<<

Annals of 322 (2007) 1466–1476 www.elsevier.com/locate/aop

Replacing by von Neumann in quantum transitions

Angela Kopp, Xun Jia, Sudip Chakravarty *

Department of Physics and Astronomy, University of California Los Angeles, Los Angeles, CA 90095-1547, USA

Received 15 July 2006; accepted 15 August 2006 Available online 28 September 2006

Abstract

We propose that quantum phase transitions are generally accompanied by non-analyticities of the von Neumann (entanglement) entropy. In particular, the entropy is non-analytic at the Anderson transition, where it exhibits unusual fractal scaling. We also examine two dissipative quantum sys- tems of considerable interest to the study of decoherence and find that non-analyticities occur if and only if the system undergoes a quantum . 2006 Elsevier Inc. All rights reserved.

Keywords: Quantum phase transition and criticality; Entanglement; Dissipative quantum system; Von Neumann entropy; Anderson localization; Nonanalyticity of free energy

In the thermodynamic limit two distinct states of matter cannot be analytic continua- tions of each other. Classical phase transitions are characterized by non-analyticities of the free energy [1]. For quantum phase transitions [2] (QPTs) the energy often assumes the role of the free energy. But in a number of important cases this criterion fails to predict a QPT, such as the three-dimensional metal-insulator transition of non-interact- ing electrons in a random potential (Anderson localization [3,4]). It is therefore essential that we find alternative criteria that can track fundamental changes in the internal corre- lations of the ground state wave function.

* Corresponding author. Fax: +1 310 825 5734. E-mail address: [email protected] (S. Chakravarty).

0003-4916/$ - see front matter 2006 Elsevier Inc. All rights reserved. doi:10.1016/j.aop.2006.08.002 A. Kopp et al. / Annals of Physics 322 (2007) 1466–1476 1467

Recently, there has been intense interest in using entanglement to study quantum phase transitions [5–10]. Since entanglement is a unique quantum phenomenon which character- izes wave functions that cannot be factored into single particle states, one expects that it should play a role in generating the correlations that exist on all length scales at a . But a state can be entangled without being critical—consider, for instance, the singlet state of two spin-1/2 particles. So it is not mere entanglement which concerns us here, but also the criticality of infinitely many interacting degrees of freedom. In an important paper on black hole entropy, Bekenstein [11] demonstrated the useful- ness of information entropy. The concept is readily applicable to a quantum mechanical ground state. If we consider the entire wave function, the conventional entropy is of course zero (assuming the state is non-degenerate), and all we can study is the analyticity of the ground state energy. In those cases, where this is a smooth analytic function of a tuning parameter that drives a QPT, we get no useful information. Yet in many cases we know that the wave function represents special internal correlations, as in Laughlin’s quantum Hall wave function [12], or in the Bijl–Dingle–Jastrow–Feynman wave function [13] describing the superfluid state of bosons. How can we quantify such correlations, in par- ticular when they change across a transition? Let us divide a system into two pieces and calculate the entanglement between them [14]. In the ground state jWæ, the of the full system is q = jWæÆWj. Now par- tition the system into two parts A and B, where A denotes the subsystem of interest and B the environment whose details are of no interest. We construct the reduced density matrix qA by tracing over the environmental degrees of freedom, akin to integrating out the microstates consistent with a given set of macroscopic thermodynamic variables. The von Neumann entropy, defined by S = Tr(qA lnqA), provides a measure of bipartite entanglement and therefore contains information about the quantum correlations present in the ground state. Note that S = Tr(qA lnqA)=Tr(qB lnqB), where the reduced den- sity matrix qB is obtained by tracing over the degrees of freedom in subsystem A. We consider three important models. In addition to demonstrating the use of von Neu- mann entropy, these models illuminate the role of disorder and dissipation in entangle- ment, inter alia in quantum computation. The two dissipative systems that we discuss have been recently addressed from this point of view [15], but the results are incorrect. Our first example, however, is Anderson localization in three-dimensions (d = 3), which has been studied extensively and is known to have a quantum critical point. A critical val- ue of disorder separates a system, where all states are localized from one where some states are extended. At the critical point the wave function exhibits a fractal character [16]. The relevant Hamiltonian for a d-dimensional lattice, expressed in terms of fermionic y creation ðci Þ and annihilation (ci) operators, is—suppressing spin indices, irrelevant in this case X X y y H ¼ Eici ci t ci cj þ h:c: ; ð1Þ i hi;ji where the sum in the second term is over distinct nearest-neighbor pairs, and the site ener- gies Ei are chosen from a uniform random distribution between W/2 and W/2. In what follows, we shall choose the unit of energy such that t = 1. Since this is a one-electron problem, the single-site density matrix in the state a is

