14
Dynamical phase transition to localized states in the two-dimensional random walk conditioned on partial currents Ricardo Guti´ errez 1 and Carlos P´ erez-Espigares 2, 3 1 Complex Systems Interdisciplinary Group (GISC), Department of Mathematics, Universidad Carlos III de Madrid, 28911 Legan´ es, Madrid, Spain 2 Departamento de Electromagnetismo y F´ ısica de la Materia, Universidad de Granada, Granada 18071, Spain 3 Institute Carlos I for Theoretical and Computational Physics, Universidad de Granada, Granada 18071, Spain The study of dynamical large deviations allows for a characterization of stationary states of lattice gas models out of equilibrium conditioned on averages of dynamical observables. The application of this framework to the two-dimensional random walk conditioned on partial currents reveals the existence of a dynamical phase transition between delocalized band dynamics and localized vortex dynamics. We present a numerical microscopic characterization of the phases involved, and provide analytical insight based on the macroscopic fluctuation theory. A spectral analysis of the microscopic generator shows that the continuous phase transition is accompanied by a spontaneous Z2-symmetry breaking whereby the stationary solution loses the reflection symmetry of the generator. Dynamical phase transitions similar to this one, which does not rely on exclusion effects or interactions, are likely to be observed in more complex non-equilibrium physics models. Introduction— The study of large fluctuations of dynami- cal observables in lattice gas models has greatly improved our understanding of non-equilibrium statistical mechan- ics, as it has shed light on rare events in transport phe- nomena [1–3], has unveiled fluctuation relations of wide applicability [4–9], and has clarified the origin of many- body effects in classical [10, 11] and quantum systems [12–14]. These microscopic analyses have been comple- mented with approaches based on the macroscopic fluc- tuation theory, where similar phenomena are studied at a coarse-grained level [15]. Both frameworks rely on sta- tistical ensembles of trajectories [16, 17], for which large- deviation functions play an analogous role to that of ther- modynamic potentials in equilibrium statistical mechan- ics [18]. Dynamical phase transitions (DPTs), characterized by sudden changes in the structure of trajectories that are reflected in the statistics of dynamical observables, are among the most interesting phenomena that have been unveiled by these approaches. DPTs manifest themselves as non-analyticities in the large deviation functions, and occur not only in systems driven far from equilibrium, but also in equilibrium settings, even in situations where the statistics of (static) configurations is trivial [11]. In recent years, the study of DPTs has been enriched by the application of the generalized Doob transform, which provides the subset of trajectories sustaining a given rare fluctuation [19–23], allowing for a characterization of dy- namical phases. The study of DPTs is frequently accomplished in one-dimensional systems [24–38], where the computa- tion of large deviations for large systems is analytically tractable. Analyses of higher-dimensional settings are challenging, since they rely on variational procedures and diagonalization problems whose complexity increases with the spatial dimension. Yet the statistics of total cur- rents (quantifying the net number of particles crossing a section that spans the whole system) or dynamical ac- tivities (counting configuration changes in a trajectory) have been studied in two dimensions, where a myriad of phenomena, including DPTs, have been reported [39– 44]. In the case of a two-dimensional random walk on a lattice, the Doob-transformed dynamics that sustains a given fluctuation of the total current amounts to the inclusion of a uniform driving field. This implies that large fluctuations are achieved by flat, i.e. lacking spa- tial structure, density and current fields, giving rise to current distributions that are trivially Gaussian. The fluctuations of so-called partial currents, i.e. currents of particles that move across some ‘wall’ or ‘slit’ that does not span the whole system, are strikingly different, how- ever, as we shall show. In this Letter we consider the statistics of partial cur- rents across a finite slit of random walks on a square lattice, as well as its continuum hydrodynamic limit. We find that, despite the simplicity of the model, there is a DPT underpinning rare fluctuations of such partial cur- rents, which is continuous and characterized by a sponta- neous breaking of the Z 2 reflection symmetry in the direc- tion of the slit. While small fluctuations are characterized by the formation of bands, rare fluctuations condense into localized vortices. This highlights how large fluctu- ations of partial currents, which are more relevant than total currents in experimental settings, are created by non-trivial emerging structures associated with a DPT. Given the simplicity and generality of the model, simi- lar DPTs are likely to be found in other systems with exclusion effects and interactions, and might shed light on intruiguing results previously reported for the two- dimensional simple exclusion process [45] Model— We consider N random walkers on a two- dimensional square lattice comprising L × L sites with periodic boundary conditions in the presence of a driv- ing field E = (E x ,E y ), which induces asymmetric jump rates between neighboring sites. The stochastic dynamics of each particle is governed by the master equation t p(n, t)= -p(n, t)+ mnnn Γ nm p(m, t), arXiv:2108.12359v1 [cond-mat.stat-mech] 27 Aug 2021

arXiv:2108.12359v1 [cond-mat.stat-mech] 27 Aug 2021

  • Upload
    others

  • View
    8

  • Download
    0

Embed Size (px)

Citation preview

Page 1: arXiv:2108.12359v1 [cond-mat.stat-mech] 27 Aug 2021

Dynamical phase transition to localized states in the two-dimensional random walkconditioned on partial currents

Ricardo Gutierrez1 and Carlos Perez-Espigares2, 3

1Complex Systems Interdisciplinary Group (GISC), Department of Mathematics,Universidad Carlos III de Madrid, 28911 Leganes, Madrid, Spain

2Departamento de Electromagnetismo y Fısica de la Materia, Universidad de Granada, Granada 18071, Spain3Institute Carlos I for Theoretical and Computational Physics, Universidad de Granada, Granada 18071, Spain

The study of dynamical large deviations allows for a characterization of stationary states of latticegas models out of equilibrium conditioned on averages of dynamical observables. The applicationof this framework to the two-dimensional random walk conditioned on partial currents reveals theexistence of a dynamical phase transition between delocalized band dynamics and localized vortexdynamics. We present a numerical microscopic characterization of the phases involved, and provideanalytical insight based on the macroscopic fluctuation theory. A spectral analysis of the microscopicgenerator shows that the continuous phase transition is accompanied by a spontaneous Z2-symmetrybreaking whereby the stationary solution loses the reflection symmetry of the generator. Dynamicalphase transitions similar to this one, which does not rely on exclusion effects or interactions, arelikely to be observed in more complex non-equilibrium physics models.

Introduction— The study of large fluctuations of dynami-cal observables in lattice gas models has greatly improvedour understanding of non-equilibrium statistical mechan-ics, as it has shed light on rare events in transport phe-nomena [1–3], has unveiled fluctuation relations of wideapplicability [4–9], and has clarified the origin of many-body effects in classical [10, 11] and quantum systems[12–14]. These microscopic analyses have been comple-mented with approaches based on the macroscopic fluc-tuation theory, where similar phenomena are studied ata coarse-grained level [15]. Both frameworks rely on sta-tistical ensembles of trajectories [16, 17], for which large-deviation functions play an analogous role to that of ther-modynamic potentials in equilibrium statistical mechan-ics [18].

Dynamical phase transitions (DPTs), characterized bysudden changes in the structure of trajectories that arereflected in the statistics of dynamical observables, areamong the most interesting phenomena that have beenunveiled by these approaches. DPTs manifest themselvesas non-analyticities in the large deviation functions, andoccur not only in systems driven far from equilibrium,but also in equilibrium settings, even in situations wherethe statistics of (static) configurations is trivial [11]. Inrecent years, the study of DPTs has been enriched bythe application of the generalized Doob transform, whichprovides the subset of trajectories sustaining a given rarefluctuation [19–23], allowing for a characterization of dy-namical phases.

The study of DPTs is frequently accomplished inone-dimensional systems [24–38], where the computa-tion of large deviations for large systems is analyticallytractable. Analyses of higher-dimensional settings arechallenging, since they rely on variational proceduresand diagonalization problems whose complexity increaseswith the spatial dimension. Yet the statistics of total cur-rents (quantifying the net number of particles crossing asection that spans the whole system) or dynamical ac-tivities (counting configuration changes in a trajectory)

have been studied in two dimensions, where a myriadof phenomena, including DPTs, have been reported [39–44]. In the case of a two-dimensional random walk ona lattice, the Doob-transformed dynamics that sustainsa given fluctuation of the total current amounts to theinclusion of a uniform driving field. This implies thatlarge fluctuations are achieved by flat, i.e. lacking spa-tial structure, density and current fields, giving rise tocurrent distributions that are trivially Gaussian. Thefluctuations of so-called partial currents, i.e. currents ofparticles that move across some ‘wall’ or ‘slit’ that doesnot span the whole system, are strikingly different, how-ever, as we shall show.

