• Aucun résultat trouvé

Reflecting slow atoms from a damped resonator

N/A
N/A
Protected

Academic year: 2021

Partager "Reflecting slow atoms from a damped resonator"

Copied!
16
0
0

Texte intégral

(1)

HAL Id: jpa-00248097

https://hal.archives-ouvertes.fr/jpa-00248097

Submitted on 1 Jan 1994

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.

Reflecting slow atoms from a damped resonator

Marco Battocletti, Berthold-Georg Englert

To cite this version:

Marco Battocletti, Berthold-Georg Englert. Reflecting slow atoms from a damped resonator. Journal

de Physique II, EDP Sciences, 1994, 4 (11), pp.1939-1953. �10.1051/jp2:1994241�. �jpa-00248097�

(2)

Classification

Physics

Abstracts

42.50 03.65N

Reflecting slow atoms from

a

damped resonator

Marco Battocletti

(*)

and

Berthold-Georg Englert (**, ***)

Sektion

Physik,

Universitit Mfinchen, Theresienstrasse

37/III,

D-80333 Mfinchen,

Germany

(Received

17

May

1994, revised 25

July1994, accepted

7

September1994)

Abstract. Previous studies of the reflection of slow excited atoms

entering

an empty mi-

crowave resonator are extended

by including

the dissipation of the

photon

energy into the de-

scription.

For

typical

parameter values, this leads to a reduction of the reflection

probability by

a factor of ten. Somewhat

unexpectedly,

we also find that the transmitted atoms may

gain

in kinetic energy as a result of the

photon damping.

1. Introduction.

The center-of-mass

(CM)

motion of very slow two-Ievel atoms, which interact

predominantly

with one mode of the

quantized electromagnetic

field in a microwave resonator of

ultrahigh quality,

was the

subject

of two recent

publications.

Since the

strength

of the interaction

changes enormously

when the atom enters the

cavity,

a

position dependent

force arises which affects the atomic CM motion. One finds

that,

if the interaction with the

privileged photon

mode is

resonant

[I],

the

cavity

may act like a

semitransparent

mirror for the atoms and reflect

50%

of

them; if, however,

the interaction is well off resonance

[2],

either all atoms could be reflected

or none,

depending

on the

sign

of the

detuning.

These

previous investigations

were facilitated

by

a number of

simplifying assumptions.

In

particular,

the

damping

of the

photon

field was

ignored,

which can

hardly

be

justified. For,

the

velocity

of the atoms must be about 5

mm/s

or less

(I],

so that it would take them a few seconds at least to traverse the resonator which is a

couple

of centimeters in

length.

This is a rather

long

time on the scale set

by

the

photon lifetime,

which may be as

long

as half a second in a resonator of very

high quality,

but more

typical

lifetimes are

considerably

shorter.

Clearly,

in a realistic calculation the

photon damping

must be taken into account

(and possibly

other

(*)

Now at Institut for

Physikalische Chemie,

Universit£t

Mfinchen,

Theresienstrasse 41, D-80333

Mfinchen, Germany.

(**)

Mailing address: Sektion

Physik,

Universitit Mfinchen, Am Coulombwall I, D-85748

Garching,

Germany.

(***)

Also at Max-Planck-Institut ffir

Quantenoptik, Hans-Kopfermann-Strasse

I, D-85748

Garching,

Germany.

(3)

additional

dynamical

contributions as

well).

It is the

objective

of the

present

paper to

report

such a more realistic treatment.

Another

simplification adopted

in references

[I]

and [2] is the

disregarding

of other atomic transitions than those between the two

masing

levels that are

strongly coupled

to the

privileged

mode of the resonator.

Clearly,

if such transitions

happen

at rates that are

large

on the scale set

by

the interaction

time,

their

dynamical

effect would be

quite significant.

In the

present

paper,

however,

we shall

pretend

in the tradition of references

[I]

and [2] that the atom is well modeled

by

two relevant internal states. More about this will be said in the summary.

2.

Equations

of motion.

We

adopt

the notational conventions of reference

[Ii.

The excited state of the two-level atom is called a, the deexcited one is

fl;

the

operators

a =

(fl) (al

and

at

=

la) (fl(

effect the atomic

transition;

the

energetic

level

spacing

is

h(Q

+

A),

where Q is

(27r times)

the

photon frequency

and A denotes the

detuning;

the

photon

ladder

operators

are a and

at;

the CM

position

and momentum operators are x and p. The state of the

system

is

specified by

a

density

operator P(t;

z, p; a,

at

a,

at

which is

a function of all

dynamical

variables. The

explicit

time

dependence

of P is determined

by

the master

equation

~ P

=

I(H, Pi

+ CP

(2.1)

The

unitary

evolution involves the Hamilton

operator

H =

(

+

hoaia

+

h(Q

+

A)aid hg(x)(ala

+

aai), (2.2)

m

where m is the mass of the atom and the

spatially dependent

Rabi

frequency g(z)

measures

the

coupling strength.

