (42)

Then, we can express the density operator in ˆx space as ρˆ(xˆ;xˆ0;t)= 1

Z(t)

n=0

hxˆjψn(t)iexp

~ω kT

n+1 2

hψn(t)jxˆ0i; (43) where k is the Boltzmann constant and T the temperature of the system at initial time.

The partition function of the system can be given by Z(t)=

n=0

hψn(t)je Iˆ(t)=(kT)jψn(t)i: (44) Using eq. (27), the partition function, eq. (44) and the density operator, eq. (43) can be calculated as

Z(t)= 1

2 sinh[~ω=(2kT)]; (45)

(7)

ρˆ(xˆ;xˆ0;t)=

ω

~α1tanh

~ω 2kT

1=2

exp

(

iα2

2~α1[(xˆ xp)2 (xˆ

0 xp)2]

ω 8~α1

(xˆ+xˆ0 2xp)2tanh

~ω 2kT

+(xˆ xˆ0)2coth

~ω 2kT

)

exp

i

~

(xˆ xˆ0)pp

: (46)

At high temperature, eq. (46) reduces to ρˆ(xˆ;xˆ0;t)'ω

2 1

(α1πkT)1=2exp

i

~

(xˆ xˆ0)pp

exp

iα2 2~α1[xˆ2 xˆ

02 2(xˆ xˆ0)xp] kT 4~2α1(xˆ xˆ

0

)

2

: (47) If the difference between ˆx and ˆx0 are sufficiently small compared to(2~α1=α2)1=2 and

~=pp, and 1=(kT)approaches zero, the density operator may be simply represented as ρˆ(xˆ;xˆ0;t)'~ω

kT δ(xˆ xˆ0): (48)

On the other hand, at low temperature, it becomes ρˆ(xˆ;xˆ0;t)'

ω

~α1

1=2

exp

i

~

pp(xˆ xˆ0)

exp

(

iα2 2~α1[xˆ2 xˆ

02 2(xˆ xˆ0)xp] ω

4~α1(xˆ2+xˆ

02

+2x2p 2 ˆxxp 2 ˆx0xp)

)

: (49)

The diagonal element of the density operator, eq. (46), can be written as f(xˆ)=

ω

~α1tanh

~ω 2kT

1=2

exp

ω

2~α1tanh

~ω 2kT

(xˆ xp)2

:

(50) The above equation represents the probability that the mass of the oscillator reside at ˆx. As temperature increases, it becomes

f(xˆ)'ω 2

1

(πα1kT)1=2exp

ω2

4kTα1(xˆ xp)2

: (51)

Equation (50) can be used to calculate the expectation value in coordinate space as

hxˆliT=

Z

xˆlf(xˆ)d ˆx: (52)

(8)

For example, we can obtain for l=1;2 as

hxˆiT=xp; (53)

hxˆ2iT=~α1 ω coth

~ω 2kT

+x2p: (54)

When considering eq. (44), the expectation value of ˆI in the thermal state can be derived from

hIˆiT=kT2

Tln Z(t): (55)

Making use of eq. (45), the above equation becomes

hIˆiT=1 2~ωcoth

~ω 2kT

: (56)

Using the same procedure in the ˆx space, the ˆp space representation of the density operator can be obtained as

ρ(pˆ;pˆ0;t)=

1ω

~π(ω2+4α22)tanh

~ω 2kT

1=2

exp

i

~

xp(pˆ pˆ0)

exp

(

α1ω

~(ω2+4α22)

(

2iα2 ω [pˆ2 pˆ

02 2pp(pˆ pˆ0)]

1 2

"

(pˆ+pˆ0 2pp)2tanh

~ω 2kT

+(pˆ pˆ0)2coth

~ω 2kT

#))

:

(57) At high temperature, the above equation becomes

ρ(pˆ;pˆ0;t)'

α1ω2 π(ω2+4α22)kT

1=2

exp

i

~

xp(pˆ pˆ0)

exp

(

α1

~(ω2+4α22)

(

2iα2[pˆ2 pˆ02 2pp(pˆ pˆ0)]

kT

~

(pˆ pˆ0)2

))

: (58)

On the other hand, at low temperature it can be expressed as ρ(pˆ;pˆ0;t)'

1ω

~π(ω2+4α22)

1=2

exp

i

~

xp(pˆ pˆ0)

exp

(

α1ω

~(ω2+4α22)

(

2iα2 ω [pˆ2 pˆ

02 2pp(pˆ pˆ0)]

(pˆ2+pˆ02 2 ˆppp 2 ˆp0pp+2p2p)

))

: (59)

(9)