In this Letter we consider the statistics of partial cur-rents across a finite slit of random walks on a squarelattice, as well as its continuum hydrodynamic limit. Wefind that, despite the simplicity of the model, there is aDPT underpinning rare fluctuations of such partial cur-rents, which is continuous and characterized by a sponta-neous breaking of the Z2 reflection symmetry in the direc-tion of the slit. While small fluctuations are characterizedby the formation of bands, rare fluctuations condenseinto localized vortices. This highlights how large fluctu-ations of partial currents, which are more relevant thantotal currents in experimental settings, are created bynon-trivial emerging structures associated with a DPT.Given the simplicity and generality of the model, simi-lar DPTs are likely to be found in other systems withexclusion effects and interactions, and might shed lighton intruiguing results previously reported for the two-dimensional simple exclusion process [45]

Model— We consider N random walkers on a two-dimensional square lattice comprising L × L sites withperiodic boundary conditions in the presence of a driv-ing field E = (Ex, Ey), which induces asymmetricjump rates between neighboring sites. The stochasticdynamics of each particle is governed by the masterequation ∂tp(n, t) = −p(n, t) +

∑m∈nnn

Γn←m p(m, t),

arX

iv:2

108.

1235

9v1

[co

nd-m

at.s

tat-

mec

h] 2

7 A

ug 2

021

Page 2: arXiv:2108.12359v1 [cond-mat.stat-mech] 27 Aug 2021

2

where p(n, t) is the probability of occupation of siten ∈ {1, 2, . . . , L2} at time t, and nnn contains its(four) nearest neighbors. Written in operatorial form,dp(t)/dt = Wp(t), with the column vector p(t) =(p(1, t), p(2, t), . . . , p(L2, t))T (T denotes transposition),and the entries of the generator Wnn′ = −δnn′ + Γn←n′(Γn←n′ = 0 if n′ /∈ nnn). The transition rates are

Γn←m = eE·(rn−rm)

L /(eExL + e−

ExL + e

EyL + e−

EyL ), (1)

so that∑n Γn←m = 1, where rn is the position vector

of site n taking the lattice constant as spatial unit . Theobservable of interest is the partial current J across avertical slit of length hL (0 < h < 1) placed in the centerof the lattice, which we take as the origin of (Cartesian)coordinates. J quantifies the number of particles thattraverse the slit in the rightward direction minus thosethat do in the leftward direction per unit of time.Microscopic large-deviation methods— The current Jfollows a probability distribution that adopts a large-deviation form for long times, P (J) ∼ e−tϕ(J), with ratefunction ϕ(J), which is non-negative and equal to zeroonly for the average current J = 〈J〉 [16]. For simplic-ity, in the following we will consider a horizontal fieldE = (E, 0) giving rise [cf. (1)] to an average partial cur-rent 〈J〉 = ρ0hL tanh[E/2L], which in the limit of large Lbecomes 〈J〉 = ρ0Eh/2. To investigate the fluctuationsof J , one would like to find the rate function ϕ(J), butthis is in general a difficult task, as we are dealing withevents that are exponentially unlikely in time. Insteadwe bias the probability distribution with a parameters, obtaining a new distribution Ps(J) = e−stJP (J)/Zs,where the dynamical partition function Zs acquires forlong times the following large-deviation form,

Zs =

∫dJ e−stJP (J) ∼ etθ(s). (2)

The scaled cumulant-generating function (SCGF) θ(s) isrelated to the rate function by a Legendre-Fenchel trans-form, θ(s) = −minJ [sJ + ϕ(J)] [16]. By choosing theappropriate value of s we find the fluctuation of inter-est, J = 〈J〉s (where 〈·〉s is the average over Ps(J)),which in turn corresponds to (minus) the first derivativeof the SCGF, 〈J〉s = −θ′(s). In general, the p-th deriva-tive of the SCGF yields the partial-current cumulant ofthe corresponding order 〈〈Jp〉〉s: limt→∞ tp−1〈〈Jp〉〉s =

(−1)p dpθ(s)dsp [11].

The SCGF can be obtained as the largest eigenvalue ofthe so-called tilted generator Ws, whose entries are [16]

Wsnn′ =

e−sWnn′ n ∈ nr andn′ ∈ nl,es Wnn′ n ∈ nl andn′ ∈ nr,Wnn′ otherwise.

(3)

where nr contains all sites that are immediately to theright of the slit, and nl those that are immediately to theleft. Notice that in Eq. (3), s = 0 corresponds to the

original dynamics, while s < 0 enhances current fluctua-tions larger that the average, and the opposite occurs fors > 0. Since Ws is not symmetric, it has different left andright eigenvectors, which satisfy lTs,iW

s = λi(s)lTs,i and

Wsrs,i = λi(s)rs,i, respectively, where i = 0, 1, . . . L2− 1.The eigenvalues λi(s) are ordered in decreasing magni-tude of the real part, and λ0(s) = θ(s). While it mightappear natural that the tilted dynamics generated by (3)must be a biased random walk, it turns out that suchdynamics is not physical, as Ws does not conserve prob-ability,

∑n Ws

nn′ 6= 0. To recover the physical dynamicssustaining the fluctuation associated with a given s atransformation is needed, namely the generalized Doobtransform [19–23],

WsDoob = LsW

sL−1s − θ(s)1 . (4)

Here Ls is the diagonal matrix whose entries are thecomponents of the leading left eigenvector, ls,0, and 1is the L×L identity matrix. The Doob generator Ws

Doobis a proper stochastic (probability conserving) genera-tor,

∑n Ws

Doob,nn′ = 0. Its eigenvalues are λDi (s) =

λi(s) − θ(s), with left eigenvectors lDs,i = L−1s ls,i and

right eigenvectors rDs,i = Lsrs,i. The largest (zero) eigen-

value is associated with the stationary state psstat = rDs,0,Ws

Doobpsstat = 0 [46]. The Doob generator (4) provides

the stochastic process whose long-time statistics of J isgiven by Ps(J).

For one random walk sustaining the fluctuation J ,the density and current fields are thus given by psstat(n)(which are the entries of psstat) and js(n) = (jsx(n), jsy(n)).The latter satisfies

∑n∈nl

jsx(n) = J , and can be ob-tained from the difference between the probability fluxesof two adjacent sites, jsα(n) = psstat(n)Ws

Doob,mn −psstat(m)Ws

Doob,nm, where m is the right (upper) neigh-

bor of n for α = x (α = y). For N particles, the den-sity field, namely the average number of particles persite, is just ρs(n) = Npsstat(n), and the current field isjs(n) = N js(n). The SCGF must also be mutiplied byN , as can be seen from the stochastic independence ofdifferent random walks, which yields ZNs as the dynami-cal partition function, Zs being the one-particle partitionfunction (2).Microscopic analysis of dynamical regimes— In Fig. 1(a) and (b) we provide a first glimpse of the fluctua-tions of the partial current J in a system of linear sizeL = 32 with a slit of relative length h = 1/2 under afield of strength E = 5. In the main panel of (a) weshow the (rescaled) SCGF L2θ(s) and the correspond-ing current fluctuation, L2〈J〉s for different values of thebiasing field s. Since N = ρ0L

2, where ρ0 is the totaldensity, we represent θ(s) and derived quantities multi-plied by L2, which corresponds to fixing the density toρ0 = 1, and will be useful in the comparison of resultsfor different size L later. When a comparison is made fora fixed number of particles N instead, we shall focus onone-particle quantities (N = 1), but then of course thedensity will vary as ρ0 = 1/L2.

Page 3: arXiv:2108.12359v1 [cond-mat.stat-mech] 27 Aug 2021

3

FIG. 1. Microscopic and macroscopic large-deviation analysis of the partial current statistics in the two-dimensional random walk. (a) (Main) Rescaled SCGF L2θ(s) (green circles) and average partial current L2〈J〉s (magentasquares) for linear size L = 32, relative slit length h = 1/2 and field strength E = 5. (Inset) Density field ρs(n) (gray colormap,darker color indicates a larger number of particles) and current field js(n) (blue arrows) for s = 0 (natural dynamics). Theslit is highlighted in red. (b) Density field ρs(n) and current field js(n) for s = −1 (corresponding to L2〈J〉s = 2.42), s = −3(L2〈J〉s = 13.09), s = 1 (L2〈J〉s = 0.23), and s = 3 (L2〈J〉s = −20.33), with format as in the inset of panel (a). (c) Macroscopicfluctuation theory functional of the flat solution Gflat(J), the inner-band solution Gi.b.(J), the outer-band solution Go.b.(J)and the vortex solution Gvort.(J). The minimum value among them corresponding to each J , denoted Gmin.(J), is also shown,as well as its convex envelope GMaxw.(J). See text for definitions, and the SM for the details [47].

For s = 0 the density and the current fields are uni-form, see inset of panel (a), and the partial current isL2〈J〉s=0 = 〈J〉 = 1.25, which corresponds to the expres-sion given above for ρ0 = 1. For s > 0, L2〈J〉s decreasesmonotonically slowly up to s ≈ 2, and then more rapidly,while a similar but opposite behavior is found for s < 0.The density and current fields {ρs(n), js(n)} for s 6= 0are displayed in Fig. 1 (b).