Inside the resonator, that is: in the range 0 < z <

L, g(x)

is

relatively large,

outside it vanishes. The

nonunitary

evolution

generated by

the Liouville

operator £,

£P

=

(aiaP 2aPai

+

Paia), (2.3)

models the

photon damping

in the usual way,

whereby

the

symbol

A denotes the

decay

rate of the

photon

number. For

simplicity,

we

opt

for a

photon

reservoir with temperature zero, so

that the resonator does not contain thermal

photons.

We shall restrict the discussion to situations in which the atom is

initially

excited and

approaches

the resonator from the

negative

x range, and there are no

photons

in the

cavity.

Thus at t = 0 we have the

density operator

Pot so(x,p)ala

:

exp(-aia)

:,

(2.4)

where so

(x, p)

is the initial

(reduced) density operator

of the CM

motion,

and the

pair

of colons indicates normal

ordering

of the

photon

variables as usual. This initial

density operator

is a

product

of three

factors,

one for each

degree

of

freedom,

because there is no

entanglement

at t = 0. Both the

probability

for

finding

the atom in the x > 0 range and the

probability

for

finding negative

momentum values have to vanish

initially,

which

imposes

the

requirements

tram

(6(x z')so(z, p))

= o for

x'

>

0,

tram j6(p p')so(x, p)j

= 0 for

p'

< 0.

j2.5)

(4)

Here and below the

primed symbols

x' and

p'

denote numbers in

particular, eigenvalues

of the

operators

z and p and trcm

signifies

the

injunction

to trace over the CM variables.

A convenient ansatz for the

density operator

is

p =

~p(o)

~

l

~(-) ~(++)

~ l

~(+) ~(--)

l

~i p(+-) /~p(-+) (2 6)

2 2 2 2 '

where

~(t,

z,

p), ~(+)(t,

z,

p),

and

~t(t,

z,

p)

account for the

dependence

on the time and on the CM

variables,

and

p~°~ =

aai exp(-a ia)

,

pl++~

=

(at

+

at )p(°~ (a

+

a) (2.7)

are time

independent

operators of the

photon

variables and the internal atomic variables.

(The

first

/second

+ on the left-hand side

corresponds

to the first

/second

one on the

right-hand

side.

This ansatz is motivated

by

the

identity

(ala

+

aai)~

=

ala

+

ala, (2.8)

which

implies

that the

coupling

term in the Hamilton

operator (2.2)

commutes with the con- tribution

proportional

to Q. One is therefore invited to

expand

P into

dyadic products

of the

eigenstates

of

at,

+

aai,

«dressed states" of one kind.

Only

the

eigenstates

to the

eigenvalues

0 and +1 are needed for the purposes of this paper, and of the nine

possible dyadic products

thereof

only

the five of

(2.7)

show up. In

particular, p(°) projects

to the state in which the atom is deexcited and no

photons

are in the resonator. For the initial

density operator (2.4),

this state can

get populated solely

as a result of the

photon damping.

By construction,

the operators

(2.7)

commute with the sum

ala

+

ala;

other

properties

that

are relevant here are listed in table I.

Upon tracing

over the two-level

degree

of freedom

(tr,)

and the

photon

variables

(tra)

we obtain the

(reduced) density operator

s,

s(t,

z,

p)

= tr,tra

P(t;

z, p; a,

at;

a,

at)

,

(2.9)

so that s is a

weighted

sum of ~,

A(+),

and

A(~),

s = ~ +

(>(+)

+

>(-)). (2.10)

This CM

density operator

is the

quantity

that we are

mostly

interested

in,

because it contains all information about the CM motion. For

instance,

the

probability

R that the atom is reflected

as a result of the

interaction,

is

given by

the

negative-momentum

content of s,

R " trcm

(Q(-P)s(T,

z,

p))

,

(2.ii)

where q is Heaviside's unit

step function,

and T is any instant after the interaction.

After

inserting

the ansatz

(2.6)

into the master

equation (2.

