arxiv:2101.07814v1 [cond-mat.stat-mech] 19 jan 2021

17
Symmetry-resolved dynamical purification in synthetic quantum matter Vittorio Vitale, 1, 2, * Andreas Elben, 3, 4, * Richard Kueng, 5 Antoine Neven, 6 Jose Carrasco, 6 Barbara Kraus, 6 Peter Zoller, 3, 4 Pasquale Calabrese, 1, 2, 7 Benoît Vermersch, 3, 4, 8 and Marcello Dalmonte 1, 2 1 The Abdus Salam International Center for Theoretical Physics, Strada Costiera 11, 34151 Trieste, Italy 2 SISSA, via Bonomea 265, 34136 Trieste, Italy 3 Center for Quantum Physics, University of Innsbruck, Innsbruck A-6020, Austria 4 Institute for Quantum Optics and Quantum Information of the Austrian Academy of Sciences, Innsbruck A-6020, Austria 5 Institute for Integrated Circuits, Johannes Kepler University Linz, Altenbergerstrasse 69, 4040 Linz, Austria 6 Institute for Theoretical Physics, University of Innsbruck, A–6020 Innsbruck, Austria 7 INFN, via Bonomea 265, 34136 Trieste, Italy 8 Univ. Grenoble Alpes, CNRS, LPMMC, 38000 Grenoble, France (Dated: 21st January 2021) When a quantum system initialized in a product state is subjected to either coherent or incoher- ent dynamics, the entropy of any of its connected partitions generically increases as a function of time, signalling the inevitable spreading of (quantum) information throughout the system. Here, we show that, in the presence of continuous symmetries and under ubiquitous experimental condi- tions, symmetry-resolved information spreading is inhibited due to the competition of coherent and incoherent dynamics: in given quantum number sectors, entropy decreases as a function of time, signalling dynamical purification. Such dynamical purification bridges between two distinct short and intermediate time regimes, characterized by a log-volume and log-area entropy law, respectively. It is generic to symmetric quantum evolution, and as such occurs for different partition geometry and topology, and classes of (local) Liouville dynamics. We then develop a protocol to measure symmetry-resolved entropies and negativities in synthetic quantum systems based on the random unitary toolbox, and demonstrate the generality of dynamical purification using experimental data from trapped ion experiments [Brydges et al., Science 364, 260 (2019)]. Our work shows that sym- metry plays a key role as a magnifying glass to characterize many-body dynamics in open quantum systems, and, in particular, in noisy-intermediate scale quantum devices. I. INTRODUCTION Symmetry and entanglement represent two corner- stones of our present understanding of many-body quantum systems. The former governs, e.g., the nature of phases of matter 13 , while the latter char- acterizes different classes of quantum dynamics in and out-of-equilibrium 46 . Perhaps surprisingly, the inter- twined role of these two pillars - falling under the um- brella of symmetry-resolved (SR) quantum information - has been relatively unexplored until comparatively re- cently 712 . Such connections are of immediate exper- imental interest in the context of quantum simulation and quantum computing. Aiming at the ultimate goal of engineering perfectly isolated quantum systems, ex- periments in synthetic quantum matter and quantum devices realize system dynamics where coupling to an ex- ternal bath, whatever weak, is ubiquitous - two paradig- matic examples being quantum simulators 13,14 and noisy intermediate-scale quantum (NISQ) devices 15 . In these settings, the microscopic dynamics is local, and is of- ten captured by a master equation with global Abelian symmetries, related to observables such as magnetiza- tion or particle number. Against this background, it is an open question whether SR quantum information can reveal novel, generic classes of quantum dynamics that emerge as a genuine effect of the competition between unitary and incoherent dynamics that is epitomized by quantum simulators and NISQ devices. In this work, we develop a theory and an experimental probe protocol for SR quantum information dynamics in synthetic quantum matter and quantum devices. We are interested in the prototypical scenario depicted in Fig. 1 a-b: an initial product state of a lattice model is subjec- ted to the evolution of a U (1) invariant dynamics, where coherent couplings (J ) are stronger than incoherent ones (γ ). Such scenarios are ubiquitous in current experiment settings. They are realized in analogue quantum sim- ulators as diverse as trapped ions 16 , cold atoms in op- tical lattices 17 , arrays of Rydberg atoms 18 , and circuit quantum-electrodynamics settings 19 . Similarly, the in- terplay between coherent U (1) dynamics and dissipation is of direct relevance to certain nascent quantum com- puters – those that implement two-qubit SWAP or phase gates with a conserved number of qubit excitations. Con- crete examples include architectures based on supercon- ducting qubits 20 and trapped ions 21 . Under these rather ubiquitous conditions, we show that a specific set of SR reduced density matrices (RDMs) undergo dynamical purification as a function of time. This phenomenon is strikingly different from purification to an uncorrelated steady state, because it does not come at the expense of quantum information: entanglement re- mains finite and sizeable over the entire purification dy- namics, both in its generic and symmetry-resolved for- mulation 22 . Furthermore, the scenario we are interested in is fundamentally different from (dissipative) state pre- paration protocols 2325 (see below), as we do not rely on reservoir engineering. arXiv:2101.07814v1 [cond-mat.stat-mech] 19 Jan 2021

Upload: others

Post on 16-Oct-2021

6 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: arXiv:2101.07814v1 [cond-mat.stat-mech] 19 Jan 2021

Symmetry-resolved dynamical purification in synthetic quantum matter

Vittorio Vitale,1, 2, ∗ Andreas Elben,3, 4, ∗ Richard Kueng,5 Antoine Neven,6 Jose Carrasco,6 BarbaraKraus,6 Peter Zoller,3, 4 Pasquale Calabrese,1, 2, 7 Benoît Vermersch,3, 4, 8 and Marcello Dalmonte1, 21The Abdus Salam International Center for Theoretical Physics, Strada Costiera 11, 34151 Trieste, Italy

2SISSA, via Bonomea 265, 34136 Trieste, Italy3Center for Quantum Physics, University of Innsbruck, Innsbruck A-6020, Austria

4Institute for Quantum Optics and Quantum Information of the Austrian Academy of Sciences, Innsbruck A-6020, Austria5Institute for Integrated Circuits, Johannes Kepler University Linz, Altenbergerstrasse 69, 4040 Linz, Austria

6Institute for Theoretical Physics, University of Innsbruck, A–6020 Innsbruck, Austria7INFN, via Bonomea 265, 34136 Trieste, Italy

8Univ. Grenoble Alpes, CNRS, LPMMC, 38000 Grenoble, France(Dated: 21st January 2021)

When a quantum system initialized in a product state is subjected to either coherent or incoher-ent dynamics, the entropy of any of its connected partitions generically increases as a function oftime, signalling the inevitable spreading of (quantum) information throughout the system. Here,we show that, in the presence of continuous symmetries and under ubiquitous experimental condi-tions, symmetry-resolved information spreading is inhibited due to the competition of coherent andincoherent dynamics: in given quantum number sectors, entropy decreases as a function of time,signalling dynamical purification. Such dynamical purification bridges between two distinct shortand intermediate time regimes, characterized by a log-volume and log-area entropy law, respectively.It is generic to symmetric quantum evolution, and as such occurs for different partition geometryand topology, and classes of (local) Liouville dynamics. We then develop a protocol to measuresymmetry-resolved entropies and negativities in synthetic quantum systems based on the randomunitary toolbox, and demonstrate the generality of dynamical purification using experimental datafrom trapped ion experiments [Brydges et al., Science 364, 260 (2019)]. Our work shows that sym-metry plays a key role as a magnifying glass to characterize many-body dynamics in open quantumsystems, and, in particular, in noisy-intermediate scale quantum devices.

I. INTRODUCTION

Symmetry and entanglement represent two corner-stones of our present understanding of many-bodyquantum systems. The former governs, e.g., thenature of phases of matter1–3, while the latter char-acterizes different classes of quantum dynamics in andout-of-equilibrium4–6. Perhaps surprisingly, the inter-twined role of these two pillars - falling under the um-brella of symmetry-resolved (SR) quantum information -has been relatively unexplored until comparatively re-cently7–12. Such connections are of immediate exper-imental interest in the context of quantum simulationand quantum computing. Aiming at the ultimate goalof engineering perfectly isolated quantum systems, ex-periments in synthetic quantum matter and quantumdevices realize system dynamics where coupling to an ex-ternal bath, whatever weak, is ubiquitous - two paradig-matic examples being quantum simulators13,14 and noisyintermediate-scale quantum (NISQ) devices15. In thesesettings, the microscopic dynamics is local, and is of-ten captured by a master equation with global Abeliansymmetries, related to observables such as magnetiza-tion or particle number. Against this background, it isan open question whether SR quantum information canreveal novel, generic classes of quantum dynamics thatemerge as a genuine effect of the competition betweenunitary and incoherent dynamics that is epitomized byquantum simulators and NISQ devices.

In this work, we develop a theory and an experimentalprobe protocol for SR quantum information dynamics insynthetic quantum matter and quantum devices. We areinterested in the prototypical scenario depicted in Fig. 1a-b: an initial product state of a lattice model is subjec-ted to the evolution of a U(1) invariant dynamics, wherecoherent couplings (J) are stronger than incoherent ones(γ). Such scenarios are ubiquitous in current experimentsettings. They are realized in analogue quantum sim-ulators as diverse as trapped ions16, cold atoms in op-tical lattices17, arrays of Rydberg atoms18, and circuitquantum-electrodynamics settings19. Similarly, the in-terplay between coherent U(1) dynamics and dissipationis of direct relevance to certain nascent quantum com-puters – those that implement two-qubit SWAP or phasegates with a conserved number of qubit excitations. Con-crete examples include architectures based on supercon-ducting qubits20 and trapped ions21.

Under these rather ubiquitous conditions, we showthat a specific set of SR reduced density matrices (RDMs)undergo dynamical purification as a function of time.This phenomenon is strikingly different from purificationto an uncorrelated steady state, because it does not comeat the expense of quantum information: entanglement re-mains finite and sizeable over the entire purification dy-namics, both in its generic and symmetry-resolved for-mulation22. Furthermore, the scenario we are interestedin is fundamentally different from (dissipative) state pre-paration protocols23–25 (see below), as we do not rely onreservoir engineering.

arX

iv:2

101.

