a modern review of the two-level approximation

16
A modern review of the two-level approximation Marco Frasca Via Erasmo Gattamelata, 3, 00176 Roma, Italy Received 25 November 2002 Abstract The paradigm of the two-level atom is revisited and its perturbative analysis is discussed in view of the principle of duality in perturbation theory. The models we consider are a two-level atom and an ensemble of two-level atoms both interacting with a single radiation mode. The aim is to see how the latter can be actually used as an amplifier of quantum fluctuations to the classical level through the thermodynamic limit of a very large ensemble of two-level atoms [M. Frasca, Phys. Lett. A 283 (2001) 271] and how can remove Schrodinger cat states. The thermodynamic limit can be very effective for producing both classical states and decoherence on a quantum system that evolves without dissipation. Decoherence without dissipation is in- deed an effect of a single two-level atom interacting with an ensemble of two-level atoms, a situation that proves to be useful to understand recent experiments on nanoscale devices show- ing unexpected disappearance of quantum coherence at very low temperatures. Ó 2003 Elsevier Science (USA). All rights reserved. 1. Introduction It is safe to say that the foundations of quantum optics are built on the concept of a few level atom. Indeed, the most important concept introduced so far in this field is the two-level atom [1]. A lot of physics can be derived by such an approximation and several recent experiments agree fairly well with a description given by the so-called Jaynes–Cumming model describing a two-level atom interacting with a single radia- tion mode [2]. Besides, by this understanding of radiation–matter interaction it has also been possible to generate Fock states of the radiation field on demand [3]. Annals of Physics 306 (2003) 193–208 www.elsevier.com/locate/aop E-mail address: [email protected]. 0003-4916/03/$ - see front matter Ó 2003 Elsevier Science (USA). All rights reserved. doi:10.1016/S0003-4916(03)00078-2

Upload: marco-frasca

Post on 02-Jul-2016

212 views

Category:

Documents


0 download

TRANSCRIPT

Annals of Physics 306 (2003) 193–208

www.elsevier.com/locate/aop

A modern review of the two-level approximation

Marco Frasca

Via Erasmo Gattamelata, 3, 00176 Roma, Italy

Received 25 November 2002

Abstract

The paradigm of the two-level atom is revisited and its perturbative analysis is discussed in

view of the principle of duality in perturbation theory. The models we consider are a two-level

atom and an ensemble of two-level atoms both interacting with a single radiation mode. The

aim is to see how the latter can be actually used as an amplifier of quantum fluctuations to the

classical level through the thermodynamic limit of a very large ensemble of two-level atoms

[M. Frasca, Phys. Lett. A 283 (2001) 271] and how can remove Schr€oodinger cat states. The

thermodynamic limit can be very effective for producing both classical states and decoherence

on a quantum system that evolves without dissipation. Decoherence without dissipation is in-

deed an effect of a single two-level atom interacting with an ensemble of two-level atoms, a

situation that proves to be useful to understand recent experiments on nanoscale devices show-

ing unexpected disappearance of quantum coherence at very low temperatures.

� 2003 Elsevier Science (USA). All rights reserved.

1. Introduction

It is safe to say that the foundations of quantum optics are built on the concept of

a few level atom. Indeed, the most important concept introduced so far in this field is

the two-level atom [1]. A lot of physics can be derived by such an approximation and

several recent experiments agree fairly well with a description given by the so-called

Jaynes–Cumming model describing a two-level atom interacting with a single radia-

tion mode [2]. Besides, by this understanding of radiation–matter interaction it has

also been possible to generate Fock states of the radiation field on demand [3].

E-mail address: [email protected].

0003-4916/03/$ - see front matter � 2003 Elsevier Science (USA). All rights reserved.

doi:10.1016/S0003-4916(03)00078-2

194 M. Frasca / Annals of Physics 306 (2003) 193–208

The radiation–matter interaction currently used is based on some relevant ap-

proximations that are still well verified in current experiments: First, it is assumed

that the dipole approximation holds, that is, the wavelength of the radiation field

is much larger than the atomic dimensions; second, the rotating wave approximation

(RWA) is always assumed, meaning by this that just near resonant terms are effectivein describing the interaction between radiation and matter, these terms being also de-

scribed as energy conserving. Indeed, it is sometimes believed that, without these two

approximations no two-level atom approximation can really holds [4]. Actually, in

the optical regime, that statement can be supported and widely justifies the success

of the Jaynes–Cumming model both theoretically and experimentally.

Actually, things are not so straightforward to describe radiation–matter interac-

tion. Infact, Cohen-Tannoudji et al. [5] were forced to introduce the concept of

dressed states for the two-level atom as, in the regime of microwaves, the RWA failsand a good description of the first experiments in this field were achieved through the

concept of dressed states without the RWA [5].

Quantum computation exploited by ionic traps has been first put out by Cirac and

Zoller [6]. A recent paper by Moya-Cessa et al. [7] proved that the standard Jaynes–

Cummings model should retain all terms for a Paul trap giving a clear example of

dismissal of the rotating wave approximation in quantum optics.

The appearance of laser sources that have large intensity has made thinkable the

possibility to extend the study of a two-level atom in such a field. Recent studies seemto indicate that such an approximation can give a viable model for such a physical

situation [8–14]. In view of this possibility, some methods have been recently devised

to approach a solution of the two-level atom in a monochromatic field (being the la-

ser field treated classically) [15–17]. These studies retain just the two-level and dipole

approximations but give up the RWA.