For negative tiltings, s < 0, the density becomeshigher in the horizontal band crossing the slit, i.e. y ∈(−hL/2, hL/2), as illustrated for s = −1. This regime,which will be referred to as the inner band, persists fora range of negative values of s, with a spatial decay for|y| > hL/2 that becomes more abrupt as the absolutevalue of s increases. The band is not perfectly uniformin the horizontal direction, however, as there is a ten-dency for particles to populate the edges of the slit andto move cyclically around them —an effective way to in-crease the partial current J— which is also enhanced asthe absolute value of s increases. Beyond a certain pointthe band practically disappears, and the (large) currentis almost completely sustained by vortices localized atthe edges, as shown for s = −3. Such cyclic, localizedbehavior, which has recently been observed in randomwalks on general graphs [48] and the zero-range processon a diamond lattice [49], presents some similarities tothe vortices of the two-dimensional simple exclusion pro-cess [45].

For positive and moderately large values of s the rel-atively low J is achieved by forming a band that avoidspassing through the slit, the bulk of the density showingin the region where |y| > hL/2, which we will call theouter band, and is illustrated for s = 1. But also herethere is a tendency of particles to populate the edges ofthe slit, and move in circles (this time in the opposite

direction, so the current decreases with the adopted signconvention). For sufficiently large s, as illustrated fors = 3, the band again practically disappears, leading toa vortex dynamics of opposite vorticity to that observedfor negative s. This phenomenology will be shown to beassociated with a DPT to a localized state. But beforeaddressing its nature, a macroscopic analysis will providefurther insight into the dynamical regimes at play.Macroscopic analysis of dynamical regimes— Followingthe macroscopic fluctuation theory, which studies fluc-tuations of dynamical observables at the macroscopic(coarse-grained) level [15], we consider the probabilityof any trajectory of duration T given by the densityand current fields {ρ(r, t), j(r, t)}T0 . For driven diffusivesystems such probability adopts a large-deviation form

P [{ρ(r, t), j(r, t)}T0 ] ∼ e−L2I[ρ,j], with the rate functional

I[ρ, j] =

∫ T

0

dt

∫Λ

dr[j +D(ρ)∇ρ− σ(ρ)E]

2

2σ(ρ). (5)

For a random walk, the diffusivity is D(ρ) = 1/4 andthe mobility σ(ρ) = ρ/2 [50]. A diffusive rescalingof the microscopic variables, whereby time is rescaledby 1/L2 and space by 1/L, so that the process takesplace over Λ = [−1/2, 1/2]× [−1/2, 1/2], is applied.From Eq. (S2) we can calculate by contraction theprobability of any observable depending on the tra-jectory. Thus the probability of having a current Jthrough the slit reads P (J) ∼ exp{−TL2G(J)}, withG(J) = limT→∞

1T min∗{ρ,j}T0

I[ρ, j]. Here ∗ means that

the minimization is subject to the constraints J =

T−1∫ T

0dt∫ h/2−h/2 dyjx(x, y; t) and ∂tρ = −∇ · j (continu-

ity equation). Notice that the probability is maximized(G(J) is zero) when j(r, t) equals the macroscopic aver-age current j = −D(ρ)∇ρ+ σ(ρ)E.

Page 4: arXiv:2108.12359v1 [cond-mat.stat-mech] 27 Aug 2021

4

FIG. 2. Characterization of the DPT (a) Density-rescaled SCGF L2θ(s). (b) Same content as in (a) but zooming aroundsmall values of |s| and also including macroscopic SCGF (red discontinuous line). (c) SCGF θ(s). (d) Average current 〈J〉s,multiplied by L2 (main panel), by L/h (top right inset) and without rescaling (lower inset). See text for justifications ofrescalings. (d) Rescaled spectral gaps L2∆1(s) (continuous lines, large symbols) and L2∆2(s) (dotted lines, small symbols),see text for definitions. (d) Inverse participation ratio, without (main panel) and with rescaling (inset). As L increases, therange of s explored decreases, due to numerical difficulties to make the eigenvalue algorithm converge. In all panels the colorsand symbols represent different system sizes L according to the legend in (a).

Since solving the two-dimensional spatiotemporal vari-ational problem for density and current fields such asthose displayed in Fig. 1 (b) is a daunting task, we shallfocus on a few idealized ansatze for {ρ(r), j(r)}T0 . Boththeir time independence and their structural features arebased on the previous microscopic results. Specifically,we consider four dynamical regimes, each one leading to adifferent form of G(J): (i) a flat solution (uniform densityand current fields) Gflat(J), (ii) an inner band solution(uniform density over the region of Λ where |y| < h/2)Gi.b.(J), (iii) an outer band solution (uniform densityover the region of Λ where |y| > h/2) Go.b.(J) and (iv)a vortex solution Gvort.(J), all of which are discussed inthe Supplemental Material (SM) [47]. Fig. 1 (c) showsGmin.(J) = min{Gflat(J), Gi.b.(J), Go.b.(J), Gvort.(J)},which maximizes the probability at each J , yielding

Gmin.(J) =

Gvort.(J) J < 0,

Go.b.(J) 0 ≤ J < 〈J〉,Gflat(J) J = 〈J〉,Gi.b.(J) 〈J〉 < J < J,

Gvort.(J) J ≤ J.

(6)

While the flat solution maximizes the probability for J =〈J〉, its Gaussian fluctuations around 〈J〉 are suppressedexponentially in time with respect to those of the bands:immediately to the right of 〈J〉 the dominant regime isthe inner band, while the outer band prevails for values ofJ between zero and 〈J〉. For larger fluctuations, Gvort.(J)dominates for J < 0, —counterclockwise rotation in theupper edge and clockwise rotation in the lower edge of theslit— and also for J > J —with vortices rotating in theopposite directions—, where J is to an extent dependenton the choice of a cutoff radius [47].

While a Legendre-Fenchel transformation of Gmin.(J)yields the macroscopic SCGF of Fig. 2 (b) (red discontin-uous line; see the discussion below), an inverse Legendre-Fenchel transformation applied on such SCGF gives the

convex envelope GMaxw.(J), which, unlike Gmin.(J) it-self, is convex, see Fig. 1 (c). This Maxwell construc-tion highlights a coexistence between dynamical phases[16], namely, that of vortices and bands, specifically theouter band for J < 〈J〉 and the inner band for J > 〈J〉.Though this is broadly in agreement with the microscopicresults of Fig. 1 (b), it should not be taken as indicativeof the existence of a first-order DPT, as Gmin.(J) is re-stricted to a few idealized cases, and is thus an upperbound of the actual macroscopic functional: such coexis-tence may well approximate dynamical regimes not fullydescribable in terms of the ansatze under consideration.

Microscopic analysis of the DPT— In order to elucidatethe nature of the DPT we again focus on the microscopicSCGF θ(s). In Fig. 2 (a) we show the (rescaled) SCGFL2θ(s) for different linear sizes ranging from L = 12to 200. A good collapse of L2θ(s) curves is found for|s| . 2, which marks the onset of the DPT, as canbe better appreciated in panel (b), which also includesthe macroscopic SCGF discussed in the previous section(red discontinuous line). Such an agreement, however,does not exist for |s| larger than critical, as L2θ(s) thengrows without bound as L increases, yielding a rate func-tion ϕ(J) that becomes linear [qualitatively similar toGMaxw.(J) in Fig. 1 (c)], and therefore an asymptoticallyexponential distribution P (J) [16]. Nevertheless, a goodcollapse is found beyond the critical points if the rescal-ing of θ(s) by L2 is not performed (i.e. if the numberof particles N , instead of the density ρ0, is kept fixed),as shown in panel (c). Similar conclusions can be drawnfrom the average current 〈J〉s, which is shown in rescaledform L2〈J〉s in the main panel of Fig. 2 (d). After furthernormalization by the slit length hL, L〈J〉s/h, displayedin the upper inset, it shows a good collapse below the crit-ical points, but certainly not for larger absolute values ofs. There the collapse is observed in the unrescaled form〈J〉s (for fixed particle number N , without normalizingby the slit size), as shown in the lower inset.

Page 5: arXiv:2108.12359v1 [cond-mat.stat-mech] 27 Aug 2021

5

We conclude that a collapse for different sizes L occursbelow the critical point for fixed density ρ0, while on theother side of the transition it is the number of particlesN that must be kept constant. Viewed from the perspec-tive of the idealized regimes of the previous section, thecurrent associated with band dynamics is proportional toρ0, while that of localized vortices is proportional to thenumber of particles N .

