<<

CHAOS 15, 026105 ͑2005͒

Fundamental aspects of quantum Brownian Peter Hänggi and Gert-Ludwig Ingold Institut für Physik, Universität Augsburg, 86135 Augsburg, Germany ͑Received 1 December 2004; accepted 9 December 2004; published online 17 June 2005͒

With this we elaborate on the physics of quantum noise in thermal equilibrium and in stationary nonequilibrium. Starting out from the celebrated quantum fluctuation-dissipation theorem we discuss some important consequences that must hold for open, dissipative quantum systems in thermal equilibrium. The issue of quantum dissipation is exemplified with the fundamental problem of a damped harmonic quantum oscillator. The role of quantum fluctuations is discussed in the context of both, the nonlinear generalized quantum Langevin equation and the path integral ap- proach. We discuss the consequences of the -reversal symmetry for an open dissipative quan- tum dynamics and, furthermore, point to a series of subtleties and possible pitfalls. The path integral methodology is applied to the decay of metastable states assisted by quantum Brownian noise. © 2005 American Institute of Physics. ͓DOI: 10.1063/1.1853631͔

This work deals with the description of quantum Brown- the substitution of the kBT from the classical equipar- ian motion in linear and nonlinear quantum systems that tition law4 by the thermally averaged quantum energy ͑but exhibit frictional influences. The symmetries of thermal leaving out the zero point energy contribution͒ of the har- equilibrium impose severe constraints on the time evolu- monic oscillator. Nyquist’s remark thus constitutes a precur- tion properties of open quantum systems. These lead to a sor of the celebrated work by Callen and Welton5 who gen- quantum generalization of the classical Einstein relation eralized the relations by Einstein, Nyquist, and Johnson to that connects friction with the strength of thermal quan- include quantum effects: In their work they put forward a tum fluctuations. There exist a variety of theoretical generally valid connection between the response function roadways to model quantum dissipation. Here, we discuss and the associated quantum fluctuations in equilibrium, the the topic for the prominent case of a damped harmonic oscillator upon combining thermodynamics and linear re- quantum fluctuation-dissipation theorem. sponse theory. A dissipative nonlinear Without doubt, quantum fluctuations constitute a promi- can be dealt with a generalized quantum Langevin equa- nent noise source in many nanoscale and biological systems. tion, a path integral formulation, or in terms of a gener- For example, the tunnelling and the transfer of electrons, alized quantum master equation for the corresponding quasiparticles, and alike, is assisted by noise for which the reduced dynamics. We illustrate the situation for the quantum nature cannot be neglected. The features of this problem of the dissipative decay out of a metastable state. noise change drastically as a function of temperature: At suf- Furthermore, we point out a series of subtleties, pitfalls ficiently high temperatures a crossover does occur to classi- and shortcomings that one must be aware of when con- cal Johnson–Nyquist noise. fronted with the world of quantum noise driven With this work we shall present various methods and phenomena. schemes of modelling quantum Brownian motion from first principles. In particular, the thermal noise must at all I. INTRODUCTION obey the quantum version of the fluctuation-dissipation theo- ͑ ͒ Albert Einstein explained the phenomenon of Brownian rem à la Callen-Welton . This latter property is necessary in motion in his annus mirabilis of 1905 by use of statistical order to be consistent with the second law of thermodynam- ͑ ͒ methods which he ingeniously combined with the laws of ics and the principle of quantum detailed balance. We thermodynamics.1 In this pioneering work he as well pro- elaborate on several alternative but equivalent methods to vided a first link between the dissipative forces and the im- describe quantum noise and quantum Brownian motion per peding thermal fluctuations, known as the Einstein relation se: These are the functional integral method for dissipative 6,7 which relates the strength of diffusion to the friction. This quantum systems and time-dependent driven quantum 8 9 intimate connection between dissipation and related fluctua- systems, the quantum Langevin ͑operator͒ approach, sto- 10,11 tions was put on a firm basis much later when Nyquist2 and chastic schemes, or the concept of stochastic Schrödinger 12 Johnson3 considered the spectral density of voltage- and equations. In doing so, we call attention to distinct differ- current-fluctuations. ences to the classical situation and, as well, identify a series What role do quantum and the associated of delicate pitfalls which must be observed when making quantum fluctuations play in this context? After the birth of even innocent looking approximations. Such pitfalls involve, in the early 1920’s we can encounter in among others, the rotating-wave approximation, the use of the very final paragraph of the 1928 paper by Nyquist for the quasiclassical Langevin forces, the quantum regression hy- first time the introduction of quantum mechanical noise via pothesis and/or the Markov approximation.6,8,13

1054-1500/2005/15͑2͒/026105/12/$22.5015, 026105-1 © 2005 American Institute of Physics

Downloaded 21 Jun 2005 to 137.250.81.48. Redistribution subject to AIP license or copyright, see http://chaos.aip.org/chaos/copyright.jsp 026105-2 P. Hanggi and G.-L. Ingold Chaos 15, 026105 ͑2005͒

II. THE QUANTUM FLUCTUATION-DISSIPATION =Y͑␻͒␦V͑␻͒ where the admittance Y͑␻͒ is identical to the THEOREM AND ITS IMPLICATIONS ␹ ͑␻͒ ˙ susceptibility IQ . As a consequence of I=Q, the symme- As already mentioned, in 1951 Callen and Welton trized power spectrum of the current fluctuations is given by ͑␻͒ ␻ ͑␻͒ proved a pivotal relation between equilibrium fluctuations SII =i SIQ so that we obtain and dissipative transport coefficients. Note also that this quantum fluctuation-dissipation relation holds true indepen- ប␻ dent of particle statistics. The following cornerstone achieve- ͑␻͒ ប␻ ͩ ͪ ͑␻͒ 5 SII = coth Re Y ments can be found in this primary work: 2kBT ͑1͒ The generalization of the classical Nyquist’s formula to ប␻ ប␻ =2ͫ + ͬRe Y͑␻͒. ͑6͒ the quantum case; 2 e␤ប␻ −1 ͑2͒ the quantum mechanical proof that susceptibilities are ӷប␻ related to the spectral densities of symmetrized correla- In the high temperature limit kBT , we recover the ͑␻͒→ ͑␻͒ tion functions. results of Nyquist and Johnson, i.e. SII 2kBT Re Y . For the Markovian limit of an Ohmic resistor, where Y͑␻͒ ͑␻͒ For a single degree of freedom, linear response theory =1/R, this result simplifies to read SII =2kBT/R. The yields for the change of the expectation value of an operator- quantum version was already anticipated by Nyquist in the valued observable B due to the of a ͑classical͒ force last paragraph of his 1928 paper.2 However, he made use of F͑t͒ that couples to the conjugate dynamical operator A the original expression of Planck which yields only the sec- ond contribution present in the lower line of ͑6͒. Nyquist t ͗␦ ͑ ͒͘ ͵ ␹ ͑ ͒ ͑ ͒ ͑ ͒ thus missed the first term arising from the energy B t = ds BA t − s F s . 1 −ϱ which already appears in a paper by Planck published in 1911.14 ␦ ͑ ͒ ͑ ͒ ͗ ͘ Here, B t =B t − B 0 denotes the difference with respect On the other hand, in the extreme quantum limit kBT to the thermal equilibrium average ͗B͘ in the absence of the Ӷប␻ ͑␻͒→ប␻ ͑␻͒ 0 , we find that SII Re Y . In particular, this force. The reaction of the system is contained in the response implies that at zero frequency the spectral weight of the cur- ␹ ͑ ͒ function BA t with a so-called dissipative part rent fluctuations vanishes in the generic case where the ad- 1 mittance does not exhibit an infrared divergence. ␹d ͑t͒ = ͓␹ ͑t͒ − ␹ ͑− t͔͒. ͑2͒ We cannot emphasize enough that the quantum BA 2i BA AB fluctuation-dissipation relation ͑5͒ and corresponding impli- ␹d ͑ ͒ ␹d ͑␻͒ The of BA t will be denoted by ˜ BA .It cations hold true for any isolated, closed quantum system. is worth noting here that only when A=B does this part in Thus, upon contracting the dynamics in full phase onto fact coincide with the imaginary part of the complex-valued a reduced description of an open quantum system exhibiting ␹Љ ͑␻͒ susceptibility ˜ BA . dissipation these relations hold true nevertheless. Therefore, The fluctuations are described by the equilibrium corre- care must be taken when invoking approximations in order to lation function avoid any violation of these rigorous relations. We next con- sider the role of quantum dissipation for an exactly solvable ͑ ͒ ͗␦ ͑ ͒␦ ͑ ͒͘ ͑ ͒ CBA t = B t A 0 ␤ 3 situation: the damped quantum harmonic oscillator dynam- ␤ ics. at inverse temperature =1/kBT. The correlation function is complex-valued because the operators B͑t͒ and A͑0͒ in gen- ͑ ͒ eral do not commute. While the antisymmetric part of CBA t is directly related to the response function by linear response theory, the power spectrum of the symmetrized correlation III. QUANTUM DISSIPATION: THE DAMPED function HARMONIC OSCILLATOR ͑ ͒ 1 ͗␦ ͑ ͒␦ ͑ ͒ ␦ ͑ ͒␦ ͑ ͒͘ ͑ ͒ A. Equilibrium correlation functions SBA t = 2 B t A 0 + A 0 B t 4 Let us next consider the most fundamental case of a depends on the Fourier transform of the dissipative part of simple open quantum system, namely, the damped harmonic the response function via oscillator. This problem could be tackled by setting up a ប␻ microscopic model describing the coupling to environmental ͑␻͒ ប ͩ ͪ␹d ͑␻͒ ͑ ͒ SBA = coth ˜ . 5 degrees of freedom to which energy can be transferred irre- 2k T BA B versibly, thus giving rise to dissipation. Such an approach This result is the quantum version of the fluctuation- will be introduced in Sec. IV A. On the other hand, the lin- dissipation theorem as it relates the fluctuations described by earity of the damped harmonic oscillator allows us as alter- ͑␻͒ ␹d ͑␻͒ SBA to the dissipative part ˜ BA of the response. native to proceed on a phenomenological level. This ap- In the spirit of the work by Nyquist and Johnson we proach is closely related to the usual classical procedure consider as an example the response of a current ␦I through where damping is frequently introduced by adding in the an electric circuit subject to a voltage change ␦V. This im- equation of motion a force proportional to the velocity. plies B=I and, because the voltage couples to the charge Q, Classically, the motion of a harmonic oscillator subject A=Q. The response of the circuit is determined by ␦I͑␻͒ to linear friction is determined by