0781

4v1

[co

nd-m

at.s

tat-

mec

h] 1

9 Ja

n 20

21

Page 2: arXiv:2101.07814v1 [cond-mat.stat-mech] 19 Jan 2021

2

Figure 1. Evolution of symmetry-resolved entropies in NISQ devices. Panel a,b): sketch of the models discussed in the maintext. a): free fermions on a square lattice, with tunneling matrix element J and one-body loss rate γ. b): spin-1/2 chainswith long-range XX exchange interactions, and single site spin relaxation rate γ. In both cases, the grey areas represent thegeometries of the A bipartition of linear length ` considered below. Panel c): time evolution of the symmetry resolved purityin the sector q = ±1, PA(±1) = Tr[ρ2A,±1], in NISQ devices undergoing three distinct regimes (indicated by different colors,see text). The purity initially increases as a function of time, signalling dynamical purification (gray dot). Panel d,e): timeevolution of the symmetry-resolved purity normalized by the partition volume, correspondent to a quantum quench from acharge-density-wave state, and with the dynamics described in panel a,b), respectively. At short times, decoherences induces auniversal scaling behavior, that corresponds to a log-volume entropy scaling, and a purity scaling with inverse of the partitionvolume. Panel f): symmetry-resolved purity for a long-range XX spin chain of L = 10 sites, with ` = 4. The lines representtheoretical simulations, with (solid) and without (dashed) decoherence. Dynamical purification is only present in the first case.Circles represent experimentally reconstructed data from for the symmetry-resolved purity in the trapped ion experiment ofRef. 26. Dynamical purification is experimentally observed for q = −1, and similarly evident for q = 1 (albeit with larger errorbars) in agreement with both theory and numerics. We refer to upcoming work27 for a thorough statistical analysis.

Entanglement-preserving dynamical purification ori-ginates from the competition between coherent and in-coherent dynamics. Indeed, in the presence of only oneof the two, no purification occurs. The effect of such com-petition is inaccessible in the absence of symmetry res-olution, as both coherent and incoherent dynamics gen-erically lead to entropy growth after a quench. From amore practical viewpoint, dynamical purification can beseen as a direct – and universal – signature of a domin-ant coherent dynamics in both quantum simulators andNISQ devices, thus providing a simple proxy to evaluatetheir functioning.

The competition between coherent and incoherent pro-cesses reflects into the existence of two distinct dynam-ical regimes in terms of SR entropy scaling as a func-tion of the partition properties: a short-time, log-volumeregime, where dissipation is the dominant effect, and

an intermediate time, log-area one, where coherent dy-namics partly overcomes dissipation and enhances purityin given quantum number sectors. While distinguish-ing log-volume versus log-area scaling is only meaningfulin one spatial dimension, the corresponding change ofdynamical behavior has dramatic consequences on theexperimentally relevant symmetry-resolved purity: thelatter quantity scales with inverse volume and inversearea (Fig. 1c-e), respectively. Hence it provides an idealproxy to diagnose scaling regimes during dynamical puri-fication. For longer time scales, thermodynamics comesback into the game and all SR-entropies show the stand-ard extensive behavior in subsystem size12.

The interplay between the two regimes can be be illus-trated in context of a simplified Markovian master equa-tion for the SR-RDM: within that framework, the pres-ence of the coherent dynamics interferes with the action

Page 3: arXiv:2101.07814v1 [cond-mat.stat-mech] 19 Jan 2021

3

of dissipation and thus leads to a transient regime whereentropy is soaked out of the SR-RDM itself. We corrob-orate our theoretical framework with numerical simula-tions on a variety of experimentally relevant scenarios.In particular, we showcase the generality of dynamicalpurification by studying both one- and two-dimensionalsystems (some of them depicted in Fig. 1a-b) with parti-tions of different topologies, including both fermionic andbosonic degrees of freedom, and using different types of(weakly-entangled) initial states.

In order to connect our results to experiments, we de-velop a protocol to access SR-RDMs building on the ran-dom measurement toolbox28–34. We show how experi-mentally demonstrated tools allow for accessing SR mo-ments of RDMs and SR Rényi entropies by means of post-selecting data. This procedure is very efficient and allowsto reach system sizes that are considerably beyond whatcan be achieved via full-state tomography, when applic-able (See Ref. 35 for a recent demonstration). We applyour protocol to the trapped ion experiment reported inRef. 26, reconstructing both SR entropies and momentaof the SR-reduced density matrix. The experiment re-veals a sharp dynamical purification (Fig. 1f) which con-firms our theoretical findings. This observation demon-strates the general applicability of our theoretical frame-work, and concretely illustrates the potential of utilizingsymmetry as an enhanced probing tool in state-of-the-artsettings.

The paper is organized as follows. In Sec. II, we setnotations and review symmetry resolved entropies andnegativities. In Sec. III, we specify the time evolution weare interested in, and develop a theory for the time evol-ution of both entropies and negativities in NISQ devices.We illustrate how entropies show distinct scaling beha-vior at short (log-volume) and intermediate (log-area)times, so that SR-purities actually increase as a func-tion of time (dynamical purification). We then arguethat, along this purification, entanglement is typicallypreserved, so that purification does not take place at theexpenses of quantum correlations. In Sec. IV, we presentnumerical results for both spin chains and fermionic sys-tems supporting our theoretical findings. In Sec. V, wediscuss the protocol for the experimental measurementof SR entropies, and present a first application in thecontext of the trapped ion experiment, that supports theobservation of dynamical purification. Finally, we drawour conclusions.

II. SYMMETRY-RESOLVED QUANTUMINFORMATION

In this section, we review definitions and propertiesof symmetry-resolved density matrices and partial trans-poses. Following those, we introduce symmetry-resolvedentropies and negativities, in order to set notation, andbriefly discuss applications of such concepts in closedquantum systems.

A. Symmetry-resolved Renyi entropies

We are interested in bipartite systems, with a partitionA∪B. In the case of a many-body pure state, the bipart-ite entanglement between A and B is fully encoded in thereduced density matrix ρA(ρB) of the given subsystemA(B), and is characterized via n-order Rényi entropies,defined as

S(n)A =

1

1− n log Tr{ρnA}. (1)

For n → 1, these reduce to the renowned von Neumannentanglement entropy

S(ρA) ≡ limn→1

S(n)A = −TrA (ρA log ρA) . (2)

The von Neumann entropy of the reduced density oper-ator is a rigorous entanglement measure for pure states,and the corresponding Rényi entropies with n > 1provide rigorous lower bounds. Both Rényi and von Neu-mann entropies have found widespread applications inthe realm of many-body physics, from the characteriz-ation of topological matter, to dynamics out of equilib-rium, to the understanding of tensor network methods -see, e.g., Ref. 6 for a review.

For a quantum system whose Hamiltonian dynamicspreserves an additive conserved charge, it is possible toidentify and compute the contributions to the entangle-ment related to each symmetry sector7,8,10,11. Here, wefocus on global symmetries.

Let Q denote such a conserved charge (Q = QA ⊗1B + 1A ⊗QB). Then, the reduced density matrix ρA isnecessarily block diagonal and each block corresponds toan eigenvalue q of QA. One can thus introduce Πq, theprojector into the eigenspace related to eigenvalue q, andthe associated density matrix ρA(q)

ρA(q) ≡ ΠqρAΠq

Tr{ρAΠq}, Tr{ρA(q)} = 1, (3)

so that

ρA = ⊕q p(q) ρA(q) (4)

with p(q) = Tr{ρAΠq} the probability of being in chargesector q. We introduce the SR-purity

PA(q) ≡ Tr{ρA(q)2

}. (5)

It quantifies how mixed the state appears in a given sym-metry sector. The symmetry resolved Rényi entropies(SRREs) are a straightforward extension of this concept:

S(n)A (q) ≡ 1

1− n log Tr{ρA(q)n} . (6)

Computing Tr{ρA(q)n} (in cases when a direct applic-ation of projectors in not feasible) requires the knowledgeof the spectral resolution in QA of ρA. As pointed out in

Page 4: arXiv:2101.07814v1 [cond-mat.stat-mech] 19 Jan 2021

4

Refs. 10 and 11, for some of the computations below, itwill be more convenient to study the charged momentsZn(α),

Zn(α) ≡ Tr{ρnAeiαQA

}, (7)

since those do not directly require spectral resolution tostart with. The charged moments have been calculatedin several cases10–12,36–49. Starting from the computationof Zn(α), it is possible to obtain Tr{ρnAΠq} by means ofa Fourier transform:

Tr{ρnAΠq} =

∫ π

−π

dα2πZn(α)e−iαq. (8)

We will exploit this last route in the fermionic simulationsin Sec. IV.

Recent studies have discussed the basic proper-ties of these symmetry-resolved contributions both in-10,11,36–47,50 and out-of-equilibrium12,48, and in presenceof disorder49. In basically all considered cases, it has beenshown that SRREs of large subsystems exhibits entangle-ment equipartition (namely all SRREs are equal) for themost relevant and populated symmetry sectors. The nonequilibrium dynamics of SRREs has been considered onlyfor isolated systems, both after a local48 and a global12quantum quench, and has revealed the presence of a uni-versal time delay for the activation of a given sector12.The investigation of SRREs is far from complete and thecharacterization of its behaviour in the presence of dis-sipation still remains an open question.

B. Symmetry-resolved entanglement negativity

In the case the system S is in a mixed state, the entrop-ies of the RDM are no longer proper measures of bipartiteentanglement, as they are also sensitive to classical cor-relations, although they still provide useful information.A more appropriate and commonly used quantity to wit-ness entanglement in these cases is the negativity51.

Considering S = A∪B, according to Peres’ criterion52,also called positive partial transpose (PPT) criterion, anecessary condition for separablity is that the eigenvaluesλi of its partial transpose ρTA (with respect to subsystemA) are exclusively nonnegative (λi ≥ 0). The entangle-ment negativity

N ≡∑i

max{0,−λi} = 12

(Tr{|ρTA |

}− 1)

(9)

quantifies the degree to which ρTA fails to be positivesemidefinite. So, a non-zero negativity implies the pres-ence of entanglement between A and B. In recent years,the negativity has been extensively studied in a largevariety of physical situation, including critical53–57 anddisordered systems58,59, topological phases60–64, and outof equilibrium65–71. It has been argued that for fermi-onic systems the partial time-reversal transpose is a more

appropriate object to characterise the entanglement inmixed states72–79, although we will not employ such aconcept here.

In analogy to entanglement entropy, one can considerthe negativity for a system possessing some additive con-served charge Q = QA⊗1B+1A⊗QB . Interestingly, ρTA

admits a block diagonal form in the quantum numbers ofthe charge imbalance Q = QA−QTA

B between A and B80.Let Πq denote the projector onto the eigenspace of Q as-sociated with eigenvalue q. We define the normalized SRpartially transposed density matrix80,81

ρTA(q) ≡ ΠqρTAΠq

Tr{ρTAΠq}, Tr

{ρTA(q)

}= 1, (10)

such that

ρTA = ⊕q p(q) ρTA(q) (11)

with p(q) = Tr{ρTAΠq

}≥ 0 the probability of being in

charge imbalance sector q. We can thus define the SRnegativity as

N (q) ≡ Tr{|ρTA(q)|

}− 1

2(12)

with N =∑q p(q)N (q). To compute the SR-negativity,

one needs the spectral resolution of ρTA as in the previouscase. Beyond the case of exact simulations, this challen-ging calculation is performed in two steps. We first focuson the moments Tr

{(ρTA(q))n

}, from which the negativ-

ity is obtained from a replica trick54. Then we considerthe charged moments75,80

Rn(α) ≡ Tr{(ρTA

)neiαQA

}(13)

and performing a Fourier transform we get the desiredTr{

(ρTA(q))n}. This way of performing the calculation

is very powerful when combined with 1 + 1D CFTs54,80,which also provided exact results for the time evolutionof the SR-negativity after a local quantum quench48.

