Intermittent many-body dynamics at equilibrium

C. Danieli,1, 2 D.K. Campbell,3 and S. Flach2, 1 1New Zealand Institute for Advanced Study, Centre for Theoretical Chemistry & , Massey University, Auckland, New Zealand 2Center for Theoretical Physics of Complex Systems, Institute for Basic Science, Daejeon, Korea 3Boston University, Department of Physics, Boston, Massachusetts 02215, USA (Dated: September 17, 2018) The equilibrium value of an observable defines a manifold in the of an ergodic and equipartitioned many-body system. A typical trajectory pierces that manifold infinitely often as time goes to infinity. We use these piercings to measure both the relaxation time of the lowest frequency eigenmode of the Fermi-Pasta-Ulam chain (FPU), as well as the fluctuations of the subsequent dynamics in equilibrium. The dynamics in equilibrium is characterized by a power-law distribution of excursion times far off equilibrium, with diverging variance. Long excursions arise from sticky dynamics close to q-breathers localized in normal mode space. Measuring the exponent allows to predict the transition into nonergodic dynamics. We generalize our method to Klein-Gordon lattices (KG) where the sticky dynamics is due to discrete breathers localized in real space.

Equipartition and thermalization have been central re- correlation functions ( which obscure the understanding search topics in many-body interacting systems since the of a detailed correspondence between the equilibrium dy- time of Maxwell, Boltzmann and Gibbs. The first com- namics and coherent structures due to event averaging). puter experiment, aimed to observe equipartition start- ing from a microscopic reversible , was Given a many-body system which possesses linearly carried out in the 1950s by , , stable coherent states, we choose an observable f (i.e., and Mary Tsingou [1]. Now famous as some function of the phase space variables) whose value the Fermi-Pasta-Ulam (FPU) paradox (for reviews see is sensitive to the excitation of such states. We assume [2–5]), this experiment failed to find equipartition but that the many-body system is thermalizing, or ergodic, instead revealed intriguing nonlinear dynamics - includ- i.e. that the phase space trajectory is evolving under the ing the celebrated FPU recurrences [1] - which has chal- constraint of fixed total energy (and perhaps other con- lenged and puzzled researchers for more than 60 years served quantities) such that the time average hfit ≡ hfi (for a recent survey of the state of the art, see [4]). In is independent of the actual chosen trajectory, up to a brief, attempts to understand the full dynamics, includ- set of measure zero (like periodic orbits, which can per- ing the recurrences, led to the observation (and nam- sist even in the strongest chaotic flows). The actual value ing) of solitons [6, 7] and important developments in of f(t) will depend upon time t along a typical trajectory. Hamiltonian chaos [8]. It is now known that these un- As time goes to infinity the trajectory is then forced to expected recurrences are linked to the choice of initial pierce infinitely often a submanifold Ff of codimension conditions used by FPU, which are set close to exact co- 1 which hosts all phase space points with f ≡ hfi. The herent time-periodic (or even quasiperiodic) trajectories, submanifold can be considered as a generalized ergodic e.g. q-breathers, which show exponential localization of Poincar´esection, which is fixed by the choice of f, the in- energy in normal mode space [9, 10]. Even if these trajec- tegrals of motion and the assumption of . The tories have support of measure zero in the phase space, time intervals between consecutive piercings will carry they might have a finite measure impact simply by being the information on whether (and when) the trajectory linearly stable [9]. Several other studies admit coher- was visiting a sticky region in phase space. Hence we will ent time-periodic states localized in real space, which are study the statistics of these time intervals. In contrast to a correlation function, these are the statistics of trackable arXiv:1611.00434v3 [nlin.CD] 12 Feb 2017 known as discrete breathers or intrinsic localized modes [11] and exist e.g. in Klein-Gordon (KG) lattices [12]. events and will always permit us to return to the event These states can also be linearly stable and thus may of interest, in order to inspect it microscopically. With have finite measure impact. Importantly, both discrete this insight we also arrive at a novel quantitative dynam- breathers and q-breathers have been experimentally ob- ical characterization of the degree of equipartition of a served in a large variety of physical settings [11, 13]. Thus given microscopical state, i.e. a point on the considered the central question becomes: How does the presence of trajectory. Rather than using an entropy-like measure such coherent states of measure zero affect the dynami- (e.g. the distance from the set Ff ), it is the time the tra- cal properties of a thermalized many-body system? How jectory needs to reach and pierce Ff which will decide do they affect the possible transition from ergodic to a whether the given configuration is close to or far from non-ergodic dynamics? Interestingly there are only a few equilibrium. recorded numerical attempts to address this complex is- We apply the above ideas to both FPU and KG sys- sue [14–21]. In our view, this is the result of the lack tems with the Hamiltonian function of the canonically of a clear strategy which can go beyond the analysis of conjugated pairs of real space momenta and coordinates 2