Downloaded 21 Jun 2005 to 137.250.81.48. Redistribution subject to AIP license or copyright, see http://chaos.aip.org/chaos/copyright.jsp 026105-3 Quantum Brownian motion Chaos 15, 026105 ͑2005͒

t tion. Quantum corrections to this term are relevant at tem- ͵ ␥͑ ͒ ͑ ͒ ␻2 ͑ ͒ Mq¨ + M ds t − s q˙ s + M 0q =0. 7 peratures of the order ប␻ /k or below, and these corrections ϱ 0 B − may be obtained from weak coupling theories like the quan- In the example of an electric circuit mentioned in the previ- tum master equation approach.17–20 However, there is an- ប␥ ␲ ous section, a damping kernel ␥͑t͒ with memory would cor- other regime at temperatures below /4 kB. Here, the sec- respond to a frequency-dependent admittance. In the special ond term may initially be small, but nevertheless it may case of ohmic friction corresponding to Y͑␻͒=1/R, the dominate the long-time behavior of the correlation function. damping force is proportional to the velocity of the harmonic This becomes particularly apparent in the limit of zero tem- oscillator, so that the equation of motion reads perature where the exponential functions in the second term in Eq. ͑10͒ sum up to an algebraic long-time behavior, i.e., ¨ ␥˙ ␻2 ͑ ͒ ͑ ͒ ͑ប␥ ␲ ␻4͒ −2 Mq + M q + M 0q =0. 8 Sqq t =− / M 0 t . Its relevance for the dynamical evo- lution of the damped harmonic oscillator depends on the de- In ͑7͒ and ͑8͒ the mass, frequency, and position of the oscil- tails of the initial preparation.21 Although the algebraic decay lator are denoted by M, ␻ , and q, respectively. Due to the 0 results from the zero temperature limit it can also be ob- , the equation of motion ͑7͒ is still valid in served at low, but finite temperatures during intermediate the quantum regime if we replace q by its expectation value. ␯ times before an exponential decay with time constant 1 sets As a consequence, the quantum mechanical dynamic suscep- 22 15,16 in. The occurrence of additional time scales besides ␥ at tibility agrees with the classical expression low temperatures leads to shortcomings with the quantum 1 1 regression hypothesis and allows for the decay of correla- ␹ ͑␻͒ = , ͑9͒ ␥ 13 qq ␻2 ␻␥͑␻͒ ␻2 tions on time scales longer than . M − − i ˜ + 0 where ˜␥͑␻͒ denotes the Fourier transform of the damping B. The reduced and the partition kernel ␥͑t͒. function As mentioned before, the response function directly In the previous section, we have seen that the dynamics yields the antisymmetric part of the position autocorrelation of a damped harmonic oscillator can be fully described in function C ͑t͒. It therefore suffices to discuss the symme- qq terms of the position autocorrelation function ͑10͒ and its trized part S ͑t͒ defined according to ͑4͒. Furthermore, our qq time derivatives as well as the ͑classical͒ response function. linear system with linear damping represents a stationary If one is interested only in equilibrium expectation values of Gaussian process so that all higher order correlation func- arbitrary operators acting in the Hilbert space of the har- tions may be expressed in terms of second order correlation monic oscillator, it is sufficient to know the reduced density functions.15 In addition, equilibrium correlation functions matrix. By means of arguments analogous to the dynamic containing momentum operators p can be reduced to position case presented in the previous section, the reduced density correlation functions by means of p=Mq˙. The dynamics of matrix can only depend on second moments of position and the damped harmonic oscillator can therefore entirely be de- 2 2 momentum, ͗q ͘␤ and ͗p ͘␤, respectively. The equilibrium scribed by the response function, i.e., the Fourier transform 6 ͑ ͒ ͑ ͒ 6,15,16 density matrix then necessarily takes the form of 9 and Sqq t . In the case of Ohmic damping, ˜␥͑␻͒=␥, the position 1 ␳␤͑q,qЈ͒ = autocorrelation can be explicitly evaluated from the ͑2␲͗q2͒͘1/2 fluctuation-dissipation theorem ͑5͒. The inverse Fourier 2 2 ͑q + qЈ͒ ͗p ͘␤ transform into the time domain is determined by the poles on ϫ ͫ ͑ Ј͒2ͬ ͑ ͒ exp − 2 − 2 q − q . 11 the right-hand side. The dissipative part of the dynamic sus- 8͗q ͘␤ ប ceptibility leads to four poles at ␻=±͑␻±i␥/2͒ with ␻ ͑␻2 ␥2 ͒1/2 The second moments are found to read = 0 − /4 which contribute to the correlation function ͑ ͒ +ϱ Sqq t at all temperatures. At sufficiently low temperatures, 2 1 1 ␻ ␯ ͗q ͘␤ = ͚ ͑12͒ the poles of the hyperbolic cotangent at =±i n with the ␤ ␻2 ␯2 ͉␯ ͉␥͉͑␯ ͉͒ ␯ ␲ ប␤ M n=−ϱ 0 + n + n ˆ n Matsubara frequencies n =2 n/ become important as well. After performing the contour integration in ͑5͒, one and arrives at6,15,16 +ϱ M ␻2 + ͉␯ ͉␥ˆ ͉͑␯ ͉͒ ͗ 2͘ 0 n n ͑ ͒ ប ␥ p ␤ = ͚ 2 2 , 13 ͑ ͒ ͩ ͉ ͉ͪ ␤ ϱ ␻ + ␯ + ͉␯ ͉␥ˆ ͉͑␯ ͉͒ S t = exp − t n=− 0 n n n qq 2M␻ 2 where we have introduced the Laplace transform of the sinh͑ប␤␻͒cos͑␻t͒ + sin͑ប␤␥/2͒sin͑␻͉t͉͒ ϫ damping kernel ͑ប␤␻͒ ͑ប␤␥ ͒ cosh − cos /2 ϱ ϱ ␥ˆ ͑ ͒ ͵ ͑ ͒␥͑ ͒ ͑ ͒ 2␥ ␯ exp͑− ␯ ͉t͉͒ z = dt exp − zt t . 14 − ͚ n n . ͑10͒ 0 ␤ 2 2 2 2 2 M n=1 ͑␯ + ␻ ͒ − ␥ ␯ n 0 n We note that for strictly Ohmic damping the second mo- In the limit of high temperatures the second term van- ment of the momentum ͑13͒ exhibits a logarithmic diver- ishes and the first term yields the classical correlation func- gence which can be removed by introducing a finite memory

Downloaded 21 Jun 2005 to 137.250.81.48. Redistribution subject to AIP license or copyright, see http://chaos.aip.org/chaos/copyright.jsp 026105-4 P. Hanggi and G.-L. Ingold Chaos 15, 026105 ͑2005͒

⌬ ͗ 2͘ ͑ ͒ ͑ ͒ FIG. 1. The weak coupling correction q to q ␤ according to 16 is FIG. 2. The density of states defined by inversion of the relation 19 for a ␥ ␻ depicted as a function of the temperature T.Fork Tӷប␻ , the correction weakly damped harmonic oscillator with =0.1 0 exhibits broadened peaks B 0 ⑀ ប␻ ⑀ becomes negligible. close to the 0 +n 0. A delta function at the energy 0 is not shown explicitly. The dotted line represents the average density of states.

