Thermalisation of a trapped one-dimensional Bose gas via diffusion

Alvise Bastianello,1 Andrea De Luca,2 Benjamin Doyon,3 and Jacopo De Nardis4 1Institute for Theoretical , University of Amsterdam, Park 904, 1098 XH Amsterdam, The Netherlands 2Laboratoire de Physique Th´eorique et Mod´elisation(CNRS UMR 8089), Universit´ede Cergy-Pontoise, F-95302 Cergy-Pontoise, France 3Department of Mathematics, King’s College London, Strand WC2R 2LS, London, U.K. 4Department of Physics and Astronomy, University of Ghent, Krijgslaan 281, 9000 Gent, Belgium For a decade the fate of a one-dimensional gas of interacting bosons in an external trapping potential remained mysterious. We here show that whenever the underlying integrability of the gas is broken by the presence of the external potential, the inevitable diffusive rearrangements between the quasiparticles, quantified by the diffusion constants of the gas, eventually lead the system to thermalise at late times. We show that the full thermalising dynamics can be described by the generalised hydrodynamics with diffusion and force terms, and we compare these predictions with numerical simulations. Finally, we provide an explanation for the slow thermalisation rates observed in numerical and experimental settings: the hydrodynamics of integrable models is characterised by a continuity of modes, which can have arbitrarily small diffusion coefficients. As a consequence, the approach to thermalisation can display pre-thermal plateau and relaxation dynamics with long polynomial finite-time corrections.

In the past two decades, low- gases of of the external force that breaks integrability and the dif- bosonic or fermionic atoms emerged as the best exper- fusion spreading that redistributes the quasiparticle mo- imental platform where to study many-body phenomena menta and increases the local thermodynamic . and where to probe their non-equilibrium dynamics [1–3]. Previous works have studied the hydrodynamics of an A great deal of attention has been dedicated to the prob- integrable gas in the presence of the external force but lem of thermalisation in isolated systems, such as a single without accounting for diffusive terms [20–24]. Here, we gas of bosonic atoms interacting via contact repulsion. show that diffusive corrections play an essential role in The experiment of 2007 dubbed “quantum Newton cra- the late-time dynamics. dle” [4] represented a turning point. The apparent lack Lieb-Liniger gas in an external potential. Our of thermalisation at late times for a non-equilibrium one- model of interest is the Lieb-Liniger model for a gas of dimensional cold atomic gas in an external trap started bosons interacting with contact repulsion. The Hamilto- numerous research directions on the role of the under- nian is given by lying integrability in the isolated dynamics. It was then 2 later clarified that integrable models do not thermalise to ~ † † † HˆLL = dx ∂xψ (x)∂xψ(x)+g ψ (x)ψ (x)ψ(x)ψ(x), standard Gibbs ensembles, but to Generalises Gibbs ones ˆ 2m [1,5,6], where entropy is maximised given the constraints (1) † of all local and quasi-local conserved quantities. As the with corresponding bosonic operators [ψ (x), ψ(y)] = integrability is weakly broken by the external trap, one δ(x − y). In the following we set the mass of the bosons 2 expects therefore to crossover from a fast relaxation to a such that ~ /2m = 1 and g = c. The model is known to generalised Gibbs state, to a slow pre-thermal decay to describe state-of-the-art cold atom experiments [4, 25– the final thermal ensemble. In the past years, the study of 32] and also to be integrable [33]. As a consequence, all pre-thermalisation dynamics has received many contribu- physical initial states [34–39] eventually relax to a GGE − P Qˆ β tions, and for homogeneous systems, a quite comprehen- density matrix% ˆ ∼ e j j j , where the set of Lagrange sive understanding has now been achieved [7–17]. How- multipliers βj is fixed by the initial condition. ever, a full description of the (pre) thermalisation mech- Its large scale dynamics can be accurately described anism in inhomogenous systems is still lacking. While via the so-called Generalised Hydrodynamics (GHD), in- a recent experimental work showed that multiple one- troduced in the past years [40, 41] and the resulting arXiv:2007.04861v2 [cond-mat.quant-gas] 23 Jul 2020 dimensional gases coupled to each other or where tran- predictions have even received an experimental confir- sitions to higher dimension are allowed, do indeed ther- mation [31]. In contrast with standard hydrodynam- malise at late times [18, 19], it remains to be understood ics, GHD is built on the infinite number of local and ˆ what is the fate of a single isolated one-dimensional gas. quasi-local conserved quantities Qj = dx qˆj(x) charac- terising integrable systems. Equivalently,´ each stationary In this letter, we show that in the presence of an ex- state of the system can be described by a set of stable ternal force and of diffusive spreading, a single gas de- quasiparticles with density ρθ, in one-to-one correspon- scribed by an integrable Hamiltonian and by an external dence with the expectation values of conserved quantities trapping potential does thermalise, due to the interplay dθρθ hj,θ = hqˆji, with hj,θ, the single-particle eigenval- ´ 2

interacting systems, describes both a diffusive spreading of the quasiparticle motion and a re-distribution of quasi- particle velocities. Both quantities are non-linear func- tional of the densities ρθ(x, t), fixed by the equation of state and by the explicit form of the diffusion kernel. The effect of an external trap is described by the fol- lowing Hamiltonian

