11
New J. Phys. 18 (2016) 023035 doi:10.1088/1367-2630/18/2/023035 PAPER Efcient simulation of non-Markovian system-environment interaction Robert Rosenbach 1 , Javier Cerrillo 2,3,4 , Susana F Huelga 1 , Jianshu Cao 2 and Martin B Plenio 1,4 1 Institute of Theoretical Physics, Universität Ulm, Albert-Einstein-Allee 11, D-89069 Ulm, Germany 2 Massachusetts Institute of Technology, 77 Massachusetts Avenue, Cambridge, Massachusetts 02139, USA 3 Institute of Theoretical Physics, Technische Universität Berlin, Hardenbergstr. 36, D-10623 Berlin, Germany 4 Authors to whom any correspondence should be addressed. E-mail: [email protected] and [email protected] Keywords: open quantum systems, quantum dissipation, non-Markovian dynamics, NakajimaZwanzig equation, density matrix renormalization group Abstract In this work, we combine an established method for open quantum systemsthe time evolving density matrix using orthogonal polynomials algorithmwith the transfer tensors formalism, a new tool for the analysis, compression and propagation of non-Markovian processes. A compact propagator is generated out of sample trajectories covering the correlation time of the bath. This enables the investigation of previously inaccessible long-time dynamics with linear effort, such as those ensuing from low temperature regimes with arbitrary, possibly highly structured, spectral densities. We briey introduce both methods, followed by a benchmark to prove viability and combination synergies. Subsequently we illustrate the capabilities of this approach at the hand of specic examples and conclude our analysis by highlighting possible further applications of our method. 1. Introduction Ranging from condensed matter physics or quantum technologies to biological chemistry, the experimental ability to accurately probe and analyze quantum systems in strong contact with highly structured environmental degrees of freedom for extended periods of time has become a reality [16]. Under certain conditions it is possible to model the observations by using approximate methods such as perturbative approaches [7, 8], frequently supplemented with Markovian assumptions [911]. Beyond their regime of validity, the task of exactly treating the dynamics of open quantum systems faces the challenge of an unfavorable scaling in the required resources. Nevertheless, many tools have been developed that address a wide variety non-Markovian scenarios for short time simulations or for specic conditions and approximations. Exact procedures such as projection operator techniques serve to derive formally exact master equations that can involve a memory kernel as in the NakajimaZwanzig formalism [12] or have a generator that is local in time, like the so-called time-convolutionless master equations [13]. The path integral formulation [14] provides an alternative perspective especially suitable for harmonic baths thanks to the FeynmannVernon inuence functional [15]. Practical implementation of these formal treatments requires perturbative expansions in terms of some specic parameter, be it weak damping, high temperature, short memory time, or short simulation time [1619]. An exhaustive list is not within the scope of this work, but some additional instances include the non- Markovian quantum state diffusion approach [20], the hierarchy of equations of motion (HEOM) [17, 21], iterative path integral resummations [22], multilayer multicongurational time-dependent Hartree methods [23], explicit computation of the NakajimaZwanzig memory kernel [24, 25] and the time-convolutionless kernel [26] or mixed quantumclassical methods [2729]. Alternatively, the harmonic bath assumption renders possible the use of stochastic Gaussian sampling of the bath operator or the inuence functional [3033]. Hybrid stochastic-deterministic methods have appeared recently as well [34]. Another option is to simulate the density OPEN ACCESS RECEIVED 17 October 2015 REVISED 21 December 2015 ACCEPTED FOR PUBLICATION 5 January 2016 PUBLISHED 8 February 2016 Original content from this work may be used under the terms of the Creative Commons Attribution 3.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI. © 2016 IOP Publishing Ltd and Deutsche Physikalische Gesellschaft

PAPER Ef cientsimulationofnon-Markoviansystem …web.mit.edu/jianshucaogroup/pdfdir/2016/Javi2016NJP.pdfD = 0.6 .Onecanobservethat,forthefastestbathJ 2,oscillationsaresustainedforalongertime,whilethis

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: PAPER Ef cientsimulationofnon-Markoviansystem …web.mit.edu/jianshucaogroup/pdfdir/2016/Javi2016NJP.pdfD = 0.6 .Onecanobservethat,forthefastestbathJ 2,oscillationsaresustainedforalongertime,whilethis

New J. Phys. 18 (2016) 023035 doi:10.1088/1367-2630/18/2/023035

PAPER

Efficient simulation of non-Markovian system-environmentinteraction

Robert Rosenbach1, Javier Cerrillo2,3,4, Susana FHuelga1, JianshuCao2 andMartin BPlenio1,4

1 Institute of Theoretical Physics, Universität Ulm, Albert-Einstein-Allee 11, D-89069Ulm,Germany2 Massachusetts Institute of Technology, 77Massachusetts Avenue, Cambridge,Massachusetts 02139, USA3 Institute of Theoretical Physics, TechnischeUniversität Berlin, Hardenbergstr. 36, D-10623Berlin, Germany4 Authors towhomany correspondence should be addressed.

E-mail: [email protected] [email protected]

Keywords: open quantum systems, quantumdissipation, non-Markovian dynamics, Nakajima–Zwanzig equation, densitymatrixrenormalization group

AbstractIn this work, we combine an establishedmethod for open quantum systems—the time evolvingdensitymatrix using orthogonal polynomials algorithm—with the transfer tensors formalism, a newtool for the analysis, compression and propagation of non-Markovian processes. A compactpropagator is generated out of sample trajectories covering the correlation time of the bath. Thisenables the investigation of previously inaccessible long-time dynamics with linear effort, such asthose ensuing from low temperature regimeswith arbitrary, possibly highly structured, spectraldensities.We briefly introduce bothmethods, followed by a benchmark to prove viability andcombination synergies. Subsequently we illustrate the capabilities of this approach at the hand ofspecific examples and conclude our analysis by highlighting possible further applications of ourmethod.