to the damping mechanism. For finite coupling to the envi- ronment, i.e., for finite damping strength ␥, the reduced den- ͑ ͒ sity matrix 11 obviously does not agree with the canonical 2 1 d ͗q ͘␤ =− ln͑Z͒. ͑17͒ density matrix exp͑−␤H ͒ at the same temperature, where H ␤␻ ␻ S S M 0 d 0 denotes the Hamiltonian of the undamped harmonic oscilla- tor. This leads to the product representation of the partition func- In order to get an idea of the deviation of the true re- tion, i.e., duced density matrix from the canonical one, we consider the leading corrections to the second moment of the position

due to the finite coupling to the environment. Expanding in ϱ orders of the damping strength ␥, we obtain for Ohmic 1 ␯2 Z = ͟ n . ͑18͒ damping ប␤␻ ␯2 ␯ ␥͑␯ ͒ ␻2 0 n=1 n + n ˆ n + 0 The properties of this partition function become more 2 ͗q ͘␤͑␥͒ ␥ transparent if one relates it to a density of states ␳͑E͒ accord- =1+ ⌬ + O͑␥2͒͑15͒ 23 ͗ 2͘ ͑␥ ͒ ␲␻ q ing to q ␤ =0 0 with ϱ Z͑␤͒ = ͵ dE␳͑E͒exp͑− ␤E͒. ͑19͒ ប␤␻ 0 Im ␺Јͩi 0 ͪ ប␤␻ ␲ 0 2 ⌬ = . ͑16͒ −1 q ␲ ប␤␻ The factor ͑ប␤␻ ͒ in ͑18͒ can then be interpreted in terms 2 0 0 cothͩ ͪ of the average density of states ͑ប␻ ͒−1 indicated in Fig. 2 as 2 0 a dotted line. We further note that the partition function di- Here, ␺Ј denotes the first derivative of the digamma func- verges for strict Ohmic damping. However, it can be shown ⌬ tion. The correction q is depicted in Fig. 1 as a function of that this divergence is entirely due to a divergence of the ⑀ temperature. We find that the leading corrections are particu- ground state energy 0 in the presence of Ohmic Ӷប␻ 24 larly important in the quantum regime, kBT 0, while in dissipation. In contrast, for a bath with a spectral density ⑀ the classical regime the corrections to the canonical density possessing a high-frequency cutoff, 0 emerges to be finite matrix are negligible. ͓see Eq. ͑1.191͒ in Ref. 24͔. For large cutoff frequencies, the As we have already mentioned, a finite memory time of poles of the partition function, which determine the density ␯2 the damping kernel or, equivalently, a finite cutoff frequency of states, can then be determined from the condition n ␻ ␥␯ ␻2 D for the environmental mode spectrum is needed in order + n + 0 =0. These poles give rise to a density of states to keep the second moment of the momentum ͑13͒ finite. If which for weak damping exhibits narrow peaks whose width ␻ ӷ␻ ␥ D 0 , , the corrections to the canonical density matrix is in agreement with the result from Fermi’s Golden Rule. ␥ ␻ for weak coupling will only be small if the temperature is Figure 2 depicts an example for / 0 =0.1. In view of the ӷប␻ larger than the cutoff frequency, i.e., kBT D. remark made before, the density of states is shifted by the The differences between the correct reduced density ma- ground state energy. In addition, a delta peak at the ground trix ͑11͒ and the canonical one are also reflected in the par- state energy has been omitted. With increasing damping tition function. Without specifying a microscopic model for strength, the peaks broaden so that for sufficiently strong the environment, the partition function Z for the damped damping a rather featureless density of states results which harmonic oscillator can be obtained by the requirement that decreases with increasing energy to the average density of it generates the second moment of position according to15 states ͑cf. Fig. 3 in Ref. 23͒.

Downloaded 21 Jun 2005 to 137.250.81.48. Redistribution subject to AIP license or copyright, see http://chaos.aip.org/chaos/copyright.jsp 026105-5 Quantum Brownian motion Chaos 15, 026105 ͑2005͒

IV. DISSIPATION IN NONLINEAR QUANTUM that this model is also applicable to strongly damped systems SYSTEMS: THE GENERALIZED QUANTUM LANGEVIN and may be employed to describe, for example, dissipative 31 EQUATION „QLE… tunnelling in solid state physics and chemical physics. A. Bath of oscillators One may convince oneself that the Hamiltonian ͑20͒ in- deed models dissipation. Making use of the solution of the For nonlinear systems the arguments given in the previ- Heisenberg equations of motion for the external degrees of ous section no longer apply. In particular, second-order cor- freedom33 one derives a reduced system operator equation of relation functions are not sufficient anymore to completely motion, the so-called generalized quantum Langevin describe the damped system. An alternative approach to equation9 quantum dissipative systems starting from a Hamiltonian at first sight does not seem feasible because the absence of t dV͑q,t͒ time-dependent forces implies energy conservation. How- Mq¨͑t͒ + M͵ ds␥͑t − s͒q˙͑s͒ + = ␰͑t͒͑21͒ dq ever, as we will see below, once it is realized that dissipation t0 arises from the coupling to other degrees of freedom, it is with the damping kernel straightforward to model a damped quantum system in terms of a Hamiltonian. N A well known technique to describe a statistical dynam- 1 c2 ␥͑t͒ = ␥͑− t͒ = ͚ i cos͑␻ t͒͑22͒ ics governed by fluctuations is given by the method of gen- ␻2 i M i=1 mi i eralized master equations and the methodology of general- ized Langevin equations. This strategy is by now well and the quantum Brownian force operator developed for thermal equilibrium systems. Here the projec- tor operator technique17–20,25 yields a clear-cut method to ob- N ͑ ͒ ␰͑ ͒ ␥͑ ͒ ͑ ͒ ͩ ͑ ͒ ͑␻ ͓ ͔͒ tain the formal equations, either the generally nonlinear t =−M t − t0 q t0 + ͚ ci xi t0 cos i t − t0 generalized quantum Langevin equation ͑QLE͒ or the gener- i=1 alized quantum master equation ͑QME͒ for the rate of p ͑t ͒ + i 0 sin͑␻ ͓t − t ͔͒ͪ. ͑23͒ change of the reduced density matrix. m ␻ i 0 Already for the case of relaxation towards a unique ther- i i mal equilibrium specified by a single temperature T, the The generalized quantum Langevin equation ͑21͒ appears 9 ͑ ͒ equivalence between the two approaches is not very first in a paper by MagalinskiŽ who started from 20 in the transparent.26 A crucial role is played by the fluctuational absence of the potential term. force which explicitly enters the equivalence, such as corre- The force operator ͑23͒ depends explicitly on the initial ͑ ͒ sponding cumulant averages to an arbitrary high order. This conditions at time t0 of the bath position operators xi t0 and ͑ ͒ fact is not appreciated generally, because one often restricts bath momenta pi t0 . The initial preparation of the total sys- the discussion to the first two cumulants only, namely the tem, which fixes the statistical properties of the bath opera- average and its autocorrelation. It is a fact that little is known tors and the system degrees of freeedom, turns the force ␰͑t͒ about the connection of the generalized master equation and into a random operator. Note that this operator depends not the corresponding generalized Langevin equation in a non- only on the bath properties but as well on the initial system 26,27 ͑ ͒ linear situation. position q t0 . To fully specify the reduced dynamics it is A popular model for the dynamics of a dissipative quan- thus of importance to specify the preparation procedure. This tum system subject to quantum Brownian noise is obtained in turn then also fixes the statistical properties of the quan- by coupling the system of interest to a bath of harmonic tum Brownian noise. Clearly, in order to qualify as a stochas- oscillators. Accordingly, we write for the total Hamiltonian tic force the random force ␰͑t͒ should not be biased; i.e. its average should be zero at all times. Moreover, this Brownian p2 quantum noise should constitute a stationary process with H = + V͑q,t͒ 2M time-homogeneous correlations. Let us also introduce the auxiliary random force ␩͑t͒, N 2 2 pi mi ci defined by + ͚ ͫ + ␻2x2 − qc x + q2 ͬ, ͑20͒ i i i i ␻2 i=1 2mi 2 2mi i ␩͑t͒ = ␰͑t͒ + M␥͑t − t ͒q͑t ͒͑24͒ where the first two terms describe the system as a particle of 0 0 mass M moving in a generally time-dependent potential which only involves bath operators. In terms of this new V͑q,t͒. The sum contains the Hamiltonian for a set of N random force the QLE ͑21͒ no longer assumes the form of an harmonic oscillators which are bilinearly coupled with ordinary generalized Langevin equation: it now contains an ␥͑ ͒ ͑ ͒ 24,27 strength ci to the system. Finally, the last term, which de- inhomogeneous term t−t0 q t0 , the initial slip term. pends only on the system coordinate, represents a potential This term is often neglected in the so-called “Markovian renormalization term which is needed to ensure that V͑q,t͒ limit” when the friction kernel assumes the Ohmic form remains the bare potential. This Hamiltonian has been stud- ␥͑t͒→2␥␦͑t͒. For a correlation-free preparation, the initial ␳ ␳ ͑ ͒␳ ied since the early 1960’s for systems which are weakly total density matrix is given by the product T = S t0 bath, 9,18–20,28–31 ␳ ͑ ͒ coupled to the environmental degrees of freedom. where S t0 is the initial system density matrix. The density Only after 1980, it was realized by Caldeira and Leggett32 matrix of the bath alone assumes canonical equilibrium, i.e.,

