exact quantum dynamics in structured environments · exact quantum dynamics in structured...

10
Exact Quantum Dynamics in Structured Environments Dominic Gribben, 1 Aidan Strathearn, 1, 2 Jake Iles-Smith, 3 Dainius Kilda, 1 Ahsan Nazir, 4 Brendon W. Lovett, 1 and Peter Kirton 5 1 SUPA, School of Physics and Astronomy, University of St Andrews, St Andrews, KY16 9SS, United Kingdom 2 School of Mathematics and Physics, The University of Queensland, St Lucia, Queensland 4072, Australia 3 Department of Physics and Astronomy, University of Sheffield, Sheffield, S3 7RH, United Kingdom 4 School of Physics and Astronomy, The University of Manchester, Oxford Road, Manchester M13 9PL, United Kingdom 5 Vienna Center for Quantum Science and Technology, Atominstitut, TU Wien, 1040 Vienna, Austria (Dated: July 8, 2019) The dynamics of a wide range of technologically important quantum systems are dominated by their interaction with just a few environmental modes. Such highly structured environments give rise to long-lived bath correlations that induce complex dynamics which are very difficult to simulate. These difficulties are further aggravated when spatial correlations between different parts of the system are important. By modeling the dynamics of a pair of two-level quantum systems in a common, structured, environment we show that a recently developed numerical approach, the time- evolving matrix product operator, is capable of accurate simulation under exactly these conditions. We find that tuning the separation to match the wavelength of the dominant environmental modes can drastically modify the system dynamics. To further explore this behavior, we show that the full dynamics of the bath can be calculated directly from those of the system, thus allowing us to develop intuition for the complex system dynamics observed. When a quantum system interacts with a structured environment in which a narrow band of modes domi- nate, the resulting dynamics can be very complex and difficult to simulate accurately. Such environments can retain a memory of their interactions with the system on a timescale comparable to that on which the state of the system changes, and in the face of such memory effects standard open system techniques can fail. How- ever, reliable simulation of such environments is vital in order to understand the behavior of an ever-increasing number of experimental platforms and quantum devices. For example, in photosynthetic systems the observation of quantum coherence comes from strong electronic and vibrational coupling in biomolecules [1], where the inter- play of these degrees of freedom has been shown to en- hance energy transport [2–5]. Similar physics can also be found in e.g. trapped ions [6] and molecular junctions [7]. Further, bath memory plays a key role in the function of micromechanical resonators, where adjustment of the en- vironmental noise spectrum is possible [8] and photonic crystals, where band gaps in the spectral density give rise to localized modes and dissipationless oscillations [9–11]. The effect of structured environments has also been ex- plored with superconducting qubits subject to different kinds of noise [12]. Memory effects can be even more pronounced when the system coupled to the structured environment de- scribed above is extended. For example if it consists of several spatially separated sites. The dynamics then be- come sensitive to groups of environmental modes with a narrower bandwidth, and so a longer correlation time. This sensitivity was recently observed in an experiment on a ‘giant’ superconducting atom [13]. Spatial correla- tion effects have also been studied in the context of quan- tum transport [14–18], and spatially correlated noise in quantum registers has been shown to affect quantum er- ror propagation [19]. In order to model environmental memory effects, we must go beyond the standard Born-Markov approxima- tions [20] which lead to a simple time-local master equa- tion for the system density matrix. The development of techniques to simulate these non-Markovian dynam- ics has therefore been the subject of much relatively re- cent theoretical effort [21]. These techniques broadly fall into two categories: First, there are approximate meth- ods which change the boundary between the system and environment such that the Born-Markov approximations are valid for the new system [22–27]; second, there are numerically exact methods which utilize some particular structure of the bath Hamiltonian to provide exact dy- namics [28–33]. These approaches usually work best for models where the system Hilbert space is quite small, or has a specific form of spectral density. However, it can often be difficult to know a priori the range of validity of some of these approaches, and so accurate and efficient benchmarking procedures are essential. In this Letter, we demonstrate that a recently devel- oped exact approach, the Time-Evolving Matrix Prod- uct Operator (TEMPO) [32] can be used to efficiently find non-local system dynamics for highly structured en- vironments. To do this, we study a model consisting of a pair of spatially separated two-level subsystems inter- acting with a common environment, in which a narrow band of modes dominate the interaction. The complex environmental structure and spatial correlations give rise to highly non-Markovian dynamics, which we are able to arXiv:1907.02932v1 [quant-ph] 5 Jul 2019

Upload: others

Post on 14-Aug-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Exact Quantum Dynamics in Structured Environments · Exact Quantum Dynamics in Structured Environments Dominic Gribben,1 Aidan Strathearn,1,2 Jake Iles-Smith,3 Dainius Kilda,1 Ahsan

Exact Quantum Dynamics in Structured Environments

Dominic Gribben,1 Aidan Strathearn,1, 2 Jake Iles-Smith,3

Dainius Kilda,1 Ahsan Nazir,4 Brendon W. Lovett,1 and Peter Kirton5

1SUPA, School of Physics and Astronomy, University of St Andrews, St Andrews, KY16 9SS, United Kingdom2School of Mathematics and Physics, The University of Queensland, St Lucia, Queensland 4072, Australia

