Significance
Understanding how statistical ensembles arise from the out-of-equilibrium dynamics of isolated pure systems has been a fascinating question since the early days of quantum mechanics. Recently, it has been proposed that the thermodynamic entropy of the long-time statistical ensemble is the stationary entanglement of a large subsystem in an infinite system. Here, we combine this concept with the quasiparticle picture of the entanglement evolution and integrability-based knowledge of the steady state to obtain exact analytical predictions for the time evolution of the entanglement in arbitrary 1D integrable models. These results explicitly show the transformation between the entanglement and thermodynamic entropy during the time evolution. Thus, entanglement is the natural witness for the generalized microcanonical principle underlying relaxation in integrable models.
Keywords: entanglement, quantum quench, integrability, thermodynamics
Abstract
Entanglement and entropy are key concepts standing at the foundations of quantum and statistical mechanics. Recently, the study of quantum quenches revealed that these concepts are intricately intertwined. Although the unitary time evolution ensuing from a pure state maintains the system at zero entropy, local properties at long times are captured by a statistical ensemble with nonzero thermodynamic entropy, which is the entanglement accumulated during the dynamics. Therefore, understanding the entanglement evolution unveils how thermodynamics emerges in isolated systems. Alas, an exact computation of the entanglement dynamics was available so far only for noninteracting systems, whereas it was deemed unfeasible for interacting ones. Here, we show that the standard quasiparticle picture of the entanglement evolution, complemented with integrability-based knowledge of the steady state and its excitations, leads to a complete understanding of the entanglement dynamics in the space–time scaling limit. We thoroughly check our result for the paradigmatic Heisenberg chain.
Since the early days of quantum mechanics, understanding how statistical ensembles arise from the unitary time evolution of an isolated quantum system has been a fascinating question (1–7). A widely accepted mechanism is that, although the entire system remains in a pure state, the reduced density matrix of an arbitrary finite compact subsystem attains a long time limit that can be described by a statistical ensemble (8). In the last decade, groundbreaking experiments with cold atoms (9–19) simulated with incredible precision the unitary time evolution of many-body quantum systems, reviving the interest in this topic. The simplest out-of-equilibrium protocol in which these ideas can be theoretically and experimentally tested is the quantum quench (20, 21). A system is prepared in an initial state , typically the ground state of a local Hamiltonian , and it evolves with a many-body Hamiltonian . At asymptotically long times, physical observables relax to stationary values, which for generic systems, are described by the Gibbs (thermal) ensemble (3–7), whereas for integrable systems, a generalized Gibbs ensemble has to be used (8, 22–45).
Although these results suggest a spectacular compression of the amount of information needed to describe steady states, state of the art numerical methods, such as the time-dependent density matrix renormalization group (46–49) (tDMRG), can only access the short-time dynamics. Physically, the origin of this conundrum is the fast growth of the entanglement entropy , with being the reduced density matrix of an interval of length embedded in an infinite system. It is well-understood that grows linearly with the time after the quench (50). This linear behavior implies an exponentially increasing amount of information manipulated during typical tDMRG simulations. Remarkably, the entanglement dynamics has been successfully observed in a very recent cold-atom experiment (19).
In this work, using a standard quasiparticle picture (50), we show that, in integrable models, the steady state and its low-lying excitations encode sufficient information to reconstruct the entanglement dynamics up to short times. According to the quasiparticle picture (50), the prequench initial state acts as a source of pairs of quasiparticle excitations. Let us first assume that there is only one type of quasiparticles identified by their quasimomentum and moving with velocity . Although quasiparticles created far apart from each other are incoherent, those emitted at the same point in space are entangled. Because these propagate ballistically throughout the system, larger regions get entangled. At time , is proportional to the total number of quasiparticle pairs that, emitted at the same point in space, are shared between A and its complement (Fig. 1A). Specifically, one obtains
[1] |
where depends on the production rate of quasiparticles. Eq. 1 holds in the space–time scaling limit at fixed. When a maximum quasiparticle velocity exists [e.g., because of the Lieb–Robinson bound (51)], for , grows linearly in time, because the second term in Eq. 1 vanishes. In contrast, for , the entanglement is extensive (i.e., ). This light-cone spreading of entanglement has been confirmed analytically only in free models (52–57), numerically in several studies (58–60), in the holographic framework (61–68), and in a recent experiment (19). The validity of the quasiparticle picture Eq. 1 for the entanglement dynamics has been proven for free models in ref. 52. In the presence of interactions, few results are known. For instance, the validity of Eq. 1 has been proven for rational CFTs (Conformal Field Theories) in ref. 69. In interacting integrable models à la Yang–Baxter, the quasiparticle picture has been used to describe the out-of-equilibrium dynamics after an inhomogeneous quench, and it is at the foundation of the integrable hydrodynamics approach for transport in integrable models (70, 71).
Results
In a generic interacting integrable model, there are different species of stable quasiparticles corresponding to bound states of an arbitrary number of elementary excitations. Integrability implies that different types of quasiparticles must be treated independently. It is then natural to conjecture that
[2] |
where the sum is over the types of particles , is their velocity, and is their entropy. To give predictive power to Eq. 2, in the following, we show how to determine and in the Bethe ansatz framework for integrable models. Eq. 2 is straighforwardly generalized to the mutual information between two intervals (see SI Materials and Methods and Fig. S1).