1 14 {pn, qn} 0.8 12 N  2  X pn

H = + V (q ) + W (q − q ) , (1) FPU n n+1 n 0.6 10

2 T n=0 η 10 1 α 0.4 8 FPU: V (q) = 0 ,W (q) = q2 + q3 , (2) 2 3 Log 0.2 6 1 1 k KG: V (q) = q2 + q4 ,W (q) = q2 . (3) 2 4 2 0 4 0 2 4 6 8 10 0.001 0.01 0.1 Both models turn into integrable sets of noninteracting Log10t ε normal modes in the limit of vanishing energies. In addi- tion the KG system turns into an integrable set of non- FIG. 1. Left: instantaneous (black) and window-averaged interacting anharmonic oscillators in the limit of diverg- (red) time evolution of the entropy η for  = 0.0566, and ing energies, due to its onsite anharmonicity (as opposed window-averaged time evolution for  = 0.145 (blue),  = to the FPU case). We use fixed boundary conditions 0.0566 (red),  = 0.0204 (green) and  = 0.0091 (magenta). p0 = pN+1 = q0 = qN+1 = 0 for the FPU chain in line Black dashed line: hηi = 0.1218. Right: TFPU (black circles) with Ref. [1],. For the KG chain we use instead periodic vs. . The blue squares are the data from Ref. [22]. The black dashed and dashed-dotted lines guide the eye and indicate a boundary conditions p1 = pN+1, q1 = qN+1 in order to keep all sites equivalent and to avoid edge effects. crossover at  ≈ 0.01. Vertical dotted line:  = 0.0023. To address the normal mode dynamics, we use the FPU system and the canonical transformation Both observables η and P −1 will fluctuate along the tem-   r N     poral evolution of a trajectory. Let us assume that their P 2 X p πnk −1 k = n sin (4) averages hηi, hP i exist and can be computed using the Qk N + 1 qn N + 1 n=1 Gibbs distribution (which follows from well-known gen- eral considerations of counting microstates or maximizing with k = 1,...,N, which diagonalizes the harmonic part the entropy) of H (α = 0 in (2)) with the normal mode momenta and coordinates {P ,Q }. The mode energies and frequencies Z k k 1 −βH −βH are WB = e ,Z = e dΓ . (10) Z Γ 2 2 2   Pk + ωkQk πk Ek = , ωk = 2 sin . (5) Here Γ denotes the whole available phase space, and β is 2 2(N + 1) the inverse temperature. At low enough energies the an- For α 6= 0 the mode energies become time-dependent and harmonic energy contribution for the FPU system will be are monitored using the normalized distribution νk(t) = a small correction and can be neglected when computing PN P averages; its relevance is reduced to the highly important Ek(t)/ k=1 Ek(t), with k νk = 1. A common tool to monitor the degree of inhomogeneity of the distribution nonlinear mode interaction which is the crucial source of is the spectral entropy [22, 23] deterministic chaos and equipartition. The final integra- tion using the Gibbs distribution (10) can be performed N X analytically [24]: S(t) = − νk(t) ln(νk(t)) (6) 1 − γ k=1 hηi = , (11) ln N − S(0) with 0 ≤ S ≤ Smax = ln N. Its rescaled analogue is

S(t) − Smax where γ ≈ 0.5772 is the Euler constant. For the KG η(t) = , 0 ≤ η ≤ 1 . (7) −1 S(0) − S system, we obtain the average hP i directly by numer- max ically averaging until the total integration time T = 108. To address the real space dynamics of the KG system, These averages define the equilibrium manifolds Fη, FP we use the energy densities which we will use for the subsequent analysis. The original FPU computation [1] was performed for p2 k X  = n + V (q ) + W (q − q ) . (8) N = 32 particles with only the lowest frequency mode n 2 n 4 n+s n s=±1 excited, Q1 6= 0 only. Then S(0) = 0 and hηi ≈ 0.1218. We will benchmark our data with the results from [22], An equally common measure of energy distribution in- who used an ad-hoc value η = 0.1. The trajectory starts homogeneity is the participation number P , which yields with η(0) = 1  hηi close to a regular periodic the number of strongly excited renormalized energies PN localized in momentum space (a q-breather) [9]. A cen- µn(t) = n(t)/ n=1 n(t): tral target of many FPU paradox studies was to quantify N the time this initial state needs to reach equipartition, −1 X 2 if it ever does (e.g. [1, 22, 23, 25, 26]). Since equiparti- P (t) = µn(t) , 1 ≤ P ≤ N. (9) n=1 tion means equal mode energies on average, we define the 3

FPU equipartition time TFPU as the time the trajectory 1 needs to reach the corresponding manifold F . We con- 3.2 η 3 -2 tinue our computations beyond this equipartition time. 10 2.8

The trajectory has to cross the manifold Fη infinitely of- δ 2.6 ten, and we record the piercing times t with i ≥ 1 (note -4 i 10 2.4 that TFPU ≡ t1). The return times 2.2