Downloaded 21 Jun 2005 to 137.250.81.48. Redistribution subject to AIP license or copyright, see http://chaos.aip.org/chaos/copyright.jsp 026105-6 P. Hanggi and G.-L. Ingold Chaos 15, 026105 ͑2005͒

N 2 B. Consequences of time-reversal symmetry 1 p m ␳ = expͩ− ␤͚ ͫ i + i ␻2x2ͬͪ, ͑25͒ bath N i i i=1 2mi 2 Let us now discuss some further properties of this QLE. If the potential V͑q,t͒ in ͑20͒ does not explicitly depend on with N denoting a normalization constant. time t, the dynamics of the full Hamiltonian ͑20͒ obeys time The statistical properties of the random force ␩͑t͒ then reversal symmetry. It is thus an immediate consequence that follow immediately: ␩͑t͒ is a stationary Gaussian operator the reduced dynamics must be invariant under time reversal noise obeying as well. This must hold true despite the fact that the QLE has been constructed to allow for a description of quantum dis- ͗␩͑t͒͘␳ =0, ͑26͒ bath sipation. It is thus instructive to see how the validity under time reversal emerges from the contracted description in ͑ ͒ 1 terms of the QLE in 21 . S␩␩͑t − s͒ = ͗␩͑t͒␩͑s͒ + ␩͑s͒␩͑t͒͘␳ 2 bath Given the time of preparation t0, reversing the time N amounts to substituting time t by t −͑t−t ͒=2t −t. Using ប c2 ប␻ 0 0 0 = ͚ i cos͑␻ ͑t − s͒͒cothͩ i ͪ. ͑27͒ again the random force ␩͑t͒ we can recast the QLE dynamics ␻ i 2 i=1 mi i 2kBT after the time reversal into the form

Being an operator-valued noise, its commutator does not 2t0−t dV͑q͒ Mq¨͑2t − t͒ + M͵ ds␥͑2t − t − s͒q˙͑s͒ + vanish 0 0 dq t0 N 2 ␰͑ ͒ ␩͑ ͒ ␥͑ ͒ ͑ ͒ ͑ ͒ c = 2t0 − t = 2t0 − t − 2t0 − t − t0 q t0 . 30 ͓␩͑t͒,␩͑s͔͒ =−iប͚ i sin͑␻ ͑t − s͒͒. ͑28͒ ␻ i ͑ ͒ ͑ ͒ ͑ ͒ i=1 mi i Setting next x t =q 2t0 −t and observing that x˙ t =−q˙͑2t −t͒,x¨͑t͒=q¨͑2t −t͒, we find after the substitution of ͑ ͒ 0 0 Setting for the initial position operator q t0 =q0, the last ͑ ͒ the integration time u=2t0 −s from 30 the result expression in ͑27͒ is also valid for the noise correlation S␰␰͑t͒ of the noise force ␰͑t͒ provided the average is now taken t dV͑q͒ ¨͑ ͒ ͵ ␥͑ ͒˙͑ ͒ with respect to a bath density matrix which contains shifted Mx t + M du u − t x u + t dq oscillators. The initial preparation of the bath is then given 0 ␳ ␰͑ ͒ ␩͑ ͒ ␥͑ ͒ ͑ ͒ ͑ ͒ by the new density matrix ˆ bath; = 2t0 − t = 2t0 − t − t0 − t x t0 . 31 Noting that the damping kernel is an even function of its 2 ␻2 2 1 p mi ci ␥͑ ͒ ␥͑ ͒ ͑ ͒ ͑ ͒ ␳ˆ = expͭ− ␤͚ ͫ i + i ͩx − q ͪ ͬͮ. argument, u−t = t−u , and that x t0 =q t0 ,wefind bath N i ␻2 0 ͑ ͒ i 2mi 2 mi i upon changing all signs of the initial momenta pi t0 → ͑ ͒ ␩͑ ͒ ␩͑ ͒ ͑29͒ −pi t0 for the noise forces the relations 2t0 −t = t ␰͑ ͒ ␰͑ ͒ and 2t0 −t = t . We conclude that the time reversed mo- ͑ ͒ ͑ ͒ In some physical situations a microscopic model for the tion x t =q 2t0 −t indeed obeys again a QLE of the form external degrees of freedom is available.31,34 Examples are ͑21͒. This even holds true in the Markovian limit where the electromagnetic modes in a resonator acting as a reser- ␥͑t−s͒=2␥␦͑t−s͒ as one can convince oneself by smearing voir or the dissipation arising from quasiparticle tunnelling out the delta function symmetrically. The QLE then reads for 35 through Josephson junctions. In the case of an electrical all times t circuit containing a resistor one may use the classical equa- dV͑q͒ tion of motion to obtain the damping kernel and model the ͑ ͒ ͑ ͒ ␥ ͑ ͒ ␰͑ ͒ ͑ ͒ Mq¨ t + sgn t − t0 M q˙ t + = t , 32 environment accordingly. This approach has been used, e.g., dq to model Ohmic dissipation in Josephson junctions in order ͑ ͒ 36 where sgn x denotes the sign of x. to study its influence on tunnelling processes, and to de- The dissipation is reflected by the fact that for times t scribe the influence of an external impedance in the charge Ͼ ͑ ͒ 37 t0 the reduced dynamics for q t exhibits a damped dynamics of ultrasmall tunnel junctions. ͑quantum-͒behavior on a time scale given by the Poincaré This scheme of the QLE can also be extended to the recurrence time;41,42 the latter reaches essentially infinity for nonequilibrium case with the system attached to two baths of 38 all practical purposes if only the bath consists of a sizable different temperature. A most recent application addresses number of bath oscillator degrees of freedom.31,41 the problem of the thermal conductance through molecular wires that are coupled to leads of different temperature. Then C. Subtleties and pitfalls the heat current assumes a form similar to the Landauer for- mula for electronic transport: The heat current is given in The use of the generally nonlinear QLE ͑21͒ is limited in terms of a transmission factor times the difference of corre- practice for several reasons. Moreover, the application of the sponding Bose functions.39 QLE bears some subtleties and pitfalls which must be ob- Furthermore, the QLE concept can also be extended to served when making approximations. Some important fea- fermionic systems coupled to electron reservoirs and which, tures are: The QLE ͑21͒ is an operator equation that acts in in addition, may be exposed to time-dependent driving.40 the full Hilbert space of system and bath. The coupling be- The corresponding Gaussian quantum noise is now com- tween system and environment also implies an entanglement posed of fermion annihilation operators. upon time evolution even for the case of an initially factor-

Downloaded 21 Jun 2005 to 137.250.81.48. Redistribution subject to AIP license or copyright, see http://chaos.aip.org/chaos/copyright.jsp 026105-7 Quantum Brownian motion Chaos 15, 026105 ͑2005͒