The eigenstates of Bethe ansatz solvable models are in correspondence with a set of pseudomomenta (rapidities) . In the thermodynamic limit, these rapidities form a continuum. One then introduces the particle densities , the hole (i.e., unoccupied rapidities) densities , and the total densities . Every set of densities identifies a thermodynamic “macrostate.” This macrostate corresponds to an exponentially large number of microscopic eigenstates, any of which can be used as a “representative” for the macrostate. The total number of representative microstates is , with as the thermodynamic Yang–Yang entropy of the macrostate
[3] |
Physically, corresponds to the total number of ways of assigning the quasimomentum label to the particles, similar to free fermion models.
In the Bethe ansatz treatment of quantum quenches (72–74), local properties of the postquench stationary state are described by a set of densities and . Calculating these densities is a challenging task that has been performed only in a few cases (75–87). From the densities, the thermodynamic entropy of the stationary ensemble Eq. 3 is . Physically, is the number of microscopic eigenstates entering in a generalized microcanonical ensemble for quenches, in which all of the microstates corresponding to the macrostate have the same probability.
We now present our predictions for the entanglement dynamics (Fig. 1B gives a survey of our theoretical scheme). First, in the stationary state, the density of thermodynamic entropy coincides with that of the entanglement entropy in Eq. 2, as it has been shown analytically for free models (52, 88, 89). This identification implies that . Moreover, it is natural to identify the entangling quasiparticles in Eq. 2 with the low-lying excitations around the stationary state . Their group velocities depend on the stationary state, because the interactions induce a state-dependent dressing of the excitations. These velocities can be calculated by Bethe ansatz techniques (90) (SI Materials and Methods and Figs. S2 and S3).
To substantiate our idea, we focus on the spin- anisotropic Heisenberg () chain, considering quenches from several low-entangled initial states, namely the tilted Néel state, the Majumdar–Ghosh (dimer) state, and the tilted ferromagnetic state (Materials and Methods). For these initial states, the densities are known analytically (77–79).
Fig. 2 summarizes the expected entanglement dynamics in the space–time scaling limit, plotting vs. . Interestingly, is always smaller than (i.e., the entropy of the maximally entangled state). For the Néel quench, because the Néel state becomes the ground state of Eq. 4 in the limit , vanishes, whereas it saturates for all of the other quenches. For the Majumdar–Ghosh state, one obtains at . For the tilted ferromagnet with (Fig. 2E), is small at any , reflecting that the ferromagnet is an eigenstate of the chain. Surprisingly, the linear growth seems to extend for . However, (Fig. 2F) is flat only for , which signals true linear regime only for . This peculiar behavior is caused by the large entanglement contribution of the slow quasiparticles. In Fig. 3, we report the bound-state resolved contributions to the entanglement dynamics. Fig. 3 A and C focuses on the steady-state entropy (second term in Eq. 2), whereas Fig. 3 B and D shows the bound-state contributions to the slope of the linear growth (first term in Eq. 2). The contribution of the bound states, although never dominant, is crucial to ensure accurate predictions.
Fig. 4A shows tDMRG results for for the quench from the symmetrized Néel state . The data are for the open chain and subsystems starting from the chain boundary. The qualitative agreement with Eq. 2 is apparent. Fig. 4B reports the steady-state entanglement entropy as a function (data at in Fig. 4A). The volume law is visible. The dashed–dotted lines in Fig. 4B are fits to , supporting the equivalence between entanglement and thermodynamic entropy. Fig. 4C focuses on the full-time dependence, plotting vs. . The dashed–dotted line in Fig. 4C is Eq. 2 with because of the open boundary conditions (58). Deviations from Eq. 2 because of the finite are visible. The diamonds in Fig. 4C are numerical extrapolations to the thermodynamic limit. The agreement with Eq. 2 is perfect. Finally, we provide a more stringent check of Eq. 2, focusing on the linear entanglement growth. Fig. 5 shows infinite time-evolving block decimation (iTEBD) results in the thermodynamic limit for , with taken from ref. 91. For all of the quenches, the agreement with Eq. 2 (horizontal lines in Fig. 5) is spectacular.
SI Materials and Methods
In this supplementary material, we provide some details on
-
i)
the Bethe ansatz solution of the chain,
-
ii)
the Bethe ansatz treatment of the postquench steady state,
-
iii)
the calculation of the group velocities of the low-lying excitations around the steady state, and
-
iv)
the entanglement and the mutual information of two disjoint intervals.
Bethe Ansatz Solution of the XXZ Chain.
In the Bethe ansatz (94) solution of the chain, the eigenstates of the chain in the sector with down spins (particles) are in correspondence with a set of rapidities . These rapidities are obtained by solving the Bethe equations (94)
[S1] |
where .
In the thermodynamic limit, the solutions of the Bethe equations (Eq. S1) form string patterns in the complex plane. The rapidities forming an string are parametrized as
[S2] |
where labels the different string components, is the “string center,” and are the so-called string deviations. Typically (i.e., for the majority of the eigenstates of the chain), one has , implying that the string deviations can be neglected (94) (string hypothesis). Physically, strings correspond to bound states of down spins. The string centers are solutions of the Bethe–Gaudin–Takahashi (BGT) equations (94):
[S3] |
For , one has . Here, we define . The scattering phases are defined as
[S4] |
Each different choice of the so-called BGT quantum numbers identifies a different set of solutions of Eq. S3 and in turn, a different eigenstate of the model. The corresponding eigenstate energy and total momentum are obtained by summing over all of the BGT rapidities (94) as and with
[S5] |
Note that depends only on the .