3Department of Physics and Astronomy, University of Sheffield, Sheffield, S3 7RH, United Kingdom4School of Physics and Astronomy, The University of Manchester,

Oxford Road, Manchester M13 9PL, United Kingdom5Vienna Center for Quantum Science and Technology, Atominstitut, TU Wien, 1040 Vienna, Austria

(Dated: July 8, 2019)

The dynamics of a wide range of technologically important quantum systems are dominated bytheir interaction with just a few environmental modes. Such highly structured environments give riseto long-lived bath correlations that induce complex dynamics which are very difficult to simulate.These difficulties are further aggravated when spatial correlations between different parts of thesystem are important. By modeling the dynamics of a pair of two-level quantum systems in acommon, structured, environment we show that a recently developed numerical approach, the time-evolving matrix product operator, is capable of accurate simulation under exactly these conditions.We find that tuning the separation to match the wavelength of the dominant environmental modescan drastically modify the system dynamics. To further explore this behavior, we show that thefull dynamics of the bath can be calculated directly from those of the system, thus allowing us todevelop intuition for the complex system dynamics observed.

When a quantum system interacts with a structuredenvironment in which a narrow band of modes domi-nate, the resulting dynamics can be very complex anddifficult to simulate accurately. Such environments canretain a memory of their interactions with the systemon a timescale comparable to that on which the stateof the system changes, and in the face of such memoryeffects standard open system techniques can fail. How-ever, reliable simulation of such environments is vital inorder to understand the behavior of an ever-increasingnumber of experimental platforms and quantum devices.For example, in photosynthetic systems the observationof quantum coherence comes from strong electronic andvibrational coupling in biomolecules [1], where the inter-play of these degrees of freedom has been shown to en-hance energy transport [2–5]. Similar physics can also befound in e.g. trapped ions [6] and molecular junctions [7].Further, bath memory plays a key role in the function ofmicromechanical resonators, where adjustment of the en-vironmental noise spectrum is possible [8] and photoniccrystals, where band gaps in the spectral density give riseto localized modes and dissipationless oscillations [9–11].The effect of structured environments has also been ex-plored with superconducting qubits subject to differentkinds of noise [12].

Memory effects can be even more pronounced whenthe system coupled to the structured environment de-scribed above is extended. For example if it consists ofseveral spatially separated sites. The dynamics then be-come sensitive to groups of environmental modes witha narrower bandwidth, and so a longer correlation time.This sensitivity was recently observed in an experimenton a ‘giant’ superconducting atom [13]. Spatial correla-

tion effects have also been studied in the context of quan-tum transport [14–18], and spatially correlated noise inquantum registers has been shown to affect quantum er-ror propagation [19].

In order to model environmental memory effects, wemust go beyond the standard Born-Markov approxima-tions [20] which lead to a simple time-local master equa-tion for the system density matrix. The developmentof techniques to simulate these non-Markovian dynam-ics has therefore been the subject of much relatively re-cent theoretical effort [21]. These techniques broadly fallinto two categories: First, there are approximate meth-ods which change the boundary between the system andenvironment such that the Born-Markov approximationsare valid for the new system [22–27]; second, there arenumerically exact methods which utilize some particularstructure of the bath Hamiltonian to provide exact dy-namics [28–33]. These approaches usually work best formodels where the system Hilbert space is quite small, orhas a specific form of spectral density. However, it canoften be difficult to know a priori the range of validity ofsome of these approaches, and so accurate and efficientbenchmarking procedures are essential.

In this Letter, we demonstrate that a recently devel-oped exact approach, the Time-Evolving Matrix Prod-uct Operator (TEMPO) [32] can be used to efficientlyfind non-local system dynamics for highly structured en-vironments. To do this, we study a model consisting ofa pair of spatially separated two-level subsystems inter-acting with a common environment, in which a narrowband of modes dominate the interaction. The complexenvironmental structure and spatial correlations give riseto highly non-Markovian dynamics, which we are able to

arX

iv:1

907.

0293

2v1

[qu

ant-

ph]

5 J

ul 2

019

Page 2: Exact Quantum Dynamics in Structured Environments · Exact Quantum Dynamics in Structured Environments Dominic Gribben,1 Aidan Strathearn,1,2 Jake Iles-Smith,3 Dainius Kilda,1 Ahsan

2

explore using TEMPO. In addition, we find that by tun-ing the separation between the subsystems, it is possibleto control their interaction with the environment.

An advantage of techniques that include some envi-ronmental modes in the system description is that theyenable the extraction of some of the bath’s dynamics,which can in turn lead to insight into the system’s be-havior through understanding system-environment cor-relations [3, 25]. However, we will show here that exactenvironmental dynamics can be easily and efficiently ex-tracted from TEMPO, leading us to a full picture of thecomplex quantum dynamics of this model.