In our recent analysis, it was shown that, treating the laser field classically in this

situation, leaves out a relevant part of the behavior of the model [18,19]. Particularly,

if one is especially interested in a resonant behavior, it is seen that some Rabi oscil-lations are neglected: These oscillations have been recently observed in an experi-

ment with Josephson junctions [20] and originate from the formation of bands for

the two levels of the atom due to the radiation field [19].

The aim of this paper is to review, using the approach of duality in perturbation

theory [21], the consequences of the validity of the two-level approximation relaxing

the RWA approximation. We will see that a single radiation mode interacting with a

large number of two-level atoms, without the RWA, provide the amplification of the

quantum fluctuations of the ground state of the radiation mode producing a classicalradiation field [22] and is able to remove macroscopic quantum superposition states.

It is important to point out that these effects arise when the initial state of the ensem-

ble of two-level atoms is properly prepared and the way of generating classical states

by unitary evolution in the thermodynamic limit of Ref.[22] is considered. Nondissi-

pative decoherence can also appear as interaction between an ensemble of two-level

systems and a quantum system interacting with it [23]. It should be said that another

approach to nondissipative decoherence has been recently proposed by Bonifacio

and coworkers [24].

M. Frasca / Annals of Physics 306 (2003) 193–208 195

The paper is so structured. In Section 2 we analyze the model from a general

perspective deriving the two-level approximation. In Section 3 we present a per-

turbative analysis of the two-level atom interacting with a single radiation mode,

by duality in perturbation theory. In Section 4 we give a brief survey of a recent

proposal of appearance of classical states and decoherence by unitary evolution inthe thermodynamic limit. In Section 5 we show how a strong radiation field can

be obtained by strong interaction of a single mode with an ensemble of two-level

atoms.

In Section 6 we present a way to approach the measurement problem in quantum

mechanics showing how, in the thermodynamic limit, Schr€oodinger cat states can be

removed leaving only a coherent state describing a classical field.

Finally, in Section 7 the conclusions are given.

2. A paradigm in quantum optics: two-level atom

In this section we want to derive the two-level approximation on a general foot-

ing. So, let us consider a system described by a Hamiltonian H0 such that we have a

complete set of eigenstates H0jni ¼ Enjni. We assume, for the sake of simplicity, that

the set is discrete. Then, we introduce a time-independent perturbation V . By using

the identity I ¼P

n jnihnj we can write the Hamiltonian H ¼ H0 þ V as

H ¼Xn

ðEn þ hnjV jniÞjnihnj þXm6¼n

jmihnjhmjV jni: ð1Þ

This Hamiltonian can be rewritten by introducing the operators

rnm ¼ jnihmj;rynm ¼ jmihnj;

r3nm ¼ 1

2ðjnihnj jmihmjÞ;

ð2Þ

and we can build the algebra of the Pauli matrices, currently named su(2), as it is

straightforward to verify that

rnm; rynm

� �¼ 2ir3

nm;

r3nm; r

ynm

� �¼ iry

nm;

r3nm; rnm

� �¼ irnm:

ð3Þ

This permits us to prove that our Hamiltonian can be rewritten as the sum of two-

level Hamiltonians. Infact, if we change to the interaction picture by the unitary

transformation (we use units �h ¼ c ¼ 1)

U0ðtÞ ¼ exp

it

Xn

~EEnjnihnj!; ð4Þ

196 M. Frasca / Annals of Physics 306 (2003) 193–208

being ~EEn ¼ En þ hnjV jni, and we rewrite the V term of the Hamiltonian as

V 0 ¼Xm 6¼n

jmihnjhmjV jni ¼Xm>n

hmjV jnirynm

�þ hnjV jmirnm

�; ð5Þ

we get

HI ¼ U y0ðtÞV 0U0ðtÞ ¼

Xm>n

eið ~EEm ~EEnÞthmjV jnirynm

hþ eið ~EEn ~EEmÞthnjV jmirnm

ið6Þ

that proves our assertion: The time evolution of a quantum system can be described by

a Hamiltonian being the sum of su(2) Hamiltonians. The reason why this problem is

not generally solvable arises from the fact that, given two su(2) parts HI ;i and HI ;j of

this Hamiltonian, it can happen that ½HI ;i;HI ;j 6¼ 0 and then, the time evolution is

not straightforward to obtain analytically. Anyhow, the Hamiltonian HI can be usedto realize some approximate study of a quantum system. The simplest way to get an

approximate solution is, indeed, the two-level approximation.

The two-level approximation can be easily justified by assuming that only nearest

levels of the unperturbed atom really counts in the time evolution, that is, the more

the separation between levels is large and the less important is the contribution to the

time evolution of the system. This means that terms with the weakest time depen-

dence in HI are the most important. Mathematically, this means that we assume a

solution by a perturbation series and recognize principally the terms where a slowertime dependence is present.