Hˆ = Hˆ + dx Vˆ (x) (2) trap LL ˆ

where the potential Vˆ (x) = V (x)ˆq0(x) and the parti- † cle densityq ˆ0(x) = ψ (x)ψ(x). The contribution of the trapping potential to the dynamics of the quasiparticle densities can be treated perturbatively in the degree of smoothness of the external potential V (x). It was first shown in [20] that at the Euler scale this amounts to adding the simple force term (∂xV (x))∂θρθ. We here

FIG. 1. Left: Quasiparticle density ρθ(x, t) at t = L claim that the addition of this single term also gives the for a Lieb-Liniger gas (2) inside the trap V (x) = 1/2(1 − correct hydrodynamic equation up to diffusive scales: cos 2πx/L), in a circle with length L = 1 (top) and L = 25.6 eff α (down) (see text for other parameter’s values). Right: Log- ∂tρθ = −∂x(vθ ρθ) + ∂x(Dθ ∂xρα) + (∂xV (x))∂θρθ. (3) plot of the thermodynamic entropy density S(t)/L (top) con- verging to the thermal entropy for the given quench and of That is, this equation holds up to order O(L−3) cor- 4 L † 2 the integrated correlator hh|ψ |ii = 0 dxh(ψ ψ) i/L. The rections, with L the minimal length of spatial varia- inset is a log-plot to show the approach´ to the thermal value tions of the system and of the external potential V (x). at late times. From a scaling analysis, at this diffusive order possible corrections to Eq. (3) could come from three different 2 sources: i) second derivative in the potential ∂xV (x); ii) second power of its first derivative (∂ V (x))2; and ues of each quasiparticle with rapidity θ. x iii) mixed derivatives, ∂xV (x)∂xρα. We now argue that Given the derivative of the scattering shift between two none of these terms contributes. First, we recall that quasiparticles with different rapidities T , any single- θ,α the definitions of densities of chargesq ˆj(x) are ambigu- particle function hj,θ is dressed via the linear transfor- ous by the addition of total derivative terms, and as dr −1 0 maton hj = (1 − T n) · h, with nθ = 2πρθ/(kθ) the shown in [49], we can make use of this freedom to en- fermionic filling function and (1−T n)θ,α = δθ,α −Tθ,αnα force PT (parity and time) symmetry. With PT sym- the dressing matrix (here and in the following we de- metry, we can argue that terms proportional to ∂2V 0 x note the derivative respect to rapidities as ∂θhθ = hθ). are excluded in (3). Specifically, consider the Heisen- The effective quasiparticle momentum kθ and energy berg time-evolution equation for any conserved density ε are obtained by integrating the dressed derivatives: θ qˆj(0), say at the origin. Taylor expanding the potential, 0 −1 bare 0 k = (1 − T n) · (k ) (and similarly for the en- V (x) = V (0) + xV 0(0) + O(x2), we argue that the O(x2) ergy). For the LL model, T = 2c/((θ − α)2 + c2), and 2 θ,α does not contribute to Eq. (3), as dx x [ˆq0(x), qˆj(0)] bare 0 bare 0 (kθ ) = 1, (εθ ) = 2θ. Because of the interactions, has a vanishing expectation on any´ stationary state by the effective energy and momentum of quasiparticles de- PT symmetry. Second, no mixed derivatives appear as pend non-trivially on the local thermodynamic state via the charge densityq ˆ0(x) is ultra-local. That is, in terms the dressing of excitations. of the field ψ(x) it does not contain any derivative, and Without external forces, the large scale evolution of as a consequence the number of particles does not gen- any local density of conserved quantities can be writ- erate any flow in space [ dxqˆ0(x), qi] = 0. This implies ten in terms of the evolution of the local density of that diffusion is not modified´ by the presence of the ex- quasiparticles ρθ(x, t). This involves a convective (Eu- ternal potential. Third, there may be additional Fermi- ler scale) and a diffusive part [42–46] and it reads ∂tρθ = golden-rule terms [13–15] leading to additional thermali- eff α −∂x(vθ ρθ) + ∂x(Dθ ∂xρα), where the repeated rapidity sation effects. These arise from a second-order perturba- index α is summed over (summation over repeated in- tion theory analysis with integrability-breaking pertur- dices is assumed in what follows, and includes an integra- bations. In our case, again concentrating on the ori- eff 0 0 tion over velocities). The effective velocity vθ = εθ/kθ gin, the leading integrability-breaking perturbation is 0 gives the convective motion of densities [47, 48] while V (0)B where B = dx xqˆ0(x). Fermi golden rule gives the non-diagonal diffusion kernel, which vanishes in non- terms proportional´ to (V 0(0))2, of the time-integrated 3

