• Aucun résultat trouvé

Radiation-reaction force on a moving mirror

N/A
N/A
Protected

Academic year: 2021

Partager "Radiation-reaction force on a moving mirror"

Copied!
10
0
0

Texte intégral

(1)

HAL Id: jpa-00246859

https://hal.archives-ouvertes.fr/jpa-00246859

Submitted on 1 Jan 1993

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Radiation-reaction force on a moving mirror

Claudia Eberlein

To cite this version:

Claudia Eberlein. Radiation-reaction force on a moving mirror. Journal de Physique I, EDP Sciences,

1993, 3 (11), pp.2151-2159. �10.1051/jp1:1993237�. �jpa-00246859�

(2)

Classification Physics Abstracts

03.70 05.40 11.10E

Radiation-reaction force

on a

moving mirror

Claudia

Eberlein(*)

School of Mathematical and Physical Sciences, University of Sussex, Brighton BNI 9QH, England

(Received 12 March 1993, accepted in final form 22 July1993)

Abstract. The radiation-reaction force

F(t)

acting on a moving mirror in a one-dimensional model is obtained by appeal to the fluctuation-dissipation theorem. Causality is found to pro- vide the link between the dissipative and reactive forces; it is implemented by requirements of

analyticity for the Fourier transform of the response function that links the mean force

F(t)

to the velocity

fl(t)

of the mirror. Particular attention is given to the issue of the divergent mass

renormalization arising from the reactive part of the force, detected in

some earlier treatments and denied in others.

1. Introduction.

Quantum theory predicts

radiation from

moving

mirrors 11, 2], as well as from

moving

di- electrics [3], in consequence of their interaction with

zero-point

fluctuations of a

quantized

field.

Conversely,

the

(macroscopic) body,

mirror or

dielectric,

is

subject

to a back-reaction from this radiation, I-e- experiences a radiation-reaction force.

Knowing

the spectrum of the

radiation one can

readily

evaluate the mean force

F(t) acting

on the

body. However,

since the calculation of the spectrum

(see

[3] for a one-dimensional scalar

model)

proves

quite

a

strenuous

task,

it is

tempting

to use the

fluctuation-dissipation

theorem [4, 5] which

yields F(t)

to first order in the

velocity fl(t) (or

its

derivatives)

from the correlation function of the

fluctuating

force on a

stationary body;

this has been done or touched on in a series of recent papers [6 or

7-9], respectively.

Though straightforward

at first

sight,

the

application

of the

fluctuation-dissipation

theorem to this kind of

problem

bears a number of subtleties

easily leading

to specious conclusions.

So,

for instance, the

dissipative

part of the force on a

moving

mirror is

manifest,

whereas its reactive part, which

gives

rise to a mass renormalization in one-dimensional models, is

missed if

divergences

are handled

improperly

and the no-cut-off limit is taken

prematurely.

This observation

gains special

interest in view of the different results for the reactive part of the force

given

in the literature

[1-3, 7-9], although

the

dissipative

part is not

subject

to

disagreement.

The present paper is intended to serve as a consistency check on previous results

(*) Present address: University of Illinois, Dept of Physics, Il10 West Green Street, Urbana IL 61801-3080,

U-S-A-

(3)

2152 JOURNAL DE PHYSIQUE I N°11

[3] and to

point

out the niceties in the interrelation of the

dissipative

and the reactive parts of the force. For the

example

of the force on a

slowly moving

dielectric

half-space

in a one-

dimensional scalar model it is shown

explicitly

that

causality intrinsically

connects

dissipative

and reactive parts.

The dielectric is

supposed

to be

non-dispersive,

and hence

non-absorptive;

its material prop- erties are described

by

a

single

parameter the refractive index n, which is chosen as an

arbitrary

constant.

Although

the limit n ~ oc of a

perfectly reflecting

mirror is aimed at in the end

results,

one should stick to finite n in

formulating

the

problem

and in any intermediate steps of

calculating

the force. The reader to whom this seems overcautious or even unfounded is reminded of the fact that no Hamiltonian formalism exists for a

moving (perfect)

mirror

since the system

changes

from one Hilbert space into another each time the mirror moves

[10],

whereas any derivation of the

fluctuation-dissipation

theorem

indispensably

involves Hamil- tonians for the

unperturbed (here: stationary)

and for the

perturbed (here: moving)

system [4, 5]. As demonstrated elsewhere [3], any

ambiguities

can be

sidestepped by keeping

the

refractive index n finite

everywhere

but in the end results.