izing full density matrix. Together with the commutator S␰␰͑t͒ = S␩␩͑t͒ ͑ ͒ property of quantum Brownian motion, see Eq. 28 ,wefind ϱ M ប␻ that the reduced, dissipative dynamics of the position opera- = ͵ d␻ Re ␥ˆ ͑− i␻ +0+͒ប␻ cothͩ ͪcos͑␻t͒. ␲ tor q͑t͒ and momentum operator p͑t͒ obey the Heisenberg 0 2kBT uncertainty relation for all times. ͑35͒ This latter feature is crucial. For example, the non- Markovian ͑colored͒ Gaussian quantum noise with real- In the classical limit this relation reduces, independent of valued correlation S␰␰͑t͒=S␰␰͑−t͒ cannot simply be substi- the preparation of the bath with ␳ or ␳ˆ, to the non-Markovian ͑ ͒ ␥͑ ͒ ͑ ͒ tuted by a classical non-Markovian Gaussian noise force Einstein relation S␰␰ t =MkBT t . The relation 35 is by no which identically obeys the correlation properties of ͑Gauss- means obvious: It implies that a modelling of quantum dis- ian͒ quantum noise ␰͑t͒. An approximation of this type sipation is possible in terms of macroscopic quantities such clearly would not satisfy the commutator property for posi- as the friction kernel ␥͑t͒ and the temperature T. For other tion and conjugate momentum of the system degrees of free- coupling schemes between system and bath we generally can dom. no longer express the correlation of quantum noise exclu- The literature is full of various attempts wherein one sively in terms of macroscopic transport coefficients. As an approximates the quantum features by corresponding colored example we mention the coupling of the system to a bath of classical noise sources, e.g., see Refs. 43–45. Such schemes two-level systems ͑spin bath͒ rather than to a bath of har- work at best near a quasiclassical limit,44,46 but even then monic oscillators.48 care must be exercised. For example, for problems that ex- hibit an exponential sensitivity, as, e.g., the dissipative decay Note, also the following differences to the classical situ- of a metastable state discussed in the next section, such an ation of a generalized Langevin equation: The quantum noise ␰͑ ͒ ͑ ͒ 49 approach gives no exact agreement with the quantum dissi- t is correlated with the initial position operator q t0 . 31,34 ͗ ͑ ͒␰͑ ͒͘ Þ pative theory. It is only in the classical high temperature This feature that q t0 t ␳ˆ 0 follows from the explicit limit, where the commutator structure of quantum mechanics form of the quantum noise ␰͑t͒. The correlation function van- no longer influences the result. Perfect agreement is only ishes only in the classical limit. Note also that the expecta- achieved in the classical limit. tion value of the system–bath interaction is finite at zero The study of quantum friction in a nonlinear quantum temperature. These features reflect the fact that at absolute system by means of the QLE ͑21͒ is plagued by the fact that zero temperature the coupling induces a nonvanishing deco- the nonlinearity forbids an explicit solution. This solution, herence via the zero-point fluctuations. ␥͑ ͒ ͑ ͒ however, is needed to obtain the statistical properties such as Moreover, the initial slip term t−t0 q t0 appears also mean values and correlation functions. This ͑unknown͒ non- in the absence of the potential renormalization in the Hamil- linear response function also determines the derivation of the tonian ͑20͒. With this initial value contribution being ab- rate of change of the reduced density operator, i.e. the QME, sorbed into the quantum fluctuation ␰͑t͒, these become sta- and its solution for the open quantum system. tionary fluctuations with respect to the initial density The very fact that the QLE acts in full Hilbert space of operator of the bath ␳ˆ given by ͑29͒. Note, however, that system and environment also needs to be distinguished from bath with respect to an average over the bare, nonshifted bath the classical case of a generalized Langevin equation. There, density operator ␳ , the quantum fluctuations ␰͑t͒ would the stochastic dynamics acts solely on the state space of the bath become nonstationary. system dynamics with the ͑classical͒ noise properties speci- fied a priori.47 It is also worthwhile to point out here that this initial value term in the QLE should not be confused with the initial The quantum noise correlations can, despite the explicit 17,18 microscopic expression given in ͑27͒, be expressed solely by value term that enters the corresponding QME. In the the macroscopic friction kernel ␥͑t͒. This result follows upon case of a classical reduced dynamics it is always noting that the Laplace transform ␥ˆ ͑z͒ of the macroscopic possible—by use of a corresponding projection operator—to friction assumes with Re zϾ0 the form formally eliminate this initial, inhomogeneous contribution in the generalized master equation.47,50 This in turn renders

N the time evolution of the reduced a truly linear 1 c2 1 1 ␥ˆ ͑z͒ = ͚ i ͫ + ͬ. ͑33͒ dynamics. This property no longer holds for the reduced ␻2 ␻ ␻ 51 2M i=1 mi i z − i i z + i i quantum dynamics. For a non-factorizing initial prepara- tion of system and bath this initial value contribution in the With help of the well known relation 1/͑x+i0+͒= P͑1/x͒ QME generally is finite and presents a true nonlinearity for −i␲␦͑x͒ we find that the time evolution law of the open quantum dynamics! There exist even further subtleties which are worthwhile Re ␥ˆ ͑z =−i␻ +0+͒ to point out. The friction enters formally the QLE just in the same way as in the classical generalized Langevin equation. N 2 ␲ c In particular, a time-dependent potential V͑q,t͒ leaves this = ͚ i ͓␦͑␻ − ␻ ͒ + ␦͑␻ + ␻ ͔͒. ͑34͒ ␻2 i i 2M i=1 mi i friction kernel invariant in the QLE. In contrast to the clas- sical Markovian case, however, where the friction enters the By means of ͑27͒ we then find the useful relation corresponding Fokker–Planck dynamics independent of the

Downloaded 21 Jun 2005 to 137.250.81.48. Redistribution subject to AIP license or copyright, see http://chaos.aip.org/chaos/copyright.jsp 026105-8 P. Hanggi and G.-L. Ingold Chaos 15, 026105 ͑2005͒

time scale of driving, this is no longer valid for the general- q͑ប␤͒=q 1 1 E ized quantum master equation dynamics of the correspond- ␳␤͑q,qЈ͒ = ͵ Dq expͩ− S ͓q͔ͪ, ͑37͒ Z␤ ͑ ͒ Ј ប ing reduced density matrix.52,53 q 0 =q The solution of the QLE involves the explicit time- where Z␤ is the partition function. This integral is called dependence of both the friction and the potential forces. imaginary-time path integral in contrast to the real-time path These in turn determine the statistical properties of the den- integral ͑36͒. Note that in ͑37͒ the action S͓q͔ has been re- sity matrix. As a consequence, the friction force enters the placed by the so-called Euclidean action SE͓q͔ which is ob- QME in a rather complex manner. This can already be veri- tained by changing the sign of the potential term as a conse- fied explicitly for a parametric dissipative oscillator dynam- quence of the transition to imaginary times. In imaginary ics, where the time-dependent driving enters the diffusive time we therefore have to consider the motion in the inverted kinetic evolution law of the reduced density operator or its potential. equivalent Wigner transform.52,53 The connection between classical and quantum mechan- For the bilinear system–bath interaction with the bath ics becomes particularly apparent in the path integral formu- composed of harmonic oscillators it was possible to integrate lation. The dominant contribution to the integrals in ͑36͒ and out the degrees of freedom of the bath explicitly. Does this ͑37͒ arise from the stationary points of the action, i.e., the hold as well for other interactions? The elimination of the classical paths. Quantum effects have their origin in fluctua- bath degrees of freedom is still possible for a nonlinear cou- tions around the classical paths. Therefore, it is useful to pling to a bath of harmonic oscillators if the system part of decompose a general path into the classical path and a fluc- the coupling is replaced by a nonlinear operator-valued func- tuation around it. Expanding the action in powers of the fluc- tion of either the momentum or position degree of freedom tuations the second order term yields the leading quantum of the system as long as the bath degrees of freedom appear corrections. Higher order terms are often neglected within a linearly. The resulting friction kernel then appears as a non- semiclassical approximation which becomes exact for linear linear friction but the influence of the bath degrees of free- systems. dom still is obtained in exact form.27 In the previous section we have derived an effective Yet another situation for which one can derive an exact equation of motion for the system variable by eliminating the QLE is when a nonlinear system, such as a spin degree of external degrees of freedom. The same procedure may of freedom, interacts with a collection of quantum ͑Bose͒ oscil- course also be carried out within the path integral lators in such a way that the interaction Hamiltonian com- formalism.6,31,34,58 The influence of the environment is then mutes with the system Hamiltonian, thus constituting a quan- contained in an effective action which has to be added to the tum nondemolition interaction. This case corresponds to pure action of the system and which in imaginary time is given dephasing and was addressed by Łuczka for the problem of a by6,31,34 spin in contact with a thermal heat bath.54 It has since been ប␤ ប␤ rederived many times, see, e.g., Ref. 55. 1 ͓ ͔ ͵ ␶͵ ␴ ͑␶ ␴͓͒ ͑␶͒ ͑␴͔͒2 ͑ ͒ We end this subsection by mentioning also the coupling Seff q =− d d k − q − q , 38 of a system to a bath of independent fermions with infinitely 4 0 0 many excitation energies. A suitable transformation then al- where lows to map the dissipation onto a bosonic environment with 31,40,56 an appropriate coupling strength. +ϱ M k͑␶͒ = ͚ ͉␯ ͉␥ˆ ͉͑␯ ͉͒exp͑i␯ ␶͒͑39͒ ប␤ n n n n=−ϱ V. PATH INTEGRALS AND EFFECTIVE ACTION and ␥ˆ ͑z͒ denotes the Laplace transform of the damping ker- A. Nonlocal effective action nel ␥͑t͒. The effective action ͑38͒ is clearly nonlocal and can A most effective approach to describe dissipation is thus not be expressed in terms of a potential. If the potential ͑ ͒ based on the path integral formulation of quantum renormalization term in the Hamiltonian 20 would be ab- ͑ ͒ mechanics.57 In the path integral formulation of quantum sent, there would have been a local contribution in 38 . The ͑ ͒ mechanics the propagator is expressed as selfinteraction of the paths induced by 38 via the kernel ͑39͒ decays for Ohmic damping only algebraically as ␶−2 and ͑ ͒ therefore represents a long range interaction. i q t =qf i ͗ ͉ ͩ ͉ͪ ͘ ͵ D ͩ ͓ ͔ͪ ͑ ͒ qf exp − Ht qi = q exp S q , 36 ប ͑ ͒ ប q 0 =qi where the integral runs over all possible paths starting at q B. Application: The dissipative decay of a metastable i state and ending after time t at qf. The paths are weighted with a phase factor which contains the classical action S͓q͔. A local potential minimum may be metastable due to the For the description of quantum dissipative systems it is environmental coupling and quantum effects. Correspond- important to realize the analogy between the propagator and ingly, there are two escape mechanisms: thermal activation the equilibrium density matrix. The latter is obtained by re- which dominates at high temperatures and quantum tunnel- placing t by −iប␤. We thus obtain from ͑36͒ the path integral ling which becomes important at low temperatures. To be representation of the equilibrium density matrix definite, we consider the cubic potential

