• Non ci sono risultati.

W. WangJ. GongCASATI, GIULIOB. Limostra contributor esterni nascondi contributor esterni 

N/A
N/A
Protected

Academic year: 2021

Condividi "W. WangJ. GongCASATI, GIULIOB. Limostra contributor esterni nascondi contributor esterni "

Copied!
4
0
0

Testo completo

(1)

Entanglement-induced decoherence and energy eigenstates

Wen-ge Wang,1,2,3,*Jiangbin Gong,1,4G. Casati,1,5,6and Baowen Li1,4

1Department of Physics and Centre for Computational Science and Engineering, National University of Singapore, Singapore 117542, Republic of Singapore

2Department of Modern Physics, University of Science and Technology of China, Hefei 230026, China

3Department of Physics, Southeast University, Nanjing 210096, China

4NUS Graduate School for Integrative Sciences and Engineering, Singapore 117597, Republic of Singapore

5Center for Nonlinear and Complex Systems, Università degli Studi dell’Insubria, Via Valleggio 11, 22100 Como, Italy

6CNR-INFM and Istituto Nazionale di Fisica Nucleare, Sezione di Milano, Milan, Italy sReceived 23 August 2007; published 22 January 2008d

Using recent results in the field of quantum chaos we derive explicit expressions for the time scale of decoherence induced by the system-environment entanglement. For a generic system-environment interaction and for a generic quantum chaotic system as environment, conditions are derived for energy eigenstates to be preferred states in the weak coupling regime. A simple model is introduced to numerically confirm our predictions. The results presented here may also help with understanding the dynamics of quantum entangle- ment generation in chaotic quantum systems.

DOI:10.1103/PhysRevA.77.012108 PACS numberssd: 03.65.Yz, 03.65.Ta, 03.67.Mn, 05.45.Mt

I. INTRODUCTION

Real physical systems are never isolated from the sur- rounding world and, as a consequence, nonclassical correla- tionssentanglementd are established between the system and the environment. This process, which leads to decoherence, has a fundamental interest since it contributes to the under- standing of the emergence of classicality in a world governed by the laws of quantum mechanicsf1g.

Remarkably, different states may decohere at drastically different rates, and a small fraction of them may be particu- larly stable under entangling interaction with the environ- ment f1,2g. Such states are “preferred states” of the system under the influence of the environment.sThey are also called

“pointer states,” a name given for the states of pointers of measurement apparatus in the study of the measurement problem f1g.d This concept is important in understanding what states will naturally emerge from a quantum system subject to decoherence.

As discussed by Paz and Zurek in an interesting paperf3g, depending on the type and strength of the system- environment interaction, different preferred states may arise during the decoherence process. In particular they consider the case of an adiabatic environment modeled by a quantum scalar field and interacting weakly with a system through a given type of coupling. In such situation they show that en- ergy eigenstates are good preferred states and hence are the natural representation of the quantum system. Notice that the adiabatic environment does not change the population of en- ergy eigenstates of the system, implying infinite relaxation time.

In realistic situations the environment is often nonadia- batic, with finite relaxation time. In this case, the relation between the relaxation time and the decoherence time is cru- cial and, typically, only when the former is much longer than

the latter, energy eigenstates can be good preferred states.

While a rough estimate of the relaxation time can be ob- tained via Fermi’s golden rule, for the decoherence time the situation is more complex. Indeed in the case of a generic type of coupling and generic environment, it is hardly pos- sible to obtain estimates by employing the master-equation approach used inf3g.

In this paper we propose an alternative approach which takes advantage of recent progress in the field of quantum chaos: namely random matrix theory and the so-called fidel- ityf4–8g which is a measure of the stability of the quantum motion under system perturbations. Using this approach we can estimate—via the fidelity decay of the environment—the decoherence time of the system for weak, generic system- environment couplings and for a broad class of environ- ments. In particular, we need not restrict ourselves to the Markovian regimesin contrast to master-equation approachd where the bath correlation decay is faster than decoherence.

Moreover, as an important result of ours, by modeling the environment by a chaotic quantum system, we determine a critical border in the coupling strength below which energy eigenstates are shown to be preferred states, while above this border, the relaxation time and decoherence time have the same scaling with the coupling strength and therefore no definite statements can be made. We also present numerical results which confirm the above picture.

II. GENERAL THEORY

Let us consider a quantum system S with a discrete spec- trum, weakly coupled to a second quantum system E as its environment. The total Hamiltonian is

H= HS+eHI+ HE, s1d where HS and HE are the Hamiltonians of S and E, respec- tively, andeHI is the weak interaction Hamiltonianse≪ 1d.