The Hamiltonian for our model is given by:

H =∑i

εi |Xi〉〈Xi|+Ω

2(|X1〉〈X2|+ |X2〉〈X1|)

+∑i

σz,i∑k

(gi,kak + g∗i,ka†k) +

∑k

ωka†kak. (1)

The two-level system (TLS) at site i and position ri hasenergy splitting εi and an excited (ground) state denoted|Xi〉 (|0i〉). The pair of TLSs form a dimer with a coher-ent coupling of strength Ω, and the environment consistsof bosons confined to one dimension: a†k creates an ex-citation in the mode with wave vector k. We also intro-duce the Pauli operators σz,i = |Xi〉〈Xi| − |0i〉〈0i|. Thecoupling gi,k = gk exp(−ikri) consists of a position inde-pendent part gk and a phase that depends on the site i.This kind of model could underpin a wide range of phys-ical systems, for example biological or molecular systemsundergoing energy transport and interacting with vibra-tional modes [4], superconducting qubits in microwaveresonators [34], or quantum dots interacting with a mi-cromechanical resonator [35].

The bath can be completely characterized by its spec-tral density

J(ω) =∑k

|gk|2δ(ω − ωk), (2)

from which we can calculate its autocorrelation functionin thermal equilibrium at temperature T :

C(t) =

∫ ∞0

dωJ(ω)(

coth( ω

2T

)cos(ωt)− i sin(ωt)

).

(3)

If this function decays slowly compared to the systemtimescales then the bath memory is important.

We now consider a bath in which a narrow band ofmodes dominate the interaction with the system, givingrise to the spectral density [36]

J0(ω) =αΓω2

(ω20 − ω2)2 + Γ2ω2

. (4)

Here α gives the coupling strength, ω0 is the frequencycharacterizing the dominant mode and Γ provides a mea-sure of the width of the dominant band.

FIG. 1. (a) Cartoon of the dimer system we consider accom-panied with schematic plots of (b) the under-damped spectraldensity J(ω) and (c) the corresponding bath auto-correlationfunction C(t).

The Hamiltonian, Eq. (1), does not couple states witha different total number of system excitations and sowe restrict ourselves to the single excitation subspace,the only one where non-trivial dynamics are observed.Within this subspace it is possible to map the problemonto that of a single spin coupled to a bosonic environ-ment [37]. The mapped Hamiltonian is then a spin-bosonHamiltonian

H =Ω

2σx +

ε

2σz + σz

∑k

(gkak + g∗ka†k) +

∑k

ωka†kak,

(5)

where ε = ε2 − ε1 and gk = 2igk sin(k(r1 − r2)/2). Wealso define new Pauli operators: σz = (σz,2−σz,1)/2 andσx = |X2〉〈X1|+ |X1〉〈X2|. This modification to the cou-pling constants results in a renormalization of the spec-tral density

J(ω) = 2J0(ω)(1− cos(ωR)) (6)

where R = |r1 − r2| is the dimer separation and we haveassumed a linear dispersion ω(k) = c|k| with c = 1. Thecosine term arises from the phase factors in the originalcoupling terms.

We begin by studying a simplified case of a single spindescribed by Eq. (5) interacting with a bath with theunmodified spectral density, Eq. (4). It is straightfor-ward to simulate this case with the reaction-coordinate(RC) mapping, which has been rigorously benchmarkedfor similar problems [25–27], allowing us to verify thatTEMPO is able to accurately simulate dynamics in struc-tured environments. This mapping takes a single collec-tive mode of the environment (the RC) into the systemdefinition; then the rest of the bath, which is assumedto be Markovian, couples to the now-augmented system.More details can be found in Ref. [38]. The RC mappingcan be performed analytically for the spectral density inEq. (4) [25], and is expected to give accurate results inthe cases we present here. To simulate more complex

Page 3: Exact Quantum Dynamics in Structured Environments · Exact Quantum Dynamics in Structured Environments Dominic Gribben,1 Aidan Strathearn,1,2 Jake Iles-Smith,3 Dainius Kilda,1 Ahsan

3

environments would require more RCs in the augmentedsystem Hilbert space, whose dimension grows exponen-tially with the number of included RCs.

An alternative, numerically exact way of simulatingthe system dynamics arising from Eq. (5) is to use aFeynman sum-over-histories approach, using an influencefunctional to capture environmental memory effects. Bywriting this path integral representation as a sum overdiscrete time steps, we can calculate the system quantumdynamics by contracting a tensor network: this is theTEMPO approach [32]. The memory effects are then en-coded in a matrix product state which contains informa-tion about the history of the system. This is propagatedforward in time by successive contraction with a ma-trix product operator. By performing singular value de-compositions and truncation after each timestep [39, 40]we can capture bath memory times orders-of-magnitudebeyond those possible with previous path integral ap-proaches [29, 30]. The similarities between TEMPO andthe process tensor [41] have recently been explored [42],and TEMPO has also been used to model dynamics inoptomechanical systems [43]. Further details on TEMPOcan be found in Ref. [38].

In Fig. 2 we compare dynamics generated by thesetwo approaches for the simplified model of the singlespin interacting with the structured spectral density, forthree different values of ω0. We find excellent agree-ment between the two algorithms for all sets of parame-ters. The RC approach is accurate for these parametersand this form of spectral density, and so able to capturethe complex dimer dynamics. In this simple case RCrequires significantly reduced numerical resources com-pared to TEMPO, though we find that both techniquestake longer to converge at lower mode frequencies andhigher temperatures. For the RC algorithm the reasonfor this is clear: the occupation of the mode includedin the system increases in these regimes and so a largersystem Hilbert space is needed for convergence. ForTEMPO we find that the number of singular values thatneed to be retained for convergence increases – i.e. moresystem paths gain a significant amplitude, as expectedwhen the environment becomes more occupied. Theseresults show both that the RC approach is able to ac-curately simulate dynamics for this model [27] and thatTEMPO is able to deal with such structured spectraldensities.