The features of the DPT can be inferred from the spec-tral properties of the Doob generator Ws

Doob [34, 38, 51,52], in particular the gaps ∆i(s) = θ(s) − Re[λi(s)] =−Re[λDi (s)] (by definition ∆0(s) = 0). The first two,∆1(s) (usually referred to as the spectral gap) and ∆2(s),are shown in rescaled form (multiplied by L2 so as to fixthe density, ρ0 = 1) in Fig. 2 (e). While L2∆2(s) onlygrows with |s|, we observe that the gap L2∆1(s) closes ata certain value, which accumulates around the critical re-gion at the onset of the DPT. As λD1 (s) is real across therange of s considered (unlike λD2 (s), which is complex forsome values of s), the closing of the gap means that thestationary distribution becomes degenerate for |s| abovethe critical value in the limit L → ∞. For |s| below thecritical value, the stationary probability psstat = rDs,0 isunique and one can check numerically that it has a re-flection symmetry y → −y also possessed by the genera-tor: U Ws

DoobUT = WsDoob, U psstat = psstat, where U is the

permutation matrix that maps the lattice site at (x,±y)into the one at (x,∓y). For the argument that follows itis useful to decompose the stationary probability into itsupper and lower halves, as psstat = (ps↑+ps↓)/2, where ps↑(ps↓) is zero for y < 0 (y > 0); both ps↑ and ps↓ are nor-

malized, and U ps↑ = ps↓ (U ps↓ = ps↑). Beyond the critical

point, the right eigenspace (corresponding to the space ofstationary distributions) is spanned by rDs,0 = (ps↑+ps↓)/2

(which by continuity shares the symmetries of psstat be-low the critical value), and an orthogonal stationary statenumerically found to be rDs,1 = (ps↑ − ps↓)/2, for which

U rDs,1 = −rDs,1. More conveniently, one can use the ba-sis formed by ps↑ and ps↓, which concentrate (more and

more sharply as |s| is increased) around the upper andlower edge of the slit, respectively, as a basis to formpossible stationary distributions. A particle starts froman initial site, moves towards one of the slit edges, andremains there, as the probability of making the journeyto the other edge is vanishingly small for L → ∞. TheZ2 reflection symmetry is spontaneously broken.

To conclude our exploration of the DPT, we use theinverse participation ratio (IPR) as an order parame-ter. This is defined as the sum of the squared probabil-ities of occupation in the stationary state over all sites,IPR(s) =

∑n[psstat(n)]2, and is widely used in the char-

acterization of localized states [53]. In idealized caseswhere the probability is evenly distributed over N sites,and is zero outside, the IPR is 1/N . For s = 0 (flat solu-tion), IPR(0) = 1/L2, as shown in Fig. 2 (f). In the caseof infinitely sharp bands we would have IPR = 1/hL2

for the inner band (IPR = 1/(1 − h)L2 for the outer

band), therefore band dynamics correspond to IPRs be-tween 1/L2 and 1/hL2 (1/(1 − h)L2), and larger valuesmust arise from localized vortices. Accordingly, we setmax

{1/hL2, 1/(1− h)L2

}as a threshold, then use psstat

to calculate the IPR for values that are smaller than suchthreshold, and otherwise base the calculation of the IPRon ps↑ = rDs,0 + rDs,1 (or ps↓ = rDs,0 − rDs,1, as they give

equivalent results), as a particle is trapped in one slitedge. The argument works in the limit of large L, but weuse our finite-size results as approximations to that limit.The IPR shown in the main panel of Fig. 2 (f) reachesa small and stable value for s below the critical point.As expected from the characteristics of band dynamics,multiplication of the IPR by L2 in this range of s leadsto an excellent collapse, see the inset of Fig. 2 (f). Forlarge values of |s|, the IPR approaches a constant valueof 1/4 for |s| → ∞. The latter (N = 4) corresponds tothe existence of a vortex around and edge slit, where theloop comprises four lattice sites. If instead of the sta-tionary state ps↑ (or ps↓) we naively take psstat,0, the IPR

approaches a value of 1/8 for large |s|, corresponding tothe existence of two vortices at the edges (N = 8), as inthe right panels of Fig. 1 (b), but sharper.Conclusions– We have shown that the two-dimensionalrandom walk conditioned on partial currents undergoesa DPT between delocalized bands and localized vortices,where particles condensate around the slit edges and a Z2

reflection symmetry is spontaneously broken. Intrigu-ing localization phenomena have also been observed inother random walks, e.g. in the maximal entropy randomwalk on a lattice [54], which can also be understood asthe result of conditioning on a particular observable [55],or in random walks on complex networks, where someDPTs have been also investigated [56]. Many interestingquestions remain open, however, regarding, for example,the existence of similar localization effects when otherprocesses (including exclusion effects or interactions) ornumber of spatial dimensions are considered. Whethersuch effects are relevant in dissipative quantum walks[57, 58] (where the spectral theory of Liouvillians canbe brought to bear [52, 59]) is also an interesting ques-tion, as an answer in the affirmative might stimulate asearch for potential relationships to well-known localiza-tion effects in quantum mechanics [60, 61].

ACKNOWLEDGMENTS

We thank Ruben Hurtado-Gutierrez and Pablo Hur-tado for insightful discussions. The research leadingto these results has received funding from the Euro-pean Union’s Horizon 2020 research and innovation pro-gramme under the Marie Sklodowska-Curie Cofund Pro-gramme Athenea3I Grant Agreement No. 754446, fromthe European Regional Development Fund, Junta de An-dalucıa-Consejerıa de Economıa y Conocimiento, Ref. A-FQM-175-UGR18. We are grateful for the the comput-ing resources and related technical support provided by

Page 6: arXiv:2108.12359v1 [cond-mat.stat-mech] 27 Aug 2021

6

PROTEUS, the super-computing center of Institute Car-los I in Granada, Spain, and by CRESCO/ENEAGRIDHigh Performance Computing infrastructure and its staff[62], which is funded by ENEA, the Italian National

Agency for New Technologies, Energy and SustainableEconomic Development and by Italian and European re-search programmes.

[1] T. Bodineau and B. Derrida, “Current fluctuations innonequilibrium diffusive systems: An additivity princi-ple,” Phys. Rev. Lett. 92, 180601 (2004).

[2] T. Bodineau and B. Derrida, “Current large deviationsfor asymmetric exclusion processes with open bound-aries,” J. Stat. Phys. 123, 277–300 (2006).

[3] B. Derrida, “Non-equilibrium steady states: fluctuationsand large deviations of the density and of the current,”J. Stat. Mech. P07023 (2007) .

[4] D. J. Evans, E. G. D. Cohen, and G. P. Morriss, “Prob-ability of 2nd law violations in shearing steady-states,”Phys. Rev. Lett. 71, 2401–2404 (1993).

[5] G. Gallavotti and E. G. D. Cohen, “Dynamical ensem-bles in nonequilibrium statistical mechanics,” Phys. Rev.Lett. 74, 2694 (1995).

[6] J. Kurchan, “Fluctuation theorem for stochastic dynam-ics,” J. Phys. A 31, 3719–3729 (1998).

[7] J. L. Lebowitz and H. Spohn, “A Gallavotti-Cohen-typesymmetry in the large deviation functional for stochasticdynamics,” J. Stat. Phys. 95, 333–365 (1999).

[8] C. Maes, “The fluctuation theorem as a gibbs property,”Journal of Statistical Physics 95, 367–392 (1999).

[9] R. J. Harris and G. M. Schuetz, “Fluctuation theoremsfor stochastic dynamics,” J. Stat. Mech. , P07020 (2007).

[10] J. P. Garrahan, R. L. Jack, V. Lecomte, E. Pitard, K. vanDuijvendijk, and F. van Wijland, “Dynamical first-order phase transition in kinetically constrained modelsof glasses,” Phys. Rev. Lett. 98, 195702 (2007).

[11] J. P. Garrahan, R. L. Jack, V. Lecomte, E. Pitard, K. vanDuijvendijk, and F. van Wijland, “First-order dynamicalphase transition in models of glasses: an approach basedon ensembles of histories,” J. Phys. A 42, 075007 (2009).

[12] J. P. Garrahan and I. Lesanovsky, “Thermodynamicsof quantum jump trajectories,” Phys. Rev. Lett. 104,160601 (2010).

[13] F. Carollo, J. P. Garrahan, I. Lesanovsky, and C. Perez-Espigares, “Fluctuating hydrodynamics, current fluctu-ations, and hyperuniformity in boundary-driven openquantum chains,” Phys. Rev. E 96, 052118 (2017).

[14] F. Carollo, J. P. Garrahan, I. Lesanovsky, and C. Perez-Espigares, “Making rare events typical in Markovianopen quantum systems,” Phys. Rev. A 98, 010103 (2018).

[15] L. Bertini, A. De Sole, D. Gabrielli, G. Jona-Lasinio, andC. Landim, “Macroscopic fluctuation theory,” Rev. Mod.Phys. 87, 593–636 (2015).