a a 2 a 2 qi ¼jwi j ðj1ih1jÞi þð1 jwi j Þðj0ih0jÞi; ð2Þ 1468 A. Kopp et al. / Annals of Physics 322 (2007) 1466–1476

a where wi is the amplitude of an energy eigenstate a at site i. The ket j0æ (j1æ) corresponds to the site being unoccupied (occupied). The von Neumann entropy, averaged over all the eigenstates A and the number of sites N,is A 1 XN X S ¼ jwaj2 ln jwaj2 þð1 jwaj2Þ lnð1 jwaj2Þ : ð3Þ AN i i i i n¼1 a¼1 For a random system S must be further averaged over realizations of the random poten- tials. The resulting quantity, denoted by S, measures the correlation between the amplitude at a given site and the rest—local entanglement [17]. We compute S by diagonalizing the Hamiltonian for finite-sized systems in d = 3 and then attempt to collapse the data using the finite size scaling hypothesis for the singular part of the entropy [18]:

y 1=m Ssing ¼ L fðL wÞ; ð4Þ 1/3 where L = N is the length of the system, w = jW Wcj/Wc,andWc is the critical dis- order at which Anderson localization occurs. The m is the exponent for the localization length, the value of which is known to be approximately 1.35; Wc is known to be approximately 16.5 [19]. The universal scaling functions f± refer to W greater or less than Wc. It has been conjectured that for dimensions d > 2, the exponent y =(d 1) [18]. Examining the data collapse in Fig. 1, we find instead that the exponent y = 2.6, a fractal dimension less than three. This is similar to the fractal scaling of the participation ratio [16,20]. The origin of the fractal dimension is the fractal nature of the wave function at the mobility edge, separating the delocalized states from the localized states. Of course, data collapse without an objective criterion can be treacherous. An interesting criterion used to justify the goodness of collapse is shown in Fig. 2. The quantitative measure is ob- tained from the minimization of a specially defined function of the residuals called Pb [21], which satisfies Pb P Pbjabs min =0.

Fig. 1. The data collapse of the calculated von Neumann entropy of the d = 3 Anderson localization problem.

The exponent m and Wc were set to the known values 1.35 and 16.5, respectively. The exponent y was then determined to be 2.6. The system sizes are indicated in the inset. The systems of sizes 73 and 93 were averaged over 10 different realizations of disorder, while the system of size 113 was averaged over 5 realizations. For the 133 lattice, it did not seem necessary to average over multiple realizations. A. Kopp et al. / Annals of Physics 322 (2007) 1466–1476 1469

Fig. 2. The minimum of Pb corresponds to the best data collapse. From the figure, we see that y = 2.6 and m = 1.35 are good choices, although the minimum with respect to m is very shallow.

The fractal dimension of the von Neumann entropy can be confirmed from another 2 measure of entanglement called the linear entropy [22], which is SL ¼ 1 TrqA. For the presentP problem, it is easy to show that the participation ratio ð2Þ 1 a 4 P ¼ A i;ajwi j ¼ 1 SL=2. This is to be averaged over the realizations of the random potential, if necessary. In the extreme localized case, only one site participates and SL = 0. In the contrasting limit of fully extended state SL =2 2/N fi 2, when N fi 1. The participation ratio, hence SL, exhibits scaling at the metal-insulator transition: (2) x 1/m P = L g±(L w), where g± is another universal function. As shown in Fig. 3, the frac- tal dimension x = 1.4, consistent with the results known in the literature. x (2) P L

Fig. 3. The data collapse, as in Fig. 1, of the participation ratio of the d = 3 Anderson localization problem. The exponent m and Wc were set to the known values 1.35 and 16.5, respectively. The exponent x was then determined to be 1.4. 1470 A. Kopp et al. / Annals of Physics 322 (2007) 1466–1476