The time evolution of the whole system is given by uCSEstdl = e−iHt/"uCSEs0dl. The initial state is set as a product

*wgwang@ustc.edu.cn

PHYSICAL REVIEW A 77, 012108s2008d

1050-2947/2008/77s1d/012108s4d 012108-1 ©2008 The American Physical Society

(2)

state uCSEs0dl = ucSs0dlufEs0dl. The reduced density matrix kaurrestdubl = kauTrErstdubl is obtained by tracing over the environment.

Consider first the case with initial state ucSs0dl = ual, whereual denotes an energy eigenstate of HSwith eigenen- ergy Ea. We define the projection operators ualkau^1E and Pa¯; obÞaublkbu^1E where 1E is the identity operator for the environment degrees of freedom. The whole Hilbert space can then be decomposed into two orthogonal sub- spaces, leading to

uCSEstdl = e−iEat/"ualufaEstdl +euxa¯stdl, s2d whereeuxa¯stdl ; Pa¯uCSEstdl. The small parameter eis intro- duced to account for the fact that, in case of weak coupling, the second term in Eq.s2d remains small inside some initial time interval ssee belowd. The normalization of uCSEstdl in Eq.s2d is unity to the first order ine.

A simple derivation shows that the evolution of the two terms in Eq.s2d is given by the coupled equations,

i"d

dtufaEstdl = HEa

effufaEstdl +e2eiEat/"kauHIuxa¯stdl, s3d

i"d

dtuha¯stdl = exp

S

"isEa− Ha¯dt

D

Pa¯HIualufaEstdl, s4d

where HEa

eff;eHIa+ HE, uha¯stdl ; expsiHa¯t /"duxa¯stdl, HIa

; kauHIual, and Ha¯; Pa¯HPa¯.

It is evident from Eq.s2d thate2kxa¯u xa¯l gives the popu- lation that has leaked to the subspace associated with Pa¯. In the case of weak coupling, this population leakage is initially very small. More precisely we have thate2kxa¯u xa¯l ≪ 1 up to times t ≪tE, wheretE is the relaxation time of the system.

Then the second term in Eq.s3d can be safely neglected and, as a result, the environment is in the state ufaEstdl

< e−itHEeffa/"ufEs0dl while the system remains in the eigenstate ual with a definite phase evolution.

Consider now as the initial state a superposition of energy eigenstatesucSs0dl = oaCaual. As in Eq. s2d we can write

uCSEstdl =

o

a

e−iEat/"CaualufaEstdl +euxstdl, s5d

where the term uxstdl = oaCauxa¯stdl contains now contribu- tions of transitions between different energy eigenstates.

Note that the first term on the right-hand side of Eq.s5d may be highly entangled even when euxstdl is small. For t ≪tE, the termeuxstdl in Eq. s5d can be neglected and, by tracing out the environment, the reduced density matrix of the sys- tem can be written as

rabre =kauTrErstdubl . e−isEa−Ebdt/"CaCbpfbastd, s6d where fbastd ; kfbEstd ufaEstdl satisfies

fbastd < kfEs0dueitsHEeffa+eVd/"e−itHEeffa/"ufEs0dl, s7d with

V; HIb− HIa=kbuHIubl − kauHIual. s8d Significantly, the quantity fbastd is simply the “fidelity am- plitude” of the environment associated with the two slightly different Hamiltonians HEaeff and sHEaeff+eVd. For nonconstant V, this fidelity amplitude usually decays with time for a Þb, then rabre also decays and therefore decoherence sets in.

In the case of constant V in Eq.s8d, Eq. s7d gives ufbastdu

< 1, hence, there is no decoherence induced by the environ- ment in the first-order perturbation theory discussed above.

Notice that Eqs. s6d and s7d become exact in the particular case in which the coupling HI commutes with HS. This case has been considered inf9g.

It turns out, from the above considerations, that for weak, but generic coupling, energy eigenstates of HSplay a special role in the sense that, only for energy superposition states snot single eigenstatesd, the decoherence process is associ- ated with the instability of the environment under perturba- tionsfidelity decayd.

III. A MODEL WITH THE ENVIRONMENT MODELED BY A QUANTUM CHAOTIC SYSTEM

Let us now turn to an explicit estimate of the decoherence time of the system. To this end we can directly apply recent results on fidelity decay f4–7g which, as shown below, al- lows us to estimate the decoherence time for a generic type of system-environment interaction and for a broad class of environments. This contrasts the situation of the master- equation approach with which only particular types of inter- action and environment have been treated f3g while exten- sion to more general situations is mathematically difficult.