We now turn to analyzing the full model including thespatial modifications to the spectral density (i.e. usingEq. 6). Previous studies have shown that spatial correla-tions in the bath can have a significant effect on popula-tion transfer between localized systems [14–16, 32]. How-ever, these studies focused on the case where the originalspectral density J0(ω) is much broader than that givenin Eq. (2) – i.e. they do not focus on systems where justa narrow band of modes dominate the system interaction– and so the modification due to the spatial structure of

FIG. 2. (a) Underdamped spectral density, given by Eq. (4),with the different frequencies used. (b) Corresponding corre-lation functions. (c)-(e) Dynamics calculated using TEMPO(dots) and RC (solid lines) for ω0 = Ω/4,Ω/2,Ω. The otherparameters are πα = 0.05Ω, Γ = 0.05Ω, T = Ω, ε = 0.5Ω.

FIG. 3. Effect of the mapping, Eq. (6), on the spectraldensity. (a) The unmapped spectral density (grey shaded)and an example of the mapped case (solid black). (b)-(d)Mapped spectral densities for the three cases considered inthe main text from top to bottom: R = 2π/ω0, 2π/1.1ω0

and 2π/0.9ω0. The dashed black lines depict the frequenciesdescribed in the main text.

the system is not as pronounced.Since we are now using the general mapped spectral

density in Eq. (6), treating the dynamics using the RCformalism becomes much more difficult: to accuratelycapture the dynamics it would be necessary to includemore modes in the system than is computationally feasi-ble. This means that for the remainder of this Letter allresults are obtained using TEMPO.

In this model there are now three possible resonanceconditions which can be met by matching two of: thebare system frequency Ω, the characteristic frequency ofthe environment ω0, or the frequency corresponding tothe separation between the TLSs ωR = 2π/R. In Fig. 3we show how choosing ωR = ω0 leads to a complete sup-pression of the main peak of the bare spectral density,and choosing ωR = (1± 0.1)ω0 leads to a very asymmet-ric lineshape. In the single line plots of Fig. 4 we showhow these differences in spectral density manifest them-selves in the system dynamics. We show results for thethree TLS spacings ωR = ω0, (1 ± 0.1)ω0 and for baresystem frequencies Ω tuned to the same three possiblevalues, resulting in nine sets of results.

The system dynamics are complex and difficult to in-terpret, with multiple oscillation frequencies appearing

Page 4: Exact Quantum Dynamics in Structured Environments · Exact Quantum Dynamics in Structured Environments Dominic Gribben,1 Aidan Strathearn,1,2 Jake Iles-Smith,3 Dainius Kilda,1 Ahsan

4

in the dynamics. These are not only due to the peakedstructure of the original spectral density (as occurredfor the simplified single spin model) but also because ofthe spatial correlations. However, we next show thatTEMPO can be used not only to extract system dynam-ics, as discussed in Ref. [32], but also to find evolutionof bath degrees of freedom. This will enable us to gainmore insight into how the interplay between the bath andsystem leads to the complex dynamics shown in Fig. 4.To do this we formally solve the Heisenberg equationsof motion [44, 45] for the expectation value of e.g. theannihilation operator for a particular bath mode, andobtain an expression entirely in terms of the expectationof the system operator which couples to the bath. Forour model, Eq. (5), we have

〈ak(t)〉 = −igke−iωkt

∫ t

0

dt′eiωkt′〈σz(t′)〉. (7)

We thus have access to the dynamics of all the bathmodes and can construct the real-space displacement ofthe bath by creating an appropriately weighted sum ofthe individual mode displacements. The quantity we con-sider is the displacement of a one-dimensional field [46]given by:

Φ(x, t) =∑k

1√ωk

(〈ak(t)〉eikx + 〈a†k(t)〉e−ikx). (8)

In the continuum limit this takes the following form:

Φ(x, t) = 8

∫ ∞0

[√J0(ω)

ωsin

(ωR

2

)sin(ωx)

×∫ t

0

dt′ sin(ω(t− t′))〈σz(t′)〉]

(9)

where we have used ω = |k|. See Ref. [38] for a derivationof this expression.

We show the real-space displacement of the bath as afunction of time and position as a color map beneath thecorresponding system dynamics in Fig. 4. The black dot-ted lines indicate the positions of the two TLSs. In thefully resonant case (the centre plot) we see that most ofthe bath excitation gets trapped between the two TLSswhich simply show oscillations which decay away at avery slow rate. We also see slow decay in the other twoplots where Ω = ωR; this is because J(Ω) = 0 in thesecases and hence a simple Markovian treatment would pre-dict no decay at all.

When at least one of the frequencies is detuned fromthis condition we see a more pronounced propagation ofbath excitations away from the system as the trapping ef-fect is reduced. In these cases the overall decay rate of theTLS is gradually enhanced as the detuning is increased,since now the value of J(Ω) is larger. The detuning canalso introduce a beat frequency into the bath dynamics,

FIG. 4. Spin dynamics (single line plots) and correspondingenvironment displacement (main panels) at position x overtime as calculated with Eq. (9) for various separations anddimer frequencies. Displacements are normalized to the max-imum displacement across all plots. The dashed black linescorrespond to the positions of the TLSs. For all of the dynam-ics we have set T = 0 and ε = 0, Γ = 0.05ω0 and πα = 0.1ω0

for the unmapped spectral density.

most clearly seen in the bath population trapped betweenthe two TLSs, and most pronounced in the most detunedcases, i.e. in the bottom right and top left panels of Fig. 4.This beating gives rise to distinctive revivals in the TLSdynamics.