2. Force correlation function for the

stationary

mirror.

The mirror at rest represents the

unperturbed

system, which is

governed by

some time-

independent

Hamiltonian Ho-

Having

in mind the

application

of the

fluctuation-dissipation

theorem, which is based on standard

time-dependent perturbation theory,

one has to

require that, subsequently,

the

moving

mirror is described

by

the same, still

time-independent

Ho

Plus

a

perturbation V(t)

that carries all

explicit time-dependence

of the system and

depends

on

a small parameter, which in fact will be the

velocity fl(t).

The correlation function for the random

fluctuating

force exerted

by

the field on the

stationary

mirror

(or

to zeroth order on the

moving mirror)

is

routinely

defined

by

C(t t')

=

101)iF(t),F(t')i10) -101F(t)10)101F(t')10)

,

(2.1)

where the force operator

F(t)

is a

Heisenberg

operator and is related to the

time-independent Schroedinger

operator F

by

F(t)

=

e'~°~~~~°~Fe~'~°~~~~°~

Although

the

dependence

of the force operator F and of the

unperturbed

Hamiltonian

Ho

on the field variables is

specific

to the

particular

system

investigated, quite generally

the

fluctuation-dissipation

theorem links the force correlation function

C(t -t')

for the

unperturbed

system to the

dissipative

part of the expectation value

F(t)

of the force operator F in the

perturbed

state

(to

first order in the

perturbation parameter).

The

implementation

of the above formulae for the case of the mirror

might

appear rather

trivial,

yet is not in fact as one will notice when

trying

to

identify

Ho- The one-dimensional scalar model used in the

following

is defined

by

the requirements that the

(classical

or

quantum)

field

4l(x, t) satisfy

the wave equation

e4l 4l"

= o

(2.2)

in the rest-frame of the dielectric

(but non-magnetic) medium,

which is described

solely by

the dielectric constant e, and that acceleration stresses on the

moving

medium be

disregarded.

Standard

Lagrangean theory

establishes the momentum H

conjugate

to 4l,

H = e 4l

(2.3)

(4)

The

quantization

of the system is conducted

by imposing

the

(equal-time)

commutation rela- tions

w(xi

,

n(x')

= i&(x xi

on the

canonically conjugate

field operators 4l and H. The Hamiltonian for the field in the presence of a stationary dielectric

half-space occupying

x >

(

reads

with the dielectric constant

e(xii)

= @(i

x)

+

n2@(x 1)

featuring

the refractive index n of the medium. Hamilton's

equations

of motion

obviously reproduce

the wave

equation (2.2)

and the

defining

relation for H

(2.3),

as

they

must. The Hamiltonian H is in fact very familiar from standard

electromagnetism,

when one

regards

4l'

and H as

loosely analogous

to the

electromagnetic

B and D. The term "dielectric"

frequently

used here also refers to this

analogy.

To see the

difficulty

in the

application

of the

fluctuation-dissipation

theorem to this system

note that when the

half-space

starts

moving

( becomes a

time-dependent

parameter, and therefore H is not

time-independent

anymore, so cannot serve as the

unperturbed

Hamiltonian Ho

The d4nouement is to accommodate the

time-dependence

of H in a suitable

time-dependent unitary

transformation of the states and operators in the

Schroedinger equation.

Such a trans-

formation is

provided by

the

unitary

translation

operator(~) U(((t))

=

exp[

I

((t)

x

fC~

dx

H(x) 4l'(x)],

whose action could be

interpreted

either as

always re-displacing

the

moving body

to the

origin

or as

making

the coordinate system travel with the

body.

For the

moving half-space

this leads to the effective interaction

Hamiltonian(~) V(t)

=

fl(t) /~

dx

~)~)(()~~

,

(2.4)

which adds to the

unperturbed

Hamiltonian

Ho-Ldzlll+~'~(zl

Note that Ho =

Ut(((t)) HU(((t))

involves

only

the dielectric constant for the

half-space

at

position (

= o and is

time-independent

as

required.

In other

words,

for the present model of

a

moving

mirror one can be sure that the

application

of the

fluctuation-dissipation

theorem is

wholly unproblematic

and

legitimate.