t eff 0 −1 −1 bare 0 form −t dsh[B(s), ··· ]i, evaluated in a stationary, homo- vθ = (kθ) [(1 − T n) (ε ) ]θ. Then one proceed geneous´ state. As, in the present case, B is the Galilean to notice that the diffusive term in (3) produces an in- boost, dB(s)/ds = P is the total momentum, thus, up to crease of the total thermodynamic entropy of the state t S = dx P β (x)hqˆ (x)i−f, with f being the total free terms which vanish at large t, this is −t ds sh[P, ··· ]i = 0 j j j by homogeneity. Any integrable model´ with these prop- energy.´ The entropy increase on a generic GGE state was erties is expected to obey (3), which is the focus of this found in [49, 52] and it can be written as letter. The above is not a full derivation, but we will ad- dS(t) 1 X = dx (∂ β (x))hθ(DC) hα(∂ β (x)) dress this, along with the general case, in an upcoming dt 2 ˆ x j j θ,α k x k publication. j,k (4) Thermalisation via diffusion. Canonical ther- with the matrix of susceptibilities [RCR−1] = malisation to a Gibbs ensemble (GE) can only be reached θ,α δ ρ (1 − n ). As the only zero eigenvalue of the D at late times by breaking the conservation of all Qˆ ex- θ,α θ θ j operator is the unit vector, the only states where en- cept for the total energy Qˆ1 + dxVˆ (x) and the total tropy does not increase are the ones where all βj>0 ˆ † particle number Q0 = dxψ ψ´, and by ensuring that are constant in space, namely any GGE states with the dynamics is irreversible.´ an arbitrary spatially modulated chemical potential ∼ Pm ˆ It can be shown that Gibbs ensembles are stationary exp(− j=1 Qjβj − dxµ˜(x)ˆq0(x)), with all βj constants solutions of the purely convective flow – the first and in space. However,´ for generic m > 1 and genericµ ˜(x), third term on the right-hand side in (3)[20, 50]. How- these states are not invariant under the convective part ever, such an evolution is completely reversible [21, 51] of (3) and therefore not stationary, leaving the thermal and does not generate entropy. In practice, the root den- stateµ ˜ = µ − V (x) and β1 = β as the only stationary sity ρθ is roughened by the convective evolution and can- solution [50]. not relax. Relaxation may happen in a weak sense as lo- Numerical solutions of the hydrodynamic equa- cal observables may become stationary [21, 51] (see also tion. The equation (3) is a typical convection-diffusion Fig.1d). The determination of such a steady state is equation. A common way to numerically solve these a non-trivial task. Physically, however, diffusion occurs, types of equation is the Crank-Nicolson method [53], which modifies this picture. Diffusion – the second term which is based on the average of the forward and back- in (3) – increases entropy while preserving all conserved word Euler step and is exact up to the second order in the quantities. We will now show that the combined effect time step. We apply this method to Eq. (3) and we solve of diffusion together with the inhomogeneous potential the mid-point problem iteratively. More specifically, let V (x) induces thermalisation. us define the right hand side of (3) as ∂tρθ(x, t) = In order to proceed we shall first write down the form Fθ[ρα(x, t)]. Note that the function Fθ is a non-linear −1 (k) of the diffusion matrix D. This is given by D = R De R, functional of ρα(x, t). Let us denote as ρθ (x, t + ∆t)  0 2 kγ the approximation of ρθ(x, t + ∆t) after k iterations. For where De θ,α = δθ,α dγ k0 Wγ,α − Wθ,α and where α the first iteration (k = 1) we use the forward Euler ´ 0 the matrix of dressings Rθ,α = 2π(δθ,α − nθTθ,α)/kθ. (1)  2 step ρθ (t + ∆t)(x, θ) = ρθ(x, t) + ∆tF [ρα(x, θ)]. Then dr 0 eff (k+1) (k) The function Wθ,α = 1/2 2πTθ,α/kθ ρθ(1 − nθ)|vθ − we obtain ρθ (x, t + ∆t) from ρθ (x, t + ∆t) using eff dr h (k) i v |, with the dressed scattering kernel T = (1 − (k+1) ρα(x,t)+ρα (x,t+∆t) α ρ = ρθ(x, t) + ∆tFθ . Conver- −1 θ 2 T n) · T . The kernel De θ,α describes a Markov pro- (k+1) (k) cess of random momentum exchanges via 2-body col- gence is reached when ||ρθ − ρθ ||  δ, with δ some lisions. However the total quasiparticle momentum is accuracy threshold. The algorithm is very stable and α θ α preserves total energy with high precision. The role of conserved, since dθDθ = 1 Dθ = 0, as it follows θ −1 0 2 the diffusion term in (3) is to smoothen the functions from 1 [R ]θ,α =´ (kα) /(2π). The simple fact that the ρθ(x, t) both in space and, primarily, in θ space, see Fig. unit vector 1θ is the only zero eigenvector of the dif- fusion operator D ensures that the dynamics induced 1, a fact that was previously noticed in [54] (see also by Eq. (3) relaxes to a thermal ensemble. Namely the [21, 22]). Therefore, the diffusion contribution not only is a necessary physical term to describe the hydrodynam- large time limit of ρθ(x, t) in (3) is such that, for any ics at order ∂2, but it constitutes the proper regulator for j, limt→∞ dθρθ(x, t)hj,θ = Tr[ˆqj(x)ˆ%] with the termal x ´  ˆ ˆ  any solution for the evolution of ρθ(x, t), by smoothing density matrix% ˆ ∝ exp −β(Htrap − µQ0) . The final out all microscopic configurations. However, whenever temperature β and chemical potential µ are fixed by the the ratio between the convective terms and the diffusive initial total density and energy. one is large (equivalently to a large Reynolds number in To proof this fact, first one notices that the total classical hydrodynamics) the profile of the fluid in (x, θ) density N = dx dθρθ(x, t) and total energy E = space becomes rapidly varying, and numerical instabili- bare dx dθρθ(x, t´)(εθ ´ + V (x)) are preserved by evolu- ties, mostly due to the discretisation of the derivatives tion´ ´ (3). The latter can be easily shown using that ∂θ, inevitably appear. 4