[16] H. Touchette, “The large deviation approach to statisti-cal mechanics,” Phys. Rep. 478, 1 (2009).

[17] V. Lecomte, C. Appert-Rolland, and F. van Wijland,“Thermodynamic formalism for systems with Markov dy-namics,” J. Stat. Phys. 127, 51–106 (2007).

[18] H. B. Callen, Thermodynamics and an Introduction toThermostatistics (John Wiley & Sons, 2nd Ed., NewYork, 1985).

[19] D. Simon, “Construction of a coordinate bethe ansatzfor the asymmetric simple exclusion process with open

boundaries,” J. Stat. Mech. P07017 (2009).[20] V. Popkov, G. M. Schutz, and D. Simon, “ASEP on

a ring conditioned on enhanced flux,” J. Stat. Mech.P10007 (2010).

[21] R. L. Jack and P. Sollich, “Large deviations and ensem-bles of trajectories in stochastic models,” Prog. Theor.Phys. Supp. 184, 304–317 (2010).

[22] R. Chetrite and H. Touchette, “Nonequilibrium micro-canonical and canonical ensembles and their equiva-lence,” Phys. Rev. Lett. 111, 120601 (2013).

[23] R. Chetrite and H. Touchette, “Nonequilibrium Markovprocesses conditioned on large deviations,” Ann. HenriPoincare 16, 2005 (2015).

[24] L. Bertini, A. De Sole, D. Gabrielli, G. Jona-Lasinio,and C. Landim, “Current fluctuations in stochastic lat-tice gases,” Phys. Rev. Lett. 94, 030601 (2005).

[25] T. Bodineau and B. Derrida, “Distribution of current innonequilibrium diffusive systems and phase transitions,”Phys. Rev. E 72, 066110 (2005).

[26] L. Bertini, A. De Sole, D. Gabrielli, G. Jona-Lasinio,and C. Landim, “Nonequilibrium current fluctuations instochastic lattice gases,” J. Stat. Phys. 123, 237–276(2006).

[27] P. I. Hurtado and P. L. Garrido, “Spontaneous symmetrybreaking at the fluctuating level,” Phys. Rev. Lett. 107,180601 (2011).

[28] C. Perez-Espigares, P. L. Garrido, and P. I. Hurtado,“Dynamical phase transition for current statistics in asimple driven diffusive system,” Phys. Rev. E 87, 032115(2013).

[29] R. L. Jack, I. R. Thompson, and P. Sollich, “Hyper-uniformity and phase separation in biased ensembles oftrajectories for diffusive systems,” Phys. Rev. Lett. 114,060601 (2015).

[30] O. Tsobgni Nyawo and H. Touchette, “A minimal modelof dynamical phase transition,” Europhys. Lett. 116,50009 (2016).

[31] Y. Baek, Y. Kafri, and V. Lecomte, “Dynamical sym-metry breaking and phase transitions in driven diffusivesystems,” Phys. Rev. Lett. 118, 030604 (2017).

[32] R. J. Harris and H. Touchette, “Phase transitions in largedeviations of reset processes,” J. Phys. A 50, 10LT01(2017).

[33] A. Lazarescu, “Generic dynamical phase transition inone-dimensional bulk-driven lattice gases with exclu-sion,” J. Phys. A 50, 254004 (2017).

[34] C. Perez-Espigares, F. Carollo, J. P. Garrahan, and P. I.Hurtado, “Dynamical criticality in open systems: Non-perturbative physics, microscopic origin, and direct ob-servation,” Phys. Rev. E 98, 060102 (2018).

[35] C. Perez-Espigares, I. Lesanovsky, J. P. Garrahan, andR. Gutierrez, “Glassy dynamics due to a trajectory phasetransition in dissipative Rydberg gases,” Phys. Rev. A98, 021804 (2018).

[36] C. Perez-Espigares and P. I. Hurtado, “Sampling rare

Page 7: arXiv:2108.12359v1 [cond-mat.stat-mech] 27 Aug 2021

7

events across dynamical phase transitions,” Chaos 29,083106 (2019).

[37] M. C. Banuls and J. P. Garrahan, “Using matrix productstates to study the dynamical large deviations of kineti-cally constrained models,” Phys. Rev. Lett. 123, 200601(2019).

[38] R. Hurtado-Gutierrez, F. Carollo, C. Perez-Espigares,and P. I. Hurtado, “Building continuous time crystalsfrom rare events,” Phys. Rev. Lett. 125, 160601 (2020).

[39] P. I. Hurtado, C. Perez-Espigares, J. J. del Pozo, andP. L. Garrido, “Symmetries in fluctuations far from equi-librium,” Proc. Natl. Acad. Sci. USA 108, 7704–7709(2011).

[40] R. Villavicencio-Sanchez, R. J. Harris, and H. Touchette,“Fluctuation relations for anisotropic systems,” Euro-phys. Lett. 105, 30009 (2014).

[41] C. Perez-Espigares, P. L. Garrido, and P. I. Hurtado,“Weak additivity principle for current statistics in d-dimensions,” Phys. Rev. E 93, 040103(R) (2016).

[42] N. Tizon-Escamilla, C. Perez-Espigares, P. L. Garrido,and P. I. Hurtado, “Order and symmetry-breaking in thefluctuations of driven systems,” Phys. Rev. Lett. 119,090602 (2017).

[43] R. L. Jack, T. Nemoto, and V. Lecomte, “Dynamicalphase coexistence in the fredrickson–andersen model,” J.Stat. Mech. 2020, 053204 (2020).

[44] C. Casert, T. Vieijra, S. Whitelam, and I. Tamblyn,“Dynamical large deviations of two-dimensional kinet-ically constrained models using a neural-network stateansatz,” arXiv:2011.08657 (2020).

[45] T. Bodineau, B. Derrida, and J.L. Lebowitz, “Vorticesin the two-dimensional simple exclusion process,” J. Stat.Phys. 131, 821 (2008).

[46] The normalization of the Doob-generator eigenvectorsis such that the largest absolute value of the compo-nents of lDs,i is one, and (lDs,i)

T rDs,i = 1. Therefore, lDs,0 =

(1, 1, . . . , 1)T and the sum of all the components of rDs,0is one (ps

stat = rDs,0 is a probability vector).[47] See the Supplemental Material for a detailed macroscopic

analysis of the different dynamical regimes under consid-eration.

[48] R. Gutierrez and C. Perez-Espigares, “Generalized opti-mal paths and weight distributions revealed through thelarge deviations of random walks on networks,” Phys.

Rev. E 103, 022319 (2021).[49] R. Villavicencio-Sanchez, R. J. Harris, and H. Touchette,

“Current loops and fluctuations in the zero-range processon a diamond lattice,” J. Stat. Mech. P07007 (2012).

[50] H. Spohn, Large Scale Dynamics of Interacting Particles,Theoretical and Mathematical Physics (Springer BerlinHeidelberg, 2012).

[51] B. Gaveau and L. S. Schulman, “Multiple phases instochastic dynamics: Geometry and probabilities,” Phys.Rev. E 73, 036124 (2006).

[52] F. Minganti, A. Biella, N. Bartolo, and C. Ciuti, “Spec-tral theory of liouvillians for dissipative phase transi-tions,” Phys. Rev. A 98, 042118 (2018).

[53] D. J Thouless, “Electrons in disordered systems and thetheory of localization,” Phys. Rep. 13, 93–142 (1974).

[54] Z. Burda, J. Duda, J. M. Luck, and B. Waclaw, “Lo-calization of the maximal entropy random walk,” Phys.Rev. Lett. 102, 160602 (2009).

[55] F. Coghi, J. Morand, and H. Touchette, “Large devia-tions of random walks on random graphs,” Phys. Rev. E99, 022137 (2019).

[56] C. De Bacco, A. Guggiola, R. Kuhn, and P. Paga, “Rareevents statistics of random walks on networks: localisa-tion and other dynamical phase transitions,” J. Phys. A49, 184003 (2016).

[57] S. Attal, F. Petruccione, and I. Sinayskiy, “Open quan-tum walks on graphs,” Phys. Lett. A 376, 1545–1548(2012).

[58] S. Garnerone, “Thermodynamic formalism for dissipativequantum walks,” Phys. Rev. A 86, 032342 (2012).

[59] K. Macieszczak, M. Guta, I. Lesanovsky, and J. P. Gar-rahan, “Towards a theory of metastability in open quan-tum dynamics,” Phys. Rev. Lett. 116, 240404 (2016).

[60] P. W. Anderson, “Absence of diffusion in certain randomlattices,” Phys. Rev. 109, 1492–1505 (1958).

[61] R. Nandkishore and D. A. Huse, “Many-body localiza-tion and thermalization in quantum statistical mechan-ics,” Annu. Rev. Condens. Matter Phys. 6, 15–38 (2015).