The Steady State: Bethe Ansatz Treatment.
Here, we provide some details on how to derive the steady-state densities in Bethe ansatz. The root densities are defined as (95)
[S6] |
Instead of working with the hole densities , it is convenient to define . For all of the initial states considered in this work, the corresponding steady-state densities obey the recursive relations (74, 80)
[S7] |
[S8] |
with and . In Eqs. S7 and S8, the initial conditions and encode the information about the prequench initial state. For all of the quenches considered in this work, these initial conditions are known analytically. For instance, for the quench from the Néel state, one has (80)
[S9] |
where [with as defined in Eq. S4]. For the dimer state, and have been calculated in ref. 81, whereas for the tilted Néel and the tilted ferromagnet, they have been derived recently (79, 87). Fig. S2A shows for for the quench from the Néel state in the chain with .
Group Velocities of the Entangling Quasiparticles.
Here, we detail the calculation of the group velocities of the entangling quasiparticles, which explain the linear entanglement growth after the quench. The low-lying excitations around the postquench steady state can be constructed as particle hole excitations over the corresponding macrostate. Notice that it has been verified in ref. 96 that the low-lying excitations around the stationary state can be used to reconstruct “back in time” the postquench out-of-equilibrium dynamics of local observables. First, one can imagine choosing among the eigenstates of the chain a representative of the macrostate, identified by some BGT quantum numbers . Second, a particle hole excitation, in each -string sector, is obtained as , where is the BGT number of the added particle (hole). Because the chain is interacting, this local change in quantum numbers affects all of the new rapidities. The excess energy of the particle hole excitation is
[S10] |
Remarkably, apart from the dressing of the “single-particle” energy (see below for its calculation), Eq. S10 is the same as for free models. The change in the total momentum is obtained from Eq. S5 as
[S11] |
Finally, the group velocity of the particle hole excitations is by definition
[S12] |
Here, we used that (94) , with , and we defined as the derivative of . Importantly, is determined by an infinite system of Fredholm integral equations of the second kind as
[S13] |
where and (compare Eq. S4 with Eq. S5). The solutions of Eq. S13 can be obtained numerically very effectively after truncating the system considering . We should mention that the method outlined above has been used to study transport properties in the chain starting from inhomogeneous initial conditions (70, 71) and the spreading of correlations after quantum quenches (20, 90, 97–105).
Importantly, in the limit , the solution of Eq. S13 and the group velocities (Eq. S12) can be obtained analytically as a power series in . For instance, for the quench from the Néel state, one obtains
[S14] |
A similar result can be obtained for the Majumdar–Ghosh state as
[S15] |
As an example of calculation of group velocities, we plot in Fig. S2B for for the quench from the Néel state in the chain with . In Fig. S1B, are plotted vs. the rapidity . The star symbols in Fig. S2B are the perturbative results (Eq. S14) in the large limit. Moreover, in Fig. S3, we plot the maximum group velocity in each string sector as a function of . The dashed–dotted lines in Fig. S3 are the analytical results (Eq. S14) in the limit . Note that, in the limit , one has , whereas the group velocities are vanishing in the other string sectors. At , is different from the low-energy spin-wave velocity (shown as the horizontal dotted line in Fig. S3).
Two Disjoint Intervals.
In this section, we investigate the entanglement of two disjoint spin blocks after a quench. This study is motivated by some recent holographic results, which also apply to irrational + CFT (64, 67, 69, 106, 107), that are in contrast with the quasiparticle picture. We focus on the behavior of the von Neumann mutual information after a global quench. We consider the tripartition of the chain as , where and are two disjoint intervals of equal length and at distance , whereas is the remainder of the chain. The mutual information is defined as
[S16] |
with and being the entanglement entropies of and , respectively.
For an infinite system and two intervals of length at a distance , it is straightforward to derive the contribution to the mutual information of each quasiparticle with velocity , namely (50)
[S17] |
which predicts for , a linear increase for followed by a linear decrease up to . In stark contrast, it has been suggested that the quasiparticle picture for the entanglement propagation does not hold in holographic contexts. The scenario of ref. 69 predicts at any time: the idea is that quasiparticles originated at the same point in space and traveling with one in and the other in , do not remain maximally entangled when they are far apart from each other (a phenomenon known as scrambling).
To clarify whether the quasiparticle picture applies to interacting integrable models, here we discuss the behavior of after the quench from the Néel state in the chain. We restrict ourselves to the open spin chain. We always consider the situation with the two intervals and at the opposite edges of the chain, because it is convenient for the numerical simulations. This geometry implies .
In this case, the contribution of each quasiparticle is given by Eq. S17 with the replacement , but the formula is valid only before the revival time . Similar to Eq. 2, the final quasiparticle prediction for the mutual information is obtained by summing (and integrating) the contribution of all quasiparticles: that is,
[S18] |
which again, it is valid before the revival time (i.e., for , such that ). In the following, we compare Eq. S18 with tDMRG results.