Downloaded 21 Jun 2005 to 137.250.81.48. Redistribution subject to AIP license or copyright, see http://chaos.aip.org/chaos/copyright.jsp 026105-9 Quantum Brownian motion Chaos 15, 026105 ͑2005͒

stant solutions q=0 in the well and qb =2q0 /3 at the barrier. Below a temperature given by the positive solution of ␯2 ͉␯ ͉␥͉͑␯ ͉͒ ␻2 ͑ ͒ 1 + 1 ˆ 1 − b =0 41 a second fluctuation mode becomes unstable, thereby indi- cating a new classical solution which performs an oscillation around the barrier.66 This new solution is associated with . Therefore, ͑41͒ defines the crossover FIG. 3. Cubic potential as defined in Eq. ͑40͒. temperature which for Ohmic damping is given by59 ប ␥2 1/2 ␥ T = ͫͩ + ␻2ͪ − ͬ. ͑42͒ 0 ␲ b M q 2 kB 4 2 ͑ ͒ ␻2 2ͩ ͪ ͑ ͒ V q = 0q 1− 40 2 q0 As discussed above, stronger damping leads to a lower cross- which is depicted in Fig. 3. The barrier height is given by over temperature and smaller quantum regime. It thus makes ͑ ͒ ␻2 2 the system more classical. A distinct feature of the dissipa- Vb = 2/27 M 0q0 and, in this special case, the barrier angu- ␻ ␻ tive quantum decay in the low temperature regime is its al- lar frequency b equals the well angular frequency 0. 31 In Fig. 4, the decay rate is shown in an Arrhenius plot. gebraic enhancement of the decay rate with temperature. At the so-called crossover temperature T , see Eq. ͑42͒ be- For the case of an Ohmic environment with a constant fric- 0 tion behavior at low frequencies one finds a universal low, there is a rather distinct transition between the thermal 2 regime on the left-hand side and the quantum regime on the T -enhancement of both, the prefactor and the effective ac- 59 tion, with the latter dominating the exponential rate right-hand side. Furthermore, we observe that the thermal 31,67 regime is larger for stronger damping, i.e., the system be- enhancement. comes more classical. While a real time approach to dissipative decay is VI. SUNDRY REMARKS AND CONCLUSIONS 61–64 feasible, a simpler alternative is provided by an imagi- With this work we elucidated the topic of quantum nary time calculation where the partition function Z␤ is con- Brownian noise which drives the dynamics of open dissipa- ͑ ͒ sidered. Since the potential 40 is not bounded from below, tive quantum systems. We have emphasized the strong im- it is no surprise that strictly speaking Z␤ does not exist. From plications that thermal equilibrium and time-reversal symme- the path integral point of view there exists an unstable fluc- try ͑leading to detailed balance symmetry͒ imposes on the tuation mode around the barrier which leads to a saddle point reduced system dynamics. We also pointed out the advanta- in function space. One can circumvent this difficulty by per- geous use of the path integral scheme for the case of nonlin- forming the integration in the direction of steepest descent. earity and strong friction. The partition function and as a consequence also the free This method seemingly is superior to any perturbative energy then acquire an imaginary part which may be related 31,32 scheme that treats the system–bath coupling to low orders to the decay rate. For details of this relation we refer the only, such as the weak coupling master equation reader to the discussion in Ref. 65. methodology.17–20 There are recent developments in the The transition between thermal and quantum regime can strong friction regime, where an alternative description in be well understood within the path integral picture by con- terms of a quantum Smoluchowski equation is promising,68 ប␤ sidering the possible classical paths of duration in the see also the contribution by Grabert, Ankerhold, and Pechu- inverted cubic potential. For high temperatures or short kas in this Focus Issue. ប␤ imaginary times the only classical solutions are the con- A consequent use of the so-called rotating-wave approxi- mations also may entail some danger. It safely can be applied only in the weak coupling regime for resonant situations. We remark that the use of the rotating-wave approximation im- plies a violation of the Ehrenfest theorem in the order of ␥2,13,69 which is clearly small only in the weak coupling ␥Ӷ␻ ␻−1 regime, i.e., for 0, with 0 denoting some typical time scale of the system dynamics. The same remarks apply to the failure of the quantum regression theorem:13,70,71 Again, the effect might be small for ͑i͒ very weak damping, ͑ii͒ not too ӷប␥ ͑ ͒ low temperatures obeying kBT , and iii not too short evolution times. The generalized quantum Langevin equation discussed in Sec. IV is formally exact for nonlinear quantum systems. Its practical use is typically restricted, however, to linear systems for which the response can be evaluated in closed FIG. 4. Arrhenius plot for the decay rate of a metastable state. The damping ␥ ␻ form. This holds true even for time-dependent linear systems strength varies from the upper to the lower curve as /2 0 =0,0.5, and 1 ͑data taken from Ref. 60͒. for which the response is still linear although the evaluation

Downloaded 21 Jun 2005 to 137.250.81.48. Redistribution subject to AIP license or copyright, see http://chaos.aip.org/chaos/copyright.jsp 026105-10 P. Hanggi and G.-L. Ingold Chaos 15, 026105 ͑2005͒