[62] F. Iannone, F. Ambrosino, G. Bracco, M. De Rosa, A. Fu-nel, G. Guarnieri, S. Migliori, F. Palombi, G. Ponti,G. Santomauro, and P. Procacci, “CRESCO ENEA HPCclusters: a working example of a multifabric GPFS Spec-trum Scale layout,” in 2019 International Conference onHigh Performance Computing Simulation (HPCS) (2019)p. 1051.

Page 8: arXiv:2108.12359v1 [cond-mat.stat-mech] 27 Aug 2021

8

SUPPLEMENTAL MATERIAL

Dynamical phase transition to localized states in the two-dimensional randomwalk conditioned on partial currents

Ricardo Gutierrez,1 and Carlos Perez-Espigares,2,3

1Complex Systems Interdisciplinary Group (GISC), Department of Mathematics, Universidad Carlos III de Madrid,28911 Leganes, Madrid, Spain

2Departamento de Electromagnetismo y Fısica de la Materia, Universidad de Granada, Granada 18071, Spain3Instituto Carlos I de Fısica Teorica y Computacional, Universidad de Granada, Granada 18071, Spain

In this Supplemental Material, we study macroscopic ansatze for the density and current fields that can be consideredas idealizations of the dynamical regimes that are revealed in the microscopic analysis of the two-dimensional randomwalk conditioned on mean values of the partial current across a slit.

S1. MACROSCOPIC FLUCTUATION THEORY OF THE TWO-DIMENSIONAL RANDOM WALK

The macroscopic fluctuation theory (MFT) is a powerful framework for the study of dynamical fluctuations ofdriven diffusive systems [15]. When such systems take place over a square region of linear size L, the probability of atrajectory of duration T , {ρ(r, t), j(r, t)}T0 , adopts a large-deviation form

P [ρ(r, t), j(r, t)] ∼ e−L2I[ρ,j] (S1)

where the so-called rate functional is as follows (the spatio-temporal dependence of ρ and j is omitted for ease ofnotation)

I[ρ, j] =

∫ T

0

dt

∫Λ

dr[j +D(ρ)∇ρ− σ(ρ)E]

2

2σ(ρ). (S2)

The spatial coordinates have been rescaled so that r ∈ Λ = [−1/2, 1/2]×[−1/2, 1/2] (i.e. a square region of unit areawith the origin (0, 0) at its center). The probability is maximized around the macroscopic average current, which inthe presence of a uniform driving field is E is j = −D(ρ)∇ρ+σ(ρ)E, where the diffusivity D(ρ) and the mobility σ(ρ)are in general functions of the density field.

For the particular case of a random walk, D = 1/4 and σ(ρ) = ρ/2, the functional can be expanded as follows :

I[ρ, j] =

∫ T

0

dt

∫Λ

dr

[j + 1

4∇ρ−12ρE

]2ρ

=

∫ T

0

dt

∫Λ

drj2 + 1

16 (∇ρ)2 + 14ρ

2E2 + 12 j · ∇ρ− ρ j ·E−

14ρ∇ρ ·E

ρ

=

∫ T

0

dt

∫Λ

drj2+ 1

16 (∇ρ)2

ρ+E2

4

∫ T

0

dt

∫Λ

dr ρ+1

2

∫ T

0

dt

∫Λ

drj · ∇ρρ−∫ T

0

dtE ·(∫

Λ

dr j− 1

4

∫Λ

dr∇ρ). (S3)

As we assume the system has periodic boundary conditions, the last term is∫

Λdr∇ρ = 0, and the following integral

can be simplified using integration by parts,∫ T

0

dt

∫Λ

drj · ∇ρρ

=

∫ T

0

dt

∫Λ

dr j · ∇ log ρ =

∫ T

0

dt

∫Λ

dr∇ · (j log ρ)−∫ T

0

dt

∫Λ

dr (∇ · j) log ρ

=

∫ T

0

dt

∫Λ

dr (∂tρ) log ρ =

∫Λ

dr

∫ T

0

dt ∂t(ρ log ρ− ρ) = A(T )−A(0). (S4)

where the density and current fields are coupled via the continuity equation ∂tρ + ∇ · j = 0. The function A(t) =∫Λdr(ρ log ρ− ρ) is bounded in time, which distinguishes the corresponding term in Eq. (S3) from the others, as they

are all time-extensive. Considering that the system is closed with a fixed number of particles N = ρ0L2, by definition

ρ0 =∫

Λdr ρ. Together with the definition of a global current JG =

∫Λdr j, this leads to a simpler functional,

I[ρ, j] =

∫ T

0

dt

∫Λ

drj2 + 1

16 (∇ρ)2

ρ+E2Tρ0

4+

1

2(A(T )−A(0))−E ·

∫ T

0

dtJG. (S5)

Page 9: arXiv:2108.12359v1 [cond-mat.stat-mech] 27 Aug 2021

9

FIG. S1. Dynamical regimes for the two-dimensional random walk conditioned on different values of the partialcurrent J across a slit (vertical strip in the center). These regimes (and superpositions of them) are idealizations ofthose observed in the microscopic analysis based on the Doob transform. (a) Flat profile. (b) Inner band. (c) Outer band. (d)Vortices.

We next calculate by contraction from the MFT functional (S5) the probability of having a current J through theslit, which reads

P (J) ∼ exp{−TL2G(J)},

with G(J) = limT→∞1T min∗{ρ,j}T0

I[ρ, j]. Here ∗ means that the minimization is subject to the constraints J =

T−1∫ T

0dt∫ h/2−h/2 dyjx(x, y; t) and ∂tρ = −∇· j. This variational problem will be applied to different dynamical regimes

that are idealizations of the time-independent density and current fields observed in the stationary states of the Doob-transformed microscopic dynamics and discussed in the main text. Sketches of these idealized regimes are providedin Fig. S1. They all refer to time-independent density and current fields, for which the rate function can be writtenas

G(J) = min∗{ρ,j}

[∫Λ

drj2 + 1

16 (∇ρ)2

ρ+E2ρ0

4−E · JG

]. (S6)

Different forms of the density and current field conditioned on a given partial current across the slit J will be evaluatedin each particular regime, always focusing on situations where E is horizontal and points to the right, E = Ex. When agiven parameter characterizes an idealized regime, as embodied in a given ansatz for {ρ(r), j(r)}, the parameter choicethat minimizes G(J) for that value of J is the one that is (overwhelmingly) more likely to be observed. When differentansatze, each one optimized for their corresponding parameters, are compared for a given J , the one that yields aminimum value of G(J) is the one that would be observed in practice. The third term in (S5) has been neglected,since the discussion focuses on the long-time limit, and, as we pointed out before, lim

T→∞(A(T )−A(0))/T = 0.

S2. FLAT SOLUTION

For a given partial current J across a slit of length h, a flat density profile that sustains a uniform current, seeFig. S1 (a), takes the following values:

ρ = ρ0, j =J

hx. (S7)

The time-rescaled MFT functional (S6) takes a simple quadratic form:

Gflat(J) =J2

h2ρ0+E2ρ0

4− EJ

h= ρ0

(J

hρ0− E

2

)2

. (S8)

The fluctuations around the mean value 〈J〉 =Ehρ0

2are thus Gaussian, with a variance that decreases as 1/T ,

Pflat(J) ∼ exp

[−(J − 〈J〉)2

/

(h2ρ0

L2T

)]. (S9)

Page 10: arXiv:2108.12359v1 [cond-mat.stat-mech] 27 Aug 2021

10

S3. INNER BAND SOLUTION

When next analyze the case when the density profile forms a horizontal band that decays for |y| > h/2, and thecurrent is proportional to the density. For simplicity, we assume an exponential decay:

ρ(x, y) =

C e−(y−h/2)/` y > h/2

C −h/2 ≤ y ≤ h/2 j(x, y) = j0ρ(x, y)x.

C e(y+h/2)/` y < −h/2(S10)

In this case, there is one parameter in the solution, which is the characteristic length of the exponential decay `. InFig. S1 (b) this regime is illustrated for `� h, there the density profile as a function of y approaches a step function.Both density and current are invariant under translations along the horizontal axis.