It is well-known that extracting the entanglement entropy of multiple disjoint intervals in density matrix renormalization group (DMRG) simulations is a formidable task in contrast with the single-interval entropy. Specifically, given the matrix product state (MPS) representation of the state, the computational cost scales as for the multiinterval, whereas it is only for the single interval, with being the bond dimension of the MPS. This issue is even more dramatic out of equilibrium, where grows exponentially with time. For this reason, we can provide reliable data for only for very small intervals and up to short times after the quench.
Our results are presented in Fig. S1. The symbols are the tDMRG data for plotted as a function of the time after the quench. We show the data for , which are the only sizes that we can simulate reliably, and only . In our simulations, for all values of , the two intervals are at fixed distance . The lines in Fig. S1 are the analytic results obtained using Eq. S18. Clearly, the system sizes and timescales accessible in our tDMRG simulations do not allow us to reach a quantitative conclusion on the validity of Eq. S18 for the mutual information. However, we observe that the DMRG data are in a good overall qualitative agreement with the quasiparticle predictions in Eq. S18. Indeed, the mutual information is zero for ; then, it starts growing (seemingly) linearly with time, up to a maximum value that is clearly visible for (but not for smaller ). The few results available in the literature for the mutual information of two disjoint intervals for free systems (55, 108) show very similar effects when the lengths of the subsystems are very small, such as in our simulations. Hence, these DMRG results provide a strong support for the validity of the quasiparticle picture for , signaling the absence of scrambling that takes place for CFTs with large central charge (69).
Conclusions
The main result of this work is the analytical prediction in Eq. 2 for the time-dependent entanglement entropy after a generic quantum quench in an integrable model. We tested our conjecture for several quantum quenches in the spin chain, although we expect Eq. 2 to be more general. Additional checks of Eq. 2 (e.g., for the Lieb–Liniger gas) are desirable. It would be also interesting to generalize Eq. 2 to quenches from inhomogeneous initial states, exploiting the recent analytical results (70, 71, 92). Although we are not able yet to provide an ab initio derivation of Eq. 2, we find it remarkable that it is possible to characterize analytically the dynamics of the entanglement entropy, whereas its equilibrium behavior is still an open challenge. Finally, we believe that Eq. 2 represents a deep conceptual breakthrough, because it shows in a single compact formula the relation between entanglement and thermodynamic entropy for integrable models. An analogous description for nonintegrable systems, where quasiparticles have finite lifetime or do not exist at all, could lead to a deeper understanding of thermalization (19).
Materials and Methods
The anisotropic spin- Heisenberg chain is defined by the Hamiltonian
[4] |
where are spin- operators, and is the anisotropy parameter. Here, we considered as prequench initial states the tilted Néel state , the Majumdar–Ghosh (dimer) state , and the tilted ferromagnetic state . The Heisenberg spin chain is the prototype of all integrable models. Moreover, for all of the initial states considered in this work, the postquench steady state can be characterized analytically via the macrostate densities . Specifically, a set of recursive relations for these densities can be obtained (SI Materials and Methods). The group velocities of the low-lying excitations around the steady state (i.e., the entangling quasiparticles) are obtained by solving numerically an infinite set of second-type Fredholm integral equations (details are in SI Materials and Methods).
The numerical data for the postquench dynamics of the entanglement entropy presented in Fig. 4 were obtained using the standard tDMRG (46–49) in the framework of matrix product states. For the implementation, we used the ITENSOR library (itensor.org/). The data presented in Fig. 5 are obtained using the iTEBD method (93) and they are a courtesy of Mario Collura.
Acknowledgments
We thank Lorenzo Piroli and Eric Vernier for sharing their results before publication. The iTEBD data presented in Fig. 5 are a courtesy of Mario Collura. We acknowledge support from European Research Council Starting Grant 279391 EDEQS. This project has received funding from the European Union’s Horizon 2020 research and innovation program under Marie Sklodowoska-Curie Grant 702612 OEMBS.
Footnotes
The authors declare no conflict of interest.
This article is a PNAS Direct Submission.
This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1703516114/-/DCSupplemental.
