This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Non-equilibrium phase transitions in active rank diffusions

Léo Touzo Laboratoire de Physique de l’École Normale Supérieure, CNRS, ENS &\& PSL University, Sorbonne Université, Université Paris Cité, 75005 Paris, France    Pierre Le Doussal Laboratoire de Physique de l’École Normale Supérieure, CNRS, ENS &\& PSL University, Sorbonne Université, Université Paris Cité, 75005 Paris, France
(September 19, 2025)
Abstract

We consider NN run and tumble particles in one dimension interacting via a linear 1D Coulomb potential, an active version of the rank diffusion problem. It was solved previously for N=2N=2 leading to a stationary bound state in the attractive case. Here the evolution of the density fields is obtained in the large NN limit in terms of two coupled Burger’s type equations. In the attractive case the exact stationary solution describes a non-trivial NN-particle bound state, which exhibits transitions between a phase where the density is smooth with infinite support, a phase where the density has finite support and exhibits ”shocks”, i.e. clusters of particles, at the edges, and a fully clustered phase. In presence of an additional linear potential, the phase diagram, obtained for either sign of the interaction, is even richer, with additional partially expanding phases, with or without shocks. Finally, a general self-consistent method is introduced to treat more general interactions. The predictions are tested through extensive numerical simulations.

The run-and-tumble particle (RTP), driven by telegraphic noise ML17 ; kac74 , is the simplest model of an active particle which mimics e.g. the motion of E. Coli bacteria Berg2004 ; TailleurCates . Interacting RTP’s, which exhibit remarkable collective effects, is a thriving area of research TailleurCates ; FM2012 ; soft ; CT2015 ; slowman ; slowman2 ; BG2021 ; SG2014 . One outstanding question is to characterize the non Gibbsian steady state reached at large time. Even in one dimension (1D), there are very few exact results, beyond two interacting RTP’s slowman ; slowman2 ; us_bound_state ; Maes_bound_state ; nonexistence ; MBE2019 ; KunduGap2020 ; LMS2019 , or for harmonic chains SinghChain2020 ; PutBerxVanderzande2019 . Recently, analytical results were obtained for some many-particle models on a 1D lattice with contact interactions Metson2022 ; MetsonLong ; Dandekar2020 ; Thom2011 , and for an active version of the Dyson Brownian motion in the continuum with logarithmic interactions TouzoDBM2023 .

An analytically tractable passive example is the Langevin dynamics of Brownian particles interacting via the linear Coulomb 1D potential. It is known as ranked diffusion since the force acting on each particle is proportional to its rank, i.e., the number of particles in front of it, minus the number behind. Introduced in finance and mathematics Banner ; Pitman , it is related to the statistical mechanics of the self-gravitating 1D gas Rybicki ; Sire (attractive case) and to the Jellium model (repulsive case, in presence of a confining potential) much studied at equilibrium Lenard ; Baxter ; Tellez ; Lewin ; SatyaJellium1 ; Chafai_edge ; Flack22 . Recently, a detailed description of the non-equilibrium dynamics was obtained PLDRankedDiffusion , using both a mapping to the delta Bose gas, solvable by the Bethe ansatz, as well as the Dean-Kawasaki approach Dean ; Kawa , which leads to a description of the density in terms of a noisy Burgers equation PLDRankedDiffusion . In the attractive case, the shock solutions of the Burgers equation correspond to macroscopic bound states, while in the repulsive case the system forms an expanding crystal at large time FlackRD .

In view of these results, it is natural to extend the model to the active case. In this Letter we study NN run-and-tumble particles (RTP) at positions xi(t)x_{i}(t) in d=1d=1, interacting via the 1D Coulomb potential and evolving via the equation of motion

dxidt=κ¯Nj=1Nsgn(xjxi)V(xi)+v0σi(t)+2Tξi(t)\frac{dx_{i}}{dt}=\frac{\bar{\kappa}}{N}\sum_{j=1}^{N}{\rm sgn}(x_{j}-x_{i})-V^{\prime}(x_{i})+v_{0}\sigma_{i}(t)+\sqrt{2T}\xi_{i}(t) (1)

Here the ξi(t)\xi_{i}(t) are unit independent white noises and the σi(t)=±1\sigma_{i}(t)=\pm 1 are independent telegraphic noises with rate γ\gamma. The particles can cross freely, i.e. there is no hard core repulsion, and by convention sgn(0)=0{\rm sgn}(0)=0. The case κ¯>0\bar{\kappa}>0 corresponds to attractive interactions and κ=κ¯>0\kappa=-\bar{\kappa}>0 to repulsive ones. The RTP’s evolve in a common external potential V(x)V(x). When V(x)=0V(x)=0 the center of mass x¯=1Nixi\bar{x}=\frac{1}{N}\sum_{i}x_{i} evolves as a sum of NN independent RTP’s with velocities v0/Nv_{0}/N, hence it diffuses at large time as x¯2DNt\bar{x}\sim\sqrt{2D_{N}t} with DN=1N(T+v022γ)D_{N}=\frac{1}{N}(T+\frac{v_{0}^{2}}{2\gamma}).