1. Introduction

Ranging from condensedmatter physics or quantum technologies to biological chemistry, the experimentalability to accurately probe and analyze quantum systems in strong contact with highly structured environmentaldegrees of freedom for extended periods of time has become a reality [1–6]. Under certain conditions it ispossible tomodel the observations by using approximatemethods such as perturbative approaches [7, 8],frequently supplementedwithMarkovian assumptions [9–11]. Beyond their regime of validity, the task ofexactly treating the dynamics of open quantum systems faces the challenge of an unfavorable scaling in therequired resources. Nevertheless,many tools have been developed that address awide variety non-Markovianscenarios for short time simulations or for specific conditions and approximations.

Exact procedures such as projection operator techniques serve to derive formally exactmaster equations thatcan involve amemory kernel as in theNakajima–Zwanzig formalism [12] or have a generator that is local intime, like the so-called time-convolutionlessmaster equations [13]. The path integral formulation [14]providesan alternative perspective especially suitable for harmonic baths thanks to the Feynmann–Vernon influencefunctional [15]. Practical implementation of these formal treatments requires perturbative expansions in termsof some specific parameter, be it weak damping, high temperature, shortmemory time, or short simulation time[16–19]. An exhaustive list is not within the scope of this work, but some additional instances include the non-Markovian quantum state diffusion approach [20], the hierarchy of equations ofmotion (HEOM) [17, 21],iterative path integral resummations [22], multilayermulticonfigurational time-dependentHartreemethods[23], explicit computation of theNakajima–Zwanzigmemory kernel [24, 25] and the time-convolutionlesskernel [26] ormixed quantum–classicalmethods [27–29]. Alternatively, the harmonic bath assumption renderspossible the use of stochastic Gaussian sampling of the bath operator or the influence functional [30–33]. Hybridstochastic-deterministicmethods have appeared recently as well [34]. Another option is to simulate the density

OPEN ACCESS

RECEIVED

17October 2015

REVISED

21December 2015

ACCEPTED FOR PUBLICATION

5 January 2016

PUBLISHED

8 February 2016

Original content from thisworkmay be used underthe terms of the CreativeCommonsAttribution 3.0licence.

Any further distribution ofthis workmustmaintainattribution to theauthor(s) and the title ofthework, journal citationandDOI.

© 2016 IOPPublishing Ltd andDeutsche PhysikalischeGesellschaft

Page 2: PAPER Ef cientsimulationofnon-Markoviansystem …web.mit.edu/jianshucaogroup/pdfdir/2016/Javi2016NJP.pdfD = 0.6 .Onecanobservethat,forthefastestbathJ 2,oscillationsaresustainedforalongertime,whilethis

matrix of both the system and the complete environment by employing an efficient description of the bath orgradually introducing select degrees of freedom [35–37]. To this class ofmethods belongs the ‘time evolvingdensitymatrix using orthogonal polynomials algorithm’ (TEDOPA) [38, 39]. Thismethod uses a stablenumerical transformation tomap the environment into a chain of harmonic oscillators, which can then besimulated togetherwith the systemusing efficient quantummany body techniques [40]. Although amorethorough discussion follows below, as compared to other simulationmethods TEDOPA is especially suitable forsimulation of quadratic harmonic bathswith arbitrarily shaped spectral-densities in the low-temperatureregime and is not restricted to small coupling orOhmic baths.

For any exact simulationmethod it is generally the case that the size of the propagator or that of thestochastic sample scales unfavorably with the time length of the simulation or the corresponding perturbativeexpansion order. Then the question arises whether there are regimeswhere this scaling can bemitigated in someform, i.e.if an effective propagator of a reduced size can be extractedwith the intention of facilitating long-timesimulations. In the present workwe address this question combining TEDOPAwith a tool quantifying the bath’sback-action on the system as in theNakajima–Zwanzig formalism. This tool is known as the transfer tensormethod (TTM) [41], which has been shown to provide considerable acceleration in the context of non-Markovian open quantum system simulations aswell as in large classical systems [42]. This is achieved byblackbox learning from sample exact trajectories for some short initial period and subsequent generation of acompactmultiplicative propagator for the systemdegrees of freedomalone. This propagator is the set of discreteelements of the integration of theNakajima–Zwanzig equationwhich, similarly to thememory kernel, decay atthe rate of the bath correlation function. This justifies the definition of amemory cutoff, corresponding to themaximum time forwhich non-Markovian effects are to be considered. Themethod does not require input of amicroscopic description of the problem and just involves straightforward analysis of the evolution of the systemdensitymatrix. In this sense, TTM is directly applicable to any state propagation simulator. For a learning periodlonger than the environment correlation time, the propagator accurately reproduces the long time systemdynamics with linear effort. Another possibility to exploit the decay of thememory effects in dissipative systemsis to explicitly calculate theNakajima–Zwanzigmemory kernel [24, 25]. Although this is in general a demandingtask, it is possible for specific systems and has been implementedwith the help of quantumMonte-Carlomethods [25, 43] ormultilayermulticonfigurational time-dependentHartreemethods [44, 45].With TTManexplicit computation of thememory kernel is avoided and a discretized propagator is directly obtained. Thispropagator grows linearly with the correlation time of the bath, improving on the size of some deterministicsimulationmethods of linear propagation effort [17, 18]. It is a general andflexible approach that does notdepend on the formof the environment or the interaction, while TEDOPA is not restricted toweak system-bathcoupling, high temperatures or specific spectral densities. Therefore, they constitute ideal partners and a study oftheir combined performance represents a natural research question.

We start the discussion in section 2 by providing a general analysis of both TEDOPA andTTM, therebyspecifying the regime inwhich their combination is expected to bemost productive. In addition, tools for theerror assessment are provided. In section 3we provide a benchmark between the proposed combination and theexact result and confirmperfect agreement. Finally, in section 4 relevant applications of the TEDOPA-TTMcombination are proposedwhich includeOhmic and non-Ohmic spectral densities, low temperaturesimulations and computations of absorption spectra.