I we find the

equations

of motion

~

~

'~

~~~~ ~ ~~~~~~ ~~~~~ ~

(ll

+

~ti

~ i ~ /~

~i)

~~

~D

A

~ 2 '

~

~ ~ ~ ~ ~~~~~~ ~

~~~~~

~ ~~ ~ ~ ~~~~~

~~~~), (2.12)

(5)

Table I, Fundamental

properties

of the

operators (2.7).

Note in

particular

the

eigenvalues

exhibited in the fourth and fifth

columns;

the

superscripts (++)

or

(0)

reflect these

eigenvalues.

p tratr«p («'«. PI ta'« +a«')p pta'« +a«') £p

pt0) 1 o o o o

~(++) i

)~~j+-)

~t-+)) +pt++) +pt++)

fjpt++)

pro)~ ~ A(pt+-) + pt-+))

p(--) i

+)(p(+-)

-p(-+)) -pi--) -p(--)

-§(p(--)

p(0)) +

f(p(+-)

+p(-+))

pt+-) o

)(p(++)

pi--)) +pt+-) -pt+-)

f(pi+-)

+ p(o)~ ~

f~~j++)

~ ~j--)~

~(-+) o

+)(p(++)

p(--)) -pt-+) +p(-+I

-~~p(-+)

+p(o)) + A(p(++) + p(--))

as well as

~'

~ =

~

j(+)

+

j(-)

+ ~ +

~i) (2 13)

Dt 4 '

where the low dot indicates a

symmetrized product,

9.ll "

(911

+

~tg), (2.14)

and

~

is defined

by

Dt

~~ ~t~~ 'm'~~

'

~~'~~~

for a free

particle

this would be the

comoving

time derivative.

Equations (2.12)

determine the

auxiliary

functions

A(+)

and ~t, which are needed in

(2.13)

for the

computation

of

~(t,x,p).

Then

s(t,

x,

p)

results from

(2.10).

The initial conditions at t

= 0 are

~(0,z,P)

"

0, A~~~(0,X,P)

"

~lL(°>X>P)

"

80(X,P) (2.16)

as found

by comparing (2.6)

with

(2A).

In

passing

we note that the reverse

situation,

in which the atom is

initially

deexcited and there is one

photon

in the resonator, is treated

analogously

with the initial conditions ~

=

0, A(+)

= ~t = so instead of

(2.16).

On resonance

(A

=

0)

and without

photon damping (A

=

0), equations (2,12)

are

decoupled.

This is the case treated in reference

iii.

The

equation obeyed by A(+)

is then a

unitary

master

equation

with a Hamilton operator

p~ /(2m)

+

hg(x).

If we take

g(x)

to be

positive,

as we shall do

throughout,

then this

component

encounters a

repulsive potential

and is

reflected, provided

that the initial kinetic energy is not sufficient to cross the barrier. The component

A(~), by

contrast, evolves

unitarily

with a Hamilton

operator p~/(2m) hg(x).

It moves

through

an

attractive

potential

and is transmitted. As mentioned

above,

the ~ channel

corresponds

to

having

the atom deexcited and no

photons present. Consequently,

there is no

potential

energy

in this

channel,

and in the absence of

photon damping

this channel cannot be

populated.

indeed,

in this A = 0

situation, equations (2.13), (2.16),

and

(2.10) imply

that ~

= 0 and

s =

j(+)

+

j(~)

at all

times,

so that we

immediately

arrive at the conclusion of reference

[1],

tram)jy

that

50%

each of the atoms are reflected and transmitted. We note further that the

(6)

nonhermitean ~t and

~ti components

are measures for the

quantum

coherence between the

A(+)

and

A(~)

channels and evolve

accordingly.

Similarly,

with

sufficiently large detuning (usually (A(

» A must

hold,

but

(A(

»

(g(x)(

is not

required)

and without

photon damping (A

=

0), equations (2.12), (2.13),

and

(2,16)

convey

the message of reference

[2].

This is more

easily

seen after

rewriting

the

equations suitably,

and we shall return to this matter in section 4.

For the numerical

treatment,

we represent the CM operator functions s, ~,

A(+),

and ~t

by

their

Wigner

functions s~, ~~,

A~f~,

and ~t~. A factor of

e~"~

is absorbed in the definition of A~f~ and ~t~, so that

sit,z,p)

-

sw(t,x',v'),

~(t, Z,P)

-

~w(t, x', p') j(+J(t,

,

z,

p)

-

A[+)jt, z', p'~~-At/2

ll(t, z,P)

-

llw(t,z~, p')e~~"~ (2,17)

is the actual

correspondence

between the

operators

and their

Wigner

functions. Since s, ~, and

A(+)

are hermitean

operators,

the

Wigner

functions s~,

~~, and A)+~ are real. The

Wigner

function of

~ti is,

of course, the

complex conjugate ~t[

of ~t~.

The action of the differential operator

(2.15)

is then

given by

&~&~h~iw ~~'~~~

For the commutators and

symmetrized products

that involve

g(x),

the WKB

approximations I[g(z), j(+Jj

_

_&~9(~')

~

j(+le-At/2

ax'

dp'

~

9(X)

11 -

9(X')llwe~~~/~ (2.19)

are very well

justified,

because the de

Broglie wavelength

(+~

10~~cm)

associated with the CM motion is very small

compared

to the

macroscopic

distance (+~

lcm)

over which

g(x') changes considerably.

The relative size of the

leading

corrections to these WKB

approximations

is

given by

the

squared

ratio of both

length

scales (+~

lo~~); consequently,

there is no need to go

beyond

the semiclassical relations

(2.19).

The

equations

of motion

(2.12)

and

(2.13)

are thus turned into

()

+

f(x') ()

Al+~ = (/1w +

/1S)

~ i

)

(/1w

/1S),

()

+

ig(x'))

~t~ =

~

(A[+)

+

Al

I ~

(A(+~ Al

,

(2.20)

t 4 2

and

~

~~ =

~e~~~/~(A(+~

+

Al

+ ~tw +

~t$), (2.21)

where

~