Until now this model was studied only for N=2N=2, in the attractive case κ¯>0\bar{\kappa}>0, and for V(x)=0V(x)=0. It was shown us_bound_state that the PDF P(z,t)P(z,t) of the relative coordinate z=x1x¯=x1x22z=x_{1}-\bar{x}=\frac{x_{1}-x_{2}}{2} reaches a stationary limit P(z)P(z) which describes a bound state. For T>0T>0, P(z)P(z) is the sum of three decaying exponentials. In the purely active limit T=0T=0, some of these exponentials become delta functions, and it was found us_bound_state that for v0κ¯>12\frac{v_{0}}{\bar{\kappa}}>\frac{1}{2}

P(z)=1c22ξ2e|z|ξ2+c2δ(z)\displaystyle P(z)=\frac{1-c_{2}}{2\xi_{2}}e^{-\frac{|z|}{\xi_{2}}}+c_{2}\delta(z) (2)
ξ2=4v02κ¯28γκ¯,c2=κ¯24v02+κ¯2,\displaystyle\xi_{2}=\frac{4v_{0}^{2}-\bar{\kappa}^{2}}{8\gamma\bar{\kappa}}~,~c_{2}=\frac{\bar{\kappa}^{2}}{4v_{0}^{2}+\bar{\kappa}^{2}}\;, (3)

while for v0κ¯<12\frac{v_{0}}{\bar{\kappa}}<\frac{1}{2} one has P(z)=δ(z)P(z)=\delta(z), i.e all particles form a single cluster (c2=1c_{2}=1). Indeed since particles at the same position do not interact (sgn(0)=0{\rm sgn}(0)=0), clusters tend to form when the interaction is strong enough. Obtaining the full stationary distribution for any number N>2N>2 of particles would be very interesting, but is quite challenging PLDGS . Here we study the time dependent density fields ρσ(x,t)\rho_{\sigma}(x,t) with σ=±1\sigma=\pm 1 and their even and odd components ρs\rho_{s} and ρd\rho_{d}, defined as

ρσ(x,t)=1Niδ(xi(t)x)δσi(t),σ\displaystyle\rho_{\sigma}(x,t)=\frac{1}{N}\sum_{i}\delta(x_{i}(t)-x)\delta_{\sigma_{i}(t),\sigma} (4)
ρs/d(x,t)=ρ+(x,t)±ρ(x,t).\displaystyle\rho_{s/d}(x,t)=\rho_{+}(x,t)\pm\rho_{-}(x,t)\,. (5)

An outstanding question is whether the delta peak in the density survives when NN increases. We first investigate this question numerically for finite NN. Next, using the Dean-Kawasaki approach Dean ; Kawa , as in PLDRankedDiffusion for the passive case, and in TouzoDBM2023 for the active Dyson Brownian motion, we show that the density fields evolve according to two coupled stochastic non-linear equations. In the large NN limit these equations become deterministic and of the Burgers type. From them we obtain the exact solution for the NN particle stationary bound state. We also extend our results in the presence of an external linear potential both in the attractive and repulsive case. This leads to a rich phase diagram at T=0T=0, with phases where the density is either smooth or contains shocks (i.e clusters) at various positions, as well as phases describing an active expanding crystal.

Refer to caption
Refer to caption
Figure 1: Left: Rank field r(x)=x𝑑yρs(y)12r(x)=\int_{-\infty}^{x}dy\,\rho_{s}(y)-\frac{1}{2} in the stationary state, obtained through simulations, as a function of xx, in the attractive case with v0=0.7v_{0}=0.7, κ¯=1\bar{\kappa}=1 and γ=1\gamma=1, for small values of NN. The analytical prediction from (2) for N=2N=2 is also shown. The density has a delta peak at x=0x=0, i.e. a jump in r(x)r(x), with a weight cNc_{N}, which decreases exponentially with NN (see inset). Right: Weight cNc_{N} as a function of v0/κ¯v_{0}/\bar{\kappa} for different values of NN. The black line corresponds to the limit N+N\to+\infty (see below).

We have performed numerical simulations of Eq. (1) at T=0T=0 for various numbers of particles NN. In the attractive case with V(x)=0V(x)=0 we have measured the density of particles in the stationary state by performing time averages over a long time window. Here the density is measured in the reference frame of the instantaneous center of mass (this distinction becomes irrelevant at large NN). The results are shown in Fig. 1. One finds that the delta function at x=0x=0 in ρs(x)\rho_{s}(x) persists for any finite NN, with an amplitude cNc_{N} which depends on v0κ¯\frac{v_{0}}{\bar{\kappa}} and on NN (which is the only dimensionless parameter in that case). The first result is that for v0κ¯<1/2\frac{v_{0}}{\bar{\kappa}}<1/2 all particles are in a single cluster at x=0x=0, i.e. cN=1c_{N}=1 for any NN. Indeed, one can show that this configuration is stable. Denote x+x_{+} (resp. xx_{-}) the position of one of the rightmost (resp. leftmost) particles and n+n_{+} (resp. nn_{-}) the number of particles at the same location. One has

d(x+x)dt2v0κ¯(2n++nN)2v0κ¯\frac{d(x_{+}-x_{-})}{dt}\leq 2v_{0}-\bar{\kappa}\left(2-\frac{n_{+}+n_{-}}{N}\right)\leq 2v_{0}-\bar{\kappa} (6)

since n++nNn_{+}+n_{-}\leq N. Hence for v0κ¯<1/2\frac{v_{0}}{\bar{\kappa}}<1/2 the width of the support always decreases to 0. For v0κ¯>1/2\frac{v_{0}}{\bar{\kappa}}>1/2 we find that the amplitude cNc_{N} decreases with v0κ¯\frac{v_{0}}{\bar{\kappa}} and decays exponentially in NN.