-6 0.01 0.1 tr(i) = ti+1 − ti , i ≥ 1 (12) 10 ε PDF -δ measure the time intervals the trajectory spends off the -8 t equilibrium manifold before piercing it again, with even 10 r and odd integers i discriminating between corresponding -10 excursions into the two different phase space subspaces 10 (e.g. η > hηi and η < hηi). -12 The computations were carried out using a symplectic 10 2 3 4 5 6 SABA2C integrator with corrector [27, 28], with a time -2 step τ = 0.1; these choices keep the relative energy er- 10 ror of the order 10−5 [29]. The system size is N = 32, 2.5 -4 and α = 0.25 in Eq.(2), and initial condition Pk(0) = 0, 10 2.4

Qk(0) = Aδk,1, which translates into a corresponding to- δ 2.3 -6 tal energy E, and energy density  = E/N (see [29] for 10 2.2 details). We follow the time dependence of observables 2.1 and also perform a window averaging over a time window -8 10 1 1.5 2ε 2.5 3 3.5

which is 100 times shorter than the actual running time. PDF -10 In Fig.1 - left plot - we show the time evolution of the 10 entropy η for different energy densities . The curves δ start at the unity at t = 0 (see Eq.(7)) and then settle -12 - 10 t to fluctuating intermediate values for a transient interval r of time that increases as the energy density  decreases. -14 10 Finally, at t = TFPU the observable transits into fluctua- tions around equilibrium at values that approximate the 2 3 4 5 6 7 8 Gibbs average hηi very well. The intermediate plateau Log t corresponds to a metastable state, where all the mode 10 r energies E are non-zero but assume an exponentially k FIG. 2. PDFs P (t ) for  = 0.0566 (FPU - upper plot) decaying profile [4, 26, 30, 31]. The second plateau cor- ± r and  = 1.867 (KG - lower plot). For both FPU and KG, responds to the regime of equipartition, confirming the the red (upper) curve corresponds to P+(tr) and the blue validity of the Gibbs distribution. (bottom) one to P−(tr). The dashed-dotted lines guide the In Fig.1 - right plot - we plot the FPU equipartition eye and indicate the algebraic tails. Inset: the exponent δ of time TFPU as a function of the density , along with the algebraic tails versus the energy density . the data from Ref.[22], which show very good agreement. We also satisfactorily compared our data to the extrap- olated equipartition times from Ponno et al [26], see de- tails in [29]. As noted previously, the equipartition time sets of return times (12) separately for the two differ- increases with decreasing energy density. Casetti et al. ent subspaces η > hηi and η < hηi. The probability predicted the equipartition time at the original FPU en- distribution functions (PDFs) of these sets P±(tr) are ergy density choice of  = 0.00226 to be of the order of shown for  = 0.0566 in Fig.2 - upper plot. In the sub- 12 TFPU ≈ 10 which currently requires about 30 days of space η > hηi, the dynamics exhibits algebraic tails in −δ CPU time with our system [29]. However, the equiparti- the PDF P+(tr) ∼ tr with an exponent 2 < δ < 3, tion time shows a crossover at  ≈ 0.01. which was not which indicates a finite average (1st moment) htri but 2 reached by previous computations. A straight-forward a diverging variance (2nd moment) htri (see [29] on the extrapolation from this crossover (see dashed-dotted line numerical details of estimating δ). The exponent δ de- 14 in In Fig.1 - left plot) increases this time to TFPU ≈ 10 creases with decreasing energy density , signalling the or about 10 years of CPU time on our system. Remark- reaching of the integrable harmonic oscillator chain limit. ably the answer to whether the original FPU trajectory Note that for δ ≤ 2 the average htri would diverge, and is thermalizing or not remains a very hard computational the ergodicity assumption would be violated, again in- problem more than six decades after the first observation diciating the transition into a nonergodic, perhaps inte- of the FPU paradox. grable, case. Therefore, our method is sensitively predict- Let us now turn to the analysis of the equilibrium dy- ing the transition from ergodic to nonergodic dynamics. namics beyond the equipartition time. We compute the In contrast, the subspace η < hηi dynamics yields tails 4 in P−(tr) with finite moments; the tails are faster than algebraic but slower than exponential, presumably expo- nentials dressed with a power law. This is due to that subspace hosting microstates for which the normal modes are even more equipartitioned than on a Gibbs average. Such microstates have small probability, and are insensi- tive for detecting nonequilibrium fluctuations. We extend the analysis of the dynamics at equiparti- tion and the distribution of the return times to the KG chain, a model known to posses discrete breather solution in the real space [12], and should show a related transi- tion to nonergodicity and integrability with increaasing energy density. At variance to the FPU case, we will FIG. 4. C(tr) (left plot) and P (tr)(right plot) for  = 0.0566 search for a gradual loss of ergodicity upon increasing (see text for details). The broad scattering of data is due to the energy density, which should favour the excitations of many independent events yielding similar return times tr. discrete breathers. We choose N = 32, k = 0.1, periodic boundary conditions, and random initial conditions with a predefined energy density. We compute the time evo- ginning of the event we observe the focusing of energy in lution of the participation ratio P until total