We now consider the case of a single radiation mode interacting with a system

having Hamiltonian H0. This means in turn that we can choose

V ¼ xayaþ ex2V

1=2ðay þ aÞx ð7Þ

being e the electron electric charge, a and ay the ladder operators of the radiation

mode with frequency x and normalization volume V , and x the coordinate where thefield is oriented, having chosen a linear polarization for it. The dipole approximation

is taken to hold. In this case we have

HI ¼ xayaþ ex2V

1=2Xm>n

eið ~EEm ~EEnÞthmjxjnirynm

hþ eið ~EEn ~EEmÞthnjxjmirnm

iðay þ aÞ

ð8Þ

and we are now in a position to obtain the Jaynes–Cummings model. Indeed, we can

apply a new unitary transformation to the interaction picture U 00 ¼ exp itxayað Þ to

get

H 0I ¼ e

x2V

1=2Xm>n

eið ~EEm ~EEnÞthmjxjnirynm

hþ eið ~EEn ~EEmÞthnjxjmirnm

iðeixtay þ eixtaÞ:

ð9Þ

Now, on the basis of the two-level approximation given above, we have to conclude

that the only terms to retain are those having no time dependence at all, and theseare the resonant terms. Here, we recover the rotating wave approximation (RWA).

M. Frasca / Annals of Physics 306 (2003) 193–208 197

So, if we have two resonant levels m ¼ 2 > n ¼ 1 and we choose the phases of the

eigenstates of the unperturbed system so that hmjxjni is real, we can finally write the

Hamiltonian of the Jaynes–Cummings model as

HJC ¼ gðr12ay þ ry12aÞ ð10Þ

being g ¼ h2jxj1ieðx=2V Þ1=2 and the resonance condition E2 E1 ¼ x. This gives a

proper understanding of the success of the two-level atom approximation in quan-

tum optics when weak fields are involved. It is important to note that also a small

detuning can be kept, in agreement with the above discussion.The Jaynes–Cummings model is good until the other terms in the Hamiltonian are

truly negligible. The higher order corrections can be computed by a quite general ap-

proach as shown in [25]. These turn out to be corrections to the the Hamiltonian at

the resonance (e.g., Bloch-Siegert shift and/or a.c. Stark shift) plus the need to add

higher orders of the small perturbation theory to the solution. In the optical regime it

is all negligible.

So, as the small perturbation theory plays a crucial role in this analysis, one may

ask what one can say if the perturbation V becomes strong. Again, by assuming thatonly a su(2) component really contributes to the Hamiltonian (9) we need to treat the

Hamiltonian

H 0S ¼ xayaþ g eið ~EE1 ~EE2Þtry

12

hþ eið ~EE2 ~EE1Þtr12

iðay þ aÞ: ð11Þ

By undoing the interaction picture transformation, this Hamiltonian can be re-

written as

HS ¼ xayaþ D2

r3 þ gr1ðay þ aÞ; ð12Þ

having set r12 þ ry12 ¼ r1, r3 ¼ 2r3

12, and D ¼ E2 E1. Neither small perturbationtheory nor rotating wave approximation apply. Our aim in the next section is to

discuss the perturbative solution of the Schr€oodinger equation with this Hamiltonian.

But, while for the case of the Jaynes–Cummings Hamiltonian we have a fully theo-

retical justification for our approximation, in the strong coupling regime, the two-level

approximation can be satisfactorily justified only by experiment, unless it is exact.

3. Perturbative analysis of an interacting two-level atom

In this section we will give a brief overview of the perturbative solution for a sys-

tem described by the Hamiltonian (12) in the strong coupling regime. This approach

has been described in [19]. To agree about what a strong coupling regime should be,

one has properly to define the weak coupling regime. Indeed, if one has the Hamil-

tonian

H ¼ H0 þ sV ð13Þ

being s an ordering parameter, the weak coupling regime is the one with s very small

(s ! 0), while the strong coupling regime is the one with s very large (s ! 1). The

198 M. Frasca / Annals of Physics 306 (2003) 193–208

duality principle in perturbation theory as devised in [21] permits to do perturbation

theory in both the cases, if one is able to find the eigenstates of V , supposing known

those of H0. Indeed, small perturbation theory by the usual Dyson series gives (we set

s ¼ 1 as this parameter is arbitrary)

jwðtÞi ¼ U0ðtÞT exp

� i

Z t

0

VIðtÞ

ð14Þ

being T the time-ordering operator,

U0ðtÞ ¼ exp ð itH0Þ ð15Þ

the time evolution of the unperturbed Hamiltonian, and

VIðtÞ ¼ U y0ðtÞVU0ðtÞ ð16Þ

the transformed perturbation. The choice of a perturbation and an unperturbed part

is absolutely arbitrary. So, we can exchange the role of H0 and V , obtaining the dual

Dyson series

jwðtÞi ¼ UF ðtÞT exp

� i

Z t

0

H0F ðtÞ

ð17Þ

being

UF ðtÞ ¼ exp ð itV Þ ð18Þ

the time evolution of the unperturbed Hamiltonian, and

H0F ðtÞ ¼ U yF ðtÞH0UF ðtÞ ð19Þ

the transformed perturbation. The duality principle states that, when this exchange is

done, restating s, the series one obtains have the ordering parameters s and 1=s,respectively. One is the inverse of the other. So, if we have the eigenstates of V as jvniand eigenvalues vn, one can write

UF ðtÞ ¼Xn

eivntjvnihvnj: ð20Þ

If V is time dependent one has formally to rewrite the above as the adiabatic series

introducing the geometric phases of the eigenvectors that now could be time de-

pendent themselves [21].Coming back to the Hamiltonian (12), we realize that small perturbation theory

can be recovered if the unperturbed part is that of the two-level atom, otherwise one

has a strong coupling perturbation series with an unperturbed Hamiltonian given by