~~~~~ ~

~~$~~~'~ ~~'~~~

is the force that a classical

particle

would

experience

in a

potential hg(z').

The initial conditions

(2.16)

are

simply

translated into

~w(o,z',p')

=

o, >1+~(o,z',p')

=

-~w(o,z',p')

=

sow(x',v'), (2.23)

(7)

and

~~ = ~~ +

-e-At/2(j[+)

+

j[-)) (2.24)

2

replaces equation (2.10).

The

requirements (2.5)

now read

m

dp'sow(x', p')

= 0 for

x'

>

0,

-m

dx'sow(x',p')

= 0 for

p'

<

0, (2.25)

or, if sow is

non-negative,

more

simply sow(x', p')

= 0 when

x'

> 0

or

p'

< 0.

(2.26)

The reflection

probability (2.ll)

is available as

m

, ,

R =

/ dx' / dv' ~~~(jj'~'

,

(2.27)

-m -m

which exhibits the convention that we are

employing

for the normalization of the

Wigner

functions.

Equations (2.20)

and

(2.21) provide

the basis for our numerical calculations. In the resonant

case,

they

can be used as

they stand;

results are

presented

in section 3. When the

detuning

is

large enough,

an adiabatic

approximation

as in reference [2] is

fitting,

which

requires

that the

equations

of motion are rewritten in a form more suitable for this purpose. This is discussed in section 4.

To be

specific,

we shall consider the

63p~/~

-

61p5/~

transition in

8~Rb

that is used in the

Garching

one-atom-maser

experiments

[3]. The mass of the atom is then m

= IA x

10~~~

g.

We take a

cavity length

of L

= 2 cm and a

photon decay

rate of A

=

2.5s~~

The initial average momentum of the atom is

(p)o

" 7 x

10~~~ gcm/s,

which

corresponds

to a

velocity

of 5 mm

Is,

to a kinetic energy of I.I x

10~~~ eV,

and to a de

Broglie wavelength

of 9.5 x

10~~

cm.

The effective Rabi

frequency g,

which is related to the

(positive) g(x')

function

by

§L

=

/dx' g(x')

,

(2.28)

is chosen to

equal

g = 44 kHz.

Indeed,

the energy

fig

= 2.9 x

10~~~

eV associated with the interaction is

markedly larger

than the kinetic energy of the

arriving

atom.

A realistic

shape

for

g(x)

would be

~jsin(7rx/L)

for 0 <

x <

L,

g(x)

= 2

(2.29)

0 elsewhere

,

if the

coupling

is to a TE mode as is the usual

experimental

situation. None of the conclusions of this paper

depends crucially

on the

specific shape

of

g(x).

We shall therefore facilitate the

numerical

integration

of the

equations

of motion

by using

the more convenient function

e/2

~ ~ "

~cosJ~2 (e(x/~

i

/~~~

(2,30~

(8)

because the

resulting

force

(2.22)

is such that the classical

equations

of motion possess an

analytical

solution. These classical

trajectories

are the characteristics of the

partial

differential

equations obeyed by

A(+~ and

Al

~, and for this reason

(2.30)

is

advantageous,

as will be

discussed below. We fix the

parameter

e

by requiring

that

g(x

=

0)

=

g(x

=

L)

is of the

27 maximal value

g(x

=

L/2)

=

ge/2;

thus

cosh(e/2)

=

/fl.

Then

g(x)

is

practically

zero outside the range 0 < x < L. Of course, similar e values are

equally good.

In reference

[I]

the

simple

mesa function

has

been

used.

It

has

its

approximate experimental ealization in TM modes.

The

the electromagnetic energy hrough the openings, through which the atom enters and leaves,

is rather

large

in this geometry and

does

not allow for the very long photon

we eed. evertheless, we

shall briefly consider (2.31)

in

order to

makedirect

contact with

3. Resonant interaction

IA

=

0).

In this section we shall deal with the case of resonant

interaction,

so that we set A

= 0 in

equations (2.20).