III. TIME-EVOLUTION OF SYMMETRYRESOLVED ENTROPIES AND NEGATIVITIES

In this section, we present a theoretical descriptionof symmetry-resolved quantum information in NISQdevices. We are specifically interested in the short- tointermediate timescales, that is, before dissipation takesover the system dynamics overwhelming coherent effects.

We shall first discuss the generic setting and sub-sequently focus on a specific example that presents thegeneric features we are interested in: the existence of dis-tinct regimes of entropy scaling, dynamical purification,and its interplay with entanglement. While, for the sakeof clarity, most of the technical discussion will be basedon illustrative examples, we point out that our conclu-sions are only relying on very generic conditions, that wenow specify in the next subsection, IIIA.

Page 5: arXiv:2101.07814v1 [cond-mat.stat-mech] 19 Jan 2021

5

A. Short-time dynamics: emergent purification

The system dynamics we are interested in features thefollowing characteristics:

• a D-dimensional system, and a ’convex’ partition Aherein with smooth boundaries82, volume VA andarea ∂VA;

• an initial state |ψ0〉 which is a product state in realspace;

• the full system dynamics shall be described by aGorini-Kossakowski-Sudarshan-Lindblad (GKSL)master equation. In particular, we will be inter-ested in Markovian time-evolution;

• the system Hamiltonian shall have a global sym-metry G. For the sake of simplicity, we will con-sider U(1) below83; most results are immediatelyextended to ZN symmetries, and might also be ap-plicable to the symmetry resolution of continuousnon-Abelian groups when sectors are labelled byAbelian subgroups. We assume local (i.e., nearest-neighbor, one- and two-body) couplings, that arehomogeneous in space. We define as J the en-ergy scale associated to these terms. Below, wewill discuss how sufficiently long-range interactionscan also be included;

• dissipation shall instead be described by local(single-site) jump operators. For the sake of sim-plicity, it is assumed that all sites are affected bythe same dissipative processes. Dissipation shall vi-olate the symmetry G. We define as γ the energyscale associated to these terms, that is, the bare in-verse decay rate. Other sources of dissipation canin principle be introduced: as it will be clear below,we expect that their effects are not particularly in-teresting for the sake of our treatment.

Such assumptions are ubiquitous in the context of syn-thetic quantum systems, such as cold atoms in opticallattices or tweezers, trapped ions, and arrays of super-conducting qubits. Engineering initial states in productstate form (up to initialization errors) is of widespreadpractice, as this can be typically carried out by manip-ulating quantum states locally. The system dynamics isoften local and associated to continuous symmetries, suchas particle number or magnetization conservation. Dis-sipation is generically violating conservation laws asso-ciated to the latter quantities: examples include particleloss in cold atom Hubbard models, and fully depolarizingnoise and spin relaxation in trapped ions and supercon-ducting circuit architectures.

Most of the present experimental settings are able toaccess parameter regimes where dissipation is weakerthan the coherent dynamics, with the ratio γ/J rangingfrom 10−3 to 10−1. We will focus explicitly on this para-meter regime, and consider dissipation as a perturbationon the top of the coherent dynamics.

Under these assumptions, one can identify three times-cales: two intrinsic, and one typical of the subsystemone is interested in. The first one tJ = 1/J is as-sociated to coherent local dynamics. The second onet2 = 1/γ is instead related to a timescale after which (onaverage) all sites within the partition have undergone aquantum jump. The last one, typical of the subsystemA, t1 = 1/(VAγ) is related to the timescale required toobserve a single quantum jump within A.

Let us mention here, that in contrast to the notion ofdissipative state preparation23–25, we study here a givenevolution of a physical system. That is, we are not en-gineering the coupling to the bath to drive the systeminto a desired state, but rather, discuss the dynamicscorresponding to naturally present quantum noise. Inaddition, whereas dissipative state preparation can beutilized to obtain as a unique stationary state a highlyentangled many-body state, or states whose subsystemscan be very pure, the situation we consider here is notrelated to long-time dynamics. We will indeed show thatdynamical purification occurs at intermediate times.

For times t � t2, ρA will be completely mixed, alsoin its symmetry-resolved sectors. For regimes whereγ � J , the system dynamics is dominated by incoher-ent processes. A promising regime to observe competi-tion between coherent and incoherent dynamics is thusVAγ > J > γ, and is the one we will consider below. Weremark that this is a rather generic situation for quantumsimulators of many-body systems, where one tries to real-ize dynamics that are as coherent as possible (J > γ)for large number of degrees of freedom (VA � 1). InSec. IV, we will discuss in more details in which experi-mental platforms such conditions are met.

We emphasize there that the presence of three dynam-ical regimes (that, as we will show below, are captured bydifferent entropy scaling) stems from purely geometricalconsiderations: while coherent dynamics is acting solelyat the boundary, incoherent processes are instead presentover the entire volume of the partition one is interested in.As such, the short-time evolution of symmetry-resolveddensity matrices will be dictated by this competition, andis expected to be largely insensitive to other characterist-ics, including the partition geometry and topology, and(to a weaker extent) the initial state. The theoretical ap-paratus discussed in the next section can be adapted toincorporate such generic features. We nevertheless optedto focus on a simple, yet paradigmatic example, and deferthe demonstration of generality of SR dynamical purific-ation to the numerical experiments discussed in Sec. IV.

1. Explicit example: hard-core Bose-Hubbard model in 2D

For the sake of clarity and to make connections to thenumerical experiments below direct, we start by focus-ing on a specific instance, and return to the general caseat the end of the subsection. We consider a model ofhard-core bosons hopping on an infinite 2D square lat-

Page 6: arXiv:2101.07814v1 [cond-mat.stat-mech] 19 Jan 2021

6

tice, described by the Hamiltonian

H =J

2

∑<i,j>

(b†i bj + h.c.). (14)

Here, bj (b†j) is the bosonic annihilation (creation) oper-ator at site j such that nj = b†jbj gives the number oper-ator for that site. The Hamiltonian dynamics conservesthe total number of bosons, and is thus U(1) invariant.The system time-evolution is described by a master equa-tion:

∂tρ = − i~

[H, ρ]+∑j

γ

[bjρb

†j + b†jρbj −

1

2{bjb†j + nj , ρ}

](15)

where the second term describes single particle loss andgain processes with decay rate γ. The full dynamics isschematically depicted in Fig. 2a. While we will keepgenerality in the theory part with respect to the possibledissipation mechanisms, in the numerical examples be-low, we will only consider loss terms, as those are morereadily accessible experimentally.

We investigate the dynamics starting from a charge-density wave (CDW), with alternating filled (blue) andempty (grey) sites (see Fig.2b). Within this state, weconsider the reduced density matrix ρA correspondingto a rectangular partition A of size Lx × Ly. Let Q =∑j∈A nj − 1

2LxLy the number of bosons in the partitionA above half-filling. Note that, while the full time evolu-tion breaks U(1) invariance, the reduced density matrixρA preserves its block-diagonal form: this is more con-veniently seen when interpreting Eq. (15) as a collectionof quantum trajectories, each corresponding to the solu-tion of a stochastic Schrödinger equation. Within eachtrajectory, the total number of particles at each time t iswell defined: a single quantum jump only changes thatvalue by an integer value. Following the previous subsec-tion, we denote such symmetry-resolved reduced densitymatrices as ρA(q), and express our quantities in ~ = 1units.

We are interested in short time evolution, where dissip-ation and coherent dynamics strongly compete. Specific-ally, we focus on timescales accessible within perturba-tion theory, that is, J2t2, tγ � 1. Therefore, we cansolve Eq. (15) in second order in t to obtain the time-evolved density matrix ρ(t) as a function of the initialstate ρ(0)84. We focus on the q = −1 sector of the RDM,that is, the one where the number of bosons in the parti-tion is decreased by 1 with respect to half-filling. At shorttimes, this is the most populated sector that does con-tribute to the initial state. We will comment on the othersectors below. Within this framework, we assume thatonly the diagonal elements of the reduced density matrixare affected by the time evolution. This assumption canbe proven for initial states that are product states in realspace.

We now divide ρA(−1) into three blocks, schematicallydepicted in Fig. 2:

1. E0(−1): states that are connected to the CDW by asingle hopping process: these states differ from theCDW by a single occupied site at the boundary.We denote the (Lx + Ly) diagonal eigenvalues ofthese states as λE0

k ;

2. E1(−1): states that are connected to the CDW bya single pump process in the bulk; these states dif-fer from the CDW by a single empty site in thebulk. We denote the (Lx − 2)(Ly − 2)/2 diagonaleigenvalues of these states as λE1

k ;

3. E2(−1): states that are not connected to the CDWby a single tunneling or pump process. We de-note the diagonal eigenvalues of these states asλE2

k . For two-body interacting Hamiltonians, thesestates will be accessed only in third order perturb-ation theory.

At second order in perturbation theory (the lowest orderrelevant to the present case), one has the following scalingof the eigenvalues of ρA(−1):

λE0

k = (J2t2 + γt)/A(t), λE1

k = γt/A(t) (16)

with λE2

k = 0, and

A(t) = γt(LxLy − 4)/2 + J2t2(Lx + Ly) (17)

We can now compute the time-evolution of the symmetryresolved entanglement entropies. At short times t < t1,only dissipation is relevant. In particular, the rank of thereduced density matrix will be (LxLy), and the corres-ponding Renyi-2 entropy results:

S(2)A (q = −1) ∝ log[LxLy] for t� t1, Lx, Ly � 1

(18)and is time-independent. The corresponding purity is:

PA(−1) ∝ 1/[LxLy]. (19)

It is worth noting that such ’log-volume’ regime is valid atarbitrarily small times, the simple reason being that theinitial state has no component in the q = −1 subspace.

At intermediate times t1 < t < tJ , tunneling affectsthe system dynamics. While unitary time evolution gen-erically leads to further information scrambling and en-tropy production, here, the opposite takes place: thesymmetry-resolved density matrix purifies as a functionof time, i.e., the purity increases and the entropy de-creases. The reason for this phenomenon stems from thenatural competition between volumetric and perimetralcontributions to the system dynamics: while dissipationhas an effect that scales with the volume of the parti-tion, and thus populates a number of eigenvalues thatare proportional to the volume itself, short-time coherentdynamics is related to boundary effects, and thus favorsa much smaller number of states within the Hilbert spaceof the partition.

Page 7: arXiv:2101.07814v1 [cond-mat.stat-mech] 19 Jan 2021

7

Figure 2. Schematics of the short-time dynamics in lattice models considered here; we show here the sector with q = −1. Panela: the system is defined on a square lattice. The initial state is a charge-density wave |Ψ〉. At short time, the evolution involvesstates belonging to the E0 and E1 subspaces only (see examples, where sites circled in red are the ones changed with respectto |Ψ〉). The influence of the rest of the Hilbert space E2 on the system dynamics is neglected, as accessing these states willrequire at least 3 proceeses starting from |Ψ〉. Panel b): same as in panel a, but for the 1D case. Note that, in the followingsections, several partition topologies will be discussed. Panel c): structure of the time evolved reduced-density matrix. Atshort times, it is further block-diagonal in both E0 and E1, where the E2 sector is traced away.