V ¼ xayaþ gr1ðay þ aÞ: ð21Þ

The dressed states originating by diagonalizing this Hamiltonian are well known [5]

and are given by

jvn;ki ¼ jkieðg=xÞkðaayÞjni ð22Þ

M. Frasca / Annals of Physics 306 (2003) 193–208 199

being r1jki ¼ kjki with k ¼ �1 and, jni the Fock number states that are displaced by

the exponential operator [26]. The eigenvalues are En ¼ nx ðg2=xÞ and are de-

generate with respect to k. So, one has

UF ðtÞ ¼Xn;k

eiEntjvn;kihvn;kj ð23Þ

and the transformed Hamiltonian becomes

H0F ¼ U yF ðtÞ

D2

r3UF ðtÞ ¼ H 00 þ H1: ð24Þ

Using the relation [26]

hljeðg=xÞkðaayÞjni ¼ffiffiffiffin!l!

rkgx

lnek2ðg2=2x2ÞLðlnÞ

n k2 g2

x2

� �ð25Þ

with l � n and LðlnÞn ðxÞ the associated Laguerre polynomial, one gets

H 00 ¼

D2

Xn

eð2g2=x2ÞLn4g2

x2

� �j½n; a1 ih½n; a1 jj1ih½ 1j

þ j½n; a1 ih½n; a1 jj 1ih1j ð26Þ

being Ln the nth Laguerre polynomial, j½n; ak i ¼ eðg=xÞkðaayÞjni, and

H1 ¼D2

Xm;n;m6¼n

eiðnmÞxt hnjeð2g=xÞðaayÞjmij½n; a1 ih½m; a1 jj1ihh

1j

þ hnjeð2g=xÞðaayÞjmij½n; a1 ih½m; a1 jj 1ih1ji: ð27Þ

The Hamiltonian H 00 can be straightforwardly diagonalized with the eigenstates

jwn; ri ¼1ffiffiffi2

p rj½n; a1 ij1i½ þ j½n; a1 ij 1i ð28Þ

and eigenvalues

En;r ¼ rD2eð2g2=x2ÞLn

4g2

x2

� �ð29Þ

being r ¼ �1. We see that two bands of levels are formed and two kind of transitions

are possible: interband (between levels of the two bands) and intraband (between the

levels of a band). This cannot happen if we consider a classical radiation mode, theintraband transitions would be neglected. So, looking for a solution in the form

jwF ðtÞi ¼Xr;n

eiEn;rtan;rðtÞjwn; ri ð30Þ

one gets the equations for the amplitudes [19]

i _aam;r0 ðtÞ ¼D2

Xn6¼m;r

an;rðtÞeiðEn;rEm;r0 ÞteiðmnÞxt

� hmjeð2g=xÞðaayÞjni r0

2

�þ hmjeð2g=xÞðaayÞjni r

2

: ð31Þ

200 M. Frasca / Annals of Physics 306 (2003) 193–208

These equations can also display Rabi oscillations between the eigenstates jwn; rithat can be seen as macroscopic quantum superposition states, both for interband

and intraband transitions [19]. States like these could prove useful for quantum

computation. These kind of Rabi oscillations in Josephson junctions have been re-

cently observed [20].At this stage it is very easy to do perturbation theory in the strong coupling re-

gime for this model. One has to rewrite the initial condition jwð0Þi by the eigenstates

jwn; ri obtaining in this way the amplitudes an;rð0Þ. Then, one has to solve Eq. (31)

perturbatively as done routinely in the weak coupling regime. In case of a resonance

one has to apply the RWA obtaining Rabi oscillations. In this way we see that the

dual Dyson series, as it should be expected, displays all the features of the standard

weak coupling expansion.

4. Classical states and decoherence by unitary evolution

An ensemble of independent two-level systems can behave classically. This has

been proven in [22]. Indeed, let us consider a Hamiltonian

Hc ¼D2

XNi¼1

r3i: ð32Þ

Assuming distinguishable systems, we can take for the initial state the one given by

jwð0Þi ¼YNi¼1

ðaij #ii þ bij "iiÞ ð33Þ

with jaij2 þ jbij2 ¼ 1, rzij #ii ¼ j #ii and rzij "ii ¼ j "ii. The time evolution gives us

jwðtÞi ¼YNi¼1

ðaieiðD=2Þtj #ii þ bie

iðD=2Þtj "iiÞ: ð34Þ

For the Hamiltonian it is easy to verify that

hHci ¼ hwð0ÞjHcjwð0Þi ¼D2

XNi¼1

ðjbij2 jaij2Þ ¼

D2kHN ð35Þ

being kH a fixed number between 1 and 1. So, in a similar way, it easy to obtain the

fluctuation

ðDHcÞ2 ¼ hwð0ÞjH 2c jwð0Þi hwð0ÞjHcjwð0Þi2 ¼ D2k0HN : ð36Þ

being k0H ¼PN

i¼1 jbij2ð1 jbij

2Þ=N and one sees that k0H is a finite number indepen-

dent on N . So, as it happens in statistical thermodynamics, in the thermodynamiclimit N ! 1 we see that quantum fluctuations are not essential, that is,