Results in the Lieb-Liniger gas. We first study the following non-equilibrium settings: we consider pe- riodic boundary conditions in x ∈ [0,L]. We prepare the gas in the thermal state with inverse temperature β = c = 0.3, with c the coupling strength in (1), and with chemical potential given by µ − V0(x), with µ = 2.5 + and V0 = 2 sin(2πx/L) and at t = 0 we switch on the potential V (x) = 1/2(1 − cos(2πx/L)). In the equation (3) one is free to rescale x → xL and t → tL, this way the convective and the force term remain unaltered, but the diffusive part is rescaled by a factor 1/L. In Figure1 we show the increase of the total entropy after the quench in the cases L = 1 and L = 25.6. For the large value of L indeed we observe a pre-thermal plateau for the evolution of local observables, see Fig.1. As discussed above, such a plateau corresponds to the stationary state attained by the purely convective evolution. The slow drift can be in- terpreted as the effect of the spatial imbalance of modes FIG. 2. Study of the dynamics predicted by Eq. (3) for a release of a cold atomic gas into a harmonic potential V (x) = with small diffusion constant, which produces a diffusive 2x2 from a double well potential (see text for all parameters’ small entropy increase in (4), and which is even more values). (a) Plot of the density profile hqˆ0(x, t)i as function of suppressed by larger values of L. We indeed stress that, t and x and (b) at different times compared with its value in even if the unit vector 1θ is the only zero eigenvector of the the thermal Gibbs ensemble. (d) Global thermodynamic the diffusion matrix, there exist a continuum of modes, entropy as function of time reaching the thermal value at late namely the ones associated to quasiparticles with large times. (c) Cartoon picture of the experimental setting. momentum, with small diffusion constants. Such pre- thermal plateau was indeed also observed in numerical simulation of a trapped hard-rods gas in [22] and it is here therefore theoretically explained. mass such that ~2/(2m) = 1, in the limit ~ → 0 the quan- We then proceed to study a non-equilibrium setting tum system is well described by the classical non-linear which is more relevant for cold atoms experiments, see Schr¨odingermodel (NLS)[58] (see also [59–61]), where Fig.2. We prepare a gas of bosons in a double-well extensive numerical simulations can be carried [62–67]. potential as done in Ref. [31]. We consider an initial We numerically solve (3) for the Lieb-Liniger gas and we gas with c = 0.5 at thermal equilibrium with temper- extrapolated the behaviour at ~ = 0 to compare with ature T = 10 and µ = 4 inside the double-well trap numerical Monte-Carlo simulations of the NLS equation 4 2 3 V0(x) = x − 2x + x /10. We then release the gas in [68], see Fig.3. We consider the same quench of Fig.1 the harmonic trap V (x) = ωx2/2 with ω = 4. The initial with L = 25.6, in order to suppress extra effects to the density and temperature at the minima of the trap V0(x) dynamics due to corrections in (3) with higher derivative −2 −1 is such that c/n ∼ O(10 ) and T/Td ∼ O(10 ), with terms. Due to the large value of L the thermalisation 2 quantum degeneracy temperature [55] Td = n (in units time-scales are large, see Fig.1. Nevertheless we observe 2 ~ /2m = 1 and Boltzmann’s constant kB = 1), in order a small drift towards the thermal value for the profile of to be within typical experimental regimes. bosonic density, correctly reproduced by Eq. (3). By simulating the dynamics of the gas by (3) we find Conclusion. In this letter we have shown that the that diffusion eventually suppresses the oscillations of the hydrodynamic equation (3) including diffusion and ex- gas inside the harmonic trap and global entropy reaches ternal force term has Gibbs thermal states as the only thermal equilibrium over around 40-50 oscillations inside stationary attractor states. This implies that interpar- the trap. Such a slow thermalisation process explains ticle diffusion is the leading mechanism for thermalisa- why strict one-dimensional experimental settings has pre- tion in the presence of integrability breaking terms, and, viously failed to observe thermalisation, where the max- as expected, no thermalisation is present in the non- imal experimental time included only 5-10 oscillations. interacting limit of hard-core bosons [69–71], where no Numerical simulations in the classical limit. diffusion is present. This is in agreement with previous Continuous quantum models are notoriously hard to sim- works on perturbed classical integrable models [72, 73], ulate with state–of–the–art tensor network techniques, where it was noticed that, while chaotic behaviour is despite recent progresses [56, 57]. We therefore bench- achieved at short times in many-body systems close to in- mark the hydrodynamic equation (3) with the classical tegrability, thermalisation requires to diffusively spread limit of the Lieb-Liniger gas. Given the rescaling of the inside the classical phase-space, and it manifests itself coupling c = ~, inverse temperature β = βNLS~ and the on much longer time scales, see also [74] for a previous 5