involves the use of numerical Floquet theory.52 The lack of 4It is worthwhile to recall that the validity of the equipartition theorem is knowledge of this generally nonlinear response function also restricted to classical statistical mechanics. 5H. B. Callen and T. A. Welton, “Irreversibility and generalized noise,” plagues the evaluation of the corresponding generalized mas- Phys. Rev. 83, 34–40 ͑1951͒. ter equations. 6H. Grabert, P. Schramm, and G.-L. Ingold, “Quantum Brownian motion: This problem of obtaining the generalized master equa- The functional integral approach,” Phys. Rep. 168, 115–207 ͑1988͒. 7T. Dittrich, P. Hänggi. G.-L. Ingold, B. Kramer, G. Schön, and W. tion from the nonlinear generalized Langevin equation is not ͑ 27 Zwerger, Quantum Transport and Dissipation Wiley-VCH, Weinheim, solved either for the classical problem with colored noise. 1998͒. It is also this very problem that limits the practical use of the 8M. Grifoni and P. Hänggi, “Driven quantum tunneling,” Phys. Rep. 304, 229–358 ͑1998͒. various variants of recently derived stochastic Schrödinger 9 12 V. B. MagalinskiŽ, “Dynamical model in the theory of the Brownian mo- approaches. tion,” Sov. Phys. JETP 9, 1381–1382 ͑1959͓͒J. Exp. Theor. Phys. 36, Likewise, the use of nonlinear, but nonstochastic 1942–1944 ͑1959͔͒; R. Benguria and M. Kac, “Quantum Langevin equa- Schrödinger equations of the type discussed and surveyed in tion,” Phys. Rev. Lett. 46,1–4͑1981͒; G. W. Ford and M. Kac, “On the Refs. 72–74 can clearly not describe the time evolution of a quantum Langevin equation,” J. Stat. Phys. 46, 803–810 ͑1987͒;G.W. Ford, J. T. Lewis, and R. F. O’Connell, “Quantum Langevin equation,” quantum mechanical mixture, nor do these nonlinear deter- Phys. Rev. A 37, 4419–4428 ͑1988͒. ministic approaches obey, in general, the Heisenberg uncer- 10E. Nelson, “Derivation of the Schrödinger equation from Newtonian me- tainty relation.73,74 chanics,” Phys. Rev. 150, 1079–1085 ͑1966͒. 11 There have been repeated attempts since the early days H. Grabert, P. Hänggi, and P. Talkner, “Is quantum mechanics equivalent to a classical stochastic process?” Phys. Rev. A 19, 2440–2445 ͑1979͒. of quantum mechanics to explain quantum phenomena in 12M. Roncadelli, “Connection between Langevin quantization and classical terms of Einstein’s theory of classical diffusion. Early efforts mechanics,” Europhys. Lett. 28, 379–384 ͑1994͒; H. Kleinert and S. V. in this direction were those of Fürth,75 Fényes,76 Weizel,77 Shabanov, “Quantum Langevin equation from forward-backward path in- ͑ ͒ and Favella.78 This credo has been popularized later by tegral,” Phys. Lett. A 200, 224–232 1995 ; L. Diósi and W. T. Strunz, 10,79 “The non-Markovian stochastic Schrödinger equation for open systems,” Nelson under the label of “Stochastic Mechanics.” It can ibid. 235, 569–573 ͑1997͒; W. T. Strunz, L. Diósi, and N. Gisin, “Open convincingly be demonstrated, however, that a quantum dy- system dynamics with non-Markovian quantum trajectories,” Phys. Rev. namics is quite distinct from a classical Markovian—or even Lett. 82, 1801–1805 ͑1999͒; H. P. Breuer, B. Kappler, and F. Petruccione, non-Markovian—stochastic dynamics.11,80 This holds even “Stochastic method for non-Markovian quantum master equations,” Phys. Rev. A 59, 1633–1643 ͑1999͒; J. T. Stockburger and H. more so, if one attempts to incorporate the quantum dissipa- Grabert, “Non-Markovian diffusion,” Chem. Phys. 268, tion for an open system. 249–256 ͑2001͒; J. Shao, “Decoupling quantum dissipation interaction via These sundry remarks thus give clear evidence that stochastic fields,” J. Chem. Phys. 120, 5053–5056 ͑2004͒. 13P. Talkner, “The failure of the quantum regression hypothesis,” Ann. Phys. the topic of quantum Brownian motion—although 100 ͑N.Y.͒ 167, 390–436 ͑1986͒; “Untersuchungen irreversibler Prozesse in years have passed since Einstein’s cornerstone quantenmechanischen Systemen,” Ph.D. thesis, Universität Stuttgart, contribution1—cannot be considered as “solved.” For ex- 1979. 14 ample, little is presently known also for the description and M. Planck, “Eine neue Strahlungshypothese,” Verh. Dtsch. Phys. Ges. 13, 138–148 ͑1911͒. the role of quantum noise acting in steady state, far from 15H. Grabert, U. Weiss, and P. Talkner, “Quantum theory of the damped equilibrium situations, i.e., when several baths of different harmonic oscillator,” Z. Phys. B: Condens. Matter 55, 87–94 ͑1984͒. nature and/or different temperature are coupled to the non- 16P. S. Riseborough, P. Hänggi, and U. Weiss, “Exact results for a damped linear system of interest. quantum-mechanical harmonic oscillator,” Phys. Rev. A 31, 471–478 ͑1985͒. The latter case is also of salient importance for the de- 17H. Grabert, Projection Operator Techniques in Nonequilibrium Statistical scription of the quantum dynamics of so-called quantum Mechanics, Springer Tracts in Modern Physics ͑Springer, Berlin, 1982͒, Brownian motors.81 In those applications quantum Brownian Vol. 95, pp. 1–164. ͑ 18F. Haake, Statistical Treatment of Open Systems by Generalized Master noise is utilized in combination with nonequilibrium classi- ͑ ͒ ͒ Equations, Springer Tracts in Modern Physics Springer, Berlin, 1973 , cal or quantum fluctuations to perform exploitable work Vol. 66, pp. 98–168. against external bias forces. In summary, the field of quan- 19H. Spohn, “Kinetic equations from Hamiltonian dynamics: Markovian limits,” Rev. Mod. Phys. 52, 569–615 ͑1980͒. tum Brownian motion is very much alive and lots of chal- 20 lenges still need to be addressed and mastered. R. Alicki, “General theory and applications to unstable particles,” in Quantum Dynamical Semigroups and Applications, Lect. Notes Phys. ͑Springer, Berlin, 1987͒, Vol. 286, Chaps. II and III. 21 ACKNOWLEDGMENTS G.-L. Ingold and H. Grabert, “Sluggish decay of preparation effects in low temperature quantum systems,” Lect. Notes Math. 1442, 219–230 ͑1990͒. 22 P.H. gratefully acknowledges financial support by the R. Jung, G.-L. Ingold, and H. Grabert, “Long-time tails in quantum ͑ Brownian motion,” Phys. Rev. A 32, 2510–2512 ͑1985͒. DAAD-KBN German-Polish project Stochastic Complex- 23A. Hanke and W. Zwerger, “Density of states of a damped quantum os- ity͒, the Foundation for Polish Science ͑Fundacja na Rzecz cillator,” Phys. Rev. E 52, 6875–6878 ͑1995͒. Nauki Polskiej͒, the Deutsche Forschungsgemeinschaft via 24G.-L. Ingold, “Path integrals and their application to dissipative quantum ͑ ͒ Grant No. HA 1517/13-4 and the collaborative research cen- systems,” Lect. Notes Phys. 611, 1–53 2002 . 25H. Mori, “Transport, collective motion and Brownian motion,” Prog. ters SFB 486 and SFB 631. Theor. Phys. 33, 423–455 ͑1965͒; K. Kawasaki, “Simple derivations of generalized linear and nonlinear Langevin equations,” J. Phys. A 6, 1289– 1A. Einstein, “Über die von der molekularkinetischen Theorie der Wärme 1295 ͑1973͒; S. Nordholm and R. Zwanzig, “Systematic derivation of geforderte Bewegung von in ruhenden Flüssigkeiten suspendierten exact generalized Brownian-motion theory,” J. Stat. Phys. 13, 347–371 Teilchen,” Ann. Phys. 17, 549–560 ͑1905͒. ͑1975͒. 2H. Nyquist, “Thermal agitation of electric charge in conductors,” Phys. 26H. Grabert, P. Hänggi, and P. Talkner, “Microdynamics and nonlinear Rev. 32, 110–113 ͑1928͒. stochastic processes of gross variables,” J. Stat. Phys. 22, 537–552 3J. B. Johnson, “Thermal agitation of in conductors,” Phys. Rev. ͑1980͒. 32, 97–109 ͑1928͒. 27P. Hänggi, “Generalized Langevin equations: A useful tool for the per-

Downloaded 21 Jun 2005 to 137.250.81.48. Redistribution subject to AIP license or copyright, see http://chaos.aip.org/chaos/copyright.jsp 026105-11 Quantum Brownian motion Chaos 15, 026105 ͑2005͒