DHc

Hc/ 1ffiffiffiffi

Np : ð37Þ

M. Frasca / Annals of Physics 306 (2003) 193–208 201

The ‘‘laws of thermodynamics’’ are obtained by Ehrenfest�s theorem and are the

classical equations of motion. That is, the variables Rx ¼PN

i¼1 rxi, Ry ¼PN

i¼1 ryi, and

Rz ¼PN

i¼1 rzi follow, without any significant deviation, the classical equations of

motion, when the thermodynamic limit is considered and the time evolution is

computed averaging with the above jwðtÞi. So, we have found a classical object outof the quantum unitary evolution. The main point here is that classical objects can be

obtained by unitary evolution in the thermodynamic limit depending on their initial

states. Actually, one cannot apply the above argument if, e.g., the state of the system

is an eigenstate of the Hamiltonian Hc. Besides, a classical state obtained by unitary

evolution, per se, does not produce decoherence. Rather, it is interesting to see what

happens when such a classical state interacts with some quantum system. This is a

relevant problem that can prove quantum mechanics and its fluctuations to be just

the bootstrap of a classical world: If by unitary evolution, in the thermodynamiclimit, some classical objects are obtained and these are permitted to interact with

other quantum objects, the latter can decohere or become classical by themselves.

As a relevant example, let us consider the interaction of the above system with a

two-level atom. This model has been considered in [23] as a possible explanation of

recent findings in some experiments with nanoscale devices that show unexpected de-

coherence in the low temperature limit [27,28]. The Hamiltonian can be written as

HD ¼ X0

2rz þ

1

2

XNi¼1

ðDxirxi þ DzirziÞ Jrx �XNi¼1

rxi; ð38Þ

where J is the coupling. The Hamiltonian of the two-level systems (second term in

Eq. (38)) is taken not diagonalized, but this does not change our argument as the

above analysis still applies. Finally, X0 is the parameter of the Hamiltonian of the

two-level atom that we want to study. We need another hypothesis to go on, that is,

we assume that the coupling J is larger than any parameter of the two-level systems

Dxi;Dzi, but not with respect to X0. By applying duality in perturbation theory, we

have the leading order solution

jwðtÞi � expitX0

2rz

þ iJtrx �

XNi¼1

rxi

!jwð0Þi: ð39Þ

Now, as already seen, we have to choose the state of the two-level systems as given

by the product of the lower eigenstates of each rxi. This can be seen as a kind of

‘‘ferromagnetic’’ state and is in agreement with our preceding discussion. So, we take

jwð0Þi ¼ j #iYNi¼1

j 1ii ð40Þ

being rzj #i ¼ j #i and, similarly, rzj "i ¼ j "i. The state of the ensemble of two-

level systems agrees fairly well with the one of Eq. (33). So, one has, by tracing awaythe state of the ensemble of two-level systems being not essential for our aims,

jw0ðtÞi � expitX0

2rz

�þ iJNtrx

�j #i ð41Þ

202 M. Frasca / Annals of Physics 306 (2003) 193–208

that defines a spin coherent state [29,30]. The point we are interested in is the

thermodynamic limit. When N is taken to be large enough, the contribution X0 can

be neglected and we have a reduced density matrix

q0ðtÞ ¼ exp iJNtrxð Þj #ih# j exp ð iJNtrxÞ ð42Þ

being

q0""ðtÞ ¼

1 cosð2NJtÞ2

; ð43Þ

q0"#ðtÞ ¼ i

1

2sin 2NJt; ð44Þ

q0#"ðtÞ ¼ i

1

2sin 2NJt; ð45Þ

q0##ðtÞ ¼

1þ cosð2NJtÞ2

; ð46Þ

where we have oscillating terms with a frequency NJ that goes to infinity in the

thermodynamic limit. The only meaning one can attach to such a frequency is by an

average in time (see [23] and references therein) and decoherence is recovered. So,

when the ensemble of two-level systems strongly interacts with a quantum system

produces decoherence and quantum behavior disappears, in the thermodynamic

limit. The ensemble of two-level systems should evolve unitarily, producing a clas-

sical behavior. Higher order corrections have also been studied in [23]. It is impor-

tant to stress that this behavior should be expected at zero temperature as quantumcoherence is lost otherwise.

Such a behavior, having a characteristic decoherence time scale depending on the

number of two-level systems that interact with the quantum one, has been recently

observed in quantum dots [28]. In this case, the ensemble of two-level systems can

be given by the spins of the electrons that are contained in the two dimensional elec-

tron gas in the dot. Another source of decoherence in quantum dots could be given

by the spins of the nuclei interacting through an hyperfine interaction with the spin

of the conduction electrons [31]. The nuclei are contained in the heterostructuresforming the dot. In this case we have a similar spin-spin interaction but isotropic.

The mechanism that produces the decay of the off-diagonal parts of the density ma-

trix, also in this case, appears to be the same, being the decoherence produced dy-

namically and dependent on the initial state.

5. Amplification of quantum fluctuations to the classical level

Spontaneous emission can be seen as a very simple example of decoherence in the

‘‘thermodynamic limit’’ of the number of radiation modes. Indeed, we can consider a

two-level system interacting with N radiation modes and being resonant with one of

it. In the limit of a small coupling between radiation and two-level system and very

M. Frasca / Annals of Physics 306 (2003) 193–208 203

few spectator modes, one has Rabi oscillations, a clear example of quantum coher-

ence. When the number of spectator modes is taken to go to infinity, a description

with continuum is possible and this gives rise to decay, i.e., spontaneous emission.

This representation of the process of decay is very well described in [5].

Here, we want to consider the opposite situation, that is, a single radiation modestrongly interacting with an ensemble of N two-level systems. We are going to show

that, when the ensemble of two-level systems behaves as a classical object if left

alone, the radiation field, supposed initially in the ground state, will have the zero

point fluctuations amplified to produce a classical field having intensity dependent

on N , the number of two-level systems.