In order to elucidate this effect, we observe that ourRDM is already normalized, and compute

PA(−1) =t2

A(t)2[(Lx + Ly)(J2t+ γ)2 +

+ (Lx − 2)(Ly − 2)γ2]

that, in the large volume limit, follows a ‘log-area’ scal-ing:

PA(−1) ' 1/(Lx + Ly) for tJ > t > t1

This implies that the transition between the two regimesis characterized by an emergent purification, that transitsthe system from a purity that is inversely proportionalto the volume of the partition, to one that is inverselyproportional to its surface. Note that the explicit timeevolution can be computed from the previous equation,and in principle, the position of the ’maximum’ of thepurification can be extracted. While the correspondingformulas reveal no more physical insight, they signal thefact that the purification time decreases as the partitionsize increases. Due to the condition tJ < t2, it is notpossible to analytically compute the VA → ∞ limit; wenevertheless expect dynamical purification to systemat-ically decrease with the partition size, as a consequenceof the area versus volume competition.

The calculation above can be straightforwardly gen-eralized to any dimension, modulo the conditions men-tioned at the beginning of the section, under the assump-

tion that dynamics is acting non-trivially at the boundary(e.g., a state with a layer of empty sites at the boundarywill not experience any meaningful coherent evolution atshort times in the q = −1 sector). The correspondingscaling behavior decomposes into three regimes:

PA(−1) ∝

1/VA t1 � t ≥ 0 (short time),1/(∂VA) tJ > t > t1 (int. time),1/2VA t� tj (long time).

(20)

This equation succinctly describes the dynamical scalingregimes depicted in Fig. 1. Starting from an unsurprisingshort time behavior (top case), the system purifies atintermediate time scales (center case) before eventuallygetting fully mixed (bottom case) due to both coherentand incoherent system dynamics.

While we have focused on the most populated sector ofthe RDM not present in the initial state, we expect dy-namical purification to occur also in other sectors - with,however, a weaker effect due to higher order perturbativeprocesses. The presence of long-range interactions thatdecay fast enough (at most as power law) shall not changethis picture at the qualitative level: however, it will leadto a renormalization of the timescale tJ . Importantly,long-range interactions will not modify the structure ofthe Hilbert subspaces discussed above.

While we have focused on purities, additional inform-ation can in principle be obtained from the population of

Page 8: arXiv:2101.07814v1 [cond-mat.stat-mech] 19 Jan 2021

8

the different sectors (denoted with A(t) above) as well.One example is equilibration at long-times: this is bey-ond the perturbative treatment we have developed, andwill be discussed in the next sections in both simulationsand experiments.

The treatment above is specific to an initial state: how-ever, the competition between volumetric and perimetralcontributions is in fact generic to a much broader set ofexperimentally relevant configurations. For the case ofpure states, dynamical purification shall occur as long asthe initial state is separable or weakly entangled, as weshow in one of the fermionic examples below. For highlyentangled initial states, the theory above is not immedi-ately applicable. Below, we will discuss a 1D numericalexample, where the initial state has log(`) entanglement:in that case, we observe no purification.

Most importantly, dynamical purification is presentalso for initial states that are globally mixed. In thosecases, this is simply due to the fact that the coherentdynamics selects a subset of states in ρA(−1) that arepopulated due to coupling to A. The extent of the dy-namical purification depends on the details of the actionof the Hamiltonian on the initial state: we will investig-ate a specific scenario below while discussing trapped ionexperiments.

Finally, we would like to comment on which type ofnoise leads to dynamical purification. The noise we haveconsidered here has two characteristics: (i) it is describedby a Markovian master equation, and (2) it is quantumnoise and described by jump operators are not Hermitian.While these conditions are typically very well satisfiedwhen describing the dynamics of cold atoms in opticallattices using a master equation, we find useful to providea short discussion of these two elements in view of pos-sible applications to other settings.

The first assumption above is delicate. Since we areinterested in intermediate time dynamics, it is reason-able to expect that our findings will not be affected by abath featuring short-lived memory effects, as long as theweak system-bath approximation (that we neverthelessconsider, since γ � J) holds. However, more complic-ated bath structures including strong memory effects -such as a low-temperature Ohmic bath - cannot be im-mediately connected to the physical picture we presenthere. We leave this interesting question - that does notpertain the experimental systems we are interested in -to future work.

The second assumption above is crucial: hermitianjump operators (such as those, for instance, describingclassical noise) would not lead to any dynamical purific-ation. This can be easily seen by considering the action ofdephasing on the various sectors of the SR RDM: for thetype of initial states we consider, the latter will not af-fect populations. This implies that entropy will be dom-inated by coherent dynamics, thus increasing with time.The relevance of the first assumption can potentially beexploited as a diagnostic in the context of quantum noisetomography; interestingly enough, such a probe would be

very sensitive, as the effects we describe can be presentfor very small values of γ, and can be tuned by changingthe volume of the partition.

B. Negativity over dynamical purification

While the system purifies at short time, due to itscoupling to the environment, it is not clear a prioriwhether this is associated to a loss of shared entangle-ment between the partition and its complement. Forinstance, dynamical purification (with or without sym-metry resolution) can also occur at long times in systemsunder the presence of dissipation only: a typical exampleis relaxation to a vacuum state, that is driven by a singlejump operator, and leads to a trivial state, with no left-over correlations between A and B, and within A. Be-low, we show explicitly how symmetry-resolved dynam-ical purification is drastically distinct from this mechan-ism: in particular, we show how not only entanglementbetween A and B is generated as a function of time, butalso that, in any given symmetry sectors (now labeledby quantum number differences), entanglement remainsfinite and sizeable (negativity of order 1) over the entirepurification process. This is a key element that charac-terizes this symmetry-resolved phenomena, and we willshow below how this is also captured within perturbationtheory.

We study the entanglement dynamics for two connec-ted partitions A and B of a spin (or hard-core boson)system, as governed by Eq. (15), in a regime where thepartition A undergoes dynamical purification. For thesake of simplicity, we will deal explicitly with the 1D caseanalog to the setup described above (see Fig. 2b), andrestrict the decoherence channels to particle loss, as thiswill allow us to keep our notations compact. Our findingsare however general, as illustrated in the next section forvarious geometries and partition configurations.

Before discussing entanglement scaling, we find it use-ful to point out that the effective dynamics describing theevolution of ρA can be interpreted as the time evolutionof a density matrix of a system coupled to a bath. Thisprovides an additional viewpoint on the phenomenon weare interested in, that could be of help to translate it toother contexts (for instance, in case the two partitions aremade of two different types of degrees of freedom, e.g.,describing light-matter interactions). A detailed discus-sion of this fact is provided in the appendix, togetherwith a proof of the fact that such effective dynamics isMarkovian.

In contrast to the situation of dynamical purification,the key features of short-time entanglement dynamics ofthe partial transpose RDM can already be captured bysolving Eq. (15) in first-order perturbation theory, i.e.by studying the dynamics of ρ(t) in first order in t �

Page 9: arXiv:2101.07814v1 [cond-mat.stat-mech] 19 Jan 2021

9

1/(γN), 1/J (� 1/γ). We rewrite this as

ρ(t) = ρ(0)− i[H, ρ(0)]t

+ γt∑j

(bjρ(0)b†j − 1

2b†jbjρ0 − 1

2ρ0b†jbj

). (21)

Consider for concreteness that the even sites 2m, m =1, . . . , N/2 are occupied and, NA = NB is even, the dens-ity matrix in first-order perturbation theory can be re-expressed as

ρ(t) =

(1− Nγt

2

)ρ(0) + Jt(−ib†NA+1bNA

ρ(0) + h.c)

+ γt

N/2∑m=1

b2mρ(0)b†2m + ... (22)

which corresponds to a diagonal part parametrized bythe decoherence rate γ, and a pair of off-diagonal ele-ments associated with the hopping J . Note that thereis no diagonal contribution due to the hopping, as thisonly appears in next-to-leading order as discussed above.Taking the partial transpose of Eq.(22) leads to

ρTA(t) =

(1− Nγt

2

)ρ(0) + Jt(−ib†NA+1ρ(0)b†NA

+ h.c)

+ γt

N/2∑m=1

b2mρ(0)b†2m, (23)

which has a 3-block structure associated with thequantum number q = qA − qB

ρTA(q = 0, t) =

(1− Nγt

2

)ρ(0)

ρTA(q = −1, t) = γt

NA/2∑m=1

b2mρ(0)b†2m +

+ Jt(−ib†NA+1ρ(0)b†NA+ h.c)

ρTA(q = 1, t) = γt

N∑m=NA/2+1

b2mρ(0)b†2m.

The sector q = 0 corresponds to the initial state com-ponent, has a weight tr(ρTA

q=0(t)) of order 1, and featuresno entanglement. The sector q = −1, corresponding tothe situation where the A partition loses one excitationw.r.t. partition B, has the richest structure, representingthe interplay between particle loss from A and coherencedynamics (hopping from A to B). Finally, the last sectorq = 1 represent decoherence events occurring in the Bpartition. In each sector, we can calculate the spectrum

λ(q = 0, t) =

(1− Nγt

2

)(24)

λ(m=1,...,NA2 −1)(q = −1, t) = γt (25)

λ(m=NA2 ,

NA2 +1)(q = −1, t) = (γ ±

√γ2 + 4J2)

t

2

≈ (γ

2± J)t (26)

λ(m=1,...,NB2 )(q = 1, t) = γt, (27)

0 1 2 3 4 5Jt

0.2

0.4

0.6

0.8

1.0

P A(°

1)

∞/J

0.00.010.020.04

0.080.160.320.64

0 1 2 3 4 5Jt

0

20

40

N(°

1) 0.01 0.04 0.16 0.64∞/J

100101102

ª 2JNA∞

0 1 2 3 4 5Jt

0.0

0.5

1.0

P A(q

)

∞/J = 0.08, q0 °1 Total

0 1 2 3 4 5Jt

0

2

4

6

N(q

)

∞/J = 0.08, q

0 °1 °2 Total

a) b)

c) d)

Figure 3. Dynamical purification and symmetry-resolved en-tanglement for one-dimensional XX spin models. We choosea system with N = 8, initialized in a Néel state |↑↓〉⊗N/2 andevolved with HXX subject to particle loss with rate γ (seemain text). We take A = [1, 2, 3, 4] and B = [5, 6, 7, 8]. Inpanels (a,c), the SR-purity of subsystem A is shown. Blackdashed lines are predictions from early-time perturbation the-ory. In panels (b,d), the normalized SR negativity N (q) isshown for various imbalance sectors q (b) and decoherencerates γ (d). The inset shows the early time value N (−1)|t=0+

as function of γ/J . Gray lines are results from perturbationtheory, Eqs. (B3) and (28), respectively.

where, in the last part of the third line, we have neglectedthe term γ2 � J2.