References
- 1.von Neumann J. Beweis des ergodensatzes und des H-theorems. Z Phys. 1929;57:30–70. [Google Scholar]
- 2.Jensen RV, Shankar R. Statistical behavior in deterministic quantum systems with few degrees of freedom. Phys Rev Lett. 1985;54:1879–1882. doi: 10.1103/PhysRevLett.54.1879. [DOI] [PubMed] [Google Scholar]
- 3.Deutsch JM. Quantum statistical mechanics in a closed system. Phys Rev A. 1991;43:2046–2049. doi: 10.1103/physreva.43.2046. [DOI] [PubMed] [Google Scholar]
- 4.Srednicki M. Chaos and quantum thermalization. Phys Rev E. 1994;50:888–901. doi: 10.1103/physreve.50.888. [DOI] [PubMed] [Google Scholar]
- 5.Rigol M, Dunjko V, Olshanii M. Thermalization and its mechanism for generic isolated quantum systems. Nature. 2008;452:854–858. doi: 10.1038/nature06838. [DOI] [PubMed] [Google Scholar]
- 6.Rigol M, Srednicki M. Alternatives to eigenstate thermalization. Phys Rev Lett. 2012;108:110601. doi: 10.1103/PhysRevLett.108.110601. [DOI] [PubMed] [Google Scholar]
- 7.D’Alessio L, Kafri Y, Polkovnikov A, Rigol M. From quantum chaos and eigenstate thermalization to statistical mechanics and thermodynamics. Adv Phys. 2016;65:239–362. [Google Scholar]
- 8.Essler FHL, Fagotti M. Quench dynamics and relaxation in isolated integrable quantum spin chains. J Stat Mech Theor Exp. 2016;2016:064002. [Google Scholar]
- 9.Kinoshita T, Wenger T, Weiss DS. A quantum Newton’s cradle. Nature. 2008;440:900–903. doi: 10.1038/nature04693. [DOI] [PubMed] [Google Scholar]
- 10.Hofferberth S, Lesanovsky I, Fischer B, Schumm T, Schiedmayer J. Non-equilibrium coherence dynamics in one-dimensional Bose gases. Nature. 2007;449:324–327. doi: 10.1038/nature06149. [DOI] [PubMed] [Google Scholar]
- 11.Trotzky S, et al. Probing the relaxation towards equilibrium in an isolated strongly correlated one-dimensional Bose gas. Nat Phys. 2012;8:325–330. [Google Scholar]
- 12.Gring M, et al. Relaxation and prethermalization in an isolated quantum system. Science. 2012;337:1318–1322. doi: 10.1126/science.1224953. [DOI] [PubMed] [Google Scholar]
- 13.Cheneau M, et al. Light-cone-like spreading of correlations in a quantum many-body system. Nature. 2012;481:484–487. doi: 10.1038/nature10748. [DOI] [PubMed] [Google Scholar]
- 14.Langen T, Geiger R, Kuhnert M, Rauer B, Schmiedmayer J. Local emergence of thermal correlations in an isolated quantum many-body system. Nat Phys. 2013;9:640–643. [Google Scholar]
- 15.Meinert F, et al. Quantum quench in an atomic one-dimensional ising chain. Phys Rev Lett. 2013;111:053003. doi: 10.1103/PhysRevLett.111.053003. [DOI] [PubMed] [Google Scholar]
- 16.Fukuhara T, et al. Quantum dynamics of a mobile spin impurity. Nat Phys. 2013;9:235–241. [Google Scholar]
- 17.Langen T, et al. Experimental observation of a generalized Gibbs ensemble. Science. 2015;348:207–211. doi: 10.1126/science.1257026. [DOI] [PubMed] [Google Scholar]
- 18.Islam R, et al. Measuring entanglement entropy in a quantum many-body system. Nature. 2015;528:77–83. doi: 10.1038/nature15750. [DOI] [PubMed] [Google Scholar]
- 19.Kaufman AM, et al. Quantum thermalization through entanglement in an isolated many-body system. Science. 2016;353:794–800. doi: 10.1126/science.aaf6725. [DOI] [PubMed] [Google Scholar]
- 20.Calabrese P, Cardy J. Time dependence of correlation functions following a quantum quench. Phys Rev Lett. 2006;96:136801. doi: 10.1103/PhysRevLett.96.136801. [DOI] [PubMed] [Google Scholar]
- 21.Polkovnikov A, Sengupta K, Silva A, Vengalattore M. Colloquium: Nonequilibrium dynamics of closed interacting quantum systems. Rev Mod Phys. 2011;83:863–883. [Google Scholar]
- 22.Rigol M, Dunjko V, Yurovsky V, Olshanii M. Relaxation in a completely integrable many-body quantum system: An ab initio study of the dynamics of the highly excited states of 1D lattice hard-core bosons. Phys Rev Lett. 2007;98:050405. doi: 10.1103/PhysRevLett.98.050405. [DOI] [PubMed] [Google Scholar]
- 23.Cazalilla MA. Effect of suddenly turning on interactions in the Luttinger model. Phys Rev Lett. 2006;97:156403. doi: 10.1103/PhysRevLett.97.156403. [DOI] [PubMed] [Google Scholar]
- 24.Barthel T, Schollwöck U. Dephasing and the steady state in quantum many-particle systems. Phys Rev Lett. 2008;100:100601. doi: 10.1103/PhysRevLett.100.100601. [DOI] [PubMed] [Google Scholar]
- 25.Cramer M, Dawson CM, Eisert J, Osborne TJ. Exact relaxation in a class of nonequilibrium quantum lattice systems. Phys Rev Lett. 2008;100:030602. doi: 10.1103/PhysRevLett.100.030602. [DOI] [PubMed] [Google Scholar]
- 26.Cramer M, Eisert J. A quantum central limit theorem for non-equilibrium systems: Exact local relaxation of correlated states. New J Phys. 2010;12:055020. [Google Scholar]