Expressing

the field operators 4l and H in terms of

photon

annihilation and creation operators

by

means of standard normal-mode

expansion,

so as to

diagonalize

the Hamiltonian Ho to

Ho =

fj~~

dk w

(a(ak

+

1/2),

allows one to rewrite the correlation function

(2.I) by

inserting

a

complete

set of

photon

states in between the two force operators; one

gets(~):

C(t t')

=

/~

dk

/~

dk'

cos [(w

+w')(t t')] [(o[F[k, k')[~

(l)Since

the details of this are not essential for the reasoning in the present article, the interested reader is referred to [3] for these; as regards the unitary translation operator U(((t)) and its utilization, especially to section 4 of reference [3].

(2)F is quadratic in the fields and therefore quadratic in the ak and

a(

operators, whence the only transitions

from the vacuum F can induce go to the vacuum or to a two-photon state.

JOURNAL DE PHYSIQUE -T 3 N"I NOVEMBER 'Q93

(5)

2154 JOURNAL DE PHYSIQUE I N°11

The fluctuation spectrum, I-e- the Fourier transform of the correlation function

C(T),

reads

x(fly

=

j~

dT

C(T) e'"~ (2.5a)

x(fly

=

j /~

dk

j~

dk'

j(0jFjk, k'jj~ jd(w

+ w' fly + &(w + w' +

fljj (2.sbj

The force operator F on the

half-space

located at x >o is

given by ill, Chap. 6],

[3,

appendix A.4]

~

~

/2 ~'~~~~

~~'~~

Although

this

expression

is

merely quoted here,

it should be

emphasized

that its derivation

requires

some care if the

half-space

is a dielectric and not a

perfect

mirror. The mechanical force

on the

(rigid) body

is obtained from the surface stress due to the

uncoupled

field

(I.e.

exclusive

of the

polarization

field inside the

dielectric) by subtracting

the momentum transferred from the

uncoupled

to the

polarization

field.

Only

for a

perfect

conductor is the force

given directly by

the surface stress of the field since the

uncoupled

field inside the conductor is

suppressed

and cannot deliver any momentum to the

polarization

field.

Normal-mode

expansion

of the field allows one to evaluate the matrix element

(o[F[k, k') and, subsequently,

the spectrum

(2.5b)

x(fl)

=

(() ~ lfll~e~°'"'

,

(2.7)

where a is a

cut-off(~)

on the sum of the

frequencies

of the two

photons.

Although

not needed in the course of the argument and recorded for

completeness only:

the correlation function reads

~~~~ ~

n

~ ~

~2

~~

(T ~ia)4

' ~~ ~~

which bears

comparison

with the correlation function for the

electromagnetic

pressure fluc- tuations at one and the same

point

on a

plane perfect

mirror

bounding

a three-dimensional

half-space

[12].

3.

Response

function and

fluctuation-dissipation

theorem.

The mean force

F(t)

on the

moving half-space

is

given by

the

expectation

value of the force operator

(2.6)

in the state that has evolved from the vacuum state under the influence of the

interaction Hamiltonian

(2.4).

To first

order,

one can write

Fit)

=

/~

dT

G(t

T)

fllT)

,

(3.1)

I-e- consider

F(t)

as a linear response to the

perturbation V(t),

equation

(2.4)(4).

(~) The need for a cut-off is not apparent here, since the expression (2.7) has an obvious, mathematically well- defined limit for a - 0. However, the correlation function (2.8) does not admit a physically sensible no-cut-off limit. Furthermore, in the following section the finite frequency cut-off will turn out the essential ingredient in the dispersion relations for the response f~nction.

(4) Since V(t) is parametrized by the velocity p(t) +

I(t)

(rather than by the displacement ((t)), the force is

more conveniently considered as a response to p(t) rather than to ((t), which will become clear by looking at the dispersion relations for the Fourier transforms of the response functions (cf. footnote 6).

(6)

The

fluctuation-dissipation

theorem involves the Fourier transform of the response function

r(n)

=

/°'

dT

G(T) ei"T (3.2)

Independently

of any

dynamical

detsdls of the system,

r(fl)

must have the

following properties:

I)

r(-fl)

=

r*(fl),

I-e-

Rer(fl)

is an even and Im

r(fl)

an odd

function,

since

G(T)

must be real as seen in

equation (3.I);

it) r(fl)

is

analytic

in the upper

half-plane

Imfl > o,

by

virtue of Titchmarsh's theorem [13,

14],

since

G(T)

is

causal,

I-e- vanishes on the

negative

real axis T < o;

furthermore,

if

r(fl)

is

square-integrable

on the real axis then

Rer(fl)

and

Imr(fl)

are connected

by dispersion

relations

(Hilbert transformation) according

to

Rer(fl)

= P

/~

dw~~~~)~

,

(3.3a)

7r -~x>

Ld

Imr(nj

= P

)°~ dw~~~ljl (3.3bj

7r -~x>

Ld

The

fluctuation-dissipation

theorem [4, 5] relates the power spectrum

x(fl)

to the Fourier transform of the response

function;

for a response function

linking

force to

velocity

it

states(~):

Rer(fl)

=

x(fl) (3.4)

(fl(

As the

dissipative

force as well as the

velocity

are odd under

time-reversal, ReT(fl)

stands for the

dissipative

part of the response. From

(2.5b)

one obtains

Rer(fl)

=

-I /~

dk

/~

dk' ~~~~~~~'

'~~~

[d(w +w'-

fl)

+

d(w

+w'+

fl)]

2

_~ -~

w + w

Then

fornJal application

of the

dispersion

relation

(3.3b) yields

It must be

emphasized

that this is

definitely only formal

in the sense that

equations (3.3a,b)

are

only

valid if

r(fl)

is

square-integrable.

However, the

square-integrability

of

r(fl) (and

with it the convergence of the

w-integrals

in the

dispersion relations) depends

on the

specific

matrix

element

[(o[F[k, k')[~

and has therefore to be checked in each

particular

case.

For the dielectric

half-space equation (2.7) implies

Rer(fl)

=

(fi) fl~e~"'"' (3.5a)

n

127r

the

dispersion

relation

(3.3b) yields

Im

r(n)

=

~

~

sgn(n) (2

'"' n2e-*'n'Ei(ajnj) n2e*'"'Ei(ajnj) (3.5b)

n 127r a

(~)If

one had chosen to write the force as response to the displacement, F(t)

=

f)

dr G(t r)((r), then the fluctuation-dissipation theorem involving the Fourier transform r(Q) of G(Q) assumes the more familiar

(zero-temperature) form

Imf(Q)

= sgn(Q)x(Q), since plainly

f(Q)

= -ion(Q).

(7)

2156 JOURNAL DE PHYSIQUE I N°11

(For

the definition and the

properties

of the

exponential integrals

see e-g- Ref.

[15].)

As

regards

the

square-integrability

of

r(fl)

in

(3.5a,b),

there is

obviously

no

problem

with

Rer(fl)

since it falls

exponentially

as

(n(

- cc. A short calculation,

using only

standard methods of

asymptotic approximation,

shows that the

expression

in the

curly

brackets in

(3.5b)

for

Imr(fl)

behaves like

-4/(a~ (n()

as

(n(

- cc. Hence

r(fl)

is indeed

square-integrable,

and satisfies the

dispersion

relations

(3.3a,b) just

as

they

are

written(~).

4.

Dissipative

and reactive forces in the no-cut-off limit.

In the no-cut-off limit o - 0

equations (3.5a,b)

entail

Rer(fl) #

'~

fl~, (4.la)

n

~

127r

~~~~~~ ~ ~

n

~

~

1~2a

~ ~~'~~~

Substitution into

~ ~

F(t)

= dr dfl

r(fl) e~~~~~~~)fl(r)

,

(4.2)

27r

~co ~co

which follows from

equation (3.I)