integration one of the modes (FPU) or sites (KG) respectively. These 10 time T = 10 , and we record the return times tr between breather-like excitations then survive over the entire du- two consecutive piercings of the equilibrium manifold FP ration of the excursion, only to dissolve their energy back again separating the phase space in P −1 > hP −1i and into the other degrees of freedom at the end of the event −1 −1 P < hP i. The PDFs P±(tr) obtained for energy (for further details, see [29]). density  = 1.867 are shown in Fig.2 - lower plot - where, To further substantiate our observation, we show the −1 −1 −δ st for P > hP i, the algebraic tail of P+(tr) ∼ tr is correlation between the 1 moment of the event-averaged N visible while P−(tr) shows exponential cut-off. In the in- P mode energy distribution C = k=1 kνk for the FPU set we plot the exponent δ, which drops below values of case versus the trapping event time tr. In Fig.4 (left δ = 3 and continues to decrease towards the critical case plot), we observe that large return times tr imply large δ = 2 with increasing energy density . Similar to the values of C ≈ N, signalling a tendency towards high FPU case, the KG system dynamics shows a transition frequency excitations. Most importantly the correspond- from ergodic into nonergodic dynamics. ing computation of the participation number P of the event-averaged mode energy distributions in Fig.4 (right Ek εn 0.20 10 plot) shows that large return times correlate with smaller values of P , a typical case of a strongly inhomogeneous 8 0.15 distributions. Therefore the equilibrium FPU dynamics 6 produces sticky excursions with long duration to strongly 0.10 excited high frequency modes. 4 0.05 The properties of fluctuations in equilibrium should 2 not depend on the choice of the trajectory, in accord 0 0 with the assumption of equipartition and ergodicity. We tested that in the FPU chain by launching various other trajectories, e.g. exciting one high frequency mode, or several modes with different frequencies (not shown FIG. 3. Left: mode energies Ek of the FPU as function of time during one of the longest trapping events for  = here). We observed that the statistics of return times is 0.0566. Right: energy densities n of the KG as function of universal and not depending on the choice of the initial time during one of the longest trapping events for  = 1.748. state. Algebraic tails in correlation functions or distributions The algebraic tails of the PDF of the return times of trapping times have been previously studied for low- with 3 > δ > 2 imply that the trajectory is with high dimensional dynamical systems with a mixed phase space probability getting trapped in some parts of phase space [32, 33], and related to the hierarchic structure of for long times, whose average is finite, but whose vari- the phase space at regular island boundaries, similar to ance diverges. We conjecture that these trapping events the phenomenological approach to understand glassy dy- are due to visiting regions of the phase space which are namics. However higher phase space dimensions destroy substantially close to some regular orbits. In order to the simple mixed phase space picture, preventing the use substantiate this conjecture, we show in Fig.3 the time of this simple argument for the observation of algebraic evolution of the mode energies Ek (FPU - left plot) and tails [34]. In the present work we derive a well-defined the energy densities n (KG - right plot) during one of sectioning at equilibrium, and a clear interpretation of their longest excursions far from equilibrium. At the be- the presence of algebraic tails in terms of temporal exci- 5 tation of coherent states, like time-periodic q-breathers. moments of the distribution indicates sticky dynamics. The large phase space dimension does not easily allow When the exponent δ < 3, the non-equilibrium excur- to connect to the physics of glasses, since the potential sions into sticky events start to dominate the dynamics. functions are smooth, and invariant regular trajectories. Finally when δ ≤ 2 the first moment diverges indicat- Instead we are in need of a new understanding how reg- ing the loss of ergodicity altogether. We expect therefore ular states of measure zero (e.g. time-periodic solutions) that our method can be used for a broad set of other can act as dynamical barriers and bottlenecks in high- cases where nonergodic fluctuations affect the dynamics dimensional phase spaces. of many-body systems , such as ultracold atomic gases We arrived at a general method to analyze the re- in optical potentials approximated by the discrete Gross- laxation from non-equilibrium states and the equilib- Pitaevsky equation, or networks of weakly interacting su- rium fluctuations of interacting many-body systems. The perconducting grains, among others. essence is to identify the relevant coherent excitations which will be the cause of stickiness, and to choose a proper observable f which can detect these events. ACKNOWLEDGMENTS The corresponding equilibrium value hfi defines the co- dimension 1 equilibrium manifolds, and the subsequent We thank Peter Jeszinszki and Ihor Vakulchyk for statistical analysis of the distributions of equilibrium fluc- helpful discussions on computational aspects. The au- tuations. When algebraic tails are observed in contrast thors acknowledge financial support from IBS (Project to exponential cutoffs, the divergence of suitably high Code:IBS-R024-D1).