The existence of a negative eigenvalue λ(m=NA/2)(q =−1) = −Jt+γt/2 ≈ −Jt < 0 demonstrates that the stateis entangled, and remains so over dynamical purification.After normalization, we obtain the SR negativity

N (q = −1) ≈ 2Jt

NAγt=

2J

NAγ(28)

that features a characteristic 1/γ scaling (that is remin-iscent of the fact that this is a perturbative effect).

IV. NUMERICAL RESULTS

A. Spin chains

In this section, we provide numerical evidence forsymmetry-resolved purification in one-dimensional spinchains. Specifically, we consider quench dynamics in theXX-model with Hamiltonian

HXX = J

N−1∑i=1

σ+i σ−i+1 + h.c. (29)

subject to spin excitation loss with rate γ, modeled viathe jump operators √γσ−i for i = 1, . . . , N . We initial-ize the system with N = 8 sites, divided into subsys-tems A = [1, 2, 3, 4] and B = [5, 6, 7, 8], in the Néel state

Page 10: arXiv:2101.07814v1 [cond-mat.stat-mech] 19 Jan 2021

10

|Ψ0〉 = |↑↓〉⊗N/2 with total magnization Sz =∑Ni=1 σ

zi =

0. While the total magnetization is conserved by theHamiltonian part of the dynamics HXX , the incoherentspin excitation loss leads to a population of various sec-tors.

In Fig. 3 a), we display the symmetry resolved purityof the subsystem A with NA = 4 sites for various sectorsq. Clearly, the sector q = −1 exhibits dynamical purific-ation at times Jt ≈ 1 which is absent in the sector q = 0and also for the purity tr

[ρ2A]of the total density matrix

ρA. Note that the second peak in panel (a) (at aroundJt = 4) is a boundary effect due to the partition size. Aspredicted by perturbation theory [Eq. (20)], purificationis pronounced most strongly for weak decoherence [seeFig. 3c)]. While the initial values PA(1)|t=0+ = 2/NAis independent of γ, the peak of the purity is approach-ing the value of the purity for unitary dynamics. On thecontrary, for γ & J , the dynamics is dominated by deco-herence, and purification is absent. In Fig. 3 (b,d), weshow the SR negativity N (q). We observe that SR entan-glement between A and B is dominated by the magnet-ization imbalance sector q = −1 sector. The magnitudeof the negativity of sector q is much larger than the totalsystem negativity. In addition, as shown in the inset, theearly time value at Jt = 0+ is decreasing as ∼ 1/γ withincreasing decoherence rate γ, as predicted by perturba-tion theory [Eq.(28)].Experimental setups. - The dynamics discussed in

this subsection is relevant for a variety of setups. Inthe next section, we will discuss and demonstrate im-plementation with trapped ions in Paul traps. Anothernatural setting is Rydberg atoms in optical tweezers oroptical lattices. Within those, the dipolar version of theXX Hamiltonian in Eq. (29) is naturally realized whenconsidering direct dipole-dipole interactions within theRydberg manifold (for a many-body demonstration, seeRef. 85). Spin excitation losses occur naturally, and canbe further enhanced via incoherently coupling the twoRydberg states. A very similar scenario (dipolar coup-lings) is also realized with superconducting qubits in 3Dcavities, and with polar molecules or magnetic atoms inoptical lattices.

B. Fermionic systems in 1D and 2D

We now provide numerical evidences of the physics de-scribed in the previous sections also using free fermionictechniques 86,87. The latter allow us to consider largersystem sizes and two-dimensional geometries. Most im-portantly, it allows us to check systematically specific fea-tures of our predictions, such as the dependence on thepartition size, dimensionality, and topology of the parti-tion (e.g.: in 1D, we will consider explicitly disconnectedpartitions).

We start from a charge-density-wave (half-filling), andlet it evolve according to a GKSL master equation master

with jump operator lj = γcj and Hamiltonian:

H = −J∑〈i,j〉

c†i cj − 2µ∑j

(c†jcj −

1

2

). (30)

The first sum runs over nearest neighbours, c†i , ci denotefermionic creation/annihilation operators, J is the hop-ping constant (that we set to unity below, J = 1) andµ is the chemical potential (µ = 0 unless stated other-wise). In free fermionic theories, at each time t one cancompute the charged-moments Zn(α) (Eq. (7)) via thetwo-point fermionic correlation matrix Cij = 〈c†i cj〉 andits evolution according to Ref. 88. We consider both 1Dchains and 2D square lattices and check numerically theanalytical predictions in the previous sections. In 1D,the tight-binding model is mapped to the XX Hamilto-nian (29) by a Jordan-Wigner transformation: the GKSLmaster equation we will consider are similar to one of theexamples discussed in Ref. 89.

In Fig. 4, we show some representative numerical res-ults. In panels a)-b) we consider PA(q), cf. Eq. (5),in 1D. The system is divided into three parts as S =A ∪B ∪A with |A| = `/2 and ` = L/2, a representationof the system is in Fig 4a). The choice of the topologyof the partition allows us to illustrate the generality ofdynamical purification, that is indeed topology independ-ent as long as `� tJ . In panel c) we compute the samequantity for a two-dimensional square lattice to highlightthat the features of the dynamics are not dependent onthe dimensionality or connectivity of the partition. Herewe consider S = A ∪ B where A is a square of lineardimension ` = L/4 at the center of the system. In pan-els d)-e)-f) we focus on the behaviour of the SR-purityin the absence of dissipation, to emphasize that the bathplays a decisive role in the dynamical purification, and onquenches starting from different states, since we expectour results to hold when the initial state is separable (cf.III A). The initial state being at half-filling, q = `/2 isthe only populated sector at t = 0. We will consider thequantity PA(q) where q is shifted by a constant `

2 , e.g.q = 1 refers to the sector `/2 + 1, i.e., one particle morethan half-filling. We omit `/2 in PA(q) to be concise.In all the following simulations we always consider openboundary conditions (OBC).

In Fig. 4a), we show PA(q) for q = 0, 1, 2, 3, L = 128.The sector q = 0 is the only one occupied at t = 0. Itis pure at the start of the evolution and does not exper-ience any purification. Oppositely, as soon as dynamicskicks in, the sector q = 1 becomes mixed. Its purityincreases at intermediate times (dynamical purification)and approaches equipartition for longer times. This ishighlighted in the inset showing the behaviour of PA(q)for Jt ∈ [1, 5] in logarithmic scale. The purification forthe other sectors is present, but less evident as it is con-nected to higher-order perturbative processes.

In Fig. 4b) we fix q = 1 and consider PA for differ-ent values of ` with L = 2`. In agreement with theory,the peak of the curves decreases, approaching zero. The

Page 11: arXiv:2101.07814v1 [cond-mat.stat-mech] 19 Jan 2021

11

0.0 0.5 1.0 1.5Jt

0.0

0.2

0.4

P A(1

)

` = 16` = 32

` = 64` = 128

101 102

`

10−1

PA(1, t = 0+)

b)

0 1 2Jt

0.0

0.5

1.0

P A(q

)

q = 0q = 1q = 2

0 2Jt

0

1

TrρA

(q)

d)

0 1 2Jt

0.0

0.1

0.2

P A(1

)

` = 16` = 32

` = 64` = 128

102

`

10−1

PA(1, t = 0+)

e)

0.0 0.5 1.0 1.5Jt

0.0

0.5

P A(1

)

` = 16` = 32

` = 64` = 128

0 1 2Jt

0.25

0.50

TrρA

(q) -q = 0 -q = 1

-q = 2

f)

Figure 4. Results of the simulation of PA(`/2 + q) for a quadratic open fermionic system. We omit `/2 and use q = q − `/2to label the symmetry sectors. Parameters: J = 1, µ = 0. First line: Symmetry resolved Rényi entropy for a) 1D system withL = 64, l = 32, γ = 0.05; b) 1D system for L = 2`, q = 1, γ = 0.05; c) 2D system with L = 4`, N = L2, q = 1, γ = 0.2. Secondline: Symmetry resolved Rényi entropy for d) 1D system with L = 64, l = 16, γ = 0, purely coherent dynamics; e) 1D systemfor L = 2`, q = 1, γ = 0.05, starting from the Majumdar-Ghosh dimer state; f)1D system for L = 2`, q = 1, γ = 0.05, startingfrom the ground state of a nearest-neighbours tight binding model with J = 1.

point at t = 0+ should approach zero as ∼ 1/`, as well,like it has been anticipated in the previous sections. Theinset shows a fit of PA(q = 1, t = 0+) as a function of `,which demonstrates the log-volume regime already dis-cussed.

The behaviour of the SR-purity for a two-dimensionalsystems is analogous. In Fig. 4c) we plot the purity, atfixed q = 1, for different values of L. The total numberof sites of the lattice is N = L2 and the subsystem Aconsists of l2 = N/16 sites picked at the center of thesquare. Studying the position of the point at t = 0+

one observes that it scales as ∼ 1/`2 as calculated inEq. (19) and shown in Fig. 1d): this confirms a 2D log-volume scaling at short times, with the correspondingSRRE displayed in the inset for the sake of completeness.

In Fig. 4d), we take into account the SR-purity in thecase of a purely coherent dynamics. This show remark-ably how the purification process is strictly related to thepresence of a bath for this class of models. We considerL = 64 and ` = 32. While q 6= 0 sectors are mixed attime t = 0+ in presence of bath, this is not the case forγ = 0. In the inset one can see the population of eachgiven sector as a function of time. As the coherent dy-namics starts playing its role, the population increasesand the purity decreases correspondingly. The q 6= 0sectors are involved in the evolution but they do not ex-perience any purification, instead their purity decreasesmonotonously to a unique value independent of q, wit-nessing information equipartition. All these results arecompatible with the exact ones reported in Ref. 12.

Finally, in Fig. 4e-f), we depict the SR-purity in thesector q = 1, in the case of a global quench starting fromtwo different states. Firstly, in e), we consider a globalquench from the Majumdar–Ghosh dimer product stateand an evolution under the Hamiltonian in Eq. (30);secondly, in f), we take as starting point the ground stateof Eq. (30) and evolve the system according to a longrange hopping Hamiltonian in the form:

H = −∑ij

J

|i− j|α c†i cj , (31)

where α = 2.The purpose of panels e) and f) is to show that the dy-

namical purification is present only in the case the initialstate is separable, as it happens for Fig. 4b)-e). In theinset of Fig. 4e) we show a fit of PA(q = 1, t = 0+) asa function of `, which exhibits a 1/` behaviour, as pre-dicted by perturbation theory. Oppositely, if the initialstate is entangled, one cannot see any emergent purific-ation during the dynamics (Fig. 4f)). This is due to thefact that the SR RDM is already mixed in all Ek sectors,and thus, local coherent dynamics is insufficient to purifythe state, as the number of non-zero eigenvalues in eachsector is exponentially large in the partition size. In theinset of the figure the populations of sectors q = 0, 1, 2for L = 256 are shown. Evidently, all the sectors areoccupied already at t = 0.Experimental setups. - The U(1) dynamics discussed

in this section is of direct relevance for various experi-

Page 12: arXiv:2101.07814v1 [cond-mat.stat-mech] 19 Jan 2021

12

mental settings. The first ones are fermionic or (hard-core) bosonic atoms trapped into optical lattices. There,one of the main sources of dissipation (in addition tospontaneous emission, that can be made small with theuse of blue detuned lattices) is single particle loss. Whilein principle loss rates due to inelastic background scat-tering are small when compared to the typical lattice dy-namics, localized losses can be engineered in a variety ofways, including weak-laser coupling to untrapped levelsor via electron beams.