The general model of a two-level system (a qubit) coupled to a dissipative heat bath is relevant to a number of interesting physical problems [23]. We will consider a specific for- mulation, the spin–boson model [23], in which the two-level system is represented by a spin-1/2 degree of freedom and the bath is a collection of (bosonic) harmonic oscillators. The Hamiltonian is 1 1 X H sb ¼ Drx þ H osc þ rz knxn; ð5Þ 2 2 n where rx and rz are Pauli matrices and Hosc represents the Hamiltonian of an infinite num- ber of harmonic oscillators. The last term couples the z-component of the spin to the coor- dinates {xn} of the oscillators. The coupling constants {kn}, together with the masses {mn} and frequencies {xn}, determine the spectral density of the heat bath, which is given by X 2 p kn JðxÞ¼ dðx xnÞ: ð6Þ 2 n mnxn

For an Ohmic bath, we can take J(x)=2pax for x < xc and J(x) = 0 for x P xc. At zero temperature, this model has a quantum critical line separating a broken-sym- metry phase with Ærzæ = M0 „ 0 from a disordered phase with Ærzæ = 0 (see Fig. 4) [24]. As with any broken-symmetry phase, its definition requires some care: the states with the order parameters +M0 and M0 are orthogonal and their Hilbert spaces are unitarily inequivalent. An infinitesimal external field can pin one of these degenerate vacuua. Thus, the broken-symmetry state has an effective classical description in which the environmen- tal degrees of freedom are relaxed around a specific state of the qubit. The corresponding uncertainty is zero and so is the von Neumann entropy. In the disordered state, by con- trast, infinitesimal field has only infinitesimal effect. Right at the quantum critical point, the qubit is maximally entangled with the environment, as will be shown below. So the phase transition can be aptly described as a classical-to-quantum transition. There are a

y = (Δ/2ω c ) 2 y = − (1−α) 2 y = (1−α)

Quantum critical line

Ordered Disordered

α > 1 α = 1 α < 1

Fig. 4. The T = 0 phase diagram of a two-level system coupled to an Ohmic heat bath and the corresponding renormalization group flows in the regime j(1 a)j1 and y 1. The renormalization group trajectories point along increasing values of the flow parameter l = ln(sxc), where s is imaginary time. The separatrix 2y = (1 a) separates a broken symmetry phase (Ærzæ„0) from the disordered phase (Ærzæ = 0). At this critical line the magnetization is discontinuous even though the correlation time (imaginary) diverges. At a = 1/2 the dynamics of the two-level system changes from an overdamped to an underdamped state, but there is no thermodynamic phase transition at this point. A. Kopp et al. / Annals of Physics 322 (2007) 1466–1476 1471 number of recent interesting proposals to observe this dissipative phase transition using a quantum dot [25,26]. The renormalization group flows for this transition are the same as those of the classical inverse-square in one-dimension; (D/2xc) plays the role of the fugacity y of the kinks in the Ising model, while a maps onto the inverse tempera- ture. Although the linearized renormalization group equations are identical to the Koster- litz–Thouless (KT) transition of the two-dimensional XY model, the physics of the ordered phase is entirely different because of the existence of a local order parameter. In the ground state of Eq. (5), the reduced density matrix of the spin degree of freedom is determined by the expectation values Ærxæ and Ærzæ 1 1 þhrzihrxi qA ¼ : ð7Þ 2 hrxi 1 hrzi

After diagonalizing qA, we can easily compute the ground state von Neumann entropy. Hence the behavior of the entanglement at the QPT follows directly from the behavior of Ærxæ and Ærzæ. Since the order parameter Ærzæ is discontinuous at the transition, the von Neumann entropy also jumps by an amount DS. As shown in Appendix A, the mag- nitude of this jump is, to leading order in y, DS ¼ ln 2 þðy=2Þ ln y: ð8Þ