The probability that the mass of the oscillator resides at ˆp is obtained taking the diagonal elements of eq. (57) as

f(pˆ)=

1ω

~π(ω2+4α22)tanh

~ω 2kT

1=2

exp

1ω

~(ω2+4α22)tanh

~ω 2kT

(pˆ pp)2

: (60)

Using eq. (60), we can calculate the expectation value in ˆp space as

hpˆiT=pp; (61)

hpˆ2iT=~(ω2+4α22)1ω coth

~ω 2kT

+p2p: (62)

We can also write the expectation value of ˆI in the thermal state as

hIˆiT=α1(hpˆ2iT p2p)+2(hx ˆˆpiT xppp)+α3(hxˆ2iT x2p): (63) Substituting eqs (54), (56) and (62) into the above equation gives

hx ˆˆpiT= ~ω 2α2

1 4

1

ω2(α22+α1α3)

coth

~ω 2kT

+xppp: (64) Then, the expectation value of the Hamiltonian, eq. (1), can be calculated as

hHˆiT= ~

ω(α3A 2α2B+α1D2)coth

~ω 2kT

+Hp(xp(t);pp(t);t) (65) and the uncertainty relation in thermal state can be calculated as

(xˆ∆pˆ)T=[hxˆ2iT hxˆi2T)(hpˆ2iT hpˆi2T)]1=2

=

~

q

ω2+4α22coth

~ω 2kT

: (66)

By comparing the above equation with eq. (40), we can confirm that the uncertainty rela- tion in thermal state varies as time goes by, with the same fashion in number state.

4. Forced Caldirola–Kanai oscillator

We can apply our theory to various kinds of time-dependent Hamiltonian systems. As an example, let us see for the forced Caldirola–Kanai oscillator [32,33]. For this system, the time-dependent coefficients in eq. (1) are given by

A(t)= 1

2me βt; (67)

D2(t)=1

2mω02eβt; (68)

D1(t)= F(t)eβt; (69)

B(t)=C(t)=D0(t)=0; (70)

(10)

so that we can rewrite the Hamiltonian as Hˆ=e βt pˆ2

2m+eβt1

2mω02xˆ2 eβtF(t)xˆ; (71) where m is the mass,β the damping constant and F(t)the arbitrary time-dependent driving force. Equation (8) becomes

ρ¨1;2+βρ˙1;2+ω02ρ1;2=0: (72) The two classical solutions of the above equation can be written as

ρ1(t)=ρ1(0)e βt=2eiωt; (73) ρ2(t)=ρ2(0)e βt=2e iωt; (74) whereωis given by

ω=

r

ω02 β2

4 : (75)

We choose c1–c3in eqs (5)–(7) as

c2= 1

2mρ1(0)ρ2(0); c1=c3=0: (76)

The particular solutions xpand ppsatisfy the following relations:

¨

xp+βx˙p+ω02xp=

F(t) m

; (77)

¨

pp βp˙p+ω02pp=eβtF˙(t): (78)

The solutions of the above equations depend on F(t). If, we choose F(t)as

F(t)=F0t; (79)

the solutions of eqs (77) and (78) will be xp(t)= F0

mω02t βF0

mω04; (80)

pp(t)= F0

ω02eβt: (81)

We will also investigate the system driven by the exponentially decaying force:

F(t)=F0e γt; (82)

whereγis an arbitrary real constant. In this case, the particular solutions are given by xp(t)= F0=m

γ2 βγ+ω02e γt; (83)

pp(t)= γF0 γ2 βγ+ω02e

(β γ)t

: (84)

References

Related documents

There are two aspects to this study: first, we consider the finiteness of each term in the perturbation theory, that is, we consider the behaviour of each of the generators Fi as a

Two cases of forced harmonic oscillators with time dependent mass for which exact propagators can be evaluated are presented.. From the exact propagators, normalized

In the present work we have investigated the most general perturbation of a three- dimensional harmonic oscillator (for l = 0 states only). the harmonic

Lie's method of differential equation is used to obtain the one-parameter Lie groups admitted by the time-dependent Schr6dinger equations for atoms, molecules and nucleons in

While raising the investment limit on the basis of some valid and generally admissible criteria, other factors like the number of employees in the enterprises and the turnover,

Integrated land-use planning at national and subnational level, carried out in consultation with relevant stakeholders, is another crucial requirement and should include scenario

SaLt MaRSheS The latest data indicates salt marshes may be unable to keep pace with sea-level rise and drown, transforming the coastal landscape and depriv- ing us of a

INDEPENDENT MONITORING BOARD | RECOMMENDED ACTION.. Rationale: Repeatedly, in field surveys, from front-line polio workers, and in meeting after meeting, it has become clear that