The second setting that is relevant to this subsectionare arrays of superconducting qubits. In the strong coup-ling limit, the dynamics of such systems can be well ap-proximated by an XX model. Qubit relaxation will thenplay the same role as single particle loss.

V. EXPERIMENTAL PROTOCOL FORMEASURING SYMMETRY-RESOLVED

PURITIES

Our protocol to extract symmetry-resolved purities isbased on randomized measurements. These methodshave been introduced and experimentally demonstratedto measure entanglement entropies26,29,30,32, and othernonlinear functions of the density matrix, such as statefidelities90, out-of-time ordered correlators91,92, topolo-gical invariants93,94, and entanglement negativity34,95. Inthe quantum information context, the moments of stat-istical correlations between randomized measurementscan also be used to define powerful entanglement wit-nesses without reference frames34,96–100.

While standard projective measurements performed ina fixed basis can only give access to expectation val-ues of a particular observable, randomized measurementsconsist instead in measuring our quantum state in dif-ferent random bases, giving access to complicated non-linear functionals of the density matrix, here symmetry-resolved purities.

As in Refs. 32 and 34, our approach is based on theidea of combining two results: randomized measurementtomography28,101, and ‘shadow’ tomography32,102. Letus consider here a spin system and show how to measurethe symmetry-resolved purity of a reduced state ρA madeof NA spins.

Randomized measurements are realized by applyingrandom local unitaries ρA → uρAu

†, u = u1⊗ · · · ⊗ uNA,

where each ui is a spin rotation that is taken, independ-ently, from a unitary 3-design103,104. After the applica-tion of random unitary, a projective measurement is real-ized in a fixed basis. This procedure is repeated with Mdifferent random unitaries, in order to obtain a list of Mmeasured bitstrings k(r), r = 1, . . . ,M .

Randomized measurementsare tomographically com-plete in expectation and can be used to provide an es-timator of the density matrix28,32,33,101,105

ρ(r)A =

⊗i∈A

[3(u

(r)i )†

∣∣∣k(r)i ⟩⟨k(r)i ∣∣∣u(r)i − I2], (32)

with the expectation value over randomized measure-ments E[ρ

(r)A ] = ρA. It is not our aim to reconstruct the

density matrix based on Eq. (32) i.e., to perform tomo-graphy, as it will be too costly in terms of measurements(and classical post-processing). However, we can makeuse of this expression Eq. (32), in order to relate dir-ectly polynomial functionals of ρ to the measured datak(r)32. For the symmetry-resolved purity, simply con-sider two independent randomized measurements r 6= r′,and define the symmetrized estimator

PA(q)(r,r′) = 1

2 tr[(ρ(r)A Πq)(ρ

(r′)A Πq)]

+ 12 tr[(ρ

(r′)A Πq)(ρ

(r)A Πq)]. (33)

Using Eq. (32), this can be seen as a simple bi-linear func-tion of the measurement data. Averaging over many pairs(r, r′), boosts convergence to the estimator’s expectationvalue

E[PA(q)(r,r′)] = 1

2 tr[(E[ρ(r)A ]Πq)(E[ρ

(r′)A ]Πq)]

+ 12 tr[(E[ρ

(r′)A ]Πq)(E[ρ

(r)A ]Πq)] = PA(q).

Here, we have used that ρ(r)A and ρ(r′)A are independ-

ent realizations of Eq. (32). This means that PA(q)(r,r′)

is an unbiased estimator of the symmetry-resolved pur-ity. This procedure can be straightforwardly extendedto higher moments with triplets of randomized meas-urements r 6= r′ 6= r′′, etc. Appropriate implementa-tion of partial transposes moreover allows for extractingSR Rényi entropies (6). This is the content of upcom-ing work27, where we also provide a thorough statist-ical analysis for estimating SR-resolved quantities basedon randomized measurements. The upshot is that es-timator (33) can be equipped with rigorous confidencebounds. Already 2NAPA(q)/ε2 measurement repetitionssuffice to estimate a given SR purity PA(q) up to accur-acy ε. This favorable scaling is a key advantage over fullquantum state tomography – especially if PA(q) is itselfvery small. Our analysis of experimental data, c.f. nextsection, support this favorable picture.

Therefore, we believe that SR-purities can be meas-ured in various NISQ platforms up to moderate parti-tion sizes NA = 10 ∼ 20, which are sufficient large toobserve many-body effects, such as dynamical purifica-tion. The second advantage of randomized measurementswith respect to tomographic-type estimations is the post-processing step. Here, the estimation of PA(q)(r,r

′) fromthe data simply consists in multiplying estimators ρ(r)Awith an efficient tensor-product representation (Eq. (32))with a projection operator with sparse-matrix structure(which can be for instance efficiently written as a Matrix-Product-Operator106).

Note finally that here randomized unitaries do nothave a symmetric structure, and therefore each estima-tion of the density matrix does not have a block-diagonalstructure. Alternatively, one can envision to performsymmetry-resolved random unitaries incorporating sym-metries28,30,31. While these random unitaries appear as

Page 13: arXiv:2101.07814v1 [cond-mat.stat-mech] 19 Jan 2021

13

0.0 2.5 5.0t(ms)

0

1

2

3

S(2

)A

(q)

Total

0.0 2.5 5.0t(ms)

0.0

0.5

1.0Tr[

¶qΩ

A]

q

°1 0 1

b)a)

Figure 5. Experimental demonstration of symmetry resolvedpurification and entanglement in a trapped ion quantum sim-ulators, using data obtained in the context of Ref. 26. We con-sider a system of N = 10 spins, with subsystems A = [4, 5, 67]and B = [1, 2, 3, 8, 9, 10]. In panels a) and b), the symmetryresolved populations and Rényi entropies of various magnet-ization sectors q = 0,±1 of the reduced density matrix ρA areshown as function of time (see Fig. 1 for symmetry resolvedpurities). Error bars have been calculated with Jackknife res-ampling. In panel b), data for the magnetization sector q = 1at Jt = 0 has been omitted due to large errorbars, resultingfrom small populations. Lines are numerical simulations ofunitary dynamics (dashed) and including decoherence (solid)as decribed in the text.

more challenging to realize experimentally compared tolocal spin rotations, one should expect a reduction ofstatistical errors in this situation101.

VI. EXPERIMENTAL OBSERVATION OFDYNAMICAL PURIFICATION IN TRAPPED

ION CHAINS

In this section, we demonstrate symmetry resolvedpurification experimentally in a trapped ion quantumsimulator, using data taken in the context of Ref.26.Here, quench dynamics with a long-range XX-model wererealized, governed by the Hamiltonian

HXX = ~∑i<j

Jij(σ+i σ−j + σ−i σ

+j ) + ~B

∑i

σzi (34)

with σzi the third spin-1/2 Pauli operator, σ+i (σ−i ) the

spin-raising (lowering) operators acting on spin i, andJij ≈ J0/|i− j|α the coupling matrix with an approx-imate power-law decay α ≈ 1.24 and J0 = 420s−1.The effective magnetic field is taken to be large B ≈2π · 1.5kHz ≈ 22J0 such that the unitary dynamics con-serve the total magnetization Sz =

∑i σ

zi .

In addition, decoherence is present in the experiment,during initial state preparation, time evolution and therandomized measurement. As detailed in Ref. 26, we canmodel these decoherence effects as follows.

The time evolution is subject to local spin-flips, andspin excitation loss (spontaneous decay). We describethe corresponding dynamics with a master equation withjump operators Ci =

√γFσ

xi for i = 1, . . . , N and

Ci+N =√γDσ

−i for i = 1, . . . , N , capturing the spin

flip and excitation loss, respectively. Here, the rates areγF ≈ γD ≈ 0.7/s.

In the experiments, the initial state is not pure,but rather it is a mixed product state ρ0 =⊗

i (pi |↑〉 〈↑|+ (1− pi) |↓〉 〈↓|) with pi ≈ 0.004 for i evenand pi ≈ 0.995 for i odd. Finally, during the applicationof the local random unitary, local depolarizing noise isacting which is modeled as

ρ(tfinal)

→ (1− pDPN)ρ(tfinal) + pDP∑i

Tri[ρ(tfinal)]⊗1i2

(35)

with pDP ≈ 0.02.In Fig. 5, we present experimental results, obtained

with the estimators defined in Eq. (33), and numericalsimulations, for unitary dynamics and including the de-coherence model described above. In panel a), the pop-ulations tr{ΠqρA} of the magnetization sectors q of thereduced density matrix ρA are shown, with A consistingof spins A = [4, 5, 6, 7]. Initially, the (q = 0)-sector ispredominantly populated, with small fractions in othersectors, due to the finite initial state preparation fidelity.With time, the population in other sectors, in particularq = ±1, increases. The symmetry resolved second Rényientropy S(2)

A (q) is shown in panel b) for various magnetiz-ation sectors (see Fig. 1 for the corresponding symmetryresolved purity). The experimental data clearly showsdynamical purification (decrease of the Rényi entropy)in the q = −1 sector. Data in the q = +1 sector arealso suggestive of dynamical purification, even if a strongstatement cannot me made here do to comparatively lar-ger error bars. In particular, this demonstrates that dy-namical purification can be observed in one-dimensionalsystems with algebraically decaying long-range interac-tions (see also Sec. IIIA). At long times, the symmetryresolved Rényi entropies approach similar values for alldisplayed sectors, consistent with expected equipartitionof the symmetry sectors10.

VII. CONCLUSIONS AND OUTLOOK

Symmetry is an ubiquitous element characterising syn-thetic quantum matter - from quantum simulators, tonoisy-intermediate scale quantum devices. In this work,we have developed a theoretical framework for the de-scription of symmetry-resolved information spreading insuch open quantum systems, focusing on the epitomecase of U(1) symmetries common to several experimentalplatforms - from cold gases in optical lattices, to trappedions and superconducting circuits. We have shown how,for various settings, specific quantum number sectors un-dergo dynamical purification under ubiquitous conditionsof weak noise and separable initial states. Such phe-nomenology is general, occurs in any dimension, is notsensitive to the partition topology, and features specific

Page 14: arXiv:2101.07814v1 [cond-mat.stat-mech] 19 Jan 2021

14

scaling scenarios for the entropy as a function of parti-tion size. Most importantly, the dynamical purificationconsidered here occurs in symmetric systems and stemsfrom the competition between coherent and incoherentdynamics that is a leitmotif of current NISQ devices.