In the limit of vanishing y, the system goes from being unentangled (Ærxæ =0,Ærzæ =1)to being maximally entangled (Ærxæ = Ærzæ = 0) as it enters the disordered state. Note that this result depends crucially on a proper treatment of the broken symmetry—without the jump in the order parameter, S would be continuous through the transition. We also emphasize that the von Neumann entropy is discontinuous (with a singular derivative) even though the correlation length in imaginary time diverges at the transition. Because this divergence takes the form of an essential singularity, it does not leave a strong signature in other quantities at the critical point. In addition to the QPT discussed above, the spin–boson model also undergoes a dynamic crossover at a = 1/2, below which the spin can exhibit damped coherent oscilla- tions on short time scales. For a > 1/2 the dynamics of the disordered phase is character- ized by an overdamped decay. If the central claim of this paper is correct, the von Neumann entropy should be analytic at a = 1/2 because no phase transition occurs at this point. This appears to be consistent with numerical renormalization group calculations of the anisotropic Kondo problem [27]. We now turn to a simpler but equally important example, the damped harmonic oscil- lator. Consider a single harmonic oscillator (momentum p, position x, mass M, and fre- quency x0) coupled to an environment that also consists of harmonic oscillators. Again we examine the case of Ohmic dissipation, where the spectral function (defined as in Eq. (6))isJ(x)=gx for small x and zero above the cutoff xc. The ground state expecta- tion values of x2 and p2 are given by [28] h hx2i¼ f ðjÞ; 2Mx0 hMx 2hMx x hp2i¼ 0 ð1 2j2Þf ðjÞþ 0 j ln c ; ð9Þ 2 p x0 where j = g/2Mx0 is the friction coefficient and 1472 A. Kopp et al. / Annals of Physics 322 (2007) 1466–1476 pffiffiffiffiffiffiffiffiffiffiffiffiffi! 1 j þ j2 1 f ðjÞ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffi ln pffiffiffiffiffiffiffiffiffiffiffiffiffi : ð10Þ p j2 1 j j2 1

Atpffiffiffiffiffiffiffiffiffiffiffiffiffij = 1 the system crosses overpffiffiffiffiffiffiffiffiffiffiffiffiffi from damped oscillatory to overdamped behavior ( j2 1 is to be replaced by i 1 j2), just as the spin–boson model does at a =1/2. The function f(j) is real for all j > 0 and has identical power series expansion whether we approach j = 1 from the positive or the negative side. Thus it is analytic at j =1. At zero temperature, the normalized reduced density matrix for the damped harmonic oscillator has matrix elements [28] pffiffiffiffiffiffiffiffiffiffi 0 00 aðx0x00Þ2bðx0þx00Þ2 hix jqAjx ¼ 4b=p e ; ð11Þ where a = Æp2æ/2h2 and b = 1/8Æx2æ. To compute the von Neumann entropy, we first note [29] that Z o 0 0 n 0 TrðqA ln qAÞ¼lim dx hx jqAjx i: ð12Þ n!1 on The subsequent derivation of S is reproduced in Appendix B; the result is ! rffiffiffi pffiffiffi pffiffiffi 1 4b 1 a a þ b S ¼ ln þ ln pffiffiffi pffiffiffi : ð13Þ 2 a b 2 b a b pNoteffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi that S fi 0asb fi a, which corresponds to the minimum uncertainty hx2ihp2i ¼ h=2. The uncertainty relation is satisfied only for b 6 a. We have verified that S is analytic at j = 1 (see Appendix B)—this follows from the analyticity of the function f(j). Despite the popularity of field-theoretic perturbative methods, involving Feynman dia- grams, some of the major breakthroughs in have come from for- mulating wave functions that reflect the unique correlations of a quantum system with infinitely many degrees of freedom. The BCS wave function is remarkable because it incor- porates the broken global gauge symmetry. The Bijl–Dingle–Jastrow–Feynman wave function for liquid 4He contains the essential ingredients of interacting superfluid matter. Similarly, the Laughlin wave function explains not only the fractional quantum Hall effect, but also the fractionalization of charge. We believe that the non-analyticity of the von Neumann entropy is a useful criterion for all quantum phase transitions at which the char- acter of the ground state wave function changes. It provides a unique insight into the inter- nal correlations of a many body wave function. The approach described in the present paper can be applied to a wide variety of QPTs for which the conventional energy criterion fails. In particular, it can be immediately applied to transitions between integer quantum Hall plateaus [30].