- 27.Calabrese P, Essler FHL, Fagotti M. Quantum quench in the transverse-field ising chain. Phys Rev Lett. 2011;106:227203. doi: 10.1103/PhysRevLett.106.227203. [DOI] [PubMed] [Google Scholar]
- 28.Cazalilla MA, Iucci A, Chung M-C. Thermalization and quantum correlations in exactly solvable models. Phys Rev E. 2012;85:011133. doi: 10.1103/PhysRevE.85.011133. [DOI] [PubMed] [Google Scholar]
- 29.Calabrese P, Essler FHL, Fagotti M. Quantum quench in the transverse field ising chain II: Stationary state properties. J Stat Mech Theor Exp. 2012;2012:P07022. [Google Scholar]
- 30.Sotiriadis S, Fioretto D, Mussardo G. Zamolodchikov-faddeev algebra and quantum quenches in integrable field theories. J Stat Mech Theor Exp. 2012;2012:P02017. [Google Scholar]
- 31.Collura M, Sotiriadis S, Calabrese P. Equilibration of a Tonks-Girardeau gas following a trap release. Phys Rev Lett. 2013;110:245301. doi: 10.1103/PhysRevLett.110.245301. [DOI] [PubMed] [Google Scholar]
- 32.Collura M, Sotiriadis S, Calabrese P. Quench dynamics of a Tonks-Girardeau gas released from a harmonic trap. J Stat Mech Theor Exp. 2013;2013:P09025. [Google Scholar]
- 33.Fagotti M, Essler FHL. Stationary behaviour of observables after a quantum quench in the spin- Heisenberg chain. J Stat Mech Theor Exp. 2013;2013:P07012. [Google Scholar]
- 34.Kormos M, Collura M, Calabrese P. Analytic results for a quantum quench from free to hard-core one-dimensional bosons. Phys Rev A. 2014;89:013609. [Google Scholar]
- 35.Sotiriadis S, Calabrese P. Validity of the GGE for quantum quenches from interacting to noninteracting models. J Stat Mech Theor Exp. 2014;2014:P07024. [Google Scholar]
- 36.Ilieveski E, et al. Complete generalized Gibbs ensembles in an interacting theory. Phys Rev Lett. 2015;115:157201. doi: 10.1103/PhysRevLett.115.157201. [DOI] [PubMed] [Google Scholar]
- 37.Alba V. 2015. Simulating the generalized Gibbs ensemble (GGE): A Hilbert space Monte Carlo approach. arXiv:1507.06994.
- 38.Essler FHL, Mussardo G, Panfil M. Generalized Gibbs ensembles for quantum field theories. Phys Rev A. 2015;91:051602. [Google Scholar]
- 39.Cardy J. Quantum quenches to a critical point in one dimension: Some further results. J Stat Mech Theor Exp. 2016;2016:023103. [Google Scholar]
- 40.Sotiriadis S. Memory-preserving equilibration after a quantum quench in a 1D critical model. Phys Rev A. 2016;94:031605. [Google Scholar]
- 41.Bastianello A, Sotiriadis S. 2016. Quasi locality of the GGE in interacting-to-free quenches in relativistic field theories. arXiv:1608.00924.
- 42.Vernier E, Cubero AC. 2016. Quasilocal charges and the complete GGE for field theories with non diagonal scattering. arXiv:1609.03220.
- 43.Vidmar L, Rigol M. Generalized Gibbs ensemble in integrable lattice models. J Stat Mech Theor Exp. 2016;2016:064007. [Google Scholar]
- 44.Gogolin C, Eisert J. Equilibration, thermalisation, and the emergence of statistical mechanics in closed quantum systems. Rep Prog Phys. 2016;79:056001. doi: 10.1088/0034-4885/79/5/056001. [DOI] [PubMed] [Google Scholar]
- 45.Calabrese P, Essler FHL, Mussardo G. Introduction to “quantum integrability in out of equilibrium systems.”. J Stat Mech Theor Exp. 2016;2016:064001. [Google Scholar]
- 46.White SR, Feiguin AE. Real-time evolution using the density matrix renormalization group. Phys Rev Lett. 2004;93:076401. doi: 10.1103/PhysRevLett.93.076401. [DOI] [PubMed] [Google Scholar]
- 47.Daley AJ, Kollath C, Schollwock U, Vidal G. Time-dependent density-matrix renormalization-group using adaptive effective Hilbert spaces. J Stat Mech Theor Exp. 2004;2004:P04005. [Google Scholar]
- 48.Schollwöck U. The density-matrix renormalization group. Rev Mod Phys. 2005;77:259. doi: 10.1098/rsta.2010.0382. [DOI] [PubMed] [Google Scholar]
- 49.Schollwöck U. The density-matrix renormalization group in the age of matrix product states. Ann Phys. 2011;326:96–192. [Google Scholar]
- 50.Calabrese P, Cardy J. Evolution of entanglement entropy in one-dimensional systems. J Stat Mech Theor Exp. 2005;2005:P04010. [Google Scholar]
- 51.Lieb EH, Robinson DW. The finite group velocity of quantum spin systems. Commun Math Phys. 1972;28:251–257. [Google Scholar]
- 52.Fagotti M, Calabrese P. Evolution of entanglement entropy following a quantum quench: Analytic results for the XY chain in a transverse magnetic field. Phys Rev A. 2008;78:010306. [Google Scholar]
- 53.Eisler V, Peschel I. Entanglement in a periodic quench. Ann Phys. 2008;17:410–423. [Google Scholar]