At very short times all of the dynamics are very simi-lar. On timescales t < ω−1R each TLS only senses its ownlocal environment and hence behaves as if it were inter-acting with bosons described by the unmapped spectraldensity. However as soon as the influence of the otherTLS is felt the dynamics become radically different tothose predicted by an independent environment model.These dynamics are highly sensitive to the relative detun-ings of the three frequencies described above, becomingsignificantly different even for the very small detuningsstudied here.

In conclusion, we have shown that TEMPO can pro-vide accurate simulations of quantum systems in struc-tured environments. We first validated the technique bysimulating the quantum dynamics of a simple model thatcan be solved accurately using the RC approach, andcomparing the results. We then presented TEMPO sim-ulations of the more complex dynamics that result fromthe interplay between a highly structured spectral func-tion and spatial correlations between different parts ofthe system. TEMPO is not limited to specific forms forthe spectral density and so provides a versatile methodfor simulating systems coupled to an environment witharbitrary structure. In addition, we have shown that it

Page 5: Exact Quantum Dynamics in Structured Environments · Exact Quantum Dynamics in Structured Environments Dominic Gribben,1 Aidan Strathearn,1,2 Jake Iles-Smith,3 Dainius Kilda,1 Ahsan

5

is possible to find the full dynamics of the environmentdirectly from the system dynamics. This enabled us toexplore how it is possible to tune the separation betweenthe spins such that one can effectively remove the cou-pling to the dominant environmental mode and how theresultant dynamics are highly sensitive to slight varia-tions in parameters around this point. Such exact envi-ronmental tracking can help to explain the behavior ofopen quantum systems in general, and so aid in the de-sign of future quantum devices. This greater insight intothe dynamics of open quantum systems could also leadto the development of new approximation schemes andto the self-consistent verification of others.

DG and DK acknowledge studentship funding fromEPSRC under grant no. EP/L015110/1. AS acknowl-edges a studentship from EPSRC under grant no.EP/L505079/1. J.I.-S. acknowledges support from theRoyal Commission for the Exhibition of 1851. ANacknowledges funding from EPSRC under grant no.EP/N008154/1. PK acknowledges support from an ESQfellowship of the Austrian Academy of Sciences (OAW).

[1] D. B. Turner, K. E. Wilk, P. M. G. Curmi, and G. D.Scholes, J. Phys. Chem. Lett. 2, 1904 (2011).

[2] M. Thorwart, J. Eckel, J. Reina, P. Nalbach, and S. Weiss,Chem. Phys. Lett. 478, 234 (2009).

[3] A. W. Chin, J. Prior, R. Rosenbach, F. Caycedo-Soler,S. F. Huelga, and M. B. Plenio, Nat. Phys. 9, 113 (2013).

[4] M. del Rey, A. W. Chin, S. F. Huelga, and M. B. Plenio,J. Phys. Chem. Lett. 4, 903 (2013).

[5] M. Mohseni, A. Shabani, S. Lloyd, and H. Rabitz, J.Chem. Phys. 140, 035102 (2014).

[6] C. Maier, T. Brydges, P. Jurcevic, N. Trautmann,C. Hempel, B. P. Lanyon, P. Hauke, R. Blatt, and C. F.Roos, Phys. Rev. Lett. 122 (2019).

[7] D. Segal and B. K. Agarwalla, Ann. Rev. Phys. Chem. 67,185 (2016).

[8] S. Groblacher, A. Trubarov, N. Prigge, G. D. Cole, M. As-pelmeyer, and J. Eisert, Nat. Commun. 6 (2015).

[9] W.-M. Zhang, P.-Y. Lo, H.-N. Xiong, M. W.-Y. Tu, andF. Nori, Phys. Rev. Lett. 109 (2012).

[10] A. Goban, C.-L. Hung, S.-P. Yu, J. Hood, J. Muniz,J. Lee, M. Martin, A. McClung, K. Choi, D. Chang,O. Painter, and H. Kimble, Nat. Commun. 5, 3808 (2014).

[11] N. M. Sundaresan, R. Lundgren, G. Zhu, A. V. Gorshkov,and A. A. Houck, Phys. Rev. X 9, 011021 (2019).

[12] A. Potocnik, A. Bargerbos, F. A. Y. N. Schroder, S. A.Khan, M. C. Collodo, S. Gasparinetti, Y. Salathe, C. Cre-atore, C. Eichler, H. E. Tureci, A. W. Chin, and A. Wall-raff, Nat. Commun. 9, 904 (2018).

[13] G. Andersson, B. Suri, L. Guo, T. Aref, and P. Delsing,(2018), arXiv:1812.01302.

[14] P. Nalbach, J. Eckel, and M. Thorwart, New J. Phys.12, 065043 (2010).