We

begin

with a reconsideration of the situation of reference

iii,

where the

mesa function

(2.31)

has been used. For this

g(x),

an

analytical

solution of the

equations

of motion

(in

WKB

approximation)

is

available, provided

that the atom is

initially

well localized and also has a rather well defined

velocity,

viz. [4]

sw

(T, x', p')

= sow

x'

+

p'T/m, -p') (3.la)

2

+

(1

+

g2 ~ll14)21~~~ Ill1

~°~

(~' l~ Al

~

?'~'/~ ?'l

(3.lb)

+1

g~

1/~~~

exP

li~ll

Sow

(ix' P'T/m) /P~)~ ~l

(3.lc)

(A/8)~ mL)

~§~ (A/4)~

~~~

2(P)o

x sow

lx'

+

~,i'-

~~

ii

L

P'T/ml IF')

~

,

P~)

,

(3. id)

where jl is a momentum measure for the interaction energy:

fig

=

jl~/(2m).

The first term

(3. la)

is the reflected

intensity.

It accounts for

50%

of the total flux. Since the A(+~

component

is here reflected at x'

= 0 without ever

entering

the resonator, the reflection

probability

is not affected

by

the

photon damping.

The three other terms in

(3.I)

are transmitted intensities of various momentum contents. The

component (3.lb)

has the same momentum distribution as

(9)

the initial CM

density operator

sow

(x', p');

this is

just

like the

A(~) component

in the

undamped

case, which

speeds

up when

entering

and slows down when

leaving

the interaction

region.

The

component (3.lc)

has

gained

the amount jl~

/(2m)

=

fig

in kinetic energy

during

the interaction.

This occurs because the

damping couples

the channels in

(2.20)

and

(2.21),

so that the

A(~)

component,

which is accelerated at x'

=

0,

feeds the ~

component,

which

experiences

no force at x'

= L when it leaves the resonator and is therefore not decelerated. The

A(+)

channel is also fed

by A(~J,

and

A(+)

is accelerated at x'

=

L,

so that a second

gain

of

fig

is

possible

and this

produces

the last contribution

(3. Id).

For the parameter values

specified

at the end of the

preceding section,

the

exponential

factor

exp(- ~AmL/(p)o)

Qi 0.01 is rather small and the

2

ratio

(A/8)~ /[j~ (A/4)~]

ci 5 x

10~~~

is

tiny,

and so the dominant transmitted

component

in

(3.I)

is

(3,lc),

which represents

99%

of the transmitted atoms. The

remaining

I$lo are in the elastic channel

(3. lb). Thus, owing

to the

photon damping

almost all atoms

gain fig

in kinetic energy when

traversing

the resonator.

This

example

demonstrates an

intriguing

constructive effect of the

photon damping:

a mech- anism for

converting

the

atom-photon

interaction energy into kinetic energy is

provided.

In

principle,

this effect is also present for a more realistic

g(x'),

such as

(2.30).

Since the

entering

and

leaving

is a much less dramatic event for such a smooth

g(x'), however,

the

periods during

which the forces of

differing signs

and

strengths

are

acting

are not well

defined,

and there- fore momentum transfers of various amounts are

possible.

As a consequence, the momentum

distribution is broadened rather than

split.

For

g(x')

of

(2.30)

we have

integrated (2.20)

and

(2.21) numerically.

The classical

phase

space

trajectories

T -

(X(+)(T; x', p'), P(+)(T; x', p'))

that are determined

by

those solutions of the classical

equations

of

motion,

~

x(+)

=

p(+)

~

p(+)

= +

f( x(+)) j3 2)

aT m ' aT '

which run

through ix', p')

at time T

=

o,

(X(+~, P(+~)

=

ix', p')

for

T =

0, (3.3)

are used to turn the

partial

differential

equations

for A)~~ into

ordinary

differential

equations,

~

A[+~

(t

+ T,

X(+~,P(+~)

=

~

~tw

+

~t$) (t

+ T,

X(+), P(+~). (3.4)

These are

supplemented by

the

equations

for ~tw and ~~,

~ r

exp

(2i /dT'g(x'

+

p'T'/m) /Jw(t

+ T,

x'

+

p'T/m, p')

dT

o

r

=

~

exp 2i

/dT'

g

(x'

+

p'T'/m) (A(+)

+

Al ) (t

+ T,

x'

+

p'T /m, p')

,

(3.5)

4

and

~

~wjt

+ T,

x'

+

p'T/m, p')

dT

~

A

~-A(t+r)/2 (j(+)

+

jj-)

+

~~ +

~j) (t

+ T,

x'

+

v'T/m, v') 13.6)

4 ~

(10)

16

,~",

b

~ ,.,~' ",

i a@@=i;[

E ".,

~

o

..:...().

~~

)

-f~~~c

lo

10 0 20 30

x' (mmj

Fig.

I. Evolution of

sw(t, x', p')

for

g(x')

of

equation (2.30)

in the absence of

photon damping

and

for resonant atoms. The plot shows

(a)

the initial distribution

(3.7)

at t

= 0 before the atom enters the resonator; the distributions

during

the interaction at

(b)

t

= 1.8s and

(c)

t = 3s; and

(d)

the

final distribution after the interaction at t = T

= 4.5s. The contour lines enclose

20%, 50il,

and 80il of the total

probability.

The dashed curves mark the classical

phase

space

trajectories

that start in

((x)o, (p)o)

at t = 0,

computed according

to

equations (3.2)

and

(3.3).

The x' range covered is

-10 mm < x' < 30 mm, and the

p'

range is -1.4

x10~~~ gcm/s

<

p'

< 2.24 x 10~~~

gcm/s,

which

corresponds

to a velocity range of

-10mm/s

<

p'/m

< 16

mm/s.

For 0 < x' < L

= 20 mm the atom

is inside the resonator.

Upon integrating

from T

= -dt to T

= 0 with the aid of the

trapezoidal rule,

one can express

A(~~,

~tw, and ~~ at time t in terms of their values at time t dt.

This,

in essence, is the

relatively simple

manner in which we have

computed

the numerical solutions. The

g(x)

of

(2.30)

is

advantageous

for this purpose because it allows for

analytical

solutions of

(3.2).

In

particular,

a numerical

integration

is not needed at all for A

= 0.

This

analytical

solution for A

= 0 is

reported

in

figure

I for reference. The

plot

shows

s~(t,x',p')

before

(t

=

0), during it

= 1.8s and t

=

3s),

and after

it

= T

=

4.5s)

the interaction. One

clearly

sees that s~ consists of two

components:

the

A(+)

branch that is reflected and the

A(~)

branch that is transmitted. Here and in the

following figures

the initial

distribution is

always

the same

Gaussian,

~°~~~" ?'~ /~6p