- 54.Nezhadhaghighi MG, Rajabpour MA. Entanglement dynamics in short and long-range harmonic oscillators. Phys Rev B. 2014;90:205438. [Google Scholar]
- 55.Coser A, Tonni E, Calabrese P. Entanglement negativity after a global quantum quench. J Stat Mech Theor Exp. 2014;2014:P12017. [Google Scholar]
- 56.Cotler JS, Hertzberg MP, Mezei M, Mueller MT. Entanglement growth after a global quench in free scalar field theory. J High Energ Phys. 2016;11:166. [Google Scholar]
- 57.Buyskikh AS, Fagotti M, Schachenmayer J, Essler FHL, Daley AJ. Entanglement growth and correlation spreading with variable-range interactions in spin and fermionic tunneling models. Phys Rev A. 2016;93:053620. [Google Scholar]
- 58.De Chiara G, Montangero S, Calabrese P, Fazio R. Entanglement entropy dynamics in Heisenberg chains. J Stat Mech Theor Exp. 2006;2006:P03001. [Google Scholar]
- 59.Läuchli AM, Kollath C. Spreading of correlations and entanglement after a quench in the one-dimensional Bose-Hubbard model. J Stat Mech Theor Exp. 2008;2008:P05018. [Google Scholar]
- 60.Kim H, Huse DA. Ballistic spreading of entanglement in a diffusive nonintegrable system. Phys Rev Lett. 2013;111:127205. doi: 10.1103/PhysRevLett.111.127205. [DOI] [PubMed] [Google Scholar]
- 61.Hubeny VE, Rangamani M, Takayanagi TA. Covariant holographic entanglement entropy proposal. J High Energ Phys. 2007;2007:062. [Google Scholar]
- 62.Abajo-Arrastia J, Aparicio J, Lopez E. Holographic evolution of entanglement entropy. J High Energ Phys. 2010;1011:149. [Google Scholar]
- 63.Albash T, Johnson CV. Evolution of holographic entanglement entropy after thermal and electromagnetic quenches. New J Phys. 2011;13:045017. [Google Scholar]
- 64.Allais A, Tonni E. Holographic evolution of the mutual information. J High Energ Phys. 2012;1201:102. [Google Scholar]
- 65.Callan R, He J-Y, Headrick M. Strong subadditivity and the covariant holographic entanglement entropy formula. J High Energ Phys. 2012;1206:081. [Google Scholar]
- 66.Liu H, Suh SJ. Entanglement tsunami: Universal scaling in holographic thermalization. Phys Rev Lett. 2014;112:011601. doi: 10.1103/PhysRevLett.112.011601. [DOI] [PubMed] [Google Scholar]
- 67.Balasubramanian V, Bernamonti A, Copland N, Craps B, Galli F. Thermalization of mutual and tripartite information in strongly coupled two dimensional conformal field theories. Phys Rev D Part Fields. 2011;84:105017. [Google Scholar]
- 68.Liu H, Suh SJ. Entanglement growth during thermalization in holographic systems. Phys Rev D Part Fields. 2014;89:066012. [Google Scholar]
- 69.Asplund CT, Bernamonti A, Galli F, Hartmann T. Entanglement scrambling in 2D conformal field theory. J High Energ Phys. 2015;9:110. [Google Scholar]
- 70.Castro-Alvaredo OA, Doyon B, Yoshimura T. Emergent hydrodynamics in integrable quantum systems out of equilibrium. Phys Rev X. 2016;6:041065. [Google Scholar]
- 71.Bertini B, Collura M, De Nardis J, Fagotti M. Transport in out-of-equilibrium XXZ chains: Exact profiles of charges and currents. Phys Rev Lett. 2016;117:207201. doi: 10.1103/PhysRevLett.117.207201. [DOI] [PubMed] [Google Scholar]
- 72.Caux J-S, Essler FHL. Time evolution of local observables after quenching to an integrable model. Phys Rev Lett. 2013;110:257203. doi: 10.1103/PhysRevLett.110.257203. [DOI] [PubMed] [Google Scholar]
- 73.Caux J-S. The quench action. J Stat Mech Theor Exp. 2016;2016:064006. [Google Scholar]
- 74.Ilievski E, Quinn E, De Nardis J, Brockmann M. String-charge duality in integrable lattice models. J Stat Mech Theor Exp. 2016;2016:063101. [Google Scholar]
- 75.Bertini B, Schuricht D, Essler FHL. Quantum quench in the sine-Gordon model. J Stat Mech Theor Exp. 2014;2014:P10035. [Google Scholar]
- 76.De Nardis J, Wouters B, Brockmann M, Caux J-S. Solution for an interaction quench in the Lieb-Liniger Bose gas. Phys Rev A. 2014;89:033601. [Google Scholar]
- 77.Wouters B, et al. Quenching anisotropic Heisenberg chain: Exact solution generalized Gibbs ensemble predictions. Phys Rev Lett. 2014;113:117202. doi: 10.1103/PhysRevLett.113.117202. [DOI] [PubMed] [Google Scholar]
- 78.Pozsgay B, et al. Correlations after quan quenches XXZ spin chain: Fail generalized Gibbs ensemble. Phys Rev Lett. 2014;113:117203. doi: 10.1103/PhysRevLett.113.117203. [DOI] [PubMed] [Google Scholar]
- 79.Piroli L, Vernier E, Calabrese P. Exact steady states for quantum quenches in integrable Heisenberg spin chains. Phys Rev B. 2016;94:054313. [Google Scholar]
- 80.Brockmann M, et al. Quench action approach for releasing the neel state into the spin-1/2 XXZ chain. J Stat Mech Theor Exp. 2014;2014:P12009. [Google Scholar]