tions with M. J. Bhaseen.

[1] T. Langen, R. Geiger, and J. Schmiedmayer, Annual Review of Condensed Matter Physics 6, 201 (2015). [2] M. A. Nichols, L. W. Cheuk, M. Okan, T. R. Hartke, E. Mendez, T. Senthil, E. Khatami, H. Zhang, and M. W. Zwierlein, Science 363, 383 (2018). [3] N. Jepsen, J. Amato-Grill, I. Dimitrova, W. W. Ho, FIG. 3. Comparison between the predictions of the hy- E. Demler, and W. Ketterle, (2020), arXiv:2005.09549. drodynamic equation (3) (continuous lines) and Monte-Carlo [4] T. Kinoshita, T. Wenger, and D. S. Weiss, Nature 440, numerical simulation of the NLS model (dots with error bars) 900 (2006). for the same quench as in Fig.1 with L = 25.6. Left: evolu- [5] M. Rigol, V. Dunjko, V. Yurovsky, and M. Olshanii, 2 † tion of the density profile h|ψ| i = h(ψ ψ)(x, t)i at some given Phys. Rev. Lett. 98, 050405 (2007). positions x as function of time. Inset shows the (diffusively) [6] P. Calabrese, F. H. L. Essler, and G. Mussardo, Journal vanishing difference from the final thermal expectation value of : Theory and Experiment 2016, 2 † † as function of time ∆h|ψ| i ≡ h(ψ ψ)(x, t)i − h(ψ ψ)(x)iGE. 064001 (2016). Right: density profile for all x ∈ [0,L] at different times. [7] M. Gring, M. Kuhnert, T. Langen, T. Kitagawa, Microscopic simulations have been performed with the same B. Rauer, M. Schreitl, I. Mazets, D. A. Smith, E. Demler, method of Ref. [68] and J. Schmiedmayer, Science 337, 1318 (2012). [8] B. Bertini, F. H. L. Essler, S. Groha, and N. J. Robinson, Phys. Rev. Lett. 115, 180601 (2015). [9] B. Bertini, F. H. L. Essler, S. Groha, and N. J. Robinson, study on the Bose gas. We stress that this scenario evades Phys. Rev. B 94, 245117 (2016). KAM theorem [75], where instead only a finite number [10] V. Alba and M. Fagotti, Phys. Rev. Lett. 119, 010601 of degrees of freedom are considered. We here moreover (2017). show that the presence of a continuous number of modes [11] T. Langen, M. Gring, M. Kuhnert, B. Rauer, R. Geiger, D. A. Smith, I. E. Mazets, and J. Schmiedmayer, in integrable models is responsible for a slow diffusion The European Physical Journal Special Topics 217, 43 spreading as there is an infinite number of modes with (2013). arbitrarily small diffusion constant. On the other hand, [12] T. Langen, T. Gasenzer, and J. Schmiedmayer, Journal thermalisation of local observables always occurs, as they of Statistical Mechanics: Theory and Experiment 2016, only depends on a finite number of relevant modes, which 064009 (2016). leads to slow thermalisation drifts with diffusive finite- [13] K. Mallayya, M. Rigol, and W. De Roeck, Phys. Rev. X time corrections. Our work gives a comprehensive expla- 9, 021027 (2019). [14] A. J. Friedman, S. Gopalakrishnan, and R. Vasseur, nation for the apparent lack of thermalisation on finite Phys. Rev. B 101, 180302 (2020). time-scales observed both in experimental [4] and numer- [15] J. Durnin, M. J. Bhaseen, and B. Doyon, (2020), ical [22] settings and provide a quantitative prediction for arXiv:2004.11030. the time-scales necessary to thermalise. In the near fu- [16] F. m. c. Huveneers and J. Lukkarinen, Phys. Rev. Re- ture we plan to study the hydrodynamic equation (3) in search 2, 022034 (2020). the presence of generic external potential and for other [17] J. S. Ferreira and M. Filippone, (2020), models than Lieb-Liniger. Moreover, while our work fo- arXiv:2006.13891. [18] Y. Tang, W. Kao, K.-Y. Li, S. Seo, K. Mallayya, cused on a purely one-dimensional gas, one can now con- M. Rigol, S. Gopalakrishnan, and B. L. Lev, Phys. Rev. sider additions to (3) to account for different experimen- X 8, 021030 (2018). tal effects as such as losses [76], dephasing noise [68], and [19] F. Mller, C. Li, I. Mazets, H.-P. Stimming, T. Zhou, crossover to 3-dimensional geometries [19, 77]. Finally, Z. Zhu, X. Chen, and J. Schmiedmayer, (2020), it is an interesting question for the future to relate the arXiv:2006.08577. growth of thermodynamic entropy of Eq. (4) to the well- [20] B. Doyon and T. Yoshimura, SciPost Phys. 2, 014 (2017). known growth of entanglement entropy after quantum [21] J.-S. Caux, B. Doyon, J. Dubail, R. Konik, and T. Yoshimura, SciPost Phys. 6, 70 (2019). quenches [78–83]. [22] X. Cao, V. B. Bulchandani, and J. E. Moore, Phys. Rev. Acknowledgments. A.B. acknowledges support from Lett. 120, 164101 (2018). the European Research Council (ERC) under ERC Ad- [23] A. Bastianello, V. Alba, and J.-S. Caux, Phys. Rev. Lett. vanced grant 743032 DYNAMINT. J.D.N acknowledges 123, 130602 (2019). King’s College London for hospitality, Wojciech De [24] A. Bastianello and A. De Luca, Phys. Rev. Lett. 122, Roeck for early collaborations on this work and numer- 240606 (2019). ous enlightening discussions. J.D.N. is supported by Re- [25] I. Bloch, Nature Physics 1, 23 (2005). [26] I. Bloch, J. Dalibard, and W. Zwerger, Rev. Mod. Phys. search foundation Flanders (FWO). B.D. acknowledges 80, 885 (2008). discussions with Joseph Durnin, and related collabora- 6