[15] A. Nazir, Phys. Rev. Lett. 103, 146404 (2009).[16] D. P. S. McCutcheon and A. Nazir, Phys. Rev. B 83,

165101 (2011).

[17] D. Abramavicius and S. Mukamel, J. Chem. Phys. 134,174504 (2011).

[18] M. Sarovar, Y.-C. Cheng, and K. B. Whaley, Phys. Rev.E 83, 011906 (2011).

[19] D. Aharonov, A. Kitaev, and J. Preskill, Phys. Rev.Lett. 96, 050504 (2006).

[20] H.-P. Breuer and F. Petruccione, The Theory of OpenQuantum Systems (Oxford University Press, 2002).

[21] I. de Vega and D. Alonso, Rev. Mod. Phys. 89, 015001(2017).

[22] R. Silbey and R. A. Harris, J. Chem. Phys. 80, 2615(1984).

[23] R. A. Harris and R. Silbey, J. Chem. Phys. 83, 1069(1985).

[24] S. Jang, Y.-C. Cheng, D. R. Reichman, and J. D. Eaves,J. Chem. Phys. 129, 101104 (2008).

[25] J. Iles-Smith, N. Lambert, and A. Nazir, Phys. Rev. A90, 032114 (2014).

[26] J. Iles-Smith, Excitation dynamics of strongly dissipativequantum systems, Ph.D. thesis, Imperial College London(2015).

[27] J. Iles-Smith, A. G. Dijkstra, N. Lambert, and A. Nazir,J. Chem. Phys. 144, 044110 (2016).

[28] Y. Tanimura and R. Kubo, J. Phys. Soc. Jap. 58, 101(1989).

[29] N. Makri and D. E. Makarov, J. Chem. Phys. 102, 4600(1995).

[30] N. Makri and D. E. Makarov, J. Chem. Phys. 102, 4611(1995).

[31] J. Prior, A. W. Chin, S. F. Huelga, and M. B. Plenio,Phys. Rev. Lett. 105, 050404 (2010).

[32] A. Strathearn, P. Kirton, D. Kilda, J. Keeling, and B. W.Lovett, Nat. Commun. 9, 3322 (2018).

[33] F. Mascherpa, A. Smirne, D. Tamascelli, P. F. Acebal,S. Donadi, S. F. Huelga, and M. B. Plenio, (2019),arXiv:1904.04822.

[34] H. Zheng and H. U. Baranger, Phys. Rev. Lett. 110,113601 (2013).

[35] I. Yeo, P.-L. de Assis, A. Gloppe, E. Dupont-Ferrier,P. Verlot, N. S. Malik, E. Dupuy, J. Claudon, J.-M.Gerard, A. Auffeves, G. Nogues, S. Seidelin, J.-P. Poizat,O. Arcizet, and M. Richard, Nature Nanotechnology 9,106 EP (2013).

[36] S. Mukamel, Principles of nonlinear optical spectroscopy,Vol. 29 (Oxford University Press, New York, 1995).

[37] T. M. Stace, A. C. Doherty, and S. D. Barrett, Phys.Rev. Lett. 95, 106801 (2005).

[38] Supplementary material which cites Refs. [47, 48] andcontains details of the RC mapping, numerical methodsused and derivation of the expression for the bath dis-placement.

[39] R. Orus, Annals of Physics 349, 117 (2014).[40] U. Schollwock, Ann. Phys. 326, 96 (2011).[41] F. A. Pollock, C. Rodrıguez-Rosario, T. Frauenheim,

M. Paternostro, and K. Modi, Phys. Rev. A 97, 012127(2018).

[42] M. R. Jørgensen and F. A. Pollock, (2019),arXiv:1902.00315.

[43] Y. Minoguchi, P. Kirton, and P. Rabl, (2019),arXiv:1904.02164.

[44] C. W. Gardiner and M. J. Collett, Phys. Rev. A 31, 3761(1985).

[45] C. Gardiner and P. Zoller, Quantum noise: a handbook ofMarkovian and non-Markovian quantum stochastic meth-

Page 6: Exact Quantum Dynamics in Structured Environments · Exact Quantum Dynamics in Structured Environments Dominic Gribben,1 Aidan Strathearn,1,2 Jake Iles-Smith,3 Dainius Kilda,1 Ahsan

6

ods with applications to quantum optics, Vol. 56 (SpringerScience & Business Media, 2004).

[46] G. D. Mahan, Many-particle physics (Springer Science &Business Media, 2013).

[47] A. Garg, J. N. Onuchic, and V. Ambegaokar, J. Chem.Phys. 83, 4491 (1985).

[48] A. J. Leggett, Phys. Rev. B 30, 1208 (1984).