A remarkable feature, as we will see below, is that in the limit N+N\to+\infty the delta peak at x=0x=0 is absent for v0κ¯>1/2\frac{v_{0}}{\bar{\kappa}}>1/2, but for v0<κ¯v_{0}<\bar{\kappa} it is replaced by a pair of peaks at positions ±xe\pm x_{e}, which mark the edge of the (finite) support. A genuine phase transition occurs at v0=κ¯v_{0}=\bar{\kappa}, as indicated in Fig. 2, beyond which the stationary density is smooth and the support extends to infinity. We now explain how these results are obtained.

Using the Dean-Kawasaki approach Dean ; Kawa , one can establish starting from (1) , as in PLDRankedDiffusion and TouzoDBM2023 , an exact stochastic evolution equation for the density fields, which takes the following form

tρσ=Tx2ρσ+xρσ(v0σ+V(x)\displaystyle\partial_{t}\rho_{\sigma}=T\partial_{x}^{2}\rho_{\sigma}+\partial_{x}\rho_{\sigma}\big{(}-v_{0}\sigma+V^{\prime}(x)
+κ¯dyρs(y,t)sgn(xy))+γρσγρσ+O(1N)\displaystyle+\bar{\kappa}\int dy\,\rho_{s}(y,t){\rm sgn}(x-y)\big{)}+\gamma\rho_{-\sigma}-\gamma\rho_{\sigma}+O(\frac{1}{\sqrt{N}})

where the O(1N)O(\frac{1}{\sqrt{N}}) term represents the passive and active noises, see SM , and Sec. III A of TouzoDBM2023 . It is convenient to define, as in PLDRankedDiffusion , the rank fields

r(x,t)=x𝑑yρs(y,t)12,s(x,t)=x𝑑yρd(y,t).r(x,t)=\int^{x}_{-\infty}dy\,\rho_{s}(y,t)\,-\,\frac{1}{2}\;,\;s(x,t)=\int_{-\infty}^{x}dy\,\rho_{d}(y,t)\;. (8)

Focusing from now on on the large NN limit, and henceforth neglecting the noise terms in (Non-equilibrium phase transitions in active rank diffusions), one obtains from (Non-equilibrium phase transitions in active rank diffusions) and (8) two coupled deterministic differential equations (see SM )

tr\displaystyle\!\!\!\!\!\!\partial_{t}r =\displaystyle= Tx2rv0xs+2κ¯rxr+V(x)xr,\displaystyle T\partial_{x}^{2}r-v_{0}\partial_{x}s+2\bar{\kappa}r\partial_{x}r+V^{\prime}(x)\partial_{x}r\;, (9)
ts\displaystyle\!\!\!\!\!\!\partial_{t}s =\displaystyle= Tx2sv0xr+2κ¯rxs+V(x)xs2γs.\displaystyle T\partial_{x}^{2}s-v_{0}\partial_{x}r+2\bar{\kappa}r\partial_{x}s+V^{\prime}(x)\partial_{x}s-2\gamma s\;. (10)

These equations are valid both for the attractive and the repulsive case. Since ρs(x,t)\rho_{s}(x,t) is positive and normalized to 1, r(x,t)r(x,t) must be an increasing function with r(±,t)=±12r(\pm\infty,t)=\pm\frac{1}{2}. One sees that ts(+,t)=2γs(+,t)\partial_{t}s(+\infty,t)=-2\gamma s(+\infty,t), hence at large time s(±,t)=0s(\pm\infty,t)=0. In the passive case v0=0v_{0}=0 the first equation recovers Burger’s equation which describes usual rank diffusion PLDRankedDiffusion . One generally expects that in the diffusive limit v0,γ+v_{0},\gamma\to+\infty with Ta=v022γT_{a}=\frac{v_{0}^{2}}{2\gamma} fixed, RTP’s behave as Brownian particles. This also holds here. Indeed, in (10) only two terms remain relevant, and one obtains sv02γxrs\simeq-\frac{v_{0}}{2\gamma}\partial_{x}r. Inserting in (9) the active term v0xs(x,t)-v_{0}\partial_{x}s(x,t) becomes Tax2r(x,t)T_{a}\partial_{x}^{2}r(x,t), i.e. a thermal term.

We now turn to the analysis of the large time limit of these equations, starting with the case V(x)=0V(x)=0.

Refer to caption
Figure 2: Phase diagram of active ranked diffusion in the attractive case (in the absence of external potential). For each phase (as well as in the marginal case v0=κ¯v_{0}=\bar{\kappa}), the density ρs\rho_{s} in the N+N\to+\infty limit is represented, for some values of the parameters (the arrows represent delta functions).