[27] T. Kinoshita, T. Wenger, and D. S. Weiss, Science 305, [60] A. Bastianello, B. Doyon, G. Watts, and T. Yoshimura, 1125 (2004). SciPost Phys. 4, 45 (2018). [28] T. Kinoshita, T. Wenger, and D. S. Weiss, Phys. Rev. [61] A. Bastianello, M. Arzamasovs, and D. M. Gangardt, Lett. 95, 190406 (2005). Phys. Rev. B 101, 245157 (2020). [29] J. Armijo, T. Jacqmin, K. V. Kheruntsyan, and I. Bou- [62] Y. Castin, R. Dum, E. Mandonnet, A. Minguzzi, and choule, Phys. Rev. Lett. 105, 230402 (2010). I. Carusotto, Journal of Modern Optics 47, 2671 (2000). [30] A. H. van Amerongen, J. J. P. van Es, P. Wicke, K. V. [63] M. Arzamasovs and D. M. Gangardt, Phys. Rev. Lett. Kheruntsyan, and N. J. van Druten, Phys. Rev. Lett. 122, 120401 (2019). 100, 090402 (2008). [64] A. Johnson, S. S. Szigeti, M. Schemmer, and I. Bou- [31] M. Schemmer, I. Bouchoule, B. Doyon, and J. Dubail, choule, Phys. Rev. A 96, 013623 (2017). Phys. Rev. Lett. 122, 090601 (2019). [65] M. Wouters, Phys. Rev. A 90, 033611 (2014). [32] J. Settino, N. L. Gullo, F. Plastina, and A. Minguzzi, [66] T. Jacqmin, B. Fang, T. Berrada, T. Roscilde, and (2020), arXiv:2005.13646. I. Bouchoule, Phys. Rev. A 86, 043626 (2012). [33] E. H. Lieb and W. Liniger, Phys. Rev. 130, 1605 (1963). [67] I. Bouchoule, M. Arzamasovs, K. V. Kheruntsyan, and [34] M. Kormos, A. Shashi, Y.-Z. Chou, J.-S. Caux, and D. M. Gangardt, Phys. Rev. A 86, 033626 (2012). A. Imambekov, Phys. Rev. B 88, 205131 (2013). [68] A. Bastianello, J. D. Nardis, and A. D. Luca, “Gen- [35] G. Mussardo, Phys. Rev. Lett. 111, 100401 (2013). eralised hydrodynamics with dephasing noise,” (2020), [36] J. De Nardis, B. Wouters, M. Brockmann, and J.-S. arXiv:2003.01702 [cond-mat.stat-mech]. Caux, Phys. Rev. A 89, 033601 (2014). [69] P. P. Mazza, M. Collura, M. Kormos, and P. Calabrese, [37] M. Kormos, M. Collura, and P. Calabrese, Phys. Rev. A Journal of Statistical Mechanics: Theory and Experi- 89, 013609 (2014). ment 2014, P11016 (2014). [38] F. H. L. Essler, G. Mussardo, and M. Panfil, Phys. Rev. [70] M. Collura, M. Kormos, and P. Calabrese, Phys. Rev. A A 91, 051602 (2015). 97, 033609 (2018). [39] T. Palmai and R. M. Konik, Phys. Rev. E 98, 052126 [71] A. Bastianello, M. Collura, and S. Sotiriadis, Phys. Rev. (2018). B 95, 174303 (2017). [40] O. A. Castro-Alvaredo, B. Doyon, and T. Yoshimura, [72] K.-D. N. T. Lam and J. Kurchan, Journal of Statistical Phys. Rev. X 6, 041065 (2016). Physics 156, 619 (2014). [41] B. Bertini, M. Collura, J. De Nardis, and M. Fagotti, [73] T. Goldfriend and J. Kurchan, “Quasi-integrable systems Phys. Rev. Lett. 117, 207201 (2016). are slow to thermalize but may be good scramblers,” [42] J. De Nardis, D. Bernard, and B. Doyon, Phys. Rev. (2019), arXiv:1909.02145. Lett. 121, 160603 (2018). [74] I. E. Mazets, The European Physical Journal D 65, 43 [43] J. De Nardis, D. Bernard, and B. Doyon, SciPost Phys. (2011). 