Page 7: Exact Quantum Dynamics in Structured Environments · Exact Quantum Dynamics in Structured Environments Dominic Gribben,1 Aidan Strathearn,1,2 Jake Iles-Smith,3 Dainius Kilda,1 Ahsan

1

Supplementary Material for: Exact Quantum Dynamics in Structured Environments

REACTION COORDINATE MAPPING

In this section we outline how the Reaction Coordinate (RC) mapping is implemented to obtain a time-local masterequation for the two-level system-RC (TLS-RC) reduced density matrix. The RC is the coordinate correspondingto the direction the bath equilibrium position is displaced when the system transitions between σz eigenstates. Itcorresponds to a superposition of normal modes of the bath, and is defined by the following mapping:

HI = σz∑k

gk(a†k + ak)→ λσz(c† + c). (S1)

Here, c† (c) create (destroy) an excitation in the reaction coordinate mode. The form of the transformation results incoupling between the RC and the residual modes, bk. The spin-boson Hamiltonian is then mapped to the followingform:

HRC = HS +HI +HB +HC (S2)

HS = Ωσx + εσz + λσz(c† + c) + ΩRCc

†c (S3)

HI = (c† + c)∑k

dk(b†k + bk) (S4)

HB =∑k

νkb†kbk (S5)

HC = (c† + c)2∑k

c2kωk

(S6)

where the last term exists to counter a renormalization of the RC’s potential that arises from the interaction term.Here the TLS-RC coupling is given by λ2 =

∑k g

2k and this form ensures that the canonical bosonic commutation

relation is satisfied. This coupling leads to an RC frequency of ΩRC = λ−2∑k ωkg

2k. The residual environment is

characterised by a new spectral density JRC(ω) =∑k |dk|2δ(ω − ωk) which can be found by replacing the TLS in

both the mapped and un-mapped Hamiltonians with a classical coordinate [S1]. The spectral density does not containany information about the system itself and therefore can be found, in both cases, through the classical equations ofmotion of this coordinate [S2, S3]. The under-damped spectral density considered in the main text leads to ΩRC = ω0,λ =

√παUDω0/2 and JRC = γω where γ = Γ/2πω0.

Once this mapping has been performed we follow the procedure outlined in Ref. [S3] applying both Born and Markovapproximations to arrive at the Schrodinger picture master equation for the reduced TLS-RC density operator, ρ(t):

∂ρ

∂t= −i[HS , ρ(t)]− γ

∫ ∞0

∫ ∞0

dωω cos(ωτ) coth( ω

2T

)[C, [C(−τ), ρ(t)]]

− γ∫ ∞0

∫ ∞0

dωω cos(ωτ)[C, [C(−τ), HS ], ρ(t)] (S7)

where we have defined C = c† + c and C(t) is the same operator in the interaction picture at time t. Truncatingthe Hilbert space of the RC down to n basis states, i.e. permitting a maximum of only n excitations, allows us tonumerically diagonalize HS . This gives us a set of basis states |φj〉 which satisfy HS |φj〉 = φj |φj〉 allowing us toexpress the interaction picture operators as:

C(t) =

2n∑j,k=1

Cjkeiωjkt |φj〉〈φk| (S8)

where Cjk = 〈φj | C |φk〉 and ωjk = φj − φk. We can then re-write Eq. (S7) as:

∂ρ

∂t= −i[HS , ρ(t)]− [C, [χ, ρ(t)]] + [C, Ξ, ρ(t)] (S9)

Page 8: Exact Quantum Dynamics in Structured Environments · Exact Quantum Dynamics in Structured Environments Dominic Gribben,1 Aidan Strathearn,1,2 Jake Iles-Smith,3 Dainius Kilda,1 Ahsan

2

with rate operators:

χ =π

2

∑jk

JRC(ξjk) coth

(ξjk2T

)Cjk |φj〉〈φk| (S10)

Ξ =π

2

∑jk

JRC(ξjk)Cjk |φj〉〈φk| . (S11)

The resulting master equation can then be solved using standard open systems approaches for Markovian systems [S4].

TIME-EVOLVING MATRIX PRODUCT OPERATORS

In this section we outline the implementation of the Time-Evolving Matrix Product Operators (TEMPO) algorithm.For full details see [S5].

The quasi-adiabatic path integral (QUAPI) method [S6] is used to calculate the non-Markovian evolution of thesystem up to time tN = N∆t by summing over all possible paths the system has taken to that point from initial timet0. In practice this can be done through an iterative tensor propagation routine where the object propagated is theaugmented density tensor (ADT). The ADT is grown up to time tN from an initial physical density matrix ρi1(t1)through iterative application of tensors:

AjN ...j1(tN ) =

N∏n=2

Bjn,jn−1,...,j1

in−1,...,i1ρi1(t1). (S12)

The reduced system density matrix at time tN is then given by ρjN (tN ) =∑j1...jN−1

AjN ...j1(tN ). For a harmonicbath linearly coupled to the system, the B tensor is composed of influences across all pairs of time points from t1 to tN ,as well as time-local components due to the coherent Hamiltonian evolution. The discretisation time-step ∆t needsto both resolve the features of the correlation function and ensure the Trotter errors are negligible. Without furtherconsiderations the exponential growth of the ADT only allows memory lengths of order ∼ 20∆t to be simulated. Thisrestricts QUAPI to baths with relatively short-lived and/or slowly varying auto-correlation functions.