Let us assume that the environment is modeled by a quan- tum chaotic systemf10g. sAnalogous strategy can be applied when the environment has a regular or mixed-type phase space structure.d Fidelity decay in such systems has been studied via semiclassical methods as well as with random matrix theory, both giving consistent results. Specifically, it turns out that for initial states chosen randomly, the fidelity has typically a Gaussian decay below a perturbative border epand an exponential decay above this borderf4–6g. If we model the environment HEby a single matrix of dimension N taken from the so-called Gaussian orthogonal ensemble sGOEd f11g, then the border ep can be explicitly estimated and is given byf5g

2pepVnd2 , svD, s9d where Vnd2 is the average of uknuVun8lu2 with n Þ n8 and V given by Eq.s8d. Here unl denotes the eigenstates of HEaeff, D is the mean level spacing of HEaeff, and sv

2 is the variance of knuVunl.

Below the perturbative border,e,ep, the fidelity ampli- tude decays asf5g

ufbastdu . e−e2sv2t2/2"2. s10d Then, the decoherence time td, characterizing the decay of off-diagonal matrix elementsfsee Eq. s6dg is given by

WANG et al. PHYSICAL REVIEW A 77, 012108s2008d

012108-2

(3)

td.Î2"/sesvd ~e−1, e,ep. s11d Notice that this dependence oftdonecoincides with the one derived in Ref. f3g, even though in our case we do not as- sume an adiabatic environment. For e.ep, ufbastdu has an exponential decayf4g,

ufbastdu , e−Gt/2" with G = 2pe2Vnd2/D, s12d and

td. "D/spe2Vnd2 d ~e−2, e.ep. s13d We stress that, like Eq.s6d, also Eqs. s11d and s13d are valid only if the time scale under consideration is much less than tE.

The next issue is if, and under what conditions, tE is sufficiently large so that significant decoherence may occur for t ≪tE. If this is the case, then energy eigenstates are much more robust than their superposition states. LetumEl be one eigenstate of HE and kHI,nd2 l be the mean square of the nondiagonal matrix elements ka8ukmE8uHIumElual saÞa8d.

One may estimate tEby using Fermi’s golden rule, i.e., tE. 1/R ~e−2, where R = 2pe2rkHI,nd2 l/" s14d is the transition rate in Fermi’s golden rule with the on-shell density-of-states r approximated by the average density of all possible final states for the whole systemf12g.

Therefore, below the perturbative borderse,epd the de- coherence timetdand the relaxation timetEscale ase−1and e−2, respectively. It follows that for small enoughe, we have tdtE and hence energy eigenstates are good preferred states regardless of the form of the coupling. On the other hand, above the perturbative border, bothtdin Eq.s13d and tE in Eq.s14d scale withe−2. In particular, in cases of very smallep, the two time scales can be comparablef13g even at weak perturbation and hence energy eigenstates may not be preferred states.

We will now introduce a simple dimensionless model which will also allow an explicit numerical evaluation of the different time scales. Let S be a qubit system with Hamil- tonian HS=oaEaualkau,a= 1 , 2 and E2− E1= 1. The environ- ment is modeled by a N-dimensional matrix in the GOE and the interaction HIis taken in the form of a random matrix as well. Specifically, in a given arbitrary basis ukl, the matrix elements kkuHEuk8l and kakuHIua8k8l are real random num- bers distributed according to a Gaussian with unit variance.

Moreover, we set the Planck constant " = 1.

We have numerically integrated the time-dependent Schrödinger equation for this model thus obtaining the ma- trix elements of the reduced density matrix rre. In Fig.1, we show the numerical results for parameters N = 200 and e= 5 3 10−4. The perturbative border can be numerically com- puted from Eq.s9d and it is found to beep, 0.04. It is clearly seen that for weak coupling the decay of r12reis well predicted by the Gaussian decay in Eq.s10d. By contrast, the change in the diagonal matrix elements of the reduced density is neg- ligible, as shown by the upper thin curve in Fig. 1.

Figure2 is drawn for the same parameters of Fig.1 but

for a larger coupling strength, comparable to the perturbative border. One notices deviations from the Gaussian decay and also an appreciable population change supper thin curved.

Interestingly, we found that deviations from the Gaussian decay always goes with an appreciable population change.

Indeed if, for example, we deliberately weaken those off- diagonal coupling terms sdiagonal coupling terms not touchedd that are responsible for the population decay, then the Gaussian decay can be recovered while the population decay becomes negligible.