2. Themethod

2.1. A numerically exact open quantum system simulatorTEDOPA [38, 39] is a numerically exact and certifiable simulationmethod for open quantum systemdynamics[46]. It applies to general systems under linear interactionwith an environmentmodeled by a set of independentharmonic oscillators such that the totalHamiltonianH can be split into the system, the environment and theinteraction between the two as

H H H H , 1sys env int ( )= + +

H x g x a ad , 2x

x xenv0

max

( ) ( )†ò=

H x h x a a Ad . 3x

x xint0

max

( )( ) ( )†ò= +

Here ax† and ax denote the bosonic creation and annihilation operators corresponding to the environmental

mode x. The coefficients g x( ) can be identified as the environmental dispersion relation. The interaction termHint assigns eachmode a coupling strengthh x( ) between its displacement a ax x

† + and a general systemoperatorA.

2

New J. Phys. 18 (2016) 023035 RRosenbach et al

Page 3: PAPER Ef cientsimulationofnon-Markoviansystem …web.mit.edu/jianshucaogroup/pdfdir/2016/Javi2016NJP.pdfD = 0.6 .Onecanobservethat,forthefastestbathJ 2,oscillationsaresustainedforalongertime,whilethis

Together with the temperature, the functions g x( ) and h x( ) characterize the harmonic environmentuniquely and define the spectral density J ( )w as

J h ggd

d. 42 1

1

( ) [ ( )] ( ) ( )w p ww

w= -

-

Here g g x g g x x1 1[ ( )] [ ( )]= =- - and g(x) ismonotonically growing. The quantitygd

d

1( )dwww

-

can be interpretedas the number of quantizedmodeswith frequencies betweenω andω+δω (for δω→ 0).We consider spectraldensities subject to a hard cut-off at frequency ;hcw this in turn defines the cut-off xmax in equations (2) and (3)as x gmax

1hc( )w= - .

TEDOPAuses a two-stage sequence to enable full treatment of the system and environment degrees offreedom. In a first step, an analytic transformation based on orthogonal polynomials converts the star-shapedsystem-environment structure into a one-dimensional geometry, where the system couples only to the first siteof a semi-infinite chain of harmonic oscillators that contains only nearest-neighbour interactions. This isaccomplished by use of a unitary transformationU, which defines newharmonic oscillators with creation andannihilation operators bn

†, bn given by

U x h x p x , 5n n( ) ( ) ( ) ( )=

b x U x ad . 6n

x

n x0

max

( ) ( )† †ò=

Here p xn ( ) are orthogonal polynomials definedwith respect to themeasure x h x xd d2( ) ( )m = and the three-term recurrence relation

p x x p x p x , 7k k k k k1 1( ) ( ) ( ) ( ) ( )a b= - -+ -

with p x 01( ) º- and k a positive integer or zero. This transformation can be performed analytically for specificspectral densities [39, 47]. For arbitrarily shaped spectral densities a numerically stable approach has beendeveloped [38, 48].Mappings using orthogonal polynomials have been shown to be exact for quadraticHamiltonians [49].

The recurrence relation (7) results in the one-dimensional configurationwith theHamiltonian

H H A b b b b

b b b b . 8

nn n n

nn n n n n

sys 0 0 00

01 1

˜ ( )

( ) ( )

† †

† †

å

å

h w

h

= + + +

+ +

=

¥

=

¥

+ +

For a linear dispersion relation g x g x( ) = ¢ , gn nw a= ¢ and gn n 1h b= ¢ + . Due to the formof the emerginglinear geometry, thisfirst stage of TEDOPA is coined ‘chainmapping’. One dimensional quantummany bodysystems can be efficiently simulatedwith thewell established time dependent densitymatrix renormalizationgroup (t-DMRG) algorithm [50–52]. Its application to evaluate the dynamics of the system and the transformedenvironment constitutes the second stage of TEDOPA. The long ranged correlations appearing in the originalstar-shaped geometry advise against the application of t-DMRG in that picture: it is the nearest neighborstructure thatmakes the numerical t-DMRGapproach particularly efficient. Recent works consider thepossibility to use generalizedmatrix product state formulations for treatment of star-shaped baths as well [53].

For a complete account of TEDOPA’s innerworkings refer to [38, 39, 54]. Suffice it to say that threemainaspects distinguish it fromother open quantum-systemmethods. First, it does not restrict the ratiol betweeninner-system couplings and system-environment couplings, unlike numerous othermethods (which assumeeither 1l or 1l ). Second, the spectral density characterizing the system-environment interactionmayassume any arbitrary shape, including awide variety of sharp features thatmay be related to long-livedvibrationalmodes [38, 55, 56]. Such spectral densities are often encountered in spectral densities reconstructedfrom experimental data, e.g. in biological settings [57]. And third, while applicable to all temperatures, due to itsscaling properties TEDOPA is inherently well-suited for simulations in the low-temperature domain.

Naturally, however, exact numericalmethods tend to be costly and an effort to save on the associatedcomputational demands is desirable.Where the transfer tensormethod can be applied, these costs can bereduced and challenging long time simulations become accessible.

2.2. Non-Markovian dynamicalmaps: transfer tensormethodThe transfer tensormethod [41] reduces the numerical effort of TEDOPA simulations for a large class of non-Markovian environments. Its key idea is to relate the initial stages of the system’s evolution to later times byefficiently determining the dynamical correlations built up between system and environment. This is achievedby the reconstruction of dynamicalmaps for short initial evolution times and their subsequent transformationinto so-called transfer tensors.