As done in [22], we modify the model of Eq. (12) to consider N two-level systems

interacting with a single radiation mode, as

HS ¼ xayaþ D2

XNi¼1

r3i þ gXNi¼1

r1iðay þ aÞ: ð47Þ

Then, the strong coupling regime amounts to consider the Hamiltonian D2

PNi¼1 r3i as

a perturbation, as already done in Section 3 for a single two-level atom. We take as

initial state of the full system jwð0Þi ¼ j0iQN

i¼1 j 1ii, so that, the ensemble of two-

level systems is again in a kind of ‘‘ferromagnetic’’ state representing its ground state.

Besides, no photon is initially present. It is a well known matter that the fluctuations

of the radiation mode are not zero in this case. The unitary evolution at the leadingorder gives us

jwðtÞi � exp

" itxaya itg

XNi¼1

r1iðay þ aÞ#j0iYNi¼1

j 1ii ð48Þ

that, by use of a known disentangling formula [30], produces

jwðtÞi ¼ einðtÞeixayatD½aðtÞ jwð0Þi; ð49Þ

being

nðtÞ ¼ N 2g2

x2ðxt sinðxtÞÞ; ð50Þ

aðtÞ ¼ Ngx

ð1 eixtÞ; ð51Þ

and

D½aðtÞ ¼ exp½aðtÞay aðtÞ�a : ð52Þ

We conclude that, at the leading order, the radiation mode evolves as a coherent

state with a parameter given by

aaðtÞ ¼ Ngx

ðeixt 1Þ ¼ aðtÞeixt: ð53Þ

In this way, we have amplified the quantum fluctuations of the field, being the

fluctuation of the number of photons proportional to N , but, as the average of

204 M. Frasca / Annals of Physics 306 (2003) 193–208

the number of photons is proportional to N 2, this ratio goes to zero as the

thermodynamic limit N ! 1 is taken. As it is well know [32], this produces a

classical field with increasing intensity as the number of two-level systems in-

creases, proving our initial assertion. We can see that the amplification of

quantum fluctuations gives rise to a classical object, as initially no radiation fieldis present.

Higher order corrections have been studied in [33], showing that are not essential

in the thermodynamic limit. So, this effect will prove to be a genuine example of pro-

duction of a classical object by unitary evolution in the thermodynamic limit with

possible technological applications.

6. Two-level systems, thermodynamic limit, and Schr€oodinger cat states

Decoherence, as currently devised, is able to remove superposition states through

interaction of the environment with a quantum system. This does not solve the mea-

surement problem in quantum mechanics as, mixed forms of the density matrix do

not give single states required by the measurement process [34]. This problem is fairly

well described by the Schr€oodinger cat paradox as we ask that the cat has a well de-

fined state at the observation. Schr€oodinger cat states have been currently produced in

laboratory in a form of superposition of coherent states (see, e.g., the second refer-ence in [1])

jwcati ¼ N ðjaei/i þ jaei/iÞ ð54Þ

being N a normalization factor, a and / real numbers. To understand the mea-

surement problem, we would like to get a single state out of such a superposition

after unitary evolution, if possible. In this way, we can show, at least in this case, thatquantum mechanics is, indeed, a self-contained theory. This possibility can be ex-

ploited by an ensemble of two-level systems interacting with a single radiation mode

in the thermodynamic limit.

In order to accomplish our aim, we consider again the Hamiltonian (47) with the

initial condition (54) multiplied by the ‘‘ferromagnetic state,’’ j/i ¼QN

i¼1 j 1ii, forthe ensemble of atoms as to have jwð0Þi ¼ jwcatij/i. The unitary evolution, assuming

the Hamiltonian of two-level atoms as a perturbation, gives in this case [35]

jwðtÞi � einðtÞN ðei/1ðtÞjbðtÞeixt þ aei/ixti þ ei/2ðtÞjbðtÞeixt þ aei/ixtiÞj/i;ð55Þ

where use has been made of the property of the displacement operator for coherent

states so to yield

nðtÞ ¼ N 2g2

x2ðxt sinðxtÞÞ; ð56Þ

bðtÞ ¼ Ngx

ð1 eixtÞ; ð57Þ

M. Frasca / Annals of Physics 306 (2003) 193–208 205

and

/1ðtÞ ¼ ia2½bðtÞei/ b�ðtÞei/ ; ð58Þ

/2ðtÞ ¼ ia2½bðtÞei/ b�ðtÞei/ ; ð59Þ

being the phases /1ðtÞ and /2ðtÞ generated by multiplication of two displacement

operators. In the thermodynamic limit N ! 1 one gets the macroscopic state

jbðtÞeixti and the cat state seems gone away.Actually, we have a couple of problems before one can claim that, effectively, the

cat state has been removed. Firstly, all we have done is a displacement to infinity and

no decoherence seems to be implied in such an operation, so all the properties of a

superposition state have to be there anyway. Secondly, we have done perturbation

theory and one has to prove that, in the thermodynamic limit, higher order correc-

tions are negligible.