~~~

l~

~'

~~~~

~

~' j~~~~ ~

~~'~~

with the

expectation

values

(x)o

" -3 mm and

(p)o

" 7 x

10~~~gcm/s

as well as their

spreads

fix

= I mm and

6p

= 7 x

10~~~ gcm/s.

This

corresponds

to an initial

velocity

of

(p)o/m

=

smm/s

with a

spread

of

6p/m

=

0.smm/s.

Of course, this sow does not vanish

exactly

when x' > 0 or

p'

<

0,

but the conditions

(2.25)

are

obeyed

with

sufficiently high

accuracy.

In

figure

2 we present

sw(t, x', p')

when the

photon damping

is taken into account. In contrast to the A

= 0 case, the reflected

intensity

is

strongly

reduced. Rather than R

= 50$lo, we obtain

(11)

16

(~)

10 0 20 30

X'

(mmj (b)

10 0 ?0 30

x'

(mni]

16

(c)

~

I i

°

10

10 0 20 30

x'(mmj

Fig.

2, Evolution of

sw(t, x', p')

with photon

damping

for resonant atoms. The initial distribution is the same as in

figure

I. We show the distributions at the instants

(a)

t = 1.8 s,

(b)

t = 3s, and

(c)

t = T

= 4.5 s.

Along

the contour

lines,

sw

equals 20il, 13il, Gil,

and 0.4il of the maximal value at t = 0. The x' and

p'

ranges are the same as in

figure

1.

(12)

a reflection

probability

of R =

4.8$l.

The situation is here different from that of

eqaution (3.1)

because the atom has to enter the resonator before it can

get exposed

to the force

generated by (2.30)

in its full

strength.

But as soon as the

atom-photon coupling produces

a

nonvanishing photon

number

(ala),

the

photon damping

becomes effective and the force-free ~ channel is fed at the expense of the

A(+)

and

A(~)

channels. As a consequence, the

repulsive

forces do not act

long enough

to reverse the atom's

velocity

with a

large probability.

At the final time t = T = 4.5 s, there is still a considerable

probability

for

finding

the atom inside the resonator.

Nevertheless,

this instant is "after the interaction" in the sense of

equation (2.11)

because

only

~~ is

nonvanishing

inside the

cavity,

so that the forces have

already

ceased to act. Please note

also how the momentum distribution is smeared out. This is the smoothened version of the

splitting

that we have observed in

(3.1).

4. Nonresonant interaction

IA

<

(A(

<

j).

The terms

proportional

to A in

(2.I)

can be included on the

right-hand

sides of

(3.4)

and

(3.5).

But when the

detuning

A is

markedly larger

than the

photon decay

rate

A,

these terms are troublesome in a numerical

integration,

because

they

lead to rather

rapid

oscillations. In other

words,

the "dressed-states basis"

(2.7)

is inconvenient under these circumstances.

Therefore,

we switch from A(~~ and ~t~ to another set of

phase

space functions

I)+~, Al

~, and

ji~

in accordance with

jj+)

+

jj-)

=

jj+)

+

jj-)

,

>[+~ Al

=

(i[+~ i[ ))

cos ~2 +

(jiw

+

ji$)

sin ~2,

~tw +

~t$

=

(jiw

+

ji$)

cos ~a

(I[+~ i(

~) sin ~a,

vw

AS

=

fiw AS

,

14.1)

where the

angle ~a(x')

is defined

by

g(x')sin(~(x'))

=

( cosl~(x'))

with

<

~(x')

<

(4.2)

The

corresponding

linear combinations of the

p(++) operators

in

(2.7)

are

dyadic products

of the x

dependent eigenstates

of

g(x) (ala

+

ala) hat,

=

g(z) [ala