3

New J. Phys. 18 (2016) 023035 RRosenbach et al

Page 4: PAPER Ef cientsimulationofnon-Markoviansystem …web.mit.edu/jianshucaogroup/pdfdir/2016/Javi2016NJP.pdfD = 0.6 .Onecanobservethat,forthefastestbathJ 2,oscillationsaresustainedforalongertime,whilethis

Adynamicalmap is defined for an initial conditionwhere the state of the system and the state of theenvironment are separable, and it fully determines the reduced state of the system t( )r when applied to an initialcondition 0( )r

t t, 0 0 . 9( ) ( ) ( ) ( )r r=

For a time independentHamiltonianH such as equation (1), equation (9) is related to the solution of the time-translationally invariantNakajima–Zwanzig equation

t t t t ti d , 10s

t

0˙ ( ) ( ) ( ) ( ) òr r r= - + ¢ - ¢ ¢

where H ,s sys[ ] r r= is the Liouvillian of the system alone and t t( ) - ¢ , is thememory kernel arising fromthe system-bath interactionHint.

The input for TTM is a set of dynamicalmaps t , 0k k( ) = , where a discretisation time step td (t k tk dº ) isused. These dynamicalmaps are obtained easily from the evolution of all distinct initial basis states of thesystem’s densitymatrix. The transfer tensors are then iteratively defined by

T T . 11n nm

n

n m m1

1

( ) å= -=

-

-

According to this definition, the transfer tensorTk then quantifies the correlation in the dynamicalmap k withthe non-Markovian effects built up during the previousk time steps.

Further, the discretization of thememory kernel is directly related to these tensors by the time step td

T t , 12k k2 ( ) d=

where tk k( ) = is the discretizedmemory kernel at time tk. The systemHamiltonian is itself accessible fromthefirst transfer tensor by

T ti . 13s1 ( ) ( ) d= -

These tensors can be used to propagate the system to arbitrary later times as long as they cover the relevantpart of thememory kernel. Assuming a finite coherence time of the bath tbath, the transfer tensors will decaysufficiently fast that a cutoffK can be defined such thatTk=0 for k K> . Then, tn( )r for n K> can beexpressed simply as

t T t . 14nk

K

k n k1

( ) ( ) ( )år r==

-

TTM is applicable in a variety of interesting cases, scaling favorably in system size and length of theenvironment’s correlation time. Due to its nature it also does not rely on assumptions about the system’sparameters or environmental couplings. Thus it is usable as a powerful black box tool which, given initialtrajectories, subsequently delivers the evolution trajectories for later times. For amore complete account of thistool refer to [41].

2.3. SynergiesThe combination of bothmethods facilitates the simulation of open quantum systems in regimes that werepreviously inaccessible. Exceptionally relevant is the ability to perform long-time simulations of low-temperature, highly structured harmonic environments atmerely a fraction of the computational cost whichwould be necessary if only TEDOPAwere applied. Evidence for this is provided by simple examination of someof the features of each of themethods. On the one hand, TEDOPA is based on amatrix product operator (MPO)description of the complete systemplus environment densitymatrix. Settings leading to lowoccupationnumbers of environmental oscillators—such as low temperatures—are especially suitable as they reduce theMPO’s number and size. In additionTEDOPA is not limited to a specific analytical formof spectral density. Onthe other hand, recurrence effects originating at the chain’s boundary limit the time forwhich accuratesimulation is possible. Because TTMonly requires sample system trajectories for as long as bath correlations arepresent, the required chain length is then not anymore determined by the target simulation time. Thecombination of TTMandTEDOPA is thereforemost useful in highly non-Markovian regimeswhere bathcorrelation times are comparable to themaximum simulation time that TEDOPA can reach before recurrencesappear.

2.4. Parameters and accuracyHerewe analyze relevant parametric cutoffs for the control of the accuracy of numerical simulations thatcombine TEDOPA andTTM.

4

New J. Phys. 18 (2016) 023035 RRosenbach et al

Page 5: PAPER Ef cientsimulationofnon-Markoviansystem …web.mit.edu/jianshucaogroup/pdfdir/2016/Javi2016NJP.pdfD = 0.6 .Onecanobservethat,forthefastestbathJ 2,oscillationsaresustainedforalongertime,whilethis

In the case of TEDOPAone can distinguish between parameters related to the chainmapping and to thet-DMRGpropagation. The semi-infinite chain of oscillators generated by themapping necessarily requires thetruncation of both the chain lengthN and theHilbert space dimensiond for each oscillator at a reasonablevalue. Those are native TEDOPA error sources and they have a direct effect on the recurrence time of thesimulation and themaximum temperature thatmay be simulated, respectively. It should be noted that the errorincurred by these two approximations can be upper-bounded rigorously by analytical expressions [46]. Theeffect of other relevant parameters, namely theMPO’smatrix size( )c c´ and the time step td , are alreadywell-known from the time-evolving block decimation (TEBD) algorithm [51]. To reach the accuracy required byTTM, care needs to be taken in adjusting these parameters to bound the total error of TEDOPA sufficiently.Some indications on how to accomplish this are provided in the present section.

Themaximum time tmax before unphysical back-actions of the environment due to reflections at the end ofthe chain appear is related to the chain lengthN. Usually all chain coefficients are of the same order ofmagnitudeand hence the simulation time tmax scales roughly linearly withN. This reveals one of the benefits of theapplication of TTMonTEDOPA: the chain length can be truncated according to the length of the bathcorrelation time, allocating the simulation resources properly and shortening simulation times considerably inmany cases of relevance (see figure 7). The exact relationship betweenN and tmax may be derived analyticallythrough the use of Lieb–Robinson bounds [46] or numerically by trial-and-error: by setting the chain in aninitial state 10 0∣ ¼ ñand following the evolution of the number operator n on thefirst site, O n = Ä Ä ¼Ä ,until a recurrence occurs.