6, 49 (2019). [75] A. Katok, R. de la Llave, Y. Pesin, and H. Weiss, eds., [44] B. Doyon, J. Dubail, R. Konik, and T. Yoshimura, Phys. Smooth Ergodic Theory and Its Applications (American Rev. Lett. 119, 195301 (2017). Mathematical Society, 2001). [45] V. B. Bulchandani, R. Vasseur, C. Karrasch, and J. E. [76] I. Bouchoule, B. Doyon, and J. Dubail, (2020), Moore, Phys. Rev. Lett. 119, 220604 (2017). arXiv:2006.03583. [46] S. Gopalakrishnan, D. A. Huse, V. Khemani, and [77] J. Lopez-Piqueres, B. Ware, S. Gopalakrishnan, and R. Vasseur, Phys. Rev. B 98, 220303 (2018). R. Vasseur, (2020), arXiv:2005.13546. [47] M. Borsi, B. Pozsgay, and L. Pristy´ak, Phys. Rev. X 10, [78] V. Alba and P. Calabrese, Proc. Natl. Acad. Sci. U.S.A. 011054 (2020). , 201703516 (2017). [48] T. Yoshimura and H. Spohn, (2020), arXiv:2004.07113. [79] B. Bertini, M. Fagotti, L. Piroli, and P. Calabrese, [49] J. De Nardis, D. Bernard, and B. Doyon, SciPost Phys. Journal of Physics A: Mathematical and Theoretical 51, 6, 49 (2019). 39LT01 (2018). [50] See supplementary material for a more detailed proof. [80] A. Nahum, J. Ruhman, S. Vijay, and J. Haah, Phys. [51] V. B. Bulchandani, Journal of Physics A: Mathematical Rev. X 7, 031016 (2017). and Theoretical 50, 435203 (2017). [81] L. Piroli, B. Bertini, J. I. Cirac, and T. c. v. Prosen, [52] B. Doyon, (2019), arXiv:1912.08496. Phys. Rev. B 101, 094304 (2020). [53] T. Cebeci, Convective Heat Transfer, Convective Heat [82] V. Alba, B. Bertini, and M. Fagotti, SciPost Phys. 7, 5 Transfer No. v. 2 (Horizons Pub., 2002). (2019). [54] M. Panfil and J. Paweczyk, SciPost Phys. Core 1, 2 [83] B. Bertini and P. Calabrese, (2020), arXiv:2007.01286. (2019). [84] M. Takahashi, of one-dimensional solv- [55] K. V. Kheruntsyan, D. M. Gangardt, P. D. Drummond, able models (Cambridge University Press, 2005). and G. V. Shlyapnikov, Phys. Rev. Lett. 91, 040403 (2003). [56] F. Verstraete and J. I. Cirac, Physical Review Letters 104 (2010), 10.1103/physrevlett.104.190405. [57] B. Tuybens, J. D. Nardis, J. Haegeman, and F. Ver- straete, (2020), arXiv:2006.01801. [58] G. Del Vecchio Del Vecchio, A. Bastianello, A. De Luca, and G. Mussardo, (2020), arXiv:2002.01423 [cond- mat.stat-mech]. [59] A. De Luca and G. Mussardo, Journal of Statistical Me- chanics: Theory and Experiment 2016, 064011 (2016). 7 Supplementary Material Thermalisation of a trapped one-dimensional Bose gas via diffusion