and the inverse of

equation (3.2), yields

for the force on the dielectric

half-space

~

~~~~ ~

n

~7r~~~~ 1~2a~~~~

~~'~~

For a

perfect

mirror, which can be modelled

by

infinite refractive index n, the

n-dependent prefactor

goes to I, and the force

equals

the

expression

in the

curly

brackets. The first term,

proportional

to

fl(t),

is the

dissipative (or "frictional")

force on the mirror; the second term,

proportional

to

fl(t),

is

divergent

for zero cut-off a and

gives

rise to a renormalization of the mirror's inertial mass. As

easily

understood from

equation (4.2)

and illustrated

by

the force

(4.3) ensuing

from equations

(4.la,b),

powers

(-I fl)P

in

r(fl),

the Fourier transform of the response

function, engender

terms

fl(P)(t), pth

time-derivatives of the

velocity,

in the mean force

F(t). So, Rer(fl)

in

(4.la) brings

about the

dissipative force,

and

Imr(fl)

in

(4.lb)

leads to the mass-renormalization term.

Coincidentally,

the force on a

slowly moving

classical point

charge [16],

handled

by

classical

electromagnetism (in

three

dimensions),

is

equivalent

to

equation (4.3),

apart from a different

prefactor featuring

the

squared

value of the

charge.

This

analogy

has wide

implications,

as for

instance the

replacement

of Newton's

equation

of motion for the mirror

by

the Lorentz-Dirac

integro-differential

equation [16] in order to avoid run-away solutions. Here the

parallelism

is called upon

merely

to corroborate the mass renormalization

required

for the mirror. It is well-known that the electron mass also gets

renormalized,

in consequence of the interaction between electron and

electromagnetic field;

moreover, the argument for the electron can be based on the

fluctuation-dissipation

theorem [17],

similarly

to the present lines of

reasoning

for the one-dimensional mirror.

The derivation of the force

(4.3)

as

presented,

where a finite

frequency

cut-off is maintained

throughout

the calculation of real and

imaginary

part of

r(fl),

calls for some

warning

about

(6)By

contrast, the Fourier transform r(n) of the response function with respect to the displacement (cf.

footnote 5) is not square-integrable, because at infinity Rer(n) tends to a non-zero constant. The subtraction of such a constant amounts to writing the dispersion relations (3.3a,b) for [r(n) r(0)]/n instead of for r(n).

(8)

the no-cut-off limit.

Suppose

one wanted to

apply dispersion

relations in the no-cut-off

limit,

I-e- try to determine

Imr(fl)

from

Rer(fl)

=

[(n I)/n]~ l/(127r) fl~, equation (4.la),

which follows from

(3A)

and

(2.7)

for zero cut-off-

Generally,

if

r(fl)

is not

square-integrable,

one must amend the

dispersion

relations

(3.3a,b) by appropriate

subtractions [14].

So,

if

Rer(fl) asymptotically approaches

a

(second-order) polynomial,

then one needs to subtract this

polynomial

and to write the

dispersion

relation

(3.3b)

for the function

[r(fl) r(0)

fl

r'(0) (fl~/2) r"(0)]/fl~,

which

yields

Im

r(n)

= Im

r(o)

+ n Im

r'(o)

+ Im

r"(o)

fl~

j" Rer(uJ) Rer(0) uJRer'(0) ~ Rer"(0)

7r

~

_~

~~°

(uJ

fl)

uJ3

This

dispersion

relation has three free parameters,

Imr(0), Imr'(0),

and Im

r"(0),

which can be determined

only through

additional

physical

information about the function

r(fl).

In the present case Re

r(fl)

is not

just

bounded

by

a

polynomial,

but is itself a

polynomial.

So subtraction of this

polynomial simply gives

zero for the remainder. The

analytic

continuation of a function

vanishing along

a finite line segment is of course zero

everywhere

in the

complex plane. Nevertheless,

one must not

jump

to the conclusion that

Imr(fl)

is zero; all that can be discovered from

dispersion

relations without cut-off is that

Imr(fl)

is a pure

polynomial

of up to second order in fl.

Only

the

dispersion

relations for

r(fl)

with finite

frequency

cut- off tell that in fact

Imr(0)

= 0

=

Imr"(0),

while

Imr'(0)

=

[(n I)/n]~ l/(67r~a).

One concludes that this kind of

approach (disregarding

any

physical

input

beyond

the

dispersion

relations without

cut-off)

is

incapable

by construction of

yielding

sufficient information about the reactive part of the

force,

or

consequently

about the mass renormalization.

There is however also a

pedestrian's argument

for

finding

the reactive part of the force in the no-cut-off

limit,

which

dispenses

with

dispersion

relations. If

Rer(fl)

=

[(n I) /n]~

l

/(127r)

fl~

then,

in the

light

of the remarks below

(4.3),

the response function