The second native TEDOPAparameter is the local dimensiond of the single oscillators constituting theenvironment. For a given temperature, the occupation of the single oscillators can be determined exactly, givinga rough scale of the necessary truncation level. Some error will be introduced necessarily but this can be upper-bounded analytically as explained in [46]. On the other hand, numerical benchmark calculationswith increasinglocal dimensionswill generally yield sufficiently accurate results.

A further subtlety in the chainmapping consists in the determination of an adequate hard cutofffrequency hcw of the spectral density. For instance, the slow approach to zero for large frequencies of theDrude–Lorentz bath imposes a careful convergence check of the resulting physical behavior. For further discussions ofthese effects refer to [54].

While some error sources (like the cutoff in the chain length) introduce, if treated correctly, virtually no errorat all, thematrix sizec necessarily does so due to the nature of theMPO.However, as already studied in thecontext of the TEBD algorithm, this error can bemonitored during the time evolution [50]. This results in aquantity very similar to the discardedweight known fromDMRG

w e1 . 15i

idiscarded2 ( )å= -

This quantifies the deviation from the targeted state using the discarded eigenvalues ei. This error propagates in anon-trivial fashion and it is advisable to perform convergence checks in the dynamics under variation of the sizeof c. An additional source of error is derivated from the Suzuki–Trotter decomposition used in the TEBDpartof TEDOPA.

It should be noted that themagnitude of the singular values kept duringMPO-procedures should not fallbelow some threshold e0. The transfer tensors determined byTTMdo decay rapidly, falling to comparatively lowmagnitudes, and singular values corrupted by numerical noise deteriorates the interpretation of results as well asthe propagation procedure.

Finally, for TTM the important quantity to keep track of in simulations is the normof thememory kernel.This corresponds to the normof the transfer tensors divided by the squared time step td . Thismagnitude shouldexhibit a sufficiently fast decay so that the remaining tail can be neglected. Additionally, the time step td must besuch that it provides a good resolution of the features of thememory kernel.

3. Benchmark

In this sectionwe verify the combination of TEDOPAwith TTMby comparing the obtained transfer tensorswith those originating from another numerically exact simulationmethod for non-Markovian systems underthe same conditions. The chosen benchmark regime is theOhmicDrude–Lorentz bath and the additionalsimulationmethod is the hierarchy of equations ofmotion (HEOM) [17].

We consider the spin-bosonmodel (SBM) and define the (monomeric) systemHamiltonian

H1

2

1

2. 16z xsys ( )s s= + D

5

New J. Phys. 18 (2016) 023035 RRosenbach et al

Page 6: PAPER Ef cientsimulationofnon-Markoviansystem …web.mit.edu/jianshucaogroup/pdfdir/2016/Javi2016NJP.pdfD = 0.6 .Onecanobservethat,forthefastestbathJ 2,oscillationsaresustainedforalongertime,whilethis

Herewe set 1 = , a conventionwewill stick to fromnowon, and express all frequencies in units of ò.Weemploy the standard SBMnotationwhere corresponds to the energy bias between ground and excited state,Dis the tunnelingmatrix element, and is i x y z, ,( )= are the Paulimatrices corresponding to the i’th spatialdirection. The system interaction operatorA is defined as the excited state projector e e∣ ∣ñá .We choose theparameter 0.6D = , and anOhmic spectral density of theDrude–Lorentz form

J , 172 2

( )( )

( )w lgw

w g=

+

with parameters l = and 10g = respectively identifying a scaling of the interaction strength and a softcutoff frequency. Thus the bath reorganization energy J dr ( )òl w w= is 5.89r l = . A large hard cutoff

320hc w = has been employed tomeet the aforementioned convergence requirements of TEDOPAunderDrude–Lorentz baths.

At an inverse temperature of 0.5b = , TEDOPA exhibits favorable cutoffs andHEOMsimulations areaccurate, which enables benchmarking. The resulting elements of thememory kernel obtained byTTMappliedto TEDOPA’s initial trajectories are comparedwith those retrieved fromHEOM [17] simulations of the samesystem.Wehave confirmed agreement in a broad range of additional regimes accessible to both TEDOPA andHEOM. Further the system’sHamiltonian has successfully been recovered from the first transfer tensor. Thisalso corroborates the ability of TTM to extract the same dynamical tensors irrespective of the simulationmethodused for the generation of the trajectories.Wewill now turn to applications on hitherto inaccessible regimes toillustrate the strengths of the TEDOPA-TTMcombination.

4. Applications

4.1. Non-Ohmic spectraBy construction, TEDOPA is inherently suited to treat spectral densities of arbitrary shape.When consideringnon-Ohmic spectral densities,Markovian approaches are well-known to anomalously suppress the effect ofpure dephasing contributions [58, 59]. In this sectionwe present an analysis of the dynamical effects of threedifferent non-Ohmic spectral densities, namely

J e , 181 13 c( ) ( )w l w= w w-

J e , 192 25 c( ) ( )w l w= w w-

J e . 203 3c( ) ( )w l w= w w-

To facilitate comparison, all of them exhibit the same exponential decaywith 0.3c w = and are subject to a hardcutoff at 10hc w = . Also the parameters 1.81 l = , 1.02 l = and 0.63 l = have been chosen insuch away that they all share the same reorganization energy 0.3r l = . Thus the average interaction strengthbetween system and environment is the same and the functional formof the spectral density is the factorresponsible for disparate dynamics. The resulting dissimilar amplitudes and decay rates of the oscillations due tothe different spectral densities are illustrated infigure 1. The tunneling strength is, as in previous section,

Figure 1.Effect on the population dynamics of the spin of three different non-Ohmic spectral densities J1,2,3 (seemain text forfunctional forms) at inverse temperature b = and verification of the predictability of trajectories by TTM.Black dashed lines areTEDOPA simulation results, colored lines are TTM’s predictions; TTM learning times are denoted by orange lines (roughly untilt 10 1= - ).