+

ala (aid aai)

tan

(~a(x))j

~

(4.3)

2

the relevant

eigenvalues

are

~

+

g(x)/I

+

tan~ (~a(x))

=

~

+

~/(g(x)]2

+

(A/2)2 (4.4)

These

eigenstates

are the "dressed states"

employed

in reference [2] for a

study

of the nonres- onant situation in an adiabatic

approximation.

The substitution

(4.I)

turns

(2.20), (2.21),

and

(2.24)

into

(D

+

/(zi)

~

j(+)

=

+~j(+)si~~

Dt

dp'

~ 2 ~

(~

cos ~a +

f(x')

sin ~a

~ +

~ ~~ (fi~

+

#*

2 2

~P'

m dX' ~ ,

(13)

()

+

i§(x'))

fiw

=

A

cos ~a +

f(x')

sin

~a

~

(I(+~

+

i( )

t 2

dP'

+

P' ~l~ j(+) j(-)) (4 5)

2 m ax' ~ ~ '

and

D

~

A

~-At/2 jj+)

~

jj-)

~

~

~

~* (4 6)

Dt ~ 4 ~ ~ ~ ~

as well as

sw = ~w +

-e~~"~

(i(+~

+

i[ )

,

(4.7)

where,

consistent with

(4A), 1(x')

=

g(x') cos(~(x'))

+

)

sin

(~(x'))

=

~/lg(x')12

+

(A/2)2 (4.8)

is the effective local Rabi

frequency,

and

f(x')

=

f(x') C°S(~2(x'))

=

-h£§(x') (4.9)

is the effective force associated with it.

The adiabatic

approximation

of reference [2] amounts to

discarding

those terms on the

right-hand

sides of

(4.5)

that involve

multiplications

with

p'

or differentiations with respect to

p'.

This is well

justified

for the

typical experimental parameters

that we have in mind.

For a

detuning

of A

=

+looHz,

say, which is

large compared

with A

= 2.5 Hz and small

compared

with

g

=

44kHz,

we

get

~a ci

lo~~

inside the

resonator,

so that the relative size of

f(x')

sin ~a

d/dp' compared

to A cos

~a is 2

where we have used that the

product

of

typical

x' and

p'

ranges is +~

lo~h

for atoms in a beam. Likewise one finds that the

(p'/m)d~a/dx'

terms are

equally

small. Once these small

contributions are

neglected

in

(4.5),

the numerical

procedure

of

equations (3A)

and

(3.5)

can

be

employed

with the necessary

changes.

Inasmuch as the

limiting

values of

~a(x')

are

~

/2

for A > 0

~~~~~

g(z')

# 0

~~'~~~

~g(X')

"

_~

/2

for A < 0

the initial conditions

(2.23)

appear here as

~w(o,z',p') =Jw(0,x',v') =o,

I(+~(0, x', p')

=

~~°~~~"?'~ )~ ~'

°~

,

~~ ~~~'~"?'~ sow~x',

p') ~~ ~~'~~~

(14)

16

(a)

~

E E

I

do

~

10

10 0 20 30

x'

[mm]

16

16)

10

10 0 20 30

x' (mm]

ic)

~

E E

~ I

~

lo

10 0 20 30

z'

(mmj

Fig.

3. Evolution of

sw(t, x',p')

with

photon damping

for nonresonant atoms with

a

positive

de-

tuning

of 6

= 100 Hz. The initial distribution is the same as in

figure

1. We show the distributions at the instants

(a)

t

= 1.8s,

(b)

t

= 3s, and

(c)

t

= T

= 4.5s. The contour lines have the same

significance

as in

figure

2. The x' and

p'

ranges are the same as in

figure

1.

(15)

The close

analogy

between

(4.5)

and

(2.20)

tells us that the

I(+~ component

is reflected and the

Al component

is transmitted for A

=

o,

and thus we see that all atoms are reflected for A > o and all transmitted for A < o if there is no

photon damping.

This

is,

of course, what

has been found in reference [2].

In

figure

3 we show the results of a numerical

integration

of

(4.5), (4.6),

and

(4.7)

in the adiabatic

approximation,

for a

positive detuning

of A

= +100 Hz.