Finally, if we further increase the coupling strength above the perturbative border, then the exponential decay of the off-diagonal matrix elements fEq. s12dg is indeed observed snot shown hered.

We have carefully studied the scaling behavior oftE and tdas functions of the coupling strengtheand the results are shown in Fig.3. Numericallytdis defined as the time scale over whichur12reu decays by a factor of 1 / e, andtEis defined as the reciprocal of the slope of r22restd in the initial interval of FIG. 1. sColor onlined Time dependence of the elements of the reduced density matrix of the qubit system S. Here N = 200 and the coupling strength e = 5 3 10−4is much smaller than the perturbative borderep, 0.04 numerically computed from Eq. s9d. The upper thin solid curve gives the diagonal matrix element r22restd for an initial state withucSs0dl = u2l. The lower thick solid curve gives the off- diagonal matrix element r12restd for an initial state which is a product state in which the system S is in an energy superposition state while the environment is in a randomly chosen state. The dotted and dashed curves give the theoretical Gaussian decay equation s10d and the exponential decay equations12d, respectively.

FIG. 2.sColor onlined Same as in Fig.1but for e = 0.01 which is still below but close to the perturbation border. It is seen that the numerically computed ur12reu begin to deviate from the predicted Gaussian decay. Moreover the population decaysthin solid curved becomes appreciable.

ENTANGLEMENT-INDUCED DECOHERENCE AND ENERGY … PHYSICAL REVIEW A 77, 012108s2008d

012108-3

(4)

time in which Fermi’s golden rule is valid. It is seen that numerical data nicely agree with analytical predictions. In particular, one can distinguish between the two scaling be- haviors oftdse−1ande−2d separated by the perturbative bor- der ep. Below this border, tdtE, implying that energy eigenstates are much more stable than energy superposition states.

IV. DISCUSSIONS AND CONCLUSIONS

Entanglement-induced decoherence within a closed total system is also referred to as “intrinsic decoherence” f14g.

The above analysis directly indicates the existence of pre- ferred states of intrinsic decoherence. As such, the dynamics of entanglement generation between two subsystems de- pends strongly on the coherence properties of the initial state. This is of interest to studies of entanglement generation in classically chaotic systemsf15g. Further, our results might also shed light on the dynamics of quantum thermalization processes within a closed systemf16g, where energy eigen- states also play a special role.

In summary, for weak but generic coupling between two quantum subsystems, energy eigenstates are shown to play a special role in the entanglement-induced decoherence pro- cess. The quality of the energy eigenstates as preferred states is also analyzed in terms of a simple dynamical model which allows for numerical analysis.

ACKNOWLEDGMENTS

The authors are very grateful to F. Haake for valuable discussions and suggestions. This work is supported in part by an Academic Research Fund of NUS and the TYIA of DSTA, Singapore under Contract No. POD0410553. One of the authorssW.W.d is partially supported by Natural Science Foundation of China Grants No. 10275011 and No.

10775123. One of the authors sJ.G.d is supported by the start-up funding and the NUS “YIA” fundingsWBS Grant No. R-144-000-195-123d. One of the authors sG.C.d is funded by the MIUR-PRIN 2005 “Quantum computation with trapped particle arrays, neutral and charged.”

f1g W. H. Zurek, Phys. Rev. D 24, 1516 s1981d; W. H. Zurek, S.

Habib, and J. P. Paz, Phys. Rev. Lett. 70, 1187s1993d.

f2g D. Braun, F. Haake, and W. T. Strunz, Phys. Rev. Lett. 86, 2913s2001d; L. Diósi and C. Kiefer, ibid. 85, 3552 s2000d; J.

Eisert, ibid. 92, 210401 s2004d; H. M. Wiseman and J. A.

Vaccaro, Phys. Rev. A 65, 043606s2002d.

f3g J. P. Paz and W. H. Zurek, Phys. Rev. Lett. 82, 5181 s1999d.

f4g R. A. Jalabert and H. M. Pastawski, Phys. Rev. Lett. 86, 2490 s2001d; Ph. Jacquod, P. G. Silvestrov, and C. W. J. Beenakker, Phys. Rev. E 64, 055203sRd s2001d; T. Prosen and M. Žni- darič, J. Phys. A 35, 1455s2002d; F. M. Cucchietti, C. H.

Lewenkopf, E. R. Mucciolo, H. M. Pastawski, and R. O. Valle- jos, Phys. Rev. E 65, 046209s2002d.

f5g N. R. Cerruti and S. Tomsovic, J. Phys. A 36, 3451 s2003d;

Phys. Rev. Lett. 88, 054103s2002d.

f6g W.-G. Wang and B. Li, Phys. Rev. E 66, 056208 s2002d; W.-G.