6

New J. Phys. 18 (2016) 023035 RRosenbach et al

Page 7: PAPER Ef cientsimulationofnon-Markoviansystem …web.mit.edu/jianshucaogroup/pdfdir/2016/Javi2016NJP.pdfD = 0.6 .Onecanobservethat,forthefastestbathJ 2,oscillationsaresustainedforalongertime,whilethis

0.6D = . One can observe that, for the fastest bath J2, oscillations are sustained for a longer time, while thisability decreases for spectral densities centered in lower frequencies J1 and almost disappears for the very slowbath represented by J3. Some brief initial time is sufficient to generate the transfer tensors and predict the furtherevolution, whereupon high-accuracy TEDOPA simulations are used to verify these predictions.

The suitability of the TEDOPA-TTMcombination is supported by the fact that these simulations require onthe order of just 100 tensors to converge to the exact results that have been obtained by full TEDOPAsimulations, as shown infigure 2. This translates into about an order ofmagnitude faster results for TTM-TEDOPA combination than for TEDOPA alone. Further improvements in simulation speed are possible and arediscussed in section 4.4.

4.2. Low temperaturesTo further illustrate the power of our approach, we present results for a broad range of low to very lowtemperatures, up to 10b = . For the super-Ohmic spectral density J2 ( )w we show infigure 3 that it is possibleto simulate the dynamics of amonomeric system at various inverse temperatures and the same systemparameters as in the previous example. For the case of spectral density J1 ( )w we employ TTM to propagate thesystemuntil the steady state is reached (figure 4) and plot the steady-state occupation of the excited state for

Figure 2.Time evolution of the excited state population subject to an environment with super-Ohmic spectral density J2 ( )w atb = . Black crosses denote the TEDOPA-only evolution, while the TTMpredictions (colored lines) show a gradual convergence

upon increased learning time. The full 100 learning steps correspond to time t 10 1= - .

Figure 3.Decay of amonomeric system’s population in the SBM, subject to an environment with super-Ohmic spectral density J2 ( )wat different inverse temperaturesb . For better clarity only the first few oscillations are shown. Solid lines correspond to TEDOPA-only simulationswith verified accuracy. The orange initial part of each curve corresponds to the learning period. The decay of thetensor norm for the learning period is shown in the inset.

7

New J. Phys. 18 (2016) 023035 RRosenbach et al

Page 8: PAPER Ef cientsimulationofnon-Markoviansystem …web.mit.edu/jianshucaogroup/pdfdir/2016/Javi2016NJP.pdfD = 0.6 .Onecanobservethat,forthefastestbathJ 2,oscillationsaresustainedforalongertime,whilethis

various inverse temperaturesb infigure 5. The insets infigures 3 and 4 show thememory kernel’s decay overseveral orders ofmagnitude for the corresponding spectral densities. It is this decaywhich certifies the possibilityto use the tensors for long-time propagation of the dynamics.

4.3. Absorption spectrumThe combination of TEDOPA andTTM is especially indicated for applications where accurate simulation oflong time dynamics is crucial. The determination of absorption spectra belongs to this class of problems andweanalyze here themore complex case of a dimeric system consisting of two coupledmonomers.

The coupled dimeric system in the single excitationmanifold consists of two excited states e1∣ ñ, e2∣ ñand acommon ground state g∣ ñ, and is described by theHamiltonian

H e e e e J e e e e , 21sys 1 1 1 2 2 2 1 2 2 1∣ ∣ ∣ ∣ (∣ ∣ ∣ ∣) ( ) = ñá + ñá + ñá + ñá

where parameters 1 º , 22 = , and exchange interaction strength J 0.6= are chosen. Each of the twosystems is coupled to a bath.Without loss of generality we assume both environments are described by the samespectral density J1 ( )w and at temperature b = .

The absorption spectrum is calculated as the Fourier transformof the two point correlation function of thedipole operator e g e g h.c.1 1 2 2ˆ ∣ ∣ ∣ ∣m m m= ñá + ñá +

C t t, 0 0 22( ) ˆ ( ) ˆ ( ) ( )m m= á ñm m-

Figure 4.Thermalization of the system’s population subject to an environmentwith spectral density J1 ( )w at different temperatures.The combination TEDOPA-TTMhas been used and verifiedwith TEDOPA-only simulations. The decay of the tensor norm for thelearning period is shown in the inset.

Figure 5.Excited state population in the steady state for amonomeric system subject to spectral density J1 ( )w , plotted over the inversetemperatureb . The steady state is determined by TTMevolution of the initial TEDOPA trajectories. The line is a guide to the eye.

8

New J. Phys. 18 (2016) 023035 RRosenbach et al

Page 9: PAPER Ef cientsimulationofnon-Markoviansystem …web.mit.edu/jianshucaogroup/pdfdir/2016/Javi2016NJP.pdfD = 0.6 .Onecanobservethat,forthefastestbathJ 2,oscillationsaresustainedforalongertime,whilethis

tr e e 0 , 23Ht Hti i[ ˆ ˆ ( )] ( )m mr= -

between times t=0 and t t= such that the steady state has been reached at τ.In the limit of weak interactionwith the environment, the absorption spectrum emerging fromHamiltonian

equation (21) exhibits two peaks in the one-exciton subspace. One of them is shown in the 0.0181 l = line(green) offigure 6, corresponding to the second excited state in the excitonicmanifold at awavelength of around0.363 c

. For higher coupling strengths with the environment, the emergence of the vibrational fine structure

splits the peak in two, which is shown in the 0.181 l = line (blue) and the 1.81 l = line (black).It will be interesting to compare the efficiency of the approach presented herewith other approaches such as

stochastic path integralmethodswhich has recently been developed to calculate absorption and emission spectra[17] specifically for low temperatures and long times.