The first question is answered immediately by computing the interference term in

the Wigner function of the state (55). In the thermodynamic limit such a term shouldbecome negligible. One has

WINT ¼ 2

pexp

24 x

þ

ffiffiffi2

pNg

xð1 cosðxtÞÞ

ffiffiffi2

pa cosð/Þ cosðxtÞ

!235

� exp

24 p

þ

ffiffiffi2

pNg

xsinðxtÞ þ

ffiffiffi2

pa cosð/Þ sinðxtÞ

!235

� cos 2ffiffiffi2

pa sinð/Þ p sinðxtÞð

� x cosðxtÞÞ þ a2 sinð2/Þ

þ 8aNgx

sinð/Þð1 cosðxtÞÞ : ð60Þ

This term has a quite interesting form as displays a term that rapidly oscillates in

time for N becoming increasingly large. If such oscillations become too rapid, we can

invoke blurring in time to have these terms averaged away. Otherwise, as the Wigner

function can be measured, we will get a way to probe, immediately and by very

simple means, Planck time physics. So, we can safely claim that we have true de-coherence in the thermodynamic limit and the cat state is effectively removed gen-

erating a macroscopic classical state. It should be said that ordinary decoherence is

generally invoked for blurring in space [36] and there is no reason to say that also

blurring in time cannot occur. We can recognize here the same argument used in

order to obtain decoherence for the model of Section 4 to explain decoherence in

quantum dots. A sound mathematical basis for such an approach is given in [37].

The second question can be straightforward answered by computing higher order

corrections through the strong coupling expansion discussed in Section 3. The proofis successfully accomplished in [35] and we do not repeat it here.

206 M. Frasca / Annals of Physics 306 (2003) 193–208

It is interesting to note that all the new points introduced so far for analyzing two-

level systems can conspire to generate a new view of the way measurement is realized

in quantum mechanics, possibly making the theory self-contained.

7. Discussion and conclusions

We have presented a brief review about some new views on two-level systems.

These appear to be even more important today with a lot of new effects to be de-

scribed and experimentally observed. Paramount importance is acquiring the deco-

herence due to an ensemble of two-level systems as it is becoming ubiquitous to

different fields of application as quantum computation and nanotechnology, fields

that maybe could merge. To face these new ways to see the two-level approxima-tion, we have exposed new mathematical approaches to analyze models in the

strong coupling regime. This regime has been pioneered by Bender and coworkers

[38] in quantum field theory in the eighties, but, with our proposal of duality in per-

turbation theory, a possible spreading of such ideas to other fields is now become

possible. Indeed, a lot of useful results, as those presented here, are obtained by this

new approach and, hopefully, the future should deserve some other interesting

results.

The idea of a non dissipative decoherence is also relevant due to the recent find-ings in the field of nanodevices, where unexpected lost of quantum coherence has ap-

peared in experiments performed at very low temperatures. In these cases, it appears

as the standard idea of decoherence, meant as interaction of a quantum system with

an external environment, seems at odds with some experimental results, even if an

interesting proposal through the use of quantum fluctuations has been put forward

by B€uuttiker and coworkers [39].

The conclusion to be drawn is that, today, a lot of exciting work at the founda-

tions of quantum mechanics is expecting us, giving insight toward new understand-ings and methods and, not less important, applications.

References

[1] L. Allen, J.H. Eberly, Optical Resonance and Two-level Atoms, Wiley, New York, 1975;

(For a more recent exposition see) W.P. Schleich, Quantum Optics in Phase Space, Wiley, Berlin,

2001.

[2] J.M. Raimond, M. Brune, S. Haroche, Rev. Mod. Phys. 73 (2001) 565, and references therein.

[3] S. Brattke, B.T.H. Varcoe, H. Walther, Phys. Rev. Lett. 86 (2001) 3534.

[4] M.O. Scully, M.S. Zubairy, Quantum Optics, Cambridge University Press, Cambridge, MA, 1997.

[5] C. Cohen-Tannoudji, J. Dupont-Roc, C. Grynberg, Atom-Photon Interactions, Wiley, New York,

1992.

[6] J.I. Cirac, P. Zoller, Phys. Rev. Lett. 74 (1995) 4091.

[7] H. Moya-Cessa, A. Vidiella-Barranco, J.A. Roversi, D.S. Freitas, S.M. Dutra, Phys. Rev. A 59 (1999)

2518.

[8] B. Sundaram, P.W. Milonni, Phys. Rev. A 41 (1990) R6571.

[9] M.Yu. Ivanov, P.B. Corkum, Phys. Rev. A 48 (1993) 580.

M. Frasca / Annals of Physics 306 (2003) 193–208 207

[10] Y. Dakhnovskii, H. Metiu, Phys. Rev. A 48 (1993) 2342;

Y. Dakhnovskii, R. Bavli, Phys. Rev. B 48 (1993) 11020.

[11] A.E. Kaplan, P.L. Shkolnikov, Phys. Rev. A 49 (1994) 1275.

[12] F.I. Gauthey, C.H. Keitel, P.L. Knight, A. Maquet, Phys. Rev. A 52 (1995) 525;

Phys. Rev. A 55 (1997) 615;

F.I. Gauthey, B.M. Garraway, P.L. Knight, Phys. Rev. A 56 (1996) 3093.

[13] M. Pons, R. Ta€iieb, A. Maquet, Phys. Rev. A 54 (1996) 3634.

[14] P.P. Corso, L. Lo Cascio, F. Persico, Phys. Rev. A 58 (1998) 1549, and references therein;

See also A. Di Piazza, E. Fiordilino, Phys. Rev. A 64 (2001) 013802;

A. Di Piazza, E. Fiordilino, M.H. Mittleman, Phys. Rev. A 64 (2001) 013414.

[15] J.C.A. Barata, W.F. Wreszinski, Phys. Rev. Lett. 84 (2000);

J.C.A. Barata, Ann. Henri Poincar�ee 2 (2001) 963.