[For

the sake of

simplicity,

we have set

§(x) equal

to the

g(x)

function of

(2.30),

so that we can make use of the same classical

trajectories

as in the resonant situation. The

g(x) corresponding

to this

§(x)

differs but little from the one in

(2.30).]

The

comparison

with

figure

2 identifies two main effects of the

detuning: ii)

it increases the reflected

intensity

to R

=

9.1$l,

that

is, roughly by

a factor of

two; iii)

it suppresses the

large p' contribution,

which would

signify

a

gain

in kinetic energy, in the transmitted

intensity.

This is as

expected.

We thus find a reduction of the reflection

probability by

a factor of

ten,

both in the resonant case

(4.8~

rather than

Soil)

and in the nonresonant case

(9.1%

rather than

loo%).

5.

Summary.

We have extended

previous

studies

[1,

2] of the reflection of slow atoms from a resonator

(that

does not contain

photons initially) by including photon damping

into the

description.

This is necessary

because,

for

typical experimental parameters,

the

atom-photon

interaction lasts for several

photon lifetimes,

so that the

dissipation

of the

electromagnetic

energy cannot be

neglected

in a realistic treatment.

We have found that the

photon damping

leads to a reduction of the reflection

probability, roughly by

a factor of ten. Rather

unexpectedly,

we have further observed that transmitted

atoms may

gain

in kinetic energy. This

phenomenon

is well understood as a

dynamical

conse-

quence of the

coupling

between otherwise elastic

channels, brought

about

by

the

photon damping.

We have

consistently

referred to the parameters of the

Garching

micromaser

experiments

[3]

in order to stay in close touch with reference [1].

Now,

since that

requires

atomic CM velocities of about 5 mm

Is,

a reflection

experiment

of the discussed kind cannot be

performed

within the

gravitational

field on the surface of the earth. As mentioned also in reference [2] a

microgravity

environment would be

mandatory.

But

perhaps

progress in

increasing

the

quality

of

optical

resonators will one

day

enable the

experimenters

to realize an

analogous experiment

with an

optical

transition. Then the much

larger

Rabi

frequencies

would allow for the use of faster atoms and the

gravitational pull

would no

longer

be

disturbing.

The use of an

optical

transition rather than a microwave transition between

Rydberg

states would offer additional benefits. The deexcited state of the

optical

transition can be identical with the true

ground

state of the atom. Then transitions that lead out of the

masing

two-level

system briefly

mentioned in the Introduction would be of no concern. Of course, the

lifetime of the excited state, which results from spontaneous emissions into other modes than that of the

optical resonator,

must not be short

compared

with the interaction time needed for the reflection. In contrast to the situation dealt with

explicitly above,

in such an

optical experiment

one would

presumably prefer ground-state

atoms

entering

a resonator

containing

one

photon (or

a few

photons perhaps).

In view of the remark after

equations (2.16)

this would not constitute a difference in

principle.

(16)

Acknowledgments.

We would like to thank G.

Rempe

for valuable discussions.

References

iii Englert

B.-G.,

Schwinger

J., Barut A-O- and

Scully

M-O-,

Europhys.

Lett. 14

(1991)

25.

[2] Haroche S., Brune M. and Raimond J-M-,

Europhys.

Lett. 14

(1991)

19.

[3] A recent review is

given by

Raithel

G., Wagner

C., Walther H., Narducci L-M- and Scully M-O-,

Cavity Quantum Electrodynamics,

P-R- Berman Ed.

(Academic Press, 1994)

pp. 57-121.

[4] Battocletti M.,

Diplomarbeit,

Universitit Mfinchen, 1993

(unpublished).

Références

Documents relatifs

Continuous cold beams differ considerably in flux, velocity and width of the velocity distribution: the atomic beam extracted using intensity imbalance [9] has a flux of 5 · 10 9

By slowing down of packets of atoms injected into a magnetic guide through their reflection on a moving mag- netic mirror, one can use a relatively high injection veloc- ity v i ,

This paper is organized as follows: In Section II we connect the photon flux, corresponding to the propaga- tion of the photonic state in free space leaking from the cavity, to

emission of quanta of the external field are correlated in stimulated transitions, the gradient force is a regular quantity.. Spontaneous emission destroys the

Abstract.-The relevance of gauge invariance in considering the question of emission of photons in the presence of external electromagnetic field from the ground state

We have considered the spectral and statistical properties in the fluorescence field from N two-level atoms interacting with an intense cavity mode. It has been shown

We introduce a family of DG methods for (1.1)–(1.3) based on the coupling of a DG approximation to the Vlasov equation (transport equation) with several mixed finite element methods

FIG. The pump and probe beams have linear polarizations. The angle between them has been exaggerated for the sake of clarity.. As expected, it is a dispersionlike curve that