Acknowledgments

We thank J. Rudnick for helpful discussions. This work was supported by the NSF under Grant DMR-0411931. A. Kopp et al. / Annals of Physics 322 (2007) 1466–1476 1473

Appendix A. Derivation of DS for the spin–boson model

To compute DS, we need to know the expectation values Ærxæ and Ærzæ atp criticality.ffiffiffiffiffiffiffiffi Adapting the results of Anderson and Yuval [31], we find that Ærzæ jumps by 1=a along the quantum critical line 2y = (1 a). The expectation value of rx is most easily obtained by differentiating the ground state energy: Ærxæ = 2(oE/oD). Since the ground state energy maps onto the free energy of the inverse-square Ising model, we can make use of the free energy scaling relation [31,32] dF ðlÞ¼y2ðlÞeldl: ðA:1Þ

Here F is the free energy in units of xc, so we have the direct correspondence Ærxæ = 2 (oE/oD) fi oF/oy. On the separatrix 2y = (1 a), the flow of the fugacity is given by y(l)=y/(1 + 2yl). Since y fi 0asl fi 1, the renormalization group equations (valid for y 1) become asymptotically exact in this limit. The fixed point at l = 1 corresponds to the fully polar- ized state of the Ising chain, for which the free energy is zero. So if we tune to criticality, we can obtain a closed-form solution for F by integrating Eq. (A.1) all the way to l = 1. Differentiating with respect to y then gives "#Z 1 t 1 1 1 1=2y e hrxi¼ 1 þ 2 e dt : ðA:2Þ 2 2y ð2yÞ 1=2y t

In the limit y fi 0, Eq. (A.2) assumes the form Ærxæ 2y, while the discontinuity in Ærzæ goes as 1 y. Using the reduced density matrix given in the text, it is easy to show that DS ¼ ln 2 þðy=2Þ ln y þ OðyÞ: ðA:3Þ

Appendix B. Derivation of S for the damped harmonic oscillator

As described in the text, the von Neumann entropy of the damped harmonic oscillator is given by Z o 0 0 n 0 S ¼TrðqA ln qAÞ¼lim dx hx jqAjx i: ðB:1Þ n!1 on After inserting n 1 complete sets of position eigenstates, we can use the expression for the matrix elements to recast the integral as Z Z 4b n=2 dx0hx0jqn jx0i¼ dx dx exiMijxj : ðB:2Þ A p 1 n

M is a tri-diagonal matrix with Mii =2(a + b) and Mi+1,i = Mi,i+1 = (a b), and there- fore has eigenvalues cm =2(a + b) 2(a b) cos(2pm/n), with m =1,...,n. So the von Neumann entropy becomes 2 !3 1=2 o Yn 4 n=2 5 S ¼lim ð4bÞ cm : ðB:3Þ n!1 on m¼1 For convenience, we write 1474 A. Kopp et al. / Annals of Physics 322 (2007) 1466–1476

Yn n n cm ¼ 2 ða bÞ P; ðB:4Þ m¼1 where the product P is defined as  Yn 2pm a þ b P ¼ a cos ; a ¼ : ðB:5Þ n a b m¼1 Our main task is to evaluate this product by performing a suitable integral transform. Note that I o ln P Xn 1 1 1 nzn1 ¼ ¼ dz ðB:6Þ oa 2pm 2pi 1 1 zn 1 m¼1 a cos n C a 2 ðz þ zÞ

i2pm/n provided the contour C encloses the poles atpzffiffiffiffiffiffiffiffiffiffiffiffiffim =e (Fig. 5). Deforming the contour 2 to pick up the additional poles at z ¼ a a 1, we obtain o ln P 2n zn zn 2n zn þ 1 ¼ þ ¼ þ ; ðB:7Þ o n n n a zþ z zþ 1 z 1 zþ z zþ 1 where the second equality follows from z+z = 1. Integrating this result over a yields a closed-form expression for P. The integral is most easily evaluated by transforming to the variable z Z+ n zn þ 1 ln P ¼ dz þ þ z zn 1 þ þ 1 zn ¼ ln þ þ n 2 þ const: ðB:8Þ zþ This is consistent with Eq. (B.5) if we set the integration constant equal to nln 2. In terms of the original parameters a and b, the product is given by

z

x x x x x

x xxx

x x x x x C

i2pm/n Fig. 5. The contourpffiffiffiffiffiffiffiffiffiffiffiffiffi C encloses the poles on the unit circle at zm =e . With a >1(b < a), the additional poles 2 at z ¼ a a 1 are located on the real axis. A. Kopp et al. / Annals of Physics 322 (2007) 1466–1476 1475

Fig. 6. Von Neumann entropy (S) of the damped harmonic oscillator with xc = 100x0. Both S and its first derivative (inset) are perfectly smooth through the dynamic crossover at j = 1. Higher-order derivatives are similarly well-behaved due to the analyticity of f(j)atj =1.