4.4. Simulation timeThe ability of the TEDOPA-TTMcombination to explore new simulation regimes is a consequence of theextraordinary savings in computational resources.Wewill explore these in terms of the ‘wall time’ tw, thephysical time required for the simulation to be executed asmeasured by an external clock.

Three factors have a direct influence on simulation time:

• bath coherence time tbath,

• chain lengthN and

• systemdimension dsys.

TTM requires the simulation of dsys2 trajectories until tbath, one for each independent initial densitymatrix.

Although this overheadmay become inconvenient for systems of large dimension, the computationmay beparallelized to avoid a scaling of twwith dsys. Evenwithout parallelization, numerical studies often requireexploration of a large number of independent initial conditions anyway.

Due to the efficiency ofmultiplicative propagationwith TTM (equation (14)), nearly the totality of thewalltime tw required for a simulation until tsim is employed in the initial generation of the tensors until tbath withTEDOPA. Therefore, onemay consider tw to be essentially independent of tsim. Thismakes the TTM-TEDOPAcombination suitable for long time simulations, i.e.cases where t tsim bath . There is an additional benefit inshortening simulationswith TEDOPA to tbath, since this reduces the necessary chain lengthN.

The scaling of thewall time tw necessary to perform aTEDOPA simulation of timestep td until tbath can beexpressed as

t Nt

tt , 24w

bath ¯ ( )d

µ

where dependence on three factors has beenmade explicit: the number of sitesN, the number of time steps t

tbath

dand a factort̄ denoting the averagewall time necessary to simulate one chain site during one time step td .

Figure 6.Peak structure of the absorption spectrumof a dimeric system for different values of the system-environment couplingstrength. The emergence of the vibrational fine structure is apparent for increasing strength of the coupling to the environment.

9

New J. Phys. 18 (2016) 023035 RRosenbach et al

Page 10: PAPER Ef cientsimulationofnon-Markoviansystem …web.mit.edu/jianshucaogroup/pdfdir/2016/Javi2016NJP.pdfD = 0.6 .Onecanobservethat,forthefastestbathJ 2,oscillationsaresustainedforalongertime,whilethis

However, in order to avoid end-of-chain recurrences, for a simulation timetbath one requires

N t v, 25bath · ¯ ( )µ

sitesN in the environment, given an average propagation speed v̄ in the chain. Thus a total wall time of

tvt

tt c t , 26w bath

2bath2¯¯

· ( )d

µ º

is neededwhere c is a scenario-dependent constant.The global speedup provided by the TTM-TEDOPA combination is illustrated in figure 7 for three instances

with different spectral densities. The near independence of tw on tsim is shown for large tsim. As shown in theinset, in reality tw increases linearly with tsim, althoughwith a negligible slope. For simulationswith TEDOPAalone, the quadratic dependence expressed in equation (26) extends beyond tbath until tsim.

5. Conclusion and outlook

In this workwe demonstrated that the combination of TEDOPA andTTMresult in an enhanced simulationmethod of general non-Markovian open-quantum-systems especially well-suited for (but not restricted to) low-temperature regimes and highly structured spectral densities. The formulation in terms of amultiplicativeoperatorwhose size is independent of the goal simulation time facilitates exploration ofmuch longer, so farinaccessible timescales.

We verified the feasibility of this combination by a benchmark and presented applications for variousspectral densities to highlight theflexibility of ourmethod. Further to the paradigmatic examples presented,even larger benefits can be expected upon application to simulationswhich are post-processed by someaveraging-typemethod. These are often noise-tolerant or noise-stable, so small deviations do not change thecharacteristic features of the final result. This type of analysis are expected to be of crucial importance forproviding accuratemicroscopicmodels of the dynamical behavior ofmesoscopic systems and therefore a betterunderstanding of how coherent effects stillmanifest in those time and length scales [60–63].

Acknowledgments

Thisworkwas supported by theAlexander vonHumboldt-Professorship, the EU Integrating project SIQS, theEUSTREP projects PAPETS,QUCHIP and EQUAM,National Science Foundation (NSF) (GrantNo. CHE-1112825) andDefense AdvancedResearch Projects Agency (DARPA) (GrantNo.N99001-10-1-4063), theMIT-Germany Seed Fund, and the ERCSynergy grant BioQ. Computational resources used included bwUniCluster,supported by theMinistry of Science, Research and the Arts Baden-Württemberg and theUniversities of theState of Baden-Württemberg, Germany, within the framework programbwHPC.

Figure 7.Data points show single-core TEDOPAwall times tw for a specific simulation time t ;sim the respective line of the same color isthe corresponding quadratic fit. Black lines show the simulation time for the same physical setting upon employing the combinationof TEDOPA andTTM. The TTM-part grows linearly as can be seen in the inset (on themain panel the slope of these lines is so smallthat they appear horizontal). Note the different scales on the vertical axis betweenmain plot and inset.

10

New J. Phys. 18 (2016) 023035 RRosenbach et al

Page 11: PAPER Ef cientsimulationofnon-Markoviansystem …web.mit.edu/jianshucaogroup/pdfdir/2016/Javi2016NJP.pdfD = 0.6 .Onecanobservethat,forthefastestbathJ 2,oscillationsaresustainedforalongertime,whilethis

References

[1] TaharaH,OgawaY,Minami F, AkahaneK and SasakiM2014Phys. Rev. Lett. 112 147404[2] Putz S, KrimerDO,AmsüssD, ValookaranA,Nöbauer T, Schmiedmayer J, Rotter S andMajer J 2014Nat. Phys. 10 720–4[3] Cai JM, Retzker A, Jelezko F and PlenioMB2013Nat. Phys. 9 168[4] Engel G, CalhounT, Read E, AhnT,Mancal T, Cheng Y, Blankenship R and FlemingG 2007Nature 446 782[5] PanitchayangkoonG,HayesD, FranstedKA,Caram JR,Harel E,Wen J, Blankenship RE and Engel G S 2010 Proc. Natl Acad. Sci. USA