Attractive case. We set here V(x)=0V(x)=0. A stationary density then exists only in the attractive case κ¯>0\bar{\kappa}>0, which we now consider. To find this stationary density, we set the time derivatives to zero in (9), (10). The resulting equations are invariant by translation, hence we choose coordinates such that the center of mass x¯=𝑑xxr(x)=0\bar{x}=\int dx\,x\,r^{\prime}(x)=0. The first equation can be integrated leading to

Tr(x)v0s(x)+κ¯(r(x)214)=0.Tr^{\prime}(x)-v_{0}s(x)+\bar{\kappa}\big{(}r(x)^{2}-\frac{1}{4}\big{)}=0\;. (11)
Refer to caption
Refer to caption
Figure 3: Left: The (odd) function f(r)f(r) in (17) for r0r\geq 0 and for 3 different values of v0κ¯\frac{v_{0}}{\bar{\kappa}}. For v0κ¯>1\frac{v_{0}}{\bar{\kappa}}>1 it diverges at r=12r=\frac{1}{2}. For v0κ¯<1\frac{v_{0}}{\bar{\kappa}}<1 a maximum appears at r=r<12r=r^{*}<\frac{1}{2}. Right: Density ρs(x)\rho_{s}(x) for v0=0.7v_{0}=0.7, κ¯=1\bar{\kappa}=1 and γ=1\gamma=1: as NN increases it takes a bimodal shape, with two smooth symmetric peaks which, for N=+N=+\infty, become delta peaks (shocks) at the two edges (shown as dotted vertical lines). The histogram shows schematically the finite NN delta peak at x=0x=0. The dashed black line shows the analytical prediction (2) for N=2N=2. The plot of ρs\rho_{s} for N+N\to+\infty is obtained using the parametric representation (35) for a=0a=0.

In the absence of activity, v0=0v_{0}=0, the solution is that of standard rank diffusion, PLDRankedDiffusion

r(x)=12tanh(κ¯x2T),ρs(x)=κ¯4Tcosh(κ¯x2T)2.r(x)=\frac{1}{2}\tanh\left(\frac{\ \bar{\kappa}x}{2T}\right)\quad,\quad\rho_{s}(x)=\frac{\bar{\kappa}}{4T\cosh(\frac{\ \bar{\kappa}x}{2T})^{2}}\;. (12)

which for T0T\to 0 becomes a shock solution of Burgers equation, r(x)=12sgn(x)r(x)=\frac{1}{2}{\rm sgn}(x), i.e a single delta peak in the density ρs(x)=r(x)=δ(x)\rho_{s}(x)=r^{\prime}(x)=\delta(x). At T>0T>0 this shock solution acquires a finite width x=O(T)x=O(T). One expects that for v0>0v_{0}>0 a similar rounding occurs at low TT, not studied here. Instead, we focus directly on T=0T=0 and v0>0v_{0}>0, i.e. pure telegraphic noise. In that case one must solve

v0s(x)=2κ¯r(x)r(x),\displaystyle v_{0}s^{\prime}(x)=2\bar{\kappa}\,r(x)r^{\prime}(x)\;, (13)
v0r(x)=2κ¯r(x)s(x)2γs(x),\displaystyle v_{0}r^{\prime}(x)=2\bar{\kappa}r(x)s^{\prime}(x)-2\gamma s(x)\;, (14)

keeping in mind the possibility that r(x)r(x) may have jumps (see below). The first equation (13) can be integrated, leading to

s(x)=κ¯v0(r(x)214).s(x)=\frac{\bar{\kappa}}{v_{0}}\big{(}r(x)^{2}-\frac{1}{4}\big{)}\;. (15)

Substituting in the second equation one obtains

(v024κ¯2r(x)2)r(x)=2γκ¯(14r(x)2).\big{(}v_{0}^{2}-4\bar{\kappa}^{2}r(x)^{2}\big{)}r^{\prime}(x)=2\gamma\bar{\kappa}\big{(}\frac{1}{4}-r(x)^{2}\big{)}\;. (16)

Since |r(x)|1/2|r(x)|\leq 1/2 for all xx, with r(±)=±1/2r(\pm\infty)=\pm 1/2, the r.h.s. is positive and bounded. Let us first consider v0>κ¯v_{0}>\bar{\kappa}, in which case r(x)r^{\prime}(x) is bounded from (16). Hence there cannot be a shock and the stationary density is smooth, with an infinite support, and r(x)r(x) is obtained by inversion of the equation

γxκ¯=f(r(x)),f(r)=2r+(v02κ¯21)arctanh(2r),\frac{\gamma x}{\bar{\kappa}}=f(r(x))\ ,\ f(r)=2r+\big{(}\frac{v_{0}^{2}}{\bar{\kappa}^{2}}-1\big{)}{\rm arctanh}(2\,r)\;, (17)

where we used again the condition x¯=0\bar{x}=0 to fix the integration constant footnoteparity . One can check that for v0>κ¯v_{0}>\bar{\kappa}, the function f(r)f(r) is indeed invertible, see Fig. 3. The total density ρs(x)=r(x)\rho_{s}(x)=r^{\prime}(x) is even in xx, while ρd(x)=s(x)\rho_{d}(x)=s^{\prime}(x) is odd in xx, and both decay for |x|+|x|\to+\infty as

ρs(x)Ase|x|xeξ,ρd(x)Adsgn(x)e|x|xeξ\displaystyle\rho_{s}(x)\simeq A_{s}\,e^{-\frac{|x|-x_{e}^{*}}{\xi_{\infty}}}~,~\rho_{d}(x)\simeq A_{d}\,{\rm sgn}(x)e^{-\frac{|x|-x_{e}^{*}}{\xi_{\infty}}} (18)
ξ=v02κ¯22γκ¯,xe=κ¯γ,As=1ξ,Ad=κ¯v0As.\displaystyle\xi_{\infty}=\frac{v_{0}^{2}-\bar{\kappa}^{2}}{2\gamma\bar{\kappa}}~,~x_{e}^{*}=\frac{\bar{\kappa}}{\gamma}~,~A_{s}=\frac{1}{\xi_{\infty}}~,~A_{d}=\frac{\bar{\kappa}}{v_{0}}A_{s}\;.