In order to determine C we perform the integral of the density over the square region Λ and equate it to ρ0:

ρ0 =

(∫ −h/2−1/2

dyρ(x, y) +

∫ h/2

−h/2dyρ(x, y) +

∫ 1/2

h/2

dyρ(x, y)

)= C

(`e(y+h/2)/`

∣∣∣−h/2−1/2

+ h− `e−(y−h/2)/`∣∣∣1/2h/2

)= C

(2`(

1− e−(1−h)/2`)

+ h)

=⇒ C =ρ0

2`(1− e−(1−h)/2`

)+ h

. (S11)

In the following it will be useful to divide the integrated density in two contributions, that corresponding to particlespassing through the slit ρin(`), which is decreasing in `, and that corresponding to particles moving across the regionwhere |y| > h/2, ρout(`), which increases with `,

ρin(`) = Ch =ρ0h

2`(1− e−(1−h)/2`

)+ h

, ρout(`) = ρ0 − ρin(`) =ρ02`

(1− e−(1−h)/2`

)2`(1− e−(1−h)/2`

)+ h

. (S12)

As for the proportionality constant j0, it can be obtained from the fact that the current integrated over the slit is bydefinition J :

J =

∫ h/2

−h/2dy j(x, y) = j0

∫ h/2

−h/2dy ρ(x, y) = j0ρin(`), j0 =

J

ρin(`). (S13)

The rate functional (S6) thus takes the form

Gi.b.(J, `)=

∫Λ

drj20ρ

2+ 116 (∂yρ)2

ρ+E2ρ0

4−Ej0ρ0 = ρ0

(j0−

E

2

)2

+

∫|y|>h/2dr ρ

16`2= ρ0

(J

ρin(`)−E

2

)2

+ρout(`)

16`2. (S14)

The first term quantifies the cost of having a density ρin(`) different from 2J/E, and the second that of forming anon-uniform density field. The flat solution previously discussed corresponds to ρin(`) = ρ0h, hence the first term isminimized for J = Ehρ0/2 = 〈J〉 in that case. As the density is then uniform, `→∞, the second term proportionalto ρout(`) also vanishes.

Of all possible band profiles, which are characterized by different values of the characteristic length `, the one thatis most likely to be observed for a given J is the one that minimizes the functional. For flucutations around the meanvalue we have two possibilities:

• if J > 〈J〉, a band is formed, whose vertical profile is steeper as J increases, as, in order to minimize the first termwe need some ρin(`) > ρ0h, which requires having a finite value of ` (non-flat solution), so ρout(`) < ρ0(1− h);

• if 0 < J ≤ 〈J〉, we still have a flat solution, as ρ0h is the lowest value that ρin(`) can achieve, and moreover thesecond term vanishes for such flat solution, `→∞.

For J > 〈J〉 there is an effective competition between the first and the second term in Eq. (S14), as the second,ρout(`)/16`2, includes the cost of creating a non-uniform density profile, and is a decreasing function of `. In thelimit of very large J , the first term clearly dominates and the system adopts an abrupt vertical profile, ` ≈ 0, but allintermediate cases between 〈J〉 and such large values of J must be determined from a minimization of Gi.b.(J, `).

Given in terms of `, Eq. (S14) is too cumbersome for a convenient analytical minimization. Instead, the values of `that minimize this function have been found numerically. In Fig. S2 (a) we provide a surface plot showing Gi.b.(J, `)as a function of J and ` for E = 5 and h = 1/2. The values of ` where the minimum is achieved for each J aredisplayed as a gray continuous line. As expected, it turns out that for 0 < J ≤ 〈J〉 the minimum is achieved for`→∞, and for J > 〈J〉 the position of the minimum gets closer and closer to zero as J increases. 〈J〉 is highlightedby a red discontinuous line.

Page 11: arXiv:2108.12359v1 [cond-mat.stat-mech] 27 Aug 2021

11

FIG. S2. Macroscopic functionals for the inner band, outer band and vortex solution as functions of thecharacteristic length ` and the partial current J. (a) Inner-band functional Gi.b.(J, `). (b) Outer-band functionalGo.b.(J, `). (a) Vortex functional Gvort.(J, `). In each case, the location of the minimum value of the functional for each J isshown as a gray continuous line. In (a) and (b) the red discontinuous line corresponds to the average partial current 〈J〉.

S4. OUTER BAND SOLUTION

We next consider the case when the density profile forms a horizontal band that avoids passing through the slit,which is constant for |y| > h/2 and decays for |y| ≤ h/2, see Fig. S1 (c), where this regime is illustrated for ` � h.As in the previous case, the current is taken to be proportional to the density. For simplicity, we again assume anexponential decay:

ρ(x, y) =

C y > h/2

C

1 + e−h/`

(e(y−h/2)/` + e−(y+h/2)/`

)−h/2 ≤ y ≤ h/2 j(x, y) = j0ρ(x, y)x.

C y < −h/2

(S15)

where prefactor of the middle term has be chosen so as to ensure continuity. In order to determine C we need tonormalize the integral of the density over the square region Λ and equate it to ρ0:

ρ0 =

∫ −h/2−1/2

dyρ(x, y) +

∫ h/2

−h/2dyρ(x, y) +

∫ 1/2

h/2

dyρ(x, y) = C

(1− h+

`

1 + e−h/l

(e(y−h/2)/` − e−(y+h/2)/`

)∣∣∣h/2−h/2

)= C

(1− h+ 2` tanh

(h

2`

))=⇒ C =

ρ0

1− h+ 2` tanh(h2`

) . (S16)

As in the previous case, it will be useful to divide the integrated density in two contributions, that corresponding toparticles passing through the slit ρin(`), which is increasing in `, and that corresponding to particles moving acrossthe region where |y| > h/2, ρout(`), which decreases with `,

ρout(`) = C(1− h) =ρ0(1− h)

1− h+ 2` tanh(h2`

) , ρin(`) = ρ0 − ρout(`) =ρ02` tanh

(h2`

)1− h+ 2` tanh

(h2`

) . (S17)

As for the proportionality constant j0, it can be obtained from the fact that the current integrated over the slit is bydefinition J :

J =

∫ h/2

−h/2dy j(x, y) = j0

∫ h/2

−h/2dy ρ(x, y) = j0ρin(`), j0 =

J

ρin(`). (S18)

Page 12: arXiv:2108.12359v1 [cond-mat.stat-mech] 27 Aug 2021

12

The rate functional is in this case

Go.b.(J, `)=

∫Λ

drj20ρ

2 + 116 (∂yρ)2

ρ+E2ρ0

4− Ej0ρ0 = ρ0

(j0 −

E

2

)2

+1

16

∫|y|<h/2

dr(∂yρ)2

ρ

= ρ0

(J

ρin(`)− E

2

)2

+C

16`2(1 + e−h/`)

∫ h/2

−h/2dy

(e(y−h/2)/` + e−(y+h/2)/` − 4e−h/`

e(y−h/2)/` + e−(y+h/2)/`

)= ρ0

(J

ρin(`)− E

2

)2

+C

16`2(1 + e−h/`)

(2`(1− e−h/`)− 4e−h/2`

∫ h/2

−h/2dy

ey/`

1 + e2y/`

)

= ρ0

(J

ρin(`)− E

2

)2

+ρ0e−h/2` (sinh

(h2`

)+ arctan e−h/2` − arctan eh/2`

)4`(1 + e−h/`)(1− h+ 2` tanh

(h2`

))

. (S19)

Again, the first term quantifies the cost of having a density ρin(`) different from 2J/E, and the second that of forminga non-uniform density field. The flat solution is given by to ρin(`) = ρ0h, hence the first term is minimized for J = 〈J〉in that case. This corresponds to a uniform density, `→∞, for which the second term also vanishes.

Of all possible band profiles, which are characterized by different values of the characteristic length `, the one thatis most likely to be observed for a given J , is the one that minimizes the functional. For flucutations around the meanvalue we have two possibilities:

• if 0 < J < 〈J〉, a band is formed, whose vertical profile is steeper as J decreases from 〈J〉, at least up to certainvalue of J (more about this later), as, in order to minimize the first term we need some ρin(`) < ρ0h, whichrequires having a finite value of ` (non-flat solution), so ρout(`) > ρ0(1− h);

• if J ≥ 〈J〉, we have a flat solution, as ρ0h is the highest value that ρin(`) can achieve, and moreover the secondterm vanishes for such flat solution, `→∞.

For 0 < J < 〈J〉 there is an effective competition between the first and the second term in Eq. (S19), as the secondincludes the cost creating a non-uniform density profile, and is a decreasing function of `. As J decreases from theaverage value 〈J〉, the first term tries to adapt to a smaller ρin(`), so ` becomes smaller, at the cost of creating anon-uniform profile, which contributes to the second term of Go.b.(J, `). However, when J gets very small, deviationsaround the average in the first term become less important than the second term, and therefore we again obtain aflat solution, `→∞.

The values of ` that minimize Go.b.(J, `) have been found numerically. In Fig. S2 (b) we provide a surface plotshowing Go.b.(J, `) as a function of J and ` for E = 5 and h = 1/2. The values of ` where the minimum value isachieved for each J are displayed as a gray continuous line. As expected, it turns out that for J > 〈J〉 the minimumis achieved for ` → ∞, and for 0 < J ≤ 〈J〉 the value of ` where the minimum is achieved has the non-monotonicbehavior discussed in the previous paragraph. 〈J〉 is highlighted by a red discontinuous line.