G(t r)

must contain a term

[(n I) /n]~

l

/(127r)

b(~)

(t r).

But this is acausal since it is non-zero for times r > t, even if

only

in an infinitesimal interval. So one has to allow for

causality by

hand and put the upper

integration

limit in

(3.I)

to t. Since this eliminates half of the

delta-peak

from under the

integral

one needs to

multiply by

two in order to retain the correct result for the

dissipative force,

which then

yields

the correct

expression

F(t)

=

'~ dr

b(~)(t r) fl(r)

n

~

67r

/~~

On

integrating by

parts twice one encounters a surface term

proportional

to

b(0) fl(t),

and one

is left with

F(t)

#

(~

~

b(0) fl(t)

+

/~

dT

b(t

T)

#(T)

,

n 67r

_~

where

again only

half of the

delta-peak

lies under the

integral. Eventually,

the Lorentzian

representation

of the delta function

b(r)

= lim

/~~

a-0 a + r~

gives b(o)

=

1/(a7r);

so one recovers the

expression (4.3)

for the force.

However,

the

simplistic

reasoning

just

presented

could do more harm than

good:

it proves fatal to

mingle

any of the

(9)

2158 JOURNAL DE PHYSIQUE I N°11

above

rough-and-ready

arguments with the

scrupulous

calculation where the finite

frequency

cut-off is

kept

until the last stage. In

particular,

the upper

integration

limit in the response relation

(3.I)

must not at a wrong

point

be

replaced by

t.

A comparison of the force

(4.3)

with results in

previous publications

on this

subject

is now in order. In reference [3]

exactly

the same

expression

is obtained

directly

from the

(perturbatively calculated)

radiated

photon

spectrum. The forces on

perfect

and on

partially

transparent

mirrors derived in references

[7-9]

have

dissipative

terms that

fully

agree with

(4.3),

but lack

terms

proportional

to

fi(t), although requirements

of

causality

seem to

explicitly

enter the

reasoning.

Also the force on a

perfect

mirror found in references 11, 2] possesses

only

the

dissipative

term, and no mass-renormalization term.

Since the derivation of the force on a

perfect

one-dimensional mirror

given

in references 11,2]

rests on field-theoretical

techniques,

mere

comparison

with the present

approach

cannot reveal

the reason for the absence of mass-renormalization terms; further

investigation

will not be

attempted

here.

A more detailed

comparison

is

possible

to the results of references

[7-9],

which also

employ

the

language

of linear response

theory.

So, for instance

just

above equation

(6a)

in [9] it is claimed that the Fourier transform xo(uJ]

(here: r(fl),

cf. footnote

(~))

of the

displacement

response "scales as uJ3 at low

frequencies".

In the present notation, this would claim that

r(fl)

is

proportional

to

fl~,

from which then had to be concluded,

along

the lines of the

paragraph following equation (4.3) above,

that there is no term

proportional

to

fl(t)

in the force.

Indeed,

no such term is

given

in references

[7-9] (at

zero

temperature).

However,

in references [7, 8]

causality

is found to lead to a difference in the mass of the mirror if

coupled

to or

uncoupled

from the

field,

which the authors of [7, 8] refer to as "low-"

or

"high-frequency

mass". This difference

equals II (67r~o),

which is

precisely

the amount

given by

the mass-renormalization term in

equation (4.3).

So

far,

the present writer has not

managed

to find out

why

the model

employed

in references

[7-9]

leads to such a mass difference but not to a reactive part in the force.

Finally,

a few remarks

concerning moving

mirrors in three dimensions. With reasonable

effort,

one can work out the

time-averaged

mean-square fluctuations of the forces on a

piston

in an infinite

plane [18],

on a

sphere

[19] and even on a flat disk

[20],

which is

practically equivalent

to

calculating

the correlation functions of the fluctuations in

time(~)

if the time-

averaging

function has been chosen to be a Lorentziau as in references

[18-20]. However,

it turns out that the

simplicity

of the calculation and as well as of the end result for the mean force

on a

moving

one-dimensional mirror is

peculiar

to one

dimension;

in the no-cut-off limit the spectrum of the fluctuations is

given by just

a power of the

frequency

fl. As known from field

theory

this is a consequence of the conformal invariance of the field

equations

for a massless scalar field in one space and one time dimension. In

simpler

words: in more than one dimension every

object

has a certain cross-section that the force is distributed over. Hence a