[16] V. Delgado, J.M. Gomez Llorente, J. Phys. B 33 (2000) 5403;

A. Santana, J.M. Gomez Llorente, V. Delgado, J. Phys. B 34 (2001) 2371.

[17] M. Frasca, Phys. Rev. A 56 (1997) 1548;

Phys. Rev. A 60 (1999) 573.

[18] M. Frasca, J. Opt. B: Quant. Semiclass. Opt. 3(2001) S15. This behavior agrees with the result given in

E. Fiordilino, Phys. Rev. A 59 (1999) 876.

[19] M. Frasca, Z. Naturforsch. 56a (2001) 197; Phys. Rev. A 66 (2002) 023810. For a generalization of

this approach see K. Fujii, quant-ph/0203135, J. Phys. A 36 (2003) 2109, and quant-ph/0210166.

[20] Y. Nakamura, Yu.A. Pashkin, J.S. Tsai, Phys. Rev. Lett. 87 (2001) 246601.

[21] M. Frasca, Phys. Rev. A 58 (1998) 3439.

[22] M. Frasca, Phys. Lett. A 283 (2001) 271; Erratum: Phys. Lett. A 306 (2003) 184. Some general

theorems about this matter are given in M. Frasca, Phys. Lett. A 308 (2003) 135.

[23] M. Frasca, Physica E 15 (2002) 252.

[24] R. Bonifacio, N. Cim. B 114 (1999) 473;

R. Bonifacio, S. Olivares, P. Tombesi, D. Vitali, Phys. Rev. A 61 (2000) 053802.

[25] M. Frasca, Phys. Rev. A 58 (1998) 771.

[26] F.A.M. de Oliveira, M.S. Kim, P.L. Knight, V. Buzzek, Phys. Rev. A 41 (1990) 2645. See also K. Fujii,

quant-ph/0202081.

[27] A.G. Huibers, J.A. Folk, S.R. Patel, C.M. Marcus, C.I. Duru~ooz, J.S. Harris Jr., Phys. Rev. Lett. 83

(1999) 5090;

P. Mohanty, E.M.Q. Jariwala, R.A. Webb, Phys. Rev. Lett. 78 (1997) 3366;

P. Mohanty, R.A. Webb, Phys. Rev. B 55 (1997) 13452.

[28] C. Prasad, D.K. Ferry, A. Shailos, M. Elhasan, J.P. Bird, L.-H. Lin, N. Aoki, Y. Ochiai, K. Ishibashi,

Y. Aoyagi, Phys. Rev. B 62 (2000) 15356;

D.P. Pivin Jr., A. Andresen, J.P. Bird, D.K. Ferry, Phys. Rev. Lett. 82 (1999) 4687;

D.K. Ferry, private communication.

[29] F.T. Arecchi, E. Courtens, R. Gilmore, H. Thomas, Phys. Rev. A 6 (1972) 2211.

[30] Wei-Min Zhang, Da Hsuan Feng, R. Gilmore, Rev. Mod. Phys. 62 (1990) 867.

[31] A.V. Khaetskii, D. Loss, L. Glazman, Phys. Rev. Lett. 88 (2002) 186802;

J. Schliemann, A.V. Khaetskii, D. Loss, Phys. Rev. B 66 (2002) 245303;

D. Loss, private communication.

[32] L. Mandel, E. Wolf, Optical Coherence and Quantum Optics, Cambridge University Press,

Cambridge, MA, 1995.

[33] M. Frasca, J. Opt. B: Quant. Semiclass. Opt. 4 (2002) S443. This is a contribution to Garda 2001

Conference Proceedings edited by R. Bonifacio and D. Vitali.

[34] S.L. Adler, quant-ph/0112095. F. Lalo€ee, Am. J. Phys. 69 (2001) 655.

[35] M. Frasca, quant-ph/0212119.

[36] M.V. Berry, Chaos and the semiclassical limit of quantum mechanics (is the moon there when

somebody looks?), in: R.J. Russell, P. Clayton, K. Wegter-McNelly, J. Polkinghorne (Eds.), Quantum

Mechanics: Scientific Perspectives on Divine Action, Vatican Observatory CTNS publications, 2001,

pp. 41–54.

208 M. Frasca / Annals of Physics 306 (2003) 193–208

[37] G.H. Hardy, in: Divergent Series, AMS Chelsea Publishing, Providence, 1991, p. 10ff.

[38] C.M. Bender, F. Cooper, G.S. Guralnik, D.H. Sharp, Phys. Rev. D 19 (1979) 1865;

C.M. Bender, F. Cooper, G.S. Guralnik, H. Moreno, R. Roskies, D.H. Sharp, Phys. Rev. Lett. 45

(1980) 501;

C.M. Bender, F. Cooper, R. Kenway, L.M. Simmons Jr., Phys. Rev. D 24 (1981) 2693.

[39] P. Cedraschi, M. B€uuttiker, Phys. Rev. B 63 (2001) 165312;

Ann. Phys. 289 (2001) 1;

M. B€uuttiker, in: A.T. Skjeltorp, T. Vicsek (Eds.), Complexity from Microscopic to Macroscopic

Scales: Coherence and Large Deviations, Kluwer, Dordrecht, 2002, pp. 21–47.