S5. VORTICES

Finally we consider a solution where vortices form around the edges of the slit, see Fig. S1 (d), which as mentionedin the main text resembles a configuration previously described in the two-dimensional simple exclusion processconditioned on partial currents [45]. Due to the rotational symmetry of the solution, the density is best representedin polar coordinates, by some decreasing function of the radial distance r from the center of the vortex. In fact thereare two vortices, each one located in the vicinity of each of the two edges of the slit. We again assume the densityprofile to be of exponential form due to its simplicity

ρ(r, ϕ) =

{0 r ≤ r,Ce−(r−r)/` r > r.

(S20)

Again the steepness of the profile is given by the characteristic length `. The role played by the cutoff radius r willbe discussed later. The normalization in this case is given by

ρ0 = 2

∫ 2π

0

∫ ∞r

dr r C e−(r−r)/` = 4πC

(− `re−(r−r)/`

∣∣∣∞r

+ `

∫ ∞r

dre−(r−r)/`)

= 4πC` (r + `) =⇒ C =ρ0

4π` (r + `). (S21)

Page 13: arXiv:2108.12359v1 [cond-mat.stat-mech] 27 Aug 2021

13

The factor of 2 multiplying the integral of the density arises from the existence of two vortices. The upper limit ofintegration in the integral over r should not exceed h/2, but at least for `� h the simplification arising from takingit to infinity is expected to compensate for any slight loss of accuracy, as we will explain below.

The current at the vortices in the microscopic results was numerically found to be roughly proportional to thedensity divided by r. For simplicity we assume a strict proportionality j(r) ∝ ρ(r)/r in our ansatz:

j(r) = j0ρ(r)

rϕ =

0 r ≤ r,j0 C e

−(r−r)/`

rϕ r > r.

(S22)

The proportionality factor j0 is as usual determined from a calculation of the partial current through the slit

J = 2

∫ ∞r

drj(r) = 2j0Cer/`

∫ ∞r

dre−r/`

r= 2j0Ce

r/`Γ(0, r/`) =⇒ j0 =Je−r/`

2C Γ(0, r/`)=J2π` (r + `) e−r/`

ρ0Γ(0, r/`), (S23)

where the final expression contains an incomplete gamma function evaluation.The rate functional is in this case

Gvort.(J, `)=

∫Λ

drj2 + 1

16 (∂rρ)2

ρ+E2ρ0

4−E · JG = 2j2

0

∫ 2π

0

∫ ∞r

drrρ(r)

r2+

1

8`2

∫ 2π

0

∫ ∞r

drr ρ(r) +E2ρ0

4

= 4πCj20er/`

∫ ∞r

dre−r/`

r+

ρ0

8`2+E2ρ0

4=J24π2` (r + `) e−r/`

ρ0Γ(0, r/`)+

ρ0

8`2+E2ρ0

4, (S24)

where we have taken into consideration that, due to the geometry of the vortices, the total current vanishes JG = 0.The values of ` that minimize Gvort.(J, `) have been found numerically. In Fig. S2 (c) we provide a surface plot

showing Gvort.(J, `) as a function of J and ` for E = 5 and h = 1/2. The values of ` where the minimum value isachieved for each J are displayed as a gray continuous line. We see that the ` that minimizes Gvort.(J, `) decreases as|J | increases. For |J | ' 2, the condition `� h that was mentioned before in order to simplify some integrals is indeedsatisfied. In the figure we have set the cutoff r = 0.001, as the qualitative behavior with ` is independent of this choice(for further details see the next section). As for the quantitative influence of the cutoff r on the minimum value thatis achieved by Gvort.(J, `) for a given J , it will be discussed next when the MFT functionals of the four dynamicalregimes that have been considered (that is, flat solution, inner band, outer band, and vortices) are compared for thesame range of J .

S6. COMPARISON OF DIFFERENT DYNAMICAL REGIMES

In order to determine the most likely form of a fluctuation away from the mean, J 6= 〈J〉, the functionals correspond-ing to the different dynamical regimes must be compared. This includes the flat solution Gflat(J), the inner-bandsolution Gi.b.(J, `

∗(J)), the outer-band solution Go.b.(J, `∗(J)) and the vortex solution Gvort.(J, `

∗(J)), where `∗(J)denotes the value of ` for which each of those functionals are minimized for a given J , which was highlighted by graycontinuous lines in the different panels of Fig. S2. The dependence on `∗(J) is omitted in the main text, where forlack of space it is not possible to discuss the role played by the characteristic length `.

A comparison of the different dynamical regimes is provided in Fig. S3. Aside from the functionals correspondingto the different ansatze, we include the functional that takes the minimum value at each J ,

Gmin.(J) = min{Gflat(J), Gi.b.(J, `∗(J)), Go.b.(J, `

∗(J)), Gvort.(J, `∗(J))},

and also its convex envelope GMaxw.(J). While Gmin.(J) gives the functional that maximizes the probability for eachvalue of J , from which the scaled cumulant-generating function of Fig. 2 (b) in the main text has been obtained viaa Legendre-Fenchel transform, it is not a convex function. If the inverse Legendre-Fenchel transform is performed onthe scaled cumulant-generating function, we obtain GMaxw.(J). This Maxwell construction highlights a coexistencebetween the different dynamical phases [16].

The two panels show the same results, but the axis ranges are very different, so as to focus on different aspects. InFig. S3 (a), which is meant to facilitate the comparison between the different dynamical regimes, we can clearly seethat the minimum value changes across the range of J as follows:

Gmin.(J) =

Gvort.(J, `∗(J)) J < 0,

Go.b.(J, `∗(J)) 0 ≤ J < 〈J〉,

Gflat(J) J = 〈J〉,Gi.b.(J, `

∗(J)) 〈J〉 < J < J,

Gvort.(J, `∗(J)) J ≤ J.

(S25)

Page 14: arXiv:2108.12359v1 [cond-mat.stat-mech] 27 Aug 2021

14

The average 〈J〉 corresponds to the absolute minimum Gmin.(J) where the flat solution and the band solutions areequal, Gmin.(〈J〉) = Gflat(〈J〉) = Gi.b.(〈J〉, `∗(〈J〉)) = Go.b.(〈J〉, `∗(〈J〉)), as expected from the discussions above. Allsolutions take the value Gflat(0) = ρ0E

2/4 for J = 0, including Gvort.(J, `∗(J)), which dominates immediately to the

left, i.e. for J < 0, —corresponding to counterclockwise rotation in the upper edge and clockwise rotation in the loweredge of the slit— and also for J > J —corresponding to clockwise rotation in the upper edge and counterclockwiserotation in the lower edge of the slit—, where the latter value is to some extent dependent on the choice of the cutoffr. Indeed Gvort.(J, `

∗(J)) for different values of r is qualitatively similar, starting from the same value at J = 0, andgrowing symmetrically to right and left. However, the steepness of the increase as |J | grows decreases as r is made

smaller, which in Gmin.(J) is only reflected in a slower growth for J < 0 or J > J , and in a leftwards shift of J .

FIG. S3. Comparison of the MFT functional for the different dynamical regimes under consideration. TheMFT functional of the flat solution Gflat(J), the inner-band solution Gi.b.(J, `

∗(J)), the outer-band solution Go.b.(J, `∗(J)) and

the vortex solution Gvort.(J, `∗(J)) are displayed. The minimum value among them corresponding to each J , which is denoted

Gmin.(J), is also shown, as well as its convex envelope GMaxw.(J). Panels (a) and (b) contain the same results with significantlydifferent axis ranges, thus making it possible to inspect different aspects of them (see text for an explanation). Fig. 1 (c) inthe main text contains the same results with yet another choice of axis ranges.

In Fig. S3 (b) GMaxw.(J) is shown to arise from a coexistence of the outer-band solution and the vortex solution forJ < 〈J〉 and a coexistence of the inner-band solution and the vortex solution for J > 〈J〉. Such type of coexistence,which seems to be consistent with the superpositions of bands and vortices observed in the Doob-transformed micro-scopic dynamics of Fig. 1 in the main text, should not be interpreted literally as signaling the existence of a first-orderdynamical phase transition in the random walks, as we are here considering a limited subset of idealized dynamicalbehaviors, and what is here described by a coexistence of them may in fact correspond to dynamical behaviors thatare not fully describable in terms of the ansatze under consideration. In fact, our Gmin.(J) is an upper bound of theactual macroscopic functional restricted to a quite limited subset of density and current fields {ρ(r), j(r)} (thoughqualitatively representative of some of the main features that are observed in the microscopic analysis), as is frequentlythe case is the analysis of mean-field solutions of dynamical large deviations of many-body systems [35]. Yet, despiteits simplicity, our study does provide a sound qualitative understanding of the dynamical regimes under consideration—if not of the nature of the phase transition, which is actually continuous, as explained in the main text.