Upon decreasing v0κ¯\frac{v_{0}}{\bar{\kappa}} one sees from (18) that a transition must occur at v0κ¯=1\frac{v_{0}}{\bar{\kappa}}=1. Indeed at this point f(r)=2rf(r)=2r, hence for v0κ¯=1\frac{v_{0}}{\bar{\kappa}}=1 the above solution (17) becomes

r(x)=12xxe,s(x)=14(1(xxe)2),|x|xer(x)=\frac{1}{2}\frac{x}{x_{e}^{*}}~,~s(x)=-\frac{1}{4}\big{(}1-(\frac{x}{x^{*}_{e}})^{2}\big{)}~,~|x|\leq x_{e}^{*} (19)

and r(x)=12sgn(x)r(x)=\frac{1}{2}{\rm sgn}(x) and s(x)=0s(x)=0 for |x|xe|x|\geq x_{e}. Hence the densities are non-zero in a finite interval [xe,xe][-x_{e},x_{e}], with

ρs(x)=γ2κ¯,ρd(x)=γ2x2κ¯2,|x|xe=κ¯γ\rho_{s}(x)=\frac{\gamma}{2\bar{\kappa}}~,~\rho_{d}(x)=\frac{\gamma^{2}x}{2\bar{\kappa}^{2}}~,~|x|\leq x_{e}^{*}=\frac{\bar{\kappa}}{\gamma} (20)

and vanish for |x|>xe|x|>x_{e}^{*}, with step discontinuities at the two edges.

Refer to caption
Refer to caption
Figure 4: Rank field r(x)r(x) in the stationary state, obtained through simulations, as a function of xx, in the attractive case with κ¯=1\bar{\kappa}=1 and γ=1\gamma=1, for increasing values of NN. The dashed black curves show the analytical prediction for N+N\to+\infty. The insets show s(x)s(x) for the same set of parameters. Left: v0=2v_{0}=2. The density is smooth and has infinite support. Right: v0=0.7v_{0}=0.7. The density has finite support and exhibits δ\delta peaks at the edges (i.e jumps in r(x)r(x) and s(x)s(x)).

For v0<κ¯v_{0}<\bar{\kappa}, f(r)f(r) becomes non-invertible, as can be seen from Fig. 3. In this case the density has a finite support with edges at ±xe\pm x_{e}. More precisely, f(r)f(r) is increasing only on an interval [r,r][-r^{*},r^{*}] with r=v02κ¯<12r^{*}=\frac{v_{0}}{2\bar{\kappa}}<\frac{1}{2}. Since r(x)r(x) should be increasing, one necessarily has xex=κ¯γf(r)x_{e}\leq x^{*}=\frac{\bar{\kappa}}{\gamma}f(r^{*}) and r(x)r(x) must have shocks at ±xe\pm x_{e}. Since the r.h.s. in (16) is bounded, r(x)r^{\prime}(x) can only diverge at a point where the prefactor in the l.h.s of (16) vanishes. Naively, this would lead to r(xe)=rr(x_{e})=r^{*}, i.e. xe=xx_{e}=x^{*}. However, at the position of a shock, the interaction term should be interpreted with care, as for the standard Burgers equation. Due to the convention sgn(0)=0{\rm sgn}(0)=0, the force exerted on a cluster of particles at position xex_{e} is κ¯(12r(xe+))κ¯(r(xe)+12)=κ¯(r(xe+)+r(xe))\bar{\kappa}(\frac{1}{2}-r(x_{e}^{+}))-\bar{\kappa}(r(x_{e}^{-})+\frac{1}{2})=-\bar{\kappa}(r(x_{e}^{+})+r(x_{e}^{-})). Integrating equations (13-14) from xex_{e}^{-} to xe+x_{e}^{+} then gives SM

Δr=κ¯v0(r(xe+)+r(xe))Δs,Δs=κ¯v0(r(xe+)+r(xe))Δr\Delta r=\frac{\bar{\kappa}}{v_{0}}(r(x_{e}^{+})+r(x_{e}^{-}))\Delta s~,~\Delta s=\frac{\bar{\kappa}}{v_{0}}(r(x_{e}^{+})+r(x_{e}^{-}))\Delta r

with Δr=r(xe+)r(xe)\Delta r=r(x_{e}^{+})-r(x_{e}^{-}) and similarly for Δs\Delta s. A non-zero Δr\Delta r then implies r(xe+)+r(xe)=v0/κ¯r(x_{e}^{+})+r(x_{e}^{-})=v_{0}/\bar{\kappa} leading to

r(xe)=v0κ¯12.r(x_{e}^{-})=\frac{v_{0}}{\bar{\kappa}}-\frac{1}{2}\;. (21)