[1] E. Fermi, J. Pasta and S. Ulam, Los Alamos Report LA- [12] M. V. Ivanchenko, O. I. Kanakov, V. D. Shalfeev and 1940 (1955) (unpublished); in E. Segr´e, Collected papers S. Flach, Discrete breathers in transient processes and of Enrico Fermi, University of Chicago 2, 978 (1965). thermal equilibrium, Physica D 198, 120 (2004). [2] Joseph Ford, The Fermi-Pasta-Ulam problem: paradox [13] Mingzhong Wu and Carl E Patton, Experimental ob- turns discovery, Phys. Rep. 213, 271 (1992). servation of Fermi-Pasta-Ulam recurrence in a nonlin- [3] T. P. Weissert, The genesis of simulation in dynamics: ear feedback ring system, Phys. Rev. Lett. 98 047202 pursuing the Fermi-Pasta-Ulam problem, Springer (New (2007); Arnaud Mussot, Alexandre Kudlinski, Maxime York 1997), Droques, Pascal Szriftgiser and Nail Akhmediev, Fermi- [4] G. Gallavotti (Ed.), The Fermi-Pasta-Ulam problem: a Pasta-Ulam recurrence in nonlinear fiber optics: the role status report, Springer (New York 2008). of reversible and irreversible losses, Phys. Rev. X 4, [5] Mason A. Porter, Norman J. Zabusky, Bambi Hu, and 011054 (2014). David K. Campbell, Fermi, Pasta, Ulam and the Birth of [14] Dominique Escande, Holger Kantz, Roberto Livi and Ste- Experimental Mathematics, American Scientist 97, 214- fano Ruffo, Self-Consistent check of the validity of Gibbs 221 (2009). calculus using dynamical variables J. Stat. Phys. 76 605 [6] N. J. Zabusky and M. D. Kruskal Interaction of ”soli- (1994). tons” in a collisionless plasma and the recurrence of ini- [15] G.P. Tsironis and S. Aubry, Slow relaxation phenom- tial states Phys. Rev. Lett. 15 240-243 (1965) ena induced by breathers in nonlinear lattices, Phys. Rev. [7] N. J. Zabusky and G. S. Deem Dynamics of nonlinear Lett. 77, 5225 (1996). lattices I. Localized optical excitations, acoustic radiation, [16] K.O. Rasmussen, T. Cretegny, P.G. Kevrekidis and N. and strong nonlinear behavior J. of Comp. Phys. 2 126- Gronbech-Jensen, Statistical mechanics of a discrete non- 153, (1967). linear system, Phys. Rev. Lett. 84, 3740 (2000). [8] F. V. Izrailev and B.V. Chirikov Statistical properties of [17] M. Eleftheriou, S. Flach and G. P. Tsironis, Breathers in a nonlinear string, Soviet Physics Doklady 11 (1) 30-32 1d nonlinear thermalized lattice with an energy gap, Phys- (1966). ica D 186, 20 (2003); M. Eleftheriou and S. Flach, Dis- [9] S. Flach, M. V. Ivanchenko and O. I. Kanakov, q- crete breathers in thermal equilibrium: distributions and Breathers and the Fermi-Pasta-Ulam problem, Phys. energy gaps, Physica D 202, 142 (2005); M. Eleftheriou Rev. Lett. 95, 064102 (2005); q-Breathers in Fermi- and S. Flach,Interaction of discrete breathers with ther- Pasta-Ulam chains: existence, localization and stability, mal fluctuations, Low Temp. Phys. —bf 34, 554 (2008). Phys. Rev. E 73, 036618 (2006). [18] Hirinori J. Matsuyama and Tetsuro Konishi, Multistage [10] H. Christodoulidi, C. Efthymiopoulos and T. Bountis, slow relaxation in a Hamiltonian system: The Fermi- Energy localization on q-tori, long term stability and the Pasta-Ulam model, Phys. Rev. E 92, 022917 (2015). interpretation of the FPU paradox, Phys. Rev. E 81 [19] Zhenjun Zhang, Chunmei Tang and Peiqing Tong, Dy- 016210 (2010). namical thermalization of Frenkel-Kontorova model in [11] S. Flach, and C.R. Willis Discrete breathers, Phys. Rep. the thermodynamic limit, Phys. Rev. E 93, 022216 295 181 (1998); S. Flach and A. Gorbach, Discrete (2016). breathers: advances in theory and applications, ibid. 467 [20] Colm Mulhern, Stephan Bialonski and Holger Kantz, Ex- 1 (2008). David K. Campbell, Sergej Flach, and Yuri treme events due to localization of energy, Phys. Rev. E S. Kivshar Localizing Energy Through Nonlinearity and 91, 012918 (2015). Discreteness, Physics Today (January 2004) 43-49. [21] Boris Gershgorin, Yuri V. Lvov and David Cai, Renor- 6