107 12766[6] Collini E,WongCY,Wilk KE, Curmi PMGand Scholes GD2010Nature 463 644[7] BreuerHP and Petruccione F 2007The Theory of OpenQuantumSystems (Oxford:OxfordUniversity Press)[8] Rivas Á andHuelga S F 2011OpenQuantumSystems: An Introduction (SpringerBriefs in Physics) (Berlin: Springer)[9] LindbladG1975Commun.Math. Phys. 40 147[10] Gorini V, Kossakowski A and Sudarshan ECG1976 J.Math. Phys. 17 821[11] Redfield A 1957 IBM J. Res. Dev. 1 19[12] Nakajima S 1958Prog. Theor. Phys. 20 948[13] SmirneA andVacchini B 2010Phys. Rev.A 82 022110[14] FeynmanRP 1948Rev.Mod. Phys. 20 367[15] FeynmanRP andVernon F L 1963Ann. Phys. 24 118[16] Cao J 1997 J. Chem. Phys. 107 3204[17] Tanimura Y andKuboR 1989 J. Phys. Soc. Japan 58 1199[18] MakriN andMakarovDE 1995 J. Chem. Phys. 102 4611[19] Leggett A J, Chakravarty S, Dorsey AT, FisherMPA,Garg A andZwergerW1987Rev.Mod. Phys. 59 1[20] Diosi L, StrunzWTandGisinN1998Phys. Rev.A 58 1699[21] TangZ,OuyangX,GongZ,WangH andWu J 2015 J. Chem. Phys. 143 224112[22] Weiss S, Eckel J, ThorwartM and Egger R 2008Phys. Rev.B 77 195316[23] MantheU2008 J. Chem. Phys. 128 164116[24] ShiQ andGeva E 2003 J. Chem. Phys. 119 12063[25] CohenG andRabani E 2011Phys. Rev.B 84 075150[26] Kidon L,Wilner EY andRabani E 2015 J. Chem. Phys. 143 234110[27] Kapral R 2015 J. Phys. Condens.Matter 27 073201[28] Kelly A andMarklandTE 2013 J. Chem. Phys. 139 014104[29] Gottwald F, Karsten S, Ivanov SD andKühnO2015 J. Chem. Phys. 142 244110[30] Cao J, Ungar LWandVothGA1996 J. Chem. Phys. 104 4189[31] Egger R,Mühlbacher L andMakCH2000Phys. Rev.E 61 5961[32] Stockburger J andGrabertH 2002Phys. Rev. Lett. 88 170407[33] Mühlbacher L andRabani E 2008Phys. Rev. Lett. 100 176403[34] Moix JM andCao J 2013 J. Chem. Phys. 139 134106[35] Baer R andKosloff R 1997 J. Chem. Phys. 106 8862[36] Gualdi G andKochCP 2013Phys. Rev.A 88 022122[37] deVega I 2014Phys. Rev.A 90 043806[38] Prior J, ChinAW,Huelga S F and PlenioMB2010Phys. Rev. Lett. 105 050404[39] ChinAW,Rivas A,Huelga S F and PlenioMB2010 J.Math. Phys. 51 092109[40] SchollwöckU2005Rev.Mod. Phys. 77 259[41] Cerrillo J andCao J 2014Phys. Rev. Lett. 112 110401[42] Mehraeen S, Cerrillo J andCao J 2015 in preparation[43] CohenG,Gull E, ReichmanDR,Millis A J andRabani E 2013Phys. Rev.B 87 195108[44] Wilner EY,WangH,CohenG, ThossM andRabani E 2013Phys. Rev.B 88 045137[45] Wilner EY,WangH, ThossM andRabani E 2014Phys. Rev.B 89 205129

Wilner EY,WangH, ThossM andRabani E 2014Phys. Rev.B 90 115145Wilner EY,WangH, ThossM andRabani E 2015Phys. Rev.B 92 195143

[46] WoodsMP,CramerMandPlenioMB2015Phys. Rev. Lett. 115 130401[47] Burkey R andCantrell C 1984 J. Opt. Soc. Am.B 1 169[48] GautschiW1994 J. Trans.Math. Softw. 20 21[49] deVega I, SchollwöckU andWolf FA 2015Phys. Rev.B 92 155126[50] SchollwöckU2011Ann. Phys. 326 96[51] Vidal G 2004Phys. Rev. Lett. 93 040502[52] White S R and FeiguinAE 2004Phys. Rev. Lett. 93 76401[53] Wolf FA,McCulloch I P and SchollwöckU2014Phys. Rev.B 90 235131[54] WoodsMP,Groux R,ChinAW,Huelga S F andPlenioMB2014 J.Math. Phys. 55 032101[55] ChinAW, Prior J, Rosenbach R,Caycedo-Soler F,Huelga S F and PlenioMB2013Nat. Phys. 9 113[56] Dijkstra AG,WangC,Cao J and FlemingGR2015 J. Phys. Chem. Lett. 6 627[57] Adolphs J andRenger T 2006Biophys. J. 91 2778[58] WilhelmFW2008New J. Phys. 10 115011[59] ChinAW,Huelga S F and PlenioMB2012Phys. Rev. Lett. 109 233601[60] Huelga S F and PlenioMB2013Contemp. Phys. 54 181[61] Levi F,Mostarda S, Rao F andMintert F 2015Rep. Prog. Phys. 78 082001[62] Scholes GD, FlemingGR,Olaya-Castro A and vanGrondelle R 2011Nat. Chem. 3 763[63] MohseniM,Omar Y, Engel G S and PlenioMB (ed) 2014QuantumEffects in Biology (Cambridge: CambridgeUniversity Press)

11

New J. Phys. 18 (2016) 023035 RRosenbach et al