since r(xe+)=1/2r(x_{e}^{+})=1/2. For |x|<xe|x|<x_{e}, the prefactor in the l.h.s of (16) is strictly positive, and thus r(x)r(x) is still given by (17), while for |x|>xe|x|>x_{e} one has r(x)=12r(x)=\frac{1}{2}. This means that ρs(x)\rho_{s}(x) has delta peaks at ±xe\pm x_{e} each containing a fraction 1v0κ¯1-\frac{v_{0}}{\bar{\kappa}} of particles. s(x)s(x) is still given by (15) and thus it also has jumps of amplitude 1v0κ¯1-\frac{v_{0}}{\bar{\kappa}} at ±xe\pm x_{e}. This implies that all the particles inside the cluster at ±xe\pm x_{e} have σ=±1\sigma=\pm 1 respectively. Note that as v0κ12\frac{v_{0}}{\kappa}\to\frac{1}{2}, xe0x_{e}\to 0 leading to a unique delta peak in ρs(x)\rho_{s}(x) at x=0x=0, consistent with the stability argument given above.

The existence of these edge clusters can be understood as follows. If there were no clusters, the rightmost particle would be subject to a force κ¯(11N)\bar{\kappa}(1-\frac{1}{N}) towards the left. Thus when κ¯>v0\bar{\kappa}>v_{0}, for large enough NN it will always move towards the left even if it has an intrinsic velocity +v0+v_{0}. It will then aggregate with other ++ particles, until the resulting cluster reaches a fraction nc/N:=12r(xe)n_{c}/N:=\frac{1}{2}-r(x_{e}^{-}) of the total number of particles large enough to be at dynamical equilibrium, i.e. v0=κ¯(1nc/N)v_{0}=\bar{\kappa}(1-n_{c}/N), from which we recover (21). For finite NN however, the size of this cluster will fluctuate, hence its position as well, and therefore we will observe a finite peak in the density instead of a delta function.

These predictions for r(x)r(x) and s(x)s(x) are compared with numerical simulations in Fig. 4. For v0>κ¯v_{0}>\bar{\kappa}, the agreement is good even at very small values of NN (N10N\sim 10). For κ¯2<v0<κ¯\frac{\bar{\kappa}}{2}<v_{0}<\bar{\kappa}, a distinct phase is predicted for N=+N=+\infty. The numerics for finite NN clearly shows precursor signatures of this phase: as one can see on Fig. 3 a bimodal form of the density is already visible for small values of N>2N>2. Furthermore the agreement on Fig. 4 at large NN for r(x)r(x) and s(x)s(x) is also very good. The behavior near the edges and analytical expression for the moments of the particle positions are obtained in SM .

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 5: Top Left: Density ρs(x,t)\rho_{s}(x,t), obtained through simulations, plotted at different times as a function of the scaled position x/tx/t, for the repulsive gas with v0=1v_{0}=1, κ=1\kappa=1 and γ=1\gamma=1, for N=1000N=1000 particles. At t=0t=0 all particles are at x=0x=0. The dashed black curve shows the analytical prediction for t+t\to+\infty. Top Right: Same plot for ρd(x,t)\rho_{d}(x,t). The inset shows the integral of the absolute value as a function of time. It is in good agreement with a 1/t1/t decay (red line). Bottom left: Plot of ρd(x,t)\rho_{d}(x,t) multiplied by t2t^{2} to compensate for the 1/t1/t decay. The dashed black line shows the analytical prediction for the boundary layer (Non-equilibrium phase transitions in active rank diffusions) for t=100t=100. Bottom right: Density ρs(x,t)\rho_{s}(x,t) near the edge of the plateau, plotted at different times with the boundary layer scaling (Non-equilibrium phase transitions in active rank diffusions). At large times the scaled density converges to ρ^s(z)\hat{\rho}_{s}(z).

Repulsive case. We now consider the repulsive case κ=κ¯>0\kappa=-\bar{\kappa}>0, still for V(x)=0V(x)=0,

tr=Tx2rv0xs2κrxr,\displaystyle\partial_{t}r=T\partial_{x}^{2}r-v_{0}\partial_{x}s-2\kappa\,r\partial_{x}r\;, (22)
ts=Tx2sv0xr2κrxs2γs.\displaystyle\partial_{t}s=T\partial_{x}^{2}s-v_{0}\partial_{x}r-2\kappa r\partial_{x}s-2\gamma s\;. (23)

Since there is no confining potential to balance the repulsion between particles, all the particles escape at infinity and there is no stationary state. As in the case of passive particles, studied in PLDRankedDiffusion ; FlackRD , one expects that the size of the cloud grows linearly with time. Hence we will look for a large time solution of (22-23), as a scaling function of y=x/ty=x/t in the form

r(x,t)=r0(y)+r1(y)t+,s(x,t)=s1(y)t+,y=xtr(x,t)=r_{0}(y)+\frac{r_{1}(y)}{t}+\dots\ ,\ s(x,t)=\frac{s_{1}(y)}{t}+\dots\ ,\ y=\frac{x}{t} (24)

At leading order 0 one has

yr0(y)=2κr0(y)r0(y).yr_{0}^{\prime}(y)=2\kappa\,r_{0}(y)r_{0}^{\prime}(y)\;. (25)