"# ! ! pffiffiffi pffiffiffi n pffiffiffi pffiffiffi n 1 a þ ffiffiffib a bffiffiffi P ¼ n pffiffiffi p þ pffiffiffi p 2 ðB:9Þ 2 a b a þ b and the rest of the calculation follows trivially from Eqs. (B.3) and (B.4). The von Neu- mann entropy is ! rffiffiffi pffiffiffi pffiffiffi 1 4b 1 a a þ b S ¼ ln þ ln pffiffiffi pffiffiffi ; ðB:10Þ 2 a b 2 b a b where the ratio b/a depends on the friction coefficient j and the ultraviolet cutoff xc (recall that the requires b 6 a). S is an analytic function of j as long as b/a is analytic—which, in turn, is guaranteed by the analyticity of the function f(j). The plots in Fig. 6 provide further proof of this point.

References

[1] C.N. Yang, T.D. Lee, Phys. Rev. 87 (1952) 404. [2] S. Sachdev, Quantum phase transitions, Cambridge University Press, Cambridge, 1999. [3] E. Abrahams, P.W. Anderson, D.C. Licciardello, T.V. Ramakrishnan, Phys. Rev. Lett. 42 (1979) 673. [4] J.T. Edwards, D.J. Thouless, J. Phys. C 4 (1971) 453. [5] A. Osterloh, L. Amico, G. Falci, R. Fazio, Nature 416 (2002) 608. [6] T.J. Osborne, M.A. Nielsen, Phys. Rev. A 66 (2002) 032110. [7] G. Refael, J.E. Moore, Phys. Rev. Lett. 93 (2004) 260602. [8] G. Vidal, J.I. Latorre, E. Rico, A. Kitaev, Phys. Rev. Lett. 90 (2003) 227902. [9] T.-C. Wei, D. Das, S. Mukhopadyay, S. Vishveshwara, P.M. Goldbart, Phys. Rev. A 71 (2005) 060305/1. [10] I. Klich, G. Refael, A. Silva, , 2006. [11] J.D. Bekenstein, Phys. Rev. D 7 (1973) 2333. [12] R.B. Laughlin, Phys. Rev. Lett. 50 (1983) 1395. [13] E. Feenberg, Theory of quantum fluids, Academic Press, New York, 1969. [14] C.H. Bennett, H.J. Bernstein, S. Popescu, B. Schumacher, Phys. Rev. A 53 (1996) 2046. [15] T. Stauber, F. Guinea, Phys. Rev. A 73 (2006) 042110. [16] F. Wegner, Z. Phys. B 36 (1980) 209. [17] P. Zanardi, Phys. Rev. A 65 (2002) 042101. [18] P. Calabrese, J. Cardy, Journal of : Theory and Experiment 2005, P04010 (2005), . [19] E. Hofstetter, M. Schreiber, Phys. Rev. B 49 (1994) 14726. 1476 A. Kopp et al. / Annals of Physics 322 (2007) 1466–1476

[20] D.A. Parshin, H.R. Schober, Phys. Rev. Lett. 83 (1999) 4590. [21] S.M. Bhattacharjee, F. Seno, J. Phys. A 34 (2001) 6375. [22] W.H. Zurek, S. Habib, J.P. Paz, Phys. Rev. Lett. 70 (1993) 1187. [23] A.J. Leggett, S. Chakravarty, A.T. Dorsey, M.P.A. Fisher, A. Garg, W. Zwerger, Rev. Mod. Phys. 59 (1987) 1. [24] S. Chakravarty, Phys. Rev. Lett. 49 (1982) 681. [25] L. Borda, G. Zarand, D. Goldhaber-Gordon, , 2006. [26] K. Le Hur, M.R. Li, Phys. Rev. B 72 (2005) 073305. [27] T.A. Costi, R.H. McKenzie, Phys. Rev. A 68 (2003) 034301. [28] U. Weiss, Quantum dissipative systems, World Scientific, Singapore, 1999. [29] C. Holzhey, F. Larsen, F. Wilczek, Nucl. Phys. B 424 (1994) 443. [30] S.L. Sondhi, S.M. Girvin, J.P. Carini, D. Shahar, Rev. Mod. Phys. 69 (1997) 315. [31] P.W. Anderson, G. Yuval, J. Phys. C 4 (1971) 607. [32] A.P. Young, T. Bohr, J. Phys. C 14 (1981) 2713.