In TEMPO the ADT is built in the form of a matrix product state (MPS) [S7, S8] and a singular value decomposition(SVD) sweep is carried out at each time-step. The B tensors are composed of a product of time-local operators andkeeping these components separate leads to the natural formation of an MPS representation for the A tensor givenby:

AjN ...j1(tN ) =∑

α1...αN

[ajN ]αN[ajN−1 ]αN ,αN−1

. . . [aj1 ]α1 . (S13)

where the tensors at the two ends of the chain are rank two, and those in between are rank three. In this notationthe superscripts are the ‘physical’ index which connects to the propagation tensor while the subscripts link adjoininga tensors. At each step of the propagation an SVD is performed on each a and singular values below a fixedprecision χ (relative to the largest) are discarded. This leads to a reduced dimension of the internal indices α anda significant improvement on the exponential scaling present in QUAPI. The truncation acts to remove the leastimportant internal degrees of freedom that are not needed for achieving converged dynamics. This proceedure hasbeen shown to reduce the scaling of the required computational resources to polynomial with memory length forsmooth spectral densities [S5].

Result Time-step, ∆t Precision, χFig. 2 0.2/Ω 10−8

Fig. 4 0.1/ω0 10−7

TABLE I. Convergence parameters used for figures in the main text.

In practice there are two convergence parameters associated with TEMPO which must be adjusted to produce theexact results in the main text:

• The size of the discretization time-step ∆t, decreased until convergence.

Page 9: Exact Quantum Dynamics in Structured Environments · Exact Quantum Dynamics in Structured Environments Dominic Gribben,1 Aidan Strathearn,1,2 Jake Iles-Smith,3 Dainius Kilda,1 Ahsan

3

• The singular value precision χ, increased until convergence.

Further to these, with TEMPO a memory cut-off is also possible such that the growth of the ADT is stoppedonce it covers enough of the system’s history to capture all relevant memory effects. For the results presented inthe main text, however, the long correlation times associated with the structured environments make this kind ofmemory cut-off impossible and as such none were used for generating any of the results. Table I gives the convergenceparameters used for each of the results.

BATH DYNAMICS

In this section we show that it is possible to calculate bath operator expectation values in terms of system dynamicsand specifically present the derivation of Eq. (9) in the main text.

Consider general interaction and bath Hamiltonians given by:

HI(t) +HB = A(t)∑k

(gka†k + g∗kak) +

∑k

ωka†kak (S14)

for a general system operator ˆA(t) in the system interaction picture. The time evolution of the ladder operator ak(t)can be found from the Heisenberg equation of motion:

d

dtak(t) = i[HI(t) +HB , ak(t)] (S15)

= −igkA(t)− iωkak(t) (S16)

eiωktd

dtak(t) + iωke

iωktak(t) = −igkeiωktA(t) (S17)

d

dt

[eiωktak(t)

]= −igkeiωktA(t). (S18)

Formally integrating this and taking the expectation with respect to the total density matrix ρ = ρS ⊗ ρB gives:

〈ak(t)〉 = −igke−iωkt

∫ t

0

dt′eiωkt′〈A(t′)〉, (S19)

〈a†k(t)〉 = ig∗keiωkt

∫ t

0

dt′e−iωkt′〈A(t′)〉. (S20)

Thus the dynamics of any bath mode can be found from knowing the exact system dynamics up to the time t.Calculations of higher order moments are much more involved requiring multiple numerical integrals to be performedand so for now we only consider quantities that can be constructed from the expressions above. The quantity analysedin the main text is given by:

Φ(x, t) =∑k

1√ωk

[〈ak(t)〉+ 〈a†−k(t)〉

]eikx. (S21)

Now for the A = σz coupling from Eq. (5) in the main text we have g∗−k = gk and ωk = |k| = ω−k such that:

Φ(x, t) =∑k

igk√ωk

[−e−iωkt

∫ t

0

dt′eiωkt′〈σz(t′)〉+ eiωkt

∫ t

0

dt′e−iωkt′〈σz(t′)〉

]eikx (S22)

= −∑k

2gk√ωkeikx

∫ t

0

dt′ sin[ωk(t− t′)]〈σz(t′)〉 (S23)

= −4i

∫ ∞−∞

dkgk√ωk

sin

(kR

2

)[cos(kx) + i sin(kx)]

∫ t

0

dt′ sin[ωk(t− t′)]〈σz(t′)〉 (S24)

= 8

∫ ∞0

√J0(ω)

ωsin

(ωR

2

)sin(ωx)

∫ t

0

dt′ sin[ω(t− t′)]〈σz(t′)〉 (S25)

Page 10: Exact Quantum Dynamics in Structured Environments · Exact Quantum Dynamics in Structured Environments Dominic Gribben,1 Aidan Strathearn,1,2 Jake Iles-Smith,3 Dainius Kilda,1 Ahsan

4

where between the last two lines we have discarded the odd component of the integrand and since everything remainingis even we can then halve the integration domain.

[S1] A. Garg, J. N. Onuchic, and V. Ambegaokar, J. Chem. Phys. 83, 4491 (1985).[S2] A. J. Leggett, Phys. Rev. B 30, 1208 (1984).[S3] J. Iles-Smith, N. Lambert, and A. Nazir, Phys. Rev. A 90, 032114 (2014).[S4] H.-P. Breuer and F. Petruccione, The Theory of Open Quantum Systems (Oxford University Press, 2002).[S5] A. Strathearn, P. Kirton, D. Kilda, J. Keeling, and B. W. Lovett, Nat. Commun. 9, 3322 (2018).[S6] N. Makri and D. E. Makarov, J. Chem. Phys. 102, 4600 (1995).[S7] R. Orus, Annals of Physics 349, 117 (2014).[S8] U. Schollwock, Ann. Phys. 326, 96 (2011).