The solution is either r0(y)=0r_{0}^{\prime}(y)=0 or r0(y)=y/(2k)r_{0}(y)=y/(2k) i.e. one has

r0(y)={12fory<κy2κforκ<y<κ12fory>κr_{0}(y)=\begin{cases}-\frac{1}{2}\quad{\rm for}\ y<-\kappa\\ \frac{y}{2\kappa}\quad{\rm for}\ -\kappa<y<\kappa\\ \frac{1}{2}\quad{\rm for}\ y>\kappa\end{cases} (26)

which leads to an expanding plateau

ρs(x,t)=xr(x,t)12κtθ(κt|x|)\rho_{s}(x,t)=\partial_{x}r(x,t)\simeq\frac{1}{2\kappa t}\,\theta(\kappa t-|x|) (27)

where θ(x)\theta(x) is the Heaviside distribution. Hence the scaled density ρs\rho_{s} takes the form of an expanding plateau, exactly as for passive particles PLDRankedDiffusion ; FlackRD . This is not too surprising since at large time the particles are far from each other hence the active noise averages out to diffusive noise. There are however non trivial fluctuations with slow decay at large time. Indeed, the second equation gives

s1(y)=v02γr0(y)=v04κγθ(κ|y|)s_{1}(y)=-\frac{v_{0}}{2\gamma}r_{0}^{\prime}(y)=-\frac{v_{0}}{4\kappa\gamma}\,\theta(\kappa-|y|) (28)

hence one finds

s(x,t)v04κγtθ(κt|x|)=v02γxr(x,t),\displaystyle s(x,t)\simeq-\frac{v_{0}}{4\kappa\gamma t}\,\theta(\kappa t-|x|)=-\frac{v_{0}}{2\gamma}\partial_{x}r(x,t)\;, (29)
ρd(x,t)v04κγt(δ(x+κt)δ(xκt)).\displaystyle\rho_{d}(x,t)\simeq-\frac{v_{0}}{4\kappa\gamma t}(\delta(x+\kappa t)-\delta(x-\kappa t))\;. (30)

Thus ρd=0\rho_{d}=0 inside the plateau, i.e. ρ+=ρ\rho_{+}=\rho_{-} (to leading order in 1/t1/t) but there is an excess of - particles on the left edge and ++ on the right edge with total weight decaying as 1/t1/t.

It is important to note that the 1/t1/t expansion (24) is valid inside the plateau footnoter1 but fails at the edges. A more careful analysis in a region of width t\sim\sqrt{t} near the edges shows SM that the solution takes a boundary layer form (at the right edge)

ρs(x,t)=1κtρ^s(z),ρd(x,t)=v02γκTefft3/2ρ^s(z)\displaystyle\rho_{s}(x,t)=\frac{1}{\kappa t}\hat{\rho}_{s}(z)\ ,\ \rho_{d}(x,t)=-\frac{v_{0}}{2\gamma\kappa\sqrt{T_{\rm eff}}\,t^{3/2}}\hat{\rho}_{s}^{\prime}(z)
ρ^s(z)=ez22(2+πez24zerfc(z2))2πerfc(z2)2,z=xκtTefft\displaystyle\hat{\rho}_{s}(z)=\frac{e^{-\frac{z^{2}}{2}}(2+\sqrt{\pi}e^{\frac{z^{2}}{4}}z\,{\rm erfc}(-\frac{z}{2}))}{2\pi\,{\rm erfc}(-\frac{z}{2})^{2}}\ ,\ z=\frac{x-\kappa t}{\sqrt{T_{\rm eff}t}} (31)

with Teff=T+Ta=T+v022γT_{\rm eff}=T+T_{a}=T+\frac{v_{0}^{2}}{2\gamma}. The boundary layer scaling function ρ^s(y)\hat{\rho}_{s}(y) is identical to the one obtained for the passive problem FlackRD . This shows that for the expanding repulsive gas the role of the active noise is not different from that of passive noise at temperature Ta=v022γT_{a}=\frac{v_{0}^{2}}{2\gamma}. Accordingly, note that the relation ρdv02γxρs\rho_{d}\simeq-\frac{v_{0}}{2\gamma}\partial_{x}\rho_{s} holds at large time both in the boundary layer and in the plateau (which we have checked numerically). These results show that the delta function in (30) has a finite width at finite time, i.e contrary to the attractive case there are no clusters. In fact the ±\pm population imbalance 1/t\sim 1/t is subdominant compared to the total population in the boundary layer 1/t\sim 1/\sqrt{t}.

We have checked these predictions numerically. Figure 5 shows the convergence of the density ρs\rho_{s} to (27) as time increases. The signature of the active noise is visible at short times: for t<1/γt<1/\gamma, the particles split into two packets with a spread in velocities xt[v0,v0+κ]\frac{x}{t}\in[v_{0},v_{0}+\kappa] (and same with minus), leading to two distinct blocks in the density, which disappear at larger time as the active noise averages out. One also sees in Figure 5 peaks forming the scaled t2ρd(x,t)t^{2}\rho_{d}(x,t) as well as a numerical check of the boundary layer forms (Non-equilibrium phase transitions in active rank diffusions).