malized Waves and Discrete Breathers in β-Fermi-Pasta- [29] See Supplemental Material at URL to be defined. Ulam Chains, Phys. Rev. Lett. 95, 264302 (2005). [30] G. Benettin and A. Ponno, Time-scales to equipartition [22] L. Casetti, M. Cerruti-Sola, M. Pettini and E.G.D. Co- in the Fermi-Pasta-Ulam problem: finite-size effects and hen, The Fermi-Pasta-Ulam problem revisited: stochas- thermodynamic limit, J. Stat. Phys. 144, 793 (2011). ticity thresholds in nonlinear Hamiltonian systems, Phys. [31] F. Fucito, F. Marchesoni, E. Marinari, G. Parisi, L. Peliti, Rev. E, 55 6566 (1997). S. Ruffo and A. Vulpiani, Approach to equilibrium in a [23] Miguel Onorato, Lara Vozella, Davide Proment and Yuri chain of nonlinear oscillators, J. Phys. France 43, 707- V. Lvov, Route to thermalization in the α-Fermi-Pasta- 713 (1982). Ulam system, PNAS 112, 4208 (2015). [32] B. V. Chirikov and D. L. Shepelyansky, in Proceedings of [24] C.G. Goedde, A.J. Lichtenberg, M.A. Lieberman, Chaos the IXth International Conference on Nonlinear Oscilla- and the approach to equilibrium in a discrete sine-Gordon tions, Kiev, 1981; [Naukova Dumka 2, 420 (1984)]; (En- equation, Physica D 59, 200-225 (1992). glish translation: Statistics of Poincare recurrences and [25] J. De Luca, A. J. Lichtenberg, and S. Ruffo, Finite times the structure of the stochastic layer of a nonlinear reso- to equipartition in the thermodynamic limit, Phys. Rev. nance, Princeton University Report No. PPPL-TRANS- E 60, 3781 (1999). 133, 1983). [26] A. Ponno, H. Christodoulidi, Ch. Skokos and S. Flach, [33] C. F. F. Karney, Long-time correlations in the stochastic The two-stage dynamics in the Fermi-Pasta-Ulam prob- regime, Physica D 8, 360 (1983). lem: from regular to diffusive behavior, Chaos 21, 043127 [34] M. Ding, T. Bountis, and E. Ott, Algebraic escape in (2011). higher dimensional hamiltonian systems, Phys. Lett. A [27] J. Laskar and P. Robutel, High order symplectic inte- 151, 395 (1990); E. G. Altmann and H. Kantz, Hypoth- grators for perturbed Hamiltonian systems, Celest. Mech. esis of strong chaos and anomalous diffusion in coupled Dyn. Astron. 80, 39 (2001). symplectic maps, Europhys. Lett. 78, 10008 (2007); S. [28] Ch. Skokos, D.O. Krimer, S. Komineas and S. Flach, Lange, A. B¨acker and R. Ketzmerick, What is the mech- Delocalization of wave packets in disordered nonlinear anism of power-law distributed Poincar´erecurrences in chains, Phys. Rev. E 79, 056211 (2009). higher-dimensional systems?, arXiv:1609.07977 (2016). 7

SUPPLEMENTAL MATERIAL

I. THE SABA2C SYMPLECTIC INTEGRATIOR

To integrate the FPU and the KG chains in time, we use a symplectic integration scheme SABA2C proposed in [27], which is an improved version of the SABA2 scheme. The SABA2 scheme consists in separating the Hamiltonian tLA tLB H in two integrable parts H = A + B, each one with solutions e and e . The SABA2 scheme of integrating the coupled differential equations of motion is defined as

c1τLA d1τLB c2τLA d1τLB c1τLA SABA2 = e e e e e (13) where c = 1 1 − √1  and c = √1 while d = 1 . The corrector term eτLC is the solution of the Hamiltonian 1 2 3 2 3 1 2 C = {{A, B},B}, which adds two more matrix operations:

3 3 −(τ g/2)LC −(τ g/2)LC SABA2C = e (SABA2)e (14) √ 2− 3 where g = 24 . Further details can be found in [27, 28]. In the FPU case, the Hamiltonians A and B are respectively

N 2 N  2 3  X p X (qn+1 − qn) (qn+1 − qn) A = n ; B = + α ; (15) 2 2 3 n=0 n=0 while the Hamiltonian C is

N X h  i2 C = 2qn − qn+1 − qn−1 1 + α(qn+1 − qn−1) . (16) n=0

The operators eτLA , eτLB and eτLC propagate the coordinates x(t) = (p(t), q(t)) = (p, q) at time t to the coordinates x(t + τ) = (p(t + τ), q(t + τ)) = (p0, q0) at time t + τ:

 0 τLA qn = pnτ + qn e : 0 ; pn = pn (17)  0 τLB qn = qn e : 0    ; pn = qn+1 + qn−1 − 2qn 1 + α qn+1 − qn−1 τ + pn