We have introduced and experimentally demonstrateda protocol to measure symmetry-resolved quantum in-formation quantities based on a combination of random-ized measurement probing and shadow tomography. Ourapproach is scalable to partition sizes that are well bey-ond what is accessible to full state tomography, and isapplicable to a broad spectrum of experimental settingswith single site control and high repetition rate. Bothscalability and applicability are of key importance in or-der to probe genuinely many-body features of entangle-ment dynamics in state-of-the-art experiments. Basedon our protocol, we have shown how the experimentsperformed in Ref. 26 have already realized dynamicalpurification in a trapped ion chain described by a long-range XX model. This observation, in full agreementwith our theory predictions, testifies for the generality ofsymmetry-resolved dynamical purification under experi-mentally realistic conditions.

The capability of addressing the combined role of sym-metry and quantum correlations in NISQ devices opensa novel interface between theory and experiments, wheremany-body effects intertwine with information theoreticapplications. The first instance of that is what role sym-metry plays in quantum information protocols, in par-ticular, error correction. Our tools may be of particularimportance here, as several error correcting codes canbe cast as gauge theories, one example being the toriccode107. In this context, the role of specific symmetrysectors is associated to the presence of excitations. Itmay thus be useful to employ the experimental tools wehave used here to access how specific perturbations accessthe reliability of a quantum memory. Going beyond that,understanding whether dynamical purification occurs inthe presence of local symmetries is an open question, thatcould be in principle addressed within the same methodspresented here.

The second possible applications of our methods con-cerns the capability of utilizing dynamical purification asa proxy of the system dynamics, in particular, to determ-ine its dissipative dynamics. One first element is thatdynamical purification is expected for a quantum noise,that is local: it is thus informative about the nature ofthe dissipation. The fact that the dissipation rates inter-twines with the partition size could also help to quantifythe relative strength of incoherent versus coherent pro-cesses, at least in cases where specific initial states couldbe realized with high fidelity. Remarkably, despite beinga short-to-intermediate time phenomenon, thanks to thearea-to-volume ratio being tunable, dynamical purifica-tion is also informative about very weak dissipation: thisis particularly important for diagnostics, as one wouldexpect that the latter requires long-time evolution to becharacterized.

On more general grounds, symmetry-resolved dynam-ical purification reveals how certain many-body phenom-ena can only be properly characterized utilizing sym-metry to emphasize or even magnify relevant informa-tion. In particular, symmetry-resolution allows to prop-erly diagnose physical phenomena that would not be ac-cessible otherwise, by amplifying the role of sectors in thereduced density matrix whose information content couldbe otherwise overwhelmed by other less informative - buthighly-weighted - sectors. In this context, the many-bodytheory we develop seems to suggest that symmetry canbe used to develop improved entanglement detection thatcould outperform their respective ’symmetry-blind’ coun-terparts27.

VIII. ACKNOWLEDGEMENTS

We acknowledge useful discussions with L. Capizzi,R. Fazio, and S. Murciano. We thank T. Brydges,P. Jurcevic, C. Maier, B. Lanyon, R. Blatt, and C. Roosfor generously sharing the experimental data of Ref. 26.This work is partly supported by the ERC under grantnumber 758329 (AGEnTh) and 771536 (NEMO), and bythe MIUR Programme FARE (MEPH). MD, AE, VVand PZ acknowledge support from the European Union’sHorizon 2020 research and innovation programme un-der grant agreement No 817482 (Pasquans). AE andPZ acknowledge funding by the European Union pro-gram Horizon 2020 under grant agreement No. 731473(QuantERA via QTFLAG), the US Air Force Officeof Scientific Research (AFOSR) via IOE Grant No.FA9550-19-1-7044 LASCEM, by the Simons Collabora-tion on UltraQuantum Matter, which is a grant from theSimons Foundation (651440, PZ), and by the Institutfür Quanteninformation. Work in Trieste has been car-ried out within the activities of TQT. BV acknowledgesfinancial support from the Austrian Science Fundation(FWF, P. 32597N), and the French National ResearchAgency (ANR, JCJC project QRand). J. C., A. N. andB. K. acknowledge financial support from the AustrianScience Fund (FWF) stand alone project: P32273-N27and the SFB BeyondC.

Appendix A: Effective Markovian dynamics for thesymmetry-resolved reduced density matrix

We now provide a simple interpretation of the effect-ive description derived in Sec III. Our main interest hereis to determine whether dynamical purification is an ef-fect that relies on a specific correlation present in an ef-fective bath (derived by applying the symmetry resolvedprojectors to the density matrix), or whether it is unre-lated to that, and thus captured entirely by an emergentMarkovian dynamics describing ρA,q.

Indeed, even though the evolution of the global dens-ity matrix ρ is governed by the Markovian master equa-

Page 15: arXiv:2101.07814v1 [cond-mat.stat-mech] 19 Jan 2021

15

tion of Eq. (15), the symmetry-resolved reduced dens-ity matrix ρA,q could have a non-Markovian time evol-ution. The dissipation rates derived in Eq. (16) can beinterpreted as an effective master equation acting dir-ectly on the symmetry-resolved reduced density matrix,with time dependent rates. We consider two arbitrarydensity-matrices product states, whose SR RDM can bewritten in diagonal form ρI , ρII , with matrix elementsajj;I and ajj;II , respectively. What we are interested inis whether the two states can become dynamically moredistinguishable as a function of time: if this is possibleeven for a finite time window, the time evolution is non-Markovian108. In order to address this point, we definethe distance between these states as:

DI,II = Tr√ρIρII . (A1)

After a few lines of algebra, and defining as N,M thetotal rank of the density matrix and the number of thestates belonging to E0, respectively, one obtains:

∂D

∂t= − 1

2(1 +Nγt+MJ2t2)2×

×

∑j∈E0

γ(ajj;I + ajj;II + 2γt)√(ajj;I + γt)(ajj;II + γt)

+

+∑j∈E1

(γ + 2J2t)(ajj;I + ajj;II + 2γt+ 2J2t2)√(ajj;I + γt+ J2t2)(ajj;II + γt+ J2t2)

+

− 2(Nγ + 2MJ2t)

(1 +Nγt+MJ2t2)3×

×

∑j∈E0

√(ajj;I + γt)(ajj;II + γt) +

+∑j∈E1

√(ajj;I + γt+ J2t2)(ajj;II + γt+ J2t2)

so that, under the condition above, one has ∂D/∂t < 0,as this is just the sum of two negative terms. This impliesthat arbitrary states of the type discussed above becomeless distinguishable as a function of time, a signature ofeffective Markovian dynamics.

The very same conclusion can be obtained on moregeneral grounds by noticing that the rate equations

above all have positive rates, thus satisfying P-divisibilitycriteria. At the physical level, this is a consequence ofthe fact that the relaxation time of the environment (inthis case, the part of the system we are tracing uponin a symmetry resolved fashion109) is much longer thanthe timescales we are interested in. For longer times(not accessible to the regime we can tackle with ourtheory, but definitely numerically accessible), we expectthat such effective Markovian description would ulti-mately break down due to the bath dynamics timescalesbeing comparable to the one characterizing the partition.

Appendix B: Symmetry resolved purification inone-dimensional systems

The system we consider here is a one-dimensional ver-sion of the system considered in Sec. III A, that we divideinto two connected partitions A ∪ B with NA and NBsites, respectively. We assume that the system is initial-ized in a charge-density wave |ψ0〉 = |↓, ↑, . . . , ↑〉 (a Néelstate), and for simplicity, take NA and NB to be even.We consider dynamics governed by Eq. (21) and focuson timescales accessible within perturbation theory, thatis, J2t2, tγ � 1. We are interested in the sector q = −1.Adapting the 2D calculations presented in the main text,we find that ρA(q = −1) is divided into two blocks (inperturbation theory):

1. E0(−1): the state that is connected to the CDWby a single hopping process, or a loss event, at theboundary with rate λE0

0

2. E1(−1): the (NA/2− 1) states that are connectedto the CDW by a single loss in the rest of the systemwith eigenvalue λE1

k ;

At the lowest order in perturbation theory, one has thefollowing scaling of the eigenvalues of ρA(−1):

λE00 = (J2t2 + γt)/A(t), λE1

k = γt/A(t) , (B1)

with normalization

A(t) = γt(NA/2) + J2t2. (B2)

This gives

PA(−1) =(NA/2− 1)γ2t2 + (J2t2 + γt)2

[(NA/2)γt+ J2t2]2. (B3)

∗ These authors contributed equally.1 I. Montvay and G. Muenster, Quantum Fields on a lattice(Cambridge Univ. Press, Cambridge, 1994).

2 B. Zeng, X. Chen, D.-L. Zhou, and X.-G. Wen, QuantumInformation Meets Quantum Matter (Springer New York,2019).

3 S. Sachdev, Quantum Phase Transitions (Cambridge Uni-

versity Press, 2000).4 L. Amico, R. Fazio, A. Osterloh, and V. Vedral, Reviewsof Modern Physics 80, 517 (2008).

5 P. Calabrese, J. Cardy, and B. Doyon, Journal of PhysicsA: Mathematical and Theoretical 42, 500301 (2009).

6 N. Laflorencie, Physics Reports 646, 1 (2016).7 P. Buividovich and M. Polikarpov, Physics Letters B 670,

Page 16: arXiv:2101.07814v1 [cond-mat.stat-mech] 19 Jan 2021

16

141 (2008).8 H. Casini, M. Huerta, and J. A. Rosabal, Physical ReviewD 89, 085012 (2014).

9 N. Laflorencie and S. Rachel, Journal of Statistical Mech-anics: Theory and Experiment 2014, P11013 (2014).

10 J. C. Xavier, F. C. Alcaraz, and G. Sierra, Physical Re-view B 98, 041106 (2018).

11 M. Goldstein and E. Sela, Physical Review Letters 120,200602 (2018).

12 G. Parez, R. Bonsignori, and P. Calabrese, Physical Re-view B 103, L041104 (2021).

13 J. I. Cirac and P. Zoller, Nature Physics 8, 264 (2012).14 I. Buluta, S. Ashhab, and F. Nori, Reports on Progress

in Physics 74, 104401 (2011).15 J. Preskill, Quantum 2, 79 (2018).16 R. Blatt and C. Roos, Nature Physics 8, 277 (2012).17 I. Bloch, J. Dalibard, and S. Nascimbène, Nature Physics

8, 267 (2012).18 A. Browaeys and T. Lahaye, Nature Physics 16, 132

(2020).19 S. Schmidt and J. Koch, Annalen der Physik 525, 395

(2013).20 J. M. Gambetta, J. M. Chow, and M. Steffen, npj

Quantum Information 3, 2 (2017).21 H. Häffner, C. Roos, and R. Blatt, Physics Reports 469,

155 (2008).22 We note here that entanglement as witnessed by the neg-

ativity requires a different symmetry-resolution with re-spect to the reduced density matrix80. This is due to thepresence of partial transposition.

23 S. Diehl, A. Micheli, A. Kantian, B. Kraus, H. P. Büchler,and P. Zoller, Nature Physics 4, 878 (2008).

24 F. Verstraete, M. M. Wolf, and J. Ignacio Cirac, NaturePhysics 5, 633 (2009).

25 B. Kraus, H. P. Büchler, S. Diehl, A. Kantian, A. Micheli,and P. Zoller, Physical Review A 78, 042307 (2008).

26 T. Brydges, A. Elben, P. Jurcevic, B. Vermersch,C. Maier, B. P. Lanyon, P. Zoller, R. Blatt, and C. F.Roos, Science 364, 260 (2019).