Wang, G. Casati, and B. Li, ibid. 69, 025201sRd s2004d; W.

Wang, G. Casati, B. Li, and T. Prosen, ibid. 71, 037202 s2005d; W.-G. Wang and B. Li, ibid. 71, 066203 s2005d.

f7g T. Gorin, T. Prosen, T. H. Seligman, and M. Žnidarič, Phys.

Rep. 435, 33s2006d.

f8g F. M. Cucchietti, D. A. R. Dalvit, J. P. Paz, and W. H. Zurek, Phys. Rev. Lett. 91, 210403s2003d; F. M. Cucchietti, J. P. Paz, and W. H. Zurek, Phys. Rev. A 72, 052113s2005d.

f9g T. Gorin, T. Prosen, T. H. Seligman, and W. T. Strunz, Phys.

Rev. A 70, 042105s2004d; H. T. Quan, Z. Song, X. F. Liu, P.

Zanardi, and C. P. Sun, Phys. Rev. Lett. 96, 140604s2006d.

f10g M. Srednicki, Phys. Rev. E 50, 888 s1994d.

f11g Decoherence effects with environments modeled by random matrices are recently considered in C. Pineda and T. H. Selig- man, Phys. Rev. A 75, 012106s2007d.

f12g The same scaling with e is derived in a more rigorous treat- ment for M-level quantum systems interacting with reservoirs in the thermodynamic limit in M. Merkli, I. M. Sigal, and G. P.

Berman, Phys. Rev. Lett. 98, 130401s2007d. We also remark that the decoherence time given in that paper also scales as e−2, where the initial state of the environment is assumed to be an equilibrium state and the off-diagonal elements of reduced density matrix are studied with respect to their equilibrium values.

f13g Similar to the well-known T2-T1relation, the relaxation time scale tEsets an upper bound to td.

f14g J. Gong and P. Brumer, Phys. Rev. A 68, 022101 s2003d.

f15g J. N. Bandyopadhyay and A. Lakshminarayan, Phys. Rev. Lett.

89, 060402 s2002d; Phys. Rev. E 69, 016201 s2004d; H.

Fujisaki, T. Miyadera, and A. Tanaka, ibid. 67, 066201s2003d;

R. Demkowicz-Dobrzański and M. Kuś, ibid. 70, 066216 s2004d; K. Furuya, M. C. Nemes, and G. Q. Pellegrino, Phys.

Rev. Lett. 80, 5524s1998d.

f16g S. Popescu, A. J. Short, and A. Winter, Nat. Phys. 2, 754 s2006d; S. Lloyd, ibid. 2, 727 s2006d; S. Goldstein, J. L. Leb- owitz, R. Tumulka, and N. Zanghi, Phys. Rev. Lett. 96, 050403s2006d.

FIG. 3. sColor onlined Dependence of times tdand tEon the coupling strength e. Full circles and empty squares represent the numerically computed td and tE, respectively. The solid line is given by the theoretical expressions11d and the dashed line is given by Eq.s14d. Notice the good agreement between numerical data and theoretical predictions. The arrow indicates the perturbative border epwhich clearly separates the two different scaling behaviors of td.

WANG et al. PHYSICAL REVIEW A 77, 012108s2008d

012108-4

Riferimenti

Documenti correlati

We aim (a) at the evaluation of the PECs starting at very short and unexplored internuclear distances (0.01 bohrs) and ending at full dissociation, (b) at the systematic prediction

Using the risk function R 1 , in Section 4.1 we de- rive some information inequalities and we prove in Section 4.2 some optimality and efficiency results for Bayes and ML

Figure 2, instead, shows the convergence properties of the local energy expectation value computed as a function of time step obtained using importance sampled DMC simula- tions on

Given the ease with which the appropriate deterministic benchmarks can be generated, and the fact that the convergence of parallel tempering on these systems has been measured

In the research, we use the financial statement data for the evaluation of the financial performance of companies through other comprehensive income and we determine which format

The structure of the resulting nonlinear sigma model allows us to recover the known spin wave results in the collinear regime, supports the presence of an Ising phase transition

Indeed, no CAT algorithm for generating plane partitions seems to be known (for instance, the problem is missing in [2,4]), although efficient algorithms for generating

We prove that if the primitive root of the smallest word of L (with respect to a lexicographic order) is prefix distinguishable in L then the centralizer of L is as simple as