plexed modeller of nonequilibrium fluctuations? ” Lect. Notes Phys. 484, tion,” J. Stat. Phys. 78, 299–310 ͑1995͒. 15–22 ͑1997͒. 56L.-D. Chang and S. Chakravarty, “Dissipative dynamics of a two-state 28I. R. Senitzky, “Dissipation in quantum mechanics. The harmonic oscilla- system coupled to a heat bath,” Phys. Rev. B 31, 154–164 ͑1985͒;F.Sols tor,” Phys. Rev. 119, 670–679 ͑1960͒. and F. Guinea, “Bulk and surface diffusion of heavy particles in metals: A 29P. Ullersma, “An exactly solvable model for Brownian motion,” Physica path-integral approach,” ibid. 36, 7775–7785 ͑1987͒; P. Hedegård and A. ͑Utrecht͒ 32, 27–55 ͑1966͒; 32,56–73͑1966͒; 32,74–89͑1966͒; 32, O. Caldeira, “Quantum dynamics of a particle in a fermionic environ- 90–96 ͑1966͒. ment,” Phys. Scr. 35, 609–622 ͑1987͒. 30R. Zwanzig, “Nonlinear generalized Langevin equations,” J. Stat. Phys. 9, 57R. P. Feynman, “Space–time approach to nonrelativistic quantum mechan- 215–220 ͑1973͒. ics,” Rev. Mod. Phys. 20, 367–387 ͑1948͒; R. P. Feynman and A. R. 31P. Hänggi, P. Talkner, and M. Borkovec, “Reaction-rate theory: Fifty years Hibbs, Quantum Mechanics and Path Integrals ͑McGraw–Hill, New York, after Kramers,” Rev. Mod. Phys. 62, 251–341 ͑1990͒. 1965͒. 32A. O. Caldeira and A. J. Leggett, “Quantum tunnelling in a dissipative 58R. P. Feynman and F. L. Vernon, “The theory of a general quantum system system,” Ann. Phys. ͑N.Y.͒ 149, 374–456 ͑1983͒; 153, 445 ͑1984͒. interacting with a linear dissipative system,” Ann. Phys. ͑N.Y.͒ 24,118– 33For the explicit details of this calculation, see Refs. 24 and 27. 173 ͑1963͒. 34U. Weiss, Quantum Dissipative Systems, 2nd ed. ͑World Scientific, Sin- 59P. Hanggi, H. Grabert, G.-L. Ingold, and U. Weiss, “Quantum theory of gapore, 1999͒. activated events in presence of long-time memory,” Phys. Rev. Lett. 55, 35U. Eckern, G. Schön, and V. Ambegaokar, “Quantum dynamics of a su- 761–764 ͑1985͒. perconducting tunnel junction,” Phys. Rev. B 30, 6419–6431 ͑1984͒. 60H. Grabert, P. Olschowski, and U. Weiss, “Quantum decay rates for dis- 36G. Schön and A. D. Zaikin, “Quantum coherent effects, phase transitions, sipative systems at finite temperatures,” Phys. Rev. B 36, 1931–1951 and the dissipative dynamics of ultra small tunnel junctions,” Phys. Rep. ͑1987͒. 198, 237–413 ͑1990͒. 61G.-L. Ingold, “Anwendung von Funktionalintegralen auf Transport- und 37G.-L. Ingold and Yu. V. Nazarov, “Charge tunneling rates in ultrasmall Relaxationsphänomene in dissipativen Quantensystemen,” Ph.D. thesis, junctions,” in Single Charge Tunneling, NATO ASI Series B, edited by H. Universität Stuttgart, 1988. Grabert and M. H. Devoret ͑Plenum, New York, 1992͒, Vol. 294, pp. 62H. Hofmann und G.-L. Ingold, “Dissipative transport across a parabolic 21–107. barrier,” Phys. Lett. B 264, 253–258 ͑1991͒. 38U. Zürcher and P. Talkner, “Quantum-mechanical harmonic chain attached 63J. Ankerhold, H. Grabert, and G.-L. Ingold, “Dissipative quantum systems to heat baths II. Nonequilibrium properties,” Phys. Rev. A 42, 3278–3290 with potential barrier. General theory and parabolic barrier,” Phys. Rev. E ͑1990͒. 51, 4267–4281 ͑1995͒; J. Ankerhold and H. Grabert, “Dissipative quan- 39D. Dvira, A. Nitzan, and P. Hänggi, “Thermal conductance through mo- tum systems with a potential barrier. II. Dynamics near the barrier top,” lecular wires,” J. Chem. Phys. 119, 6840–6855 ͑2003͒. ibid. 52, 4704–4723 ͑1995͒; Dissipative quantum systems with a potential 40S. Camalet, J. Lehmann, S. Kohler, and P. Hänggi, “Current noise in barrier. III. Steady state nonequilibrium flux and reaction rate,” ibid. 55, ac-driven nanoscale conductors,” Phys. Rev. Lett. 90, 210602 ͑2003͒;S. 1355–1374 ͑1997͒. Camalet, S. Kohler, and P. Hänggi, “Shot-noise control in ac-driven 64J. Ankerhold and H. Grabert, “Quantum tunneling and the semiclassical nanoscale conductors,” Phys. Rev. B 70, 155326 ͑2004͒. real time dynamics of the density matrix,” Europhys. Lett. 47, 285–291 41P. C. Hemmer, L. C. Maximon, and H. Wergeland, “Recurrence time of a ͑1999͒. dynamical system,” Phys. Rev. 111, 689–694 ͑1958͒. 65P. Hänggi and W. Hontscha, “Periodic orbit approach to the quantum- 42P. Mazur and E. Montroll, “Poincaré cycles, ergodicity, and irreversibility Kramers-rate,” Ber. Bunsenges. Phys. Chem. 95, 379–385 ͑1991͒; “Uni- in assemblies of coupled harmonic oscillators,” J. Math. Phys. 1, 70–84 fied approach to the quantum-Kramers reaction rate,” J. Chem. Phys. 88, ͑1960͒. 4094–4095 ͑1988͒. 43S. A. Adelman, “Quantum generalized Langevin equation approach to 66P. S. Riseborough, P. Hänggi, and E. Freidkin, “Quantum tunneling in gas/solid collisions,” Chem. Phys. Lett. 40, 495–499 ͑1976͒. dissipative media: Intermediate-coupling-strength results,” Phys. Rev. A 44A. Schmid, “On a quasiclassical Langevin equation,” J. Low Temp. Phys. 32, 489–499 ͑1985͒. 49, 609–626 ͑1982͒. 67H. Grabert, U. Weiss, and P. Hanggi, “Quantum tunneling in dissipative 45S. K. Banik, B. C. Bag, and D. S. Ray, “Generalized quantum Fokker– systems at finite temperatures,” Phys. Rev. Lett. 52, 2193–2196 ͑1984͒. Planck, diffusion and Smoluchowski equations with true probability dis- 68J. Ankerhold, P. Pechukas, and H. Grabert, “Strong friction limit in quan- tribution functions,” Phys. Rev. E 65, 051106 ͑2002͒; D. Banerjee, B. C. tum mechanics: The quantum Smoluchowski equation,” Phys. Rev. Lett. Bag, S. K. Banik, and D. S. Ray, “Solution of quantum Langevin equa- 87, 086802 ͑2001͒; J. Ankerhold, “Quantum decay rates for driven barrier tion: Approximations, theoretical and numerical aspects,” J. Chem. Phys. potentials in the strong friction limit,” Phys. Rev. E 64, 060102 ͑2001͒;L. 120, 8960–8972 ͑2004͒. Machura, M. Kostur, P. Hänggi, P. Talkner, and J. Łuczka, “Consistent 46U. Eckern, W. Lehr, A. Menzel-Dorwarth, F. Pelzer, and A. Schmid, “The description of quantum Brownian motors operating at strong friction,” quasiclassical Langevin equation and its application to the decay of a ibid. 70, 031107 ͑2004͒. metastable state and to quantum fluctuations,” J. Stat. Phys. 59, 885–934 69G. W. Ford and R. F. O’Connell, “Inconsistency of the rotating wave ͑1990͒. approximation with the Ehrenfest theorem,” Phys. Lett. A 215, 245–246 47P. Hänggi and H. Thomas, “Stochastic processes: Time evolution, symme- ͑1996͒. tries and linear response,” Phys. Rep. 88, 207–319 ͑1982͒. 70H. Grabert, “Non-linear relaxation and fluctuations of damped quantum 48J. Shao and P. Hänggi, “Decoherent dynamics of a two-level system systems,” Z. Phys. B: Condens. Matter 42, 161–172 ͑1982͒. coupled to a sea of spins,” Phys. Rev. Lett. 81, 5710–5713 ͑1998͒. 71G. W. Ford and R. F. O’Connell, “There is no quantum regression theo- 49P. Schramm, R. Jung, and H. Grabert, “A closer look at the quantum rem,” Phys. Rev. Lett. 77, 798–801 ͑1996͒. Langevin equation: Fokker–Planck equation and quasiprobabilities,” Phys. 72M. D. Kostin, “Friction and dissipative phenomena in quantum mechan- Lett. 107A, 385–389 ͑1985͒. ics,” J. Stat. Phys. 12,145͑1975͒. 50H. Grabert, P. Talkner, and P. Hänggi, “Microdynamics and time-evolution 73J. Messer, “Friction in quantum mechanics,” Acta Phys. Austriaca 50, of macroscopic non-Markovian systems,” Z. Phys. B 26, 389–395 ͑1977͒; 75–91 ͑1979͒, and references therein. H. Grabert, P. Talkner, P. Hänggi, and H. Thomas, “Microdynamics and 74H. Dekker, “Classical and quantum mechanics of the damped harmonic time-evolution of macroscopic non-Markovian systems II,” ibid. 29, 273– oscillator,” Phys. Rep. 80,1–112͑1981͒. 280 ͑1978͒. 75R. Fürth, “Über einige Beziehungen zwischen klassischer Statistik und 51K. M. F. Romero, P. Talkner, and P. Hänggi, “Is the dynamics of open Quantenmechanik,” Z. Phys. 81, 143–162 ͑1933͒. quantum systems always linear?” Phys. Rev. A 69, 052109 ͑2004͒. 76I. Fényes, “Eine wahrscheinlichkeitstheoretische Begründung und Inter- 52C. Zerbe and P. Hänggi, “Brownian parametric quantum oscillator with pretation der Quantenmechanik,” Z. Phys. 132, 81–106 ͑1952͒. dissipation,” Phys. Rev. E 52, 1533–1543 ͑1995͒. 77W. Weizel, “Ableitung der Quantentheorie aus einem klassischen, kausal 53S. Kohler, T. Dittrich, and P. Hänggi, “Floquet-Markovian description of determinierten Modell,” Z. Phys. 134, 264–285 ͑1953͒; “Ableitung der parametrically driven, dissipative harmonic oscillator,” Phys. Rev. E 55, Quantentheorie aus einem klassischen Modell. II,” ibid. 135, 270–273 300–313 ͑1997͒. ͑1953͒. 54J. Łuczka, “Spin in contact with thermostat: Exact reduced dynamics,” 78L. F. Favella, “Brownian and quantum mechanics,” Ann. Inst. Physica A 167, 919–934 ͑1990͒. Henri Poincare, Sect. A 7, 77–94 ͑1967͒. 55N. G. van Kampen, “A soluble model for quantum mechanical dissipa- 79G. C. Ghirardi, C. Omero, and A. Rimini, “Stochastic interpretation of

Downloaded 21 Jun 2005 to 137.250.81.48. Redistribution subject to AIP license or copyright, see http://chaos.aip.org/chaos/copyright.jsp 026105-12 P. Hanggi and G.-L. Ingold Chaos 15, 026105 ͑2005͒

quantum mechanics—Critical review,” Riv. Nuovo Cimento 1, 1–34 33–39 ͑2002͒; P. Reimann and P. Hänggi, “Quantum features of Brownian ͑1978͒. motors and stochastic resonance,” Chaos 8, 629–642 ͑1998͒; P. Reimann, 80 M. S. Wang and W. K. Liang, “Comment on repeated measurements in M. Grifoni, and P. Hänggi, “Quantum ratchets,” Phys. Rev. Lett. 79, ͑ ͒ stochastic mechanics,” Phys. Rev. D 48, 1875–1877 1993 . 10–13 ͑1997͒. 81R. D. Astumian and P. Hänggi, “Brownian motors,” Phys. Today 55 ͑11͒,

Downloaded 21 Jun 2005 to 137.250.81.48. Redistribution subject to AIP license or copyright, see http://chaos.aip.org/chaos/copyright.jsp