 0  qn = qn  0 n  2     p = 2 2 q2 − 2qn 1 + αq2 − q1 1 + αq1 1 + 2αq1  1    o  − q3 + q1 − 2q2 1 + α q3 − q1 1 − 2α q1 − q2 τ + p1   0 n  2  pn = 2 2 qn+1 + qn−1 − 2qn 1 + α qn+1 − qn−1 n = 2,...,N − 1 eτLC :    (18)  − qn+2 + qn − 2qn+1 1 + α qn+2 − qn 1 − 2α qn − qn+1    o  − qn−2 + qn − 2qn−1 1 + α qn − qn−2 1 − 2α qn−1 − qn τ + pn   0 n  2     p = 2 2 qN−1 − 2qN 1 − αqN−1 − qN 1 − αqN 1 − 2αqN  N    o  − qN−2 + qN − 2qN−1 1 + α qN − qN−2 1 − 2α qN−1 − qN τ + pN

A detailed presentation of the symplectic integration scheme SABA2C can be found in the appendix of [28]. In our simulations we used τ = 0.1 for both FPU and KG. With this time step, the SABA2C yields a relative error of the energy density ∆ = |(t) − (0)|/(0) of the order of 10−5 (Fig.5). In KG case (right plot), we notice that even if it remains well below 10−5, the energy error ∆ shows an increase of the order of its fluctuations.

II. RELATION BETWEEN THE INITIAL CONDITIONS AND THE ENERGY DENSITY 

The initial condition of the FPU system chosen in our work corresponds to the FPU choice [1]. It consists in exciting the lowest frequency mode in the space coordinate qn while keeping the conjugate momenta pn identically 8

-5 -5 -5 -5 8x10 8x10 1x10 2x10 (a) (b) (e) (f) 6 6 0.75 1.5 4 4 0.5 1 ∆ε ∆ε 2 2 0.25 0.5 0 0 0 0 0 2 4 6 8 10 0 2 4 6 8 10 0 2 4 6 8 10 0 2 4 6 8 10 -5 -5 -5 -5 8x10 8x10 2x10 6x10 (c) (d) (g) (h) 6 6 1.5 4 4 4 1 ∆ε ∆ε 2 2 2 0.5 0 0 0 0 0 2 4 6 8 10 0 2 4 6 8 10 0 2 4 6 8 10 0 2 4 6 8 10 Log10t Log10t Log10t Log10t

FIG. 5. Left plot: Time evolution of the relative energy error ∆ for the FPU chain, with  = 0.0091 (a),  = 0.0204 (b),  = 0.0566 (c) and  = 0.145 (d) and time step τ = 0.1. Right plot: Time evolution of the relative energy error ∆ for the KG chain, with  = 1.371 (e),  = 1.629 (f),  = 1.897 (g) and  = 3.501 (h) and time step τ = 0.1. zero:  πn  q (0) = A sin , p (0) = 0 . (19) n N + 1 n

The total energy E = Hα(pn(0), qn(0)) reads

A2ω2  π  E = N 1 , ω = 2 sin . (20) 4 1 2(N + 1) which yields the energy density 

E A2ω2  = = 1 , (21) N 4 In order to help the interested reader to compare data from different publications using different notations, we show in Tab.I the values of the energy density  we have considered in our work, and their corresponding values of the initial amplitude A:

Amplitude Energy density A  1.0 0.0023 1.5 0.0051 1.6 0.0058 1.7 0.0065 1.8 0.0073 1.9 0.0082 2.0 0.0091 2.5 0.0142 3.0 0.0204 4.0 0.0362 5.0 0.0566 6.0 0.0815 7.0 0.111 8.0 0.145 9.0 0.183

TABLE I. Conversion table between the values of the amplitude A and the energy density  for the FPU chain. 9

For the KG chain, we choose random initial conditions at all sites

pn(0) = c · ξ2n−1 , qn(0) = c · ξ2n , ξn ∈ [−0.5, 0.5] ; (22) where ξn are uniformly distributed random values over the interval [−0.5, 0.5], and c > 0. The energy density  = E/N depends on the choice of the parameter c and different disorder realizations.

III. COMPARISON OF TFPU WITH THE DATA FROM PONNO ET AL. [26]

In Ref.[26], the nonlinear parameter is α = 0.33. A simple rescaling relates the corresponding data to our results obtained for α = 0.25. Consider (pn(t), qn(t)) to be a solution of the FPU chain for a given α. Consider a different nonlinear parameterα ¯. Given the ratio ϑ =α/α ¯ , we define a rescaling of the solutions

p (t) q (t) p¯ (t) = n ;q ¯ (t) = n . (23) n ϑ n ϑ It follows that this trajectory solves the FPU Hamiltonian equations withα ¯ as the nonlinear parameter. The amplitude of the initial state is rescaled as A¯ = A/θ and the energy (and the energy density) is rescaled as E¯ = E/θ2. In Ref.[26]

9 1000 8

FPU 7 T 10 6 Log

FPU 5

T 4 10 100 -0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 Log10E Log

10

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 E

FIG. 6. Equipartition time Log10TFPU on a logarithmic scale vs. energy E = N. Inset: equipartition time Log10TFPU vs. Log10E. Blue data courtesy of H. Christodoulidi (published in [26]) obtained for α = 0.33. Red data computed by us for α = 0.25. System size N = 31. a system size of N = 31 was used; thus we recomputed our equipartition times TFPU for that case. The corresponding data in Ref.[26] were obtained from an extrapolation of the energy flow from strongly to weakly excited normal modes. In Fig.6 we compare our equipartition times TFPU for α = 0.25 with those from Ponno et.al. for α = 0.33. The blue triangles (Ponno et.al. data) and the red squares (our data) are in very good agreement.