- 81.Mestyán M, Pozsgay B, Takács G, Werner MA. Quenching the XXZ spin chain: Quench action approach versus generalized Gibbs ensemble. J Stat Mech Theor Exp. 2015;2015:P04001. [Google Scholar]
- 82.Bucciantini L. Stationary state after a quench to the Lieb-Liniger from rotating BECs. J Stat Phys. 2016;164:621–644. [Google Scholar]
- 83.Bertini B, Piroli L, Calabrese P. Quantum quenches in the sinh-Gordon model: Steady state and one-point correlation functions. J Stat Mech Theor Exp. 2016;2016:063102. [Google Scholar]
- 84.Alba V, Calabrese P. The quench action approach in finite integrable spin chains. J Stat Mech Theor Exp. 2016;2016:043105. [Google Scholar]
- 85.Piroli L, Calabrese P, Essler FHL. Multiparticle bound-state formation following a quantum quench to the one-dimensional Bose gas with attractive interactions. Phys Rev Lett. 2016;116:070408. doi: 10.1103/PhysRevLett.116.070408. [DOI] [PubMed] [Google Scholar]
- 86.Piroli L, Calabrese P, Essler FHL. Quantum quenches to the attractive one-dimensional Bose gas: Exact results. SciPost Phys. 2016;1:001. [Google Scholar]
- 87.Piroli L, Vernier E, Calabrese P, Rigol M. Correlations and diagonal entropy after quantum quenches in XXZ chains. Phys Rev B. 2017;95:054308. [Google Scholar]
- 88.Collura M, Kormos M, Calabrese P. Stationary entanglement entropies following an interaction quench in 1D Bose gas. J Stat Mech Theor Exp. 2014;2014:P01009. [Google Scholar]
- 89.Kormos M, Bucciantini L, Calabrese P. Stationary entropies after a quench from excited states in the Ising chain. Europhys Lett. 2014;107:40002. [Google Scholar]
- 90.Bonnes L, Essler FHL, Läuchli AM. Light-cone dynamics after quantum quenches in spin chains. Phys Rev Lett. 2014;113:187203. doi: 10.1103/PhysRevLett.113.187203. [DOI] [PubMed] [Google Scholar]
- 91.Fagotti M, Collura M, Essler FHL, Calabrese P. Relaxation after quantum quenches in the spin- Heisenberg XXZ chain. Phys Rev B. 2014;89:125101. [Google Scholar]
- 92.Dubail J, Stéphan J-M, Viti J, Calabrese P. Conformal field theory for inhomogeneous one-dimensional quantum systems: The example of non-interacting Fermi gases. SciPost Phys. 2017;2:002. [Google Scholar]
- 93.Vidal G. Classical simulation of infinite-size quantum lattice systems in one spatial dimension. Phys Rev Lett. 2007;98:070201. doi: 10.1103/PhysRevLett.98.070201. [DOI] [PubMed] [Google Scholar]
- 94.Takahashi M. Thermodynamics of One-Dimensional Solvable Models. Cambridge Univ Press; Cambridge, UK: 1999. [Google Scholar]
- 95.Yang CN, Yang CP. Thermodynamics of a one-dimensional system of bosons with repulsive delta function interaction. J Math Phys. 1969;10:1115–1122. [Google Scholar]
- 96.De Nardis J, Piroli L, Caux J-S. Relaxation dynamics of local observables in integrable systems. J Phys A. 2015;48:43FT01. [Google Scholar]
- 97.Calabrese P, Cardy J. Quantum quenches in extended systems. J Stat Mech Theor Exp. 2007;2007:P06008. [Google Scholar]
- 98.Gambassi A, Calabrese P. Quantum quenches as classical critical films. Europhys Lett. 2011;95:66007. [Google Scholar]
- 99.Delfino G. Quantum quenches with integrable pre-quench dynamics. J Phys A. 2014;47:402001. [Google Scholar]
- 100.Collura M, Calabrese P, Essler FHL. Quantum quench within the gapless phase of the spin- Heisenberg spin-chain. Phys Rev B. 2015;92:125131. [Google Scholar]
- 101.Calabrese P, Cardy J. Quantum quenches in + dimensional conformal field theories. J Stat Mech Theor Exp. 2016;2016:064003. [Google Scholar]
- 102.Barbiero L, Dell’Anna L. 2016. Spreading of correlations in quenched repulsive and attractive one dimensional lattice system. arXiv:1609.07387.
- 103.Kormos M, Collura M, Takács G, Calabrese P. Real time confinement following a quantum quench to a non-integrable model. Nat Phys. 2016;13:246–249. [Google Scholar]
- 104.Bertini B, Essler FHL, Groha S, Robinson N. Thermalization and light cones in a model with weak integrability breaking. Phys Rev B. 2016;94:245117. doi: 10.1103/PhysRevLett.115.180601. [DOI] [PubMed] [Google Scholar]
- 105.Chiocchetta A, Tavora M, Gambassi A, Mitra A. Short-time universal scaling and light-cone dynamics after a quench in an isolated quantum system in d spatial dimensions. Phys Rev B. 2016;94:134311. [Google Scholar]
- 106.Asplund CT, Bernamonti A. Mutual information after a local quench in conformal field theory. Phys Rev D Part Fields. 2014;89:066015. [Google Scholar]
- 107.Leichenauer S, Moosa M. Entanglement tsunami in (1+1)-dimensions. Phys Rev D Part Fields. 2015;92:126004. [Google Scholar]
- 108.Fagotti M, Calabrese P. Entanglement entropy of two disjoint blocks in XY chains. J Stat Mech Theor Exp. 2010;2010:P04016. [Google Scholar]