length

scale

(other

than the

cut-off)

enters the

problem,

in comparison to whose inverse the

frequency

fl in the fluctuation spectrum

x(fl)

can be small or

large.

Then

x(fl)

is a

complicated function,

even for

simple geometries,

and can at best be

approximated by (different)

powers of fl for very small or very

large frequencies.

This

implies

that the

expression

for the force is not as

simple

as in equation

(4.3)

for the one-dimensional mirror.

Generally,

it

depends

not

only

on various derivatives of the

velocity

at the present time t but also on

integrals

over the

history

before t. Nevertheless, under certain assumptions the expressions for the force can be

approximated

so as to

depend only

on derivatives of the current

velocity;

but such

approximations

are not

(7)The

calculation of the force correlation function at different points on the surface of the object under consideration is far more demanding; it is done for an infinite plane mirror in reference [12], but seems beyond manageability for spheres or even disks.

(10)

straightforward

to discover. In reference [2]

dissipative

and reactive terms on an infinite three- dimensional mirror, but

only

when

interacting

with a scalar

field,

have been found. In references [6, 21] the issue of

dissipative

forces on mirrors

moving through

the Maxwell vacuum in three

dimensions has been

addressed,

but reactive forces are not considered. The present writer

hopes

to tackle the three-dimensional

problem

in a

forthcoming publication.

Acknowledgments.

It is a

pleasure

to

acknowledge helpful

discussions with Gabriel

Barton,

an

illuminating

remark

by

Martin

Reuter, correspondence

with

Marc-Thierry

Jaekel and

Serge Reynaud,

and financial support from the Commission of the

European

Communities.

References

ill

Fulling S. A, Davies P. C. W., Proc. R. Sac. London A 348

(1976)

393.

[2] Ford L. H, Vilenkin A., Phys. Rev. D 25

(1982)

2569.

[3] Barton G, Eberlein C., Ann. Phys.

(1993)

to appear.

[4] Landau L. D, Lifschitz E. M., Statistical Physics, part 1, 3rd ed.

(Pergamon,

1985) pp. 122-126.

[5] Col~en-Tannoudji C., Dupont-Roc J, Grynberg G., Atom-Photon Interactions

(Wiley,

1992) com- plement AIV.

[6] Braginsky V. B, Kl~alili F. Ya., Phys. Lent. A161

(1991)

197.

[7] Jaekel M. T, Reynaud S., J. Phys. I France 3

(1993)

1.

[8] Jaekel M. T, Reynaud S., Phys. Lent. A 167

(1992)

227.

[9] Jaekel M. T, Reynaud S., Phys. Lent. A 172

(1993)

319.

[10] Moore G., J. Math. Phys. Ii

(1970)

2679.

[iii

Van Bladel J., Relativity and Engineering

(Springer, 1984).

[12] Barton G., J. Phys. A 24

(1991)

5533.

[13] Titcl~marsh E. C., Introduction to the Theory of Fourier Integrals, 2nd Ed.

(Clarendon,

1948)

section V.

[14]

Hilgevoord

J., Dispersion Relations and Causal Description

(North-Holland, 1960)

chapter 3.

[15] Handbook of Mathematical Functions, M. Abramowitz, I. A. Stegun Eds.

(US

Govt. Printing Office,

1964).

[16] Rol~rlich F., Classical Charged Particles

(Addison-Wesley, 1965).

[17] Dalibard J., Dupont-Roc J, Cohen-Tannoudji C., J. Phys. France 43

(1982)

1617;

J. Phys. France 45

(1984)

637.

[18] Barton G., J. Phys. A 24

(1991)

991.

[19] Eberlein C., J. Phys. A 25

(1992)

3015.

[20] Eberlein C., J. Phys. A 25

(1992)

3039.

[21] Barton G., New aspects of the Casimir effect, "Cavity Quantum Electrodynamics" Supplement:

Advances in Atomic, Molecular and Optical Physics, P. Berman Ed.

(Academic Press)

to appear.

Références

Documents relatifs

Calibrated force measurement in Atomic Force Microscopy using the Transient Fluctuation Theorem 6 applied voltage V (t). This gives a reference

Thereby, a consistent picture of the out-of-equilibrium state of colloidal gels appears, with no violation of the FDT for a fast-relaxing observable, like velocity, and an

[r]

By using the ACFD formula we gain a deeper understanding of the approximations made in the interatomic pairwise approaches, providing a powerful formalism for further development

Baker et al. [2004] were pioneers to show that radiation belt populations are sensitive to the core plasmasphere distribution and specifically to the position of the

Prove the Inverse Function Theorem (the statement is below). (Hint: reduce to the case considered in the

Prove the Inverse Function Theorem (the statement is below). (Hint: reduce to the case considered in the

The relative solvation free energies of these ions were calculated using thermodynamic integration for the non-polarizable, Drude, and QM/MM methods. The non-polarizable MM method is