IV. CPU TYPE

Numerical simulations presented in the paper has been performed on the Massey Cluster Simurg, which uses Intel Xeon CPU E5-2670 processors; and on the PCS-IBS Cluster, which uses Intel E5-2680v3 processors.

V. STICKINESS TO REGULAR ORBITS

In the upper row of Fig.7 we show the time evolution of the observables η and P in correspondence of the excursion far from equilibrium shown in Fig.3 of the main text for the FPU (left plots) and the KG (right plots). At the beginning and the end of the return times marked by the blue ticks in Fig.3 of the main text, the formation and the 10 dissolution of the inhomogeneous distributions takes place. In the lower plot of Fig.7 we plot the time average of the energies over the return times

1 Z ti+1 hEkii = Ek(t)dt , k = 1,...,N tr(i) ti (24) 1 Z ti+1 hnii = n(t)dt , n = 1,...,N. tr(i) ti These averages highlight that the inhomogeneous excitations shown in Fig.3 of the main text are not sparse peaks, which otherwise would be erased on time average, but instead are consistent breather-like excitations that persist for the whole time of the excursion. 20 0.3 η 0.2 15 P

0.1 10 7 7 8 8 7.87x10 t 7.88x10 7.898x10 t 7.9x10 0.3 15

0.2 10 i i > > n k ε

0 0 0 10 20 30 0 10 20 30 k n

FIG. 7. Upper plots: Time evolution of the observables η (left) and P (right) of the trapping events shown in Fig.3 of the main text for FPU (left) and KG (right). The red dashed lines correspond to the equipartition values of the entropy hηi = 0.1218 and hP i = 16.7473, and the blue ticks mark the consecutive crossings of the manifolds Fη and FP respectively. Bottom plots: The mode energies hEkii (left) and energy densities hnii (right) averaged over the two events in the upper plots versus mode number k and lattice site n respectively.

VI. FITTING δ IN THE TAILS OF P+(tr)

We obtain the PDF using bins equispaced on a logarithmic scale. This choice allows a more precise evaluation of the exponent δ. Note that the conversion to the correct equidistant binning on a linear scale changes the exponent by 1. The correct function is obtained by multiplying the log-binned PDF by 1/tr. We estimate the exponent δ by a power-law regression of the function P±(tr) performed on intervals of decreasing length I in the PDF tail, by fixing MAX MIN the upper interval edge tr and varying the lower edge tr . In Fig.8 we show the computations of the estimates for the FPU chain for energy densities  ranging from 0.0204 to 0.145. In Fig.9 we show the computations of the estimates for the KG chain for energy densities  ranging from 1.184 to 3.166. 11

3.15 (a) 2.9 (b) 2.35 (e) 2.3 (f) 3.1 2.25 δ 2.85 δ 2.3 3.05 2.2 2.8 2.25 3 2.15 18000 19000 20000 25000 27500 30000 60000 80000 100000 950000 1000000 2.75 I (c) 2.5 (d) (g) 2.7 2.3 2.65 2.25 δ 2.45 δ 2.6 2.2 2.55 2.4 2.15 40000 45000 50000 90000 100000 950000 1000000 I I I

FIG. 8. Estimating the exponent δ for  = 0.145 (a),  = 0.111 (b),  = 0.0815 (c),  = 0.0566 (d),  = 0.0362 (e),  = 0.0204 (f) and  = 0.0091 (g). The blue lines indicate the average values of the exponent δ, while the red shaded areas denote the standard deviations.

2.55 (a) (b) 2.2 2.4 (e) 2.3 (f) 2.5 2.15 2.2

δ 2.45 2.3 δ 2.1 2.4 2.1 2 2.2 1.9 2.35 2.05 9600 9800 10000 900000 1000000 9800000 10000000 500000 1000000 I 2.3 (c) 2.2 (d) 2.15 (g) 2.1 δ

2.25 2.15 δ 2.05

2.2 2.1 2 998000 1000000 5980000 6000000 1800000 1900000 2000000 I I I

FIG. 9. Estimating the exponent δ for  = 1.184 (a),  = 1.303 (b),  = 1.416 (c),  = 1.867 (d),  = 2.149 (e),  = 2.655 (f) and  = 3.166 (g). The blue lines indicate the average values of the exponent δ, while the red shaded areas denote the standard deviations.