The thermal state as the only stationary state in the presence of diffusion

Here we report a precise proof of what we briefly discussed in the main text, namely that the diffusive dynamics in the presence of an external trap necessarily leads to thermalisation. Let us consider the diffusive hydrodynamic equation Eq again. (3). For the sake of clarity, it is useful to introduce a global lengthscale L to tune the inhomogeneity. Hence, we assume the potential has the scaling form V (x) = U(xL−1) and we use a rescaled spatial coordinate y = xL−1. After this rescaling, the hydrodynamic equation becomes

−1 eff −2 α −1 −3 ∂tρθ = −L ∂y(vθ ρθ) + L ∂y(Dθ ∂yρα) + L (∂yV (x))∂θρθ + O(L ) (S1)

With O(L−3) terms coming from corrections beyond diffusion. We are ultimately interested in the regime of large 0 L with a suitable rescaling of time. Up to a timescale t ∼ O(L ), the system does not show any dynamics ∂tρθ ' 0, since GGEs are stationary states of the local integrable model. Timescales t ∼ O(L1) display Eulerian dynamics, while finally t ∼ O(L2) will be affected by diffusion. Now, we show that the only stationary state that can be attained 2 3 in the time window O . L  t  O(L ) is a thermal state in the local density approximation: since thermal states are expected to be stationary for the exact dynamics, we conclude that the state will remain thermal also for t  L3 where corrections beyond diffusion (not present in Eq. (S1)) cannot be neglected any longer. As we already discussed in the main text, diffusion generally implies a growth in the Yang-Yang entropy: in the rescaled coordinates we have

dS = L−2 dy ∂ w (DC) ∂ w (S2) dt ˆ y θ θ,α y α

P 2 where we introduced the short-hand notation wθ = βj(y)hj,θ. The entropy grows on a timescale t ∼ O(L ): we assume a stationary state is reached, therefore the state must be such that dS/dt = 0. Since the only eigenvector of the diffusion matrix with zero eigenvalue is the constant function (in the rapidity space), we must have ∂θ∂ywθ = 0, which leads to the general form

wθ(x) = −µ(y) + ν(θ) (S3) with µ(y) an inhomogeneous local chemical potential (constant in the rapidity space) and ν(θ) a function of only the rapidities, but constant in the coordinate space. We now require that such an inhomogeneous GGE is also stationary for Eq. (S1), i.e. we impose ∂tρθ = 0. Since the diffusion part in (S1) vanishes for the states of the form (S3), we can focus on the stationarity with respect to the Eulerian dynamics. As explained in the main text (see also [20]), this is enough to impose a thermal state. In order to proceed, we first change the variable in the GHD equation from the root density to the generalised effective wθ. To this end, we use the standard TBA results [84] connecting w with the filling function

−1 log(nθ − 1) = wθ − Tθ,α log(1 − nα) (S4) Using this relation in Eq. (S1) and imposing the stationarity with respect to the Eulerian part, standard manipulations leads to the simple equation

eff dr dr vθ (∂yw) − (∂yU)(∂θwθ) = 0 . (S5)

dr dr 0 dr Now we use Eq. (S3) and notice that (∂yw) = −(∂yµ(y)) = −∂yµ(y)(kθ) . In the last passage, we used that ∂yµ(y) is constant in the rapidity space and therefore it can be moved out of the dressing. Using this last result, we bare dr reach the equality (∂θεθ )(∂yµ) = (∂yU)(∂θwθ) . Finally, using again that constants in the rapidity space commute bare with the dressing operation and inverting the dressing, we ultimately reach the identity ∂θε ∂yµ = −∂yU∂θwθ, which can be rewritten as

∂θwθ ∂yµ bare = − . (S6) ∂θε ∂yU 8

The l.h.s. is a function only of the rapidities, while the r.h.s. is a function only of the spatial coordinate. The only bare possibility for the above identity to hold is that both ratios are constants. In particular ∂θwθ/∂θε = −∂yµ/∂yU = β: the constant β can be immediately identified with the inverse temperature. Indeed, we can now write

bare wθ = β(εθ + U(y)) + const. . (S7) and finally conclude that the stationary state is thermal (within the local density approximation) with the undeter- mined constant playing the role of a global chemical potential.