27 unpublished.28 M. Ohliger, V. Nesme, and J. Eisert, New Journal of

Physics 15, 015024 (2013).29 S. J. van Enk and C. W. J. Beenakker, Physical Review

Letters 108, 110503 (2012).30 A. Elben, B. Vermersch, M. Dalmonte, J. I. Cirac, and

P. Zoller, Physical Review Letters 120, 050406 (2018).31 B. Vermersch, A. Elben, M. Dalmonte, J. I. Cirac, and

P. Zoller, Physical Review A 97, 023604 (2018).32 H.-Y. Huang, R. Kueng, and J. Preskill, Nature Physics

(2020).33 M. Paini and A. Kalev, arXiv:1910.10543.34 A. Elben, R. Kueng, H.-Y. R. Huang, R. van Bijnen,

C. Kokail, M. Dalmonte, P. Calabrese, B. Kraus,J. Preskill, P. Zoller, and B. Vermersch, Physical ReviewLetters 125, 200501 (2020).

35 D. Azses, R. Haenel, Y. Naveh, R. Raussendorf, E. Sela,and E. G. Dalla Torre, Physical Review Letters 125,120502 (2020).

36 L. Capizzi, P. Ruggiero, and P. Calabrese, Journal ofStatistical Mechanics: Theory and Experiment 2020,073101 (2020).

37 M. T. Tan and S. Ryu, Physical Review B 101, 235169(2020).

38 S. Fraenkel and M. Goldstein, Journal of Statistical Mech-

anics: Theory and Experiment 2020, 033106 (2020).39 R. Bonsignori, P. Ruggiero, and P. Calabrese, Journal

of Physics A: Mathematical and Theoretical 52, 475302(2019).

40 S. Murciano, G. Di Giulio, and P. Calabrese, Journal ofHigh Energy Physics 2020, 1 (2020).

41 S. Murciano, G. D. Giulio, and P. Calabrese, SciPostPhysics 8, 46 (2020).

42 P. Calabrese, M. Collura, G. Di Giulio, and S. Murciano,Europhysics Letters 129, 60007 (2020).

43 S. Murciano, P. Ruggiero, and P. Calabrese, Journalof Statistical Mechanics: Theory and Experiment 2020,083102 (2020).

44 E. Cornfeld, L. A. Landau, K. Shtengel, and E. Sela,Physical Review B 99, 115429 (2019).

45 D. X. Horváth and P. Calabrese, Journal of High EnergyPhysics 2020, 1 (2020).

46 D. Azses and E. Sela, Physical Review B 102, 235157(2020).

47 R. Bonsignori and P. Calabrese, Journal of Physics A:Mathematical and Theoretical 54, 015005 (2020).

48 N. Feldman and M. Goldstein, Physical Review B 100,235146 (2019).

49 X. Turkeshi, P. Ruggiero, V. Alba, and P. Calabrese,Physical Review B 102, 014455 (2020).

50 B. Estienne, Y. Ikhlef, and A. Morin-Duchesne,arXiv:2010.10515.

51 G. Vidal and R. F. Werner, Physical Review A 65, 032314(2002).

52 A. Peres, Physical Review Letters 77, 1413 (1996).53 P. Ruggiero, V. Alba, and P. Calabrese, Physical Review

B 94, 195121 (2016).54 P. Calabrese, J. Cardy, and E. Tonni, Physical Review

Letters 109, 130502 (2012).55 P. Calabrese, L. Tagliacozzo, and E. Tonni, Journal

of Statistical Mechanics: Theory and Experiment 2013,P05002 (2013).

56 P. Calabrese, J. Cardy, and E. Tonni, Journal of PhysicsA: Mathematical and Theoretical 48, 015006 (2014).

57 O. Blondeau-Fournier, O. A. Castro-Alvaredo, andB. Doyon, Journal of Physics A: Mathematical and The-oretical 49, 125401 (2016).

58 P. Ruggiero, V. Alba, and P. Calabrese, Physical ReviewB 94, 035152 (2016).

59 X. Turkeshi, P. Ruggiero, and P. Calabrese, Physical Re-view B 101, 064207 (2020).

60 X.Wen, P.-Y. Chang, and S. Ryu, Journal of High EnergyPhysics 2016, 12 (2016).

61 C. Castelnovo, Physical Review A 88, 042319 (2013).62 O. Hart and C. Castelnovo, Physical Review B 97, 144410

(2018).63 Y. A. Lee and G. Vidal, Physical Review A 88, 042318

(2013).64 T.-C. Lu, T. H. Hsieh, and T. Grover, Physical Review

Letters 125, 116801 (2020).65 A. Coser, E. Tonni, and P. Calabrese, Journal of Stat-

istical Mechanics: Theory and Experiment 2014, P12017(2014).

66 V. Eisler and Z. Zimborás, New Journal of Physics 16,123020 (2014).

67 V. Alba and P. Calabrese, Europhysics Letters 126, 60001(2019).

68 M. Hoogeveen and B. Doyon, Nuclear Physics B 898, 78(2015).

Page 17: arXiv:2101.07814v1 [cond-mat.stat-mech] 19 Jan 2021

17

69 X. Wen, P.-Y. Chang, and S. Ryu, Physical Review B92, 075109 (2015).

70 M. J. Gullans and D. A. Huse, Physical Review X 9,021007 (2019).

71 J. Kudler-Flam, M. Nozaki, S. Ryu, and M. T. Tan,Journal of High Energy Physics 2020, 31 (2020).

72 V. Eisler and Z. Zimborás, New Journal of Physics 17,053048 (2015).

73 C. P. Herzog and Y. Wang, Journal of Statistical Mech-anics: Theory and Experiment 2016, 073102 (2016).

74 H. Shapourian, K. Shiozaki, and S. Ryu, Physical ReviewB 95, 165101 (2017).

75 H. Shapourian, P. Ruggiero, S. Ryu, and P. Calabrese,SciPost Physics 7, 37 (2019).

76 J. Eisert, V. Eisler, and Z. Zimborás, Physical Review B97, 1 (2018).

77 H. Shapourian and S. Ryu, Journal of Statistical Mech-anics: Theory and Experiment 2019, 043106 (2019).

78 A. Coser, E. Tonni, and P. Calabrese, Journal of Stat-istical Mechanics: Theory and Experiment 2016, 033116(2016).

79 H. Shapourian and S. Ryu, Physical Review A 99, 022310(2019).

80 E. Cornfeld, M. Goldstein, and E. Sela, Physical ReviewA 98, 032302 (2018).

81 S. Murciano, R. Bonsignori, and P. Calabrese, to appear.82 This assumption is only needed to a have simple count of

the coherent processes. Essentially, we do not want to havesites of the complement that are accessible from two sitesof the partition within first order perturbation theory.

83 The treatment can be extended to local symmetries, andthus gauge theories. The latter case is more complicateddue to the definition of reduced density matrices in Hilbertspaces without tensor-product structure. We leave this tofuture work.

84 A. Messiah, Quantum Mechanics, Quantum MechanicsNo. vol. 2 (Elsevier Science, 1981).

85 S. de Léséleuc, V. Lienhard, P. Scholl, D. Barredo,S. Weber, N. Lang, H. P. Büchler, T. Lahaye, andA. Browaeys, Science 365, 775 (2019).

86 I. Peschel, Journal of Physics A: Mathematical and Gen-eral 36, L205–L208 (2003).

87 I. Peschel and V. Eisler, Journal of Physics A: Mathem-atical and Theoretical 42, 504003 (2009).

88 P. Kos and T. Prosen, Journal of Statistical Mechanics:Theory and Experiment 2017, 123103 (2017).

89 V. Alba and F. Carollo, “Spreading of correla-tions in markovian open quantum systems,” (2020),arXiv:2002.09527.

90 A. Elben, B. Vermersch, R. van Bijnen, C. Kokail, T. Bry-

dges, C. Maier, M. K. Joshi, R. Blatt, C. F. Roos, andP. Zoller, Physical Review Letters 124, 010504 (2020).

91 B. Vermersch, A. Elben, L. M. Sieberer, N. Y. Yao, andP. Zoller, Physical Review X 9, 021061 (2019).

92 M. K. Joshi, A. Elben, B. Vermersch, T. Brydges,C. Maier, P. Zoller, R. Blatt, and C. F. Roos, PhysicalReview Letters 124, 240505 (2020).

93 A. Elben, J. Yu, G. Zhu, M. Hafezi, F. Pollmann,P. Zoller, and B. Vermersch, Science Advances 6,eaaz3666 (2020).

94 Z.-P. Cian, H. Dehghani, A. Elben, B. Vermersch,G. Zhu, M. Barkeshli, P. Zoller, and M. Hafezi, (2020),arXiv:2005.13543.

95 Y. Zhou, P. Zeng, and Z. Liu, Physical Review Letters125, 200502 (2020).

96 M. C. Tran, B. Dakić, F. m. c. Arnault, W. Laskowski,and T. Paterek, Physical Review A 92, 050301 (2015).

97 M. C. Tran, B. Dakić, W. Laskowski, and T. Paterek,Physical Review A 94, 042302 (2016).

98 A. Ketterer, N. Wyderka, and O. Gühne, Physical ReviewLetters 122, 120505 (2019).

99 W.-H. Zhang, C. Zhang, Z. Chen, X.-X. Peng, X.-Y. Xu,P. Yin, S. Yu, X.-J. Ye, Y.-J. Han, J.-S. Xu, G. Chen,C.-F. Li, and G.-C. Guo, Physical Review Letters 125,030506 (2020).

100 L. Knips, J. Dziewior, W. Kłobus, W. Laskowski,T. Paterek, P. J. Shadbolt, H. Weinfurter, andJ. D. A. Meinecke, npj Quantum Information 6 (2020),10.1038/s41534-020-0281-5.

101 A. Elben, B. Vermersch, C. F. Roos, and P. Zoller, Phys-ical Review A 99, 052323 (2019).

102 S. Aaronson, Proceedings of the 50th Annual ACMSIGACT Symposium on Theory of Computing , STOC2018 (Association for Computing Machinery, New York,NY, USA, 2018) p. 325–338.

103 D. Gross, K. Audenaert, and J. Eisert, Journal of Math-ematical Physics 48, 052104 (2007).

104 C. Dankert, R. Cleve, J. Emerson, and E. Livine, PhysicalReview A 80, 012304 (2009).

105 M. Guţă, J. Kahn, R. Kueng, and J. A. Tropp, J. Phys.A 53, 204001 (2020).

106 U. Schollwöck, Annals of Physics 326, 96 (2011),arXiv:1008.3477.

107 A. Kitaev, Annals of Physics 321, 2 (2006).108 H.-P. Breuer, E.-M. Laine, J. Piilo, and B. Vacchini, Re-

views of Modern Physics 88, 021002 (2016).109 We emphasize that we are dealing with a specific

symmetry-resolved sector, as other sectors may featureinformation backflow - a proxy of non-Markovianity - atearlier times.