Linear potential. We now consider the case where the particles are submitted to an attractive linear potential V(x)=a|x|V(x)=a|x|, a>0a>0. For non-interacting particles κ¯=0\bar{\kappa}=0 there is a steady state with two phases, either a<v0a<v_{0} and

ρs(x)=γav02a2exp(2γav02a2|x|)\rho_{s}(x)=\frac{\gamma a}{v_{0}^{2}-a^{2}}\,\exp\left(-\frac{2\gamma a}{v_{0}^{2}-a^{2}}|x|\right) (32)

or a>v0a>v_{0} and ρs(x)=δ(x)\rho_{s}(x)=\delta(x) DKM19 .

In presence of interactions, the large time behavior for large NN is richer and represented on the phase diagram in Fig. 6. There are six distinct phases which we briefly describe, four of them being stationary, while the other two are expanding (see SM for details and some plots of the densities).

Refer to caption
Figure 6: Phase diagram of active ranked diffusion with a linear external potential V(x)=a|x|V(x)=a|x|. The phases are named as follows: the large letter is either EE for an expanding phase, II for a stationary phase with infinite support, FF if the support is finite support, or δ\delta for a clustered phase. The subscript indicates the position of the shocks: 0 if there is a shock at x=0x=0, ee if there are shocks at the edges and ss if the density is smooth. The dotted lines denote a finite NN transition.

δ0\delta_{0} phase. It exists both in the attractive case, for a>v0κ¯2a>v_{0}-\frac{\bar{\kappa}}{2}, and in the repulsive case, a>v0+κa>v_{0}+\kappa. The density is formed of a single cluster at x=0x=0, ρs(x)=δ(x)\rho_{s}(x)=\delta(x). Note that outside the dotted lines in Fig. 6 it already exists for finite NN, while inside the density has a finite width for finite NN.

IsI_{s} phase, both in the attractive case, for a<v0κ¯a<v_{0}-\bar{\kappa}, and in the repulsive case, for κ<a<v0\kappa<a<v_{0}. This phase is quite similar to the one for a=0a=0. The function r=r(x)r=r(x) is odd and is given for x0x\geq 0 as the solution of

γx=𝖿a(r)=2κ¯r+v02(κ¯+a)22(κ¯+a)log(κ¯+2a+2κ¯r(κ¯+2a)(12r))\gamma x={\sf f}_{a}(r)=2\bar{\kappa}r+\frac{v_{0}^{2}-(\bar{\kappa}+a)^{2}}{2(\bar{\kappa}+a)}\log\left(\frac{\bar{\kappa}+2a+2\bar{\kappa}r}{(\bar{\kappa}+2a)(1-2r)}\right) (33)

an extension of (17) above. From r(x)r(x) one obtains for any xx\in\mathbb{R}

s(x)=κ¯v0(r(x)214)+av0(|r(x)|12).s(x)=\frac{\bar{\kappa}}{v_{0}}\big{(}r(x)^{2}-\frac{1}{4}\big{)}+\frac{a}{v_{0}}\big{(}|r(x)|-\frac{1}{2}\big{)}\;. (34)

The densities are given in parametric form for x0x\geq 0 and r[0,1/2]r\in[0,1/2] as

x=𝖿a(r)γ,ρs=γ𝖿a(r),ρd=(a+2κ¯r)γv0𝖿a(r).x=\frac{{\sf f}_{a}(r)}{\gamma}~,~\rho_{s}=\frac{\gamma}{{\sf f}^{\prime}_{a}(r)}~,~\rho_{d}=\frac{(a+2\bar{\kappa}r)\gamma}{v_{0}{\sf f}^{\prime}_{a}(r)}\;. (35)

They have infinite support and are smooth except at x=0x=0 where ρd(x)\rho_{d}(x) has a jump and ρs(x)\rho_{s}(x) has a linear cusp.

FeF_{e} phase, in the attractive case, for v0κ¯<a<v0κ¯2v_{0}-\bar{\kappa}<a<v_{0}-\frac{\bar{\kappa}}{2}. Here 𝖿a(r){\sf f}_{a}(r) becomes non-invertible. The densities have a finite support [xe,xe][-x_{e},x_{e}] with shocks at the two edges ±xe\pm x_{e} such that

r(xe)=v0aκ¯12,r(xe+)=12r(x_{e}^{-})=\frac{v_{0}-a}{\bar{\kappa}}-\frac{1}{2}\quad,\quad r(x_{e}^{+})=\frac{1}{2} (36)

with γxe=𝖿a(r(xe))\gamma x_{e}={\sf f}_{a}(r(x_{e}^{-})). It corresponds to clusters with a fraction κ¯+av0κ¯\frac{\bar{\kappa}+a-v_{0}}{\bar{\kappa}} of the particles (all being ++ at x=xex=x_{e} and - at x=xex=-x_{e}). Inside the support, r(x)r(x) is again given by (33), s(x)s(x) by (34) and the densities by (35), with the same singular behavior at x=0x=0.

I0I_{0} phase, in the repulsive case, for κ<a<v0+κ\kappa<a<v_{0}+\kappa and a>v0a>v_{0}. This phase is similar to IsI_{s} but has a shock at x=0x=0 corresponding to a cluster with equal fractions of ±\pm particles and of total weight av0κ\frac{a-v_{0}}{\kappa}. The function r(x)r(x) for x>0x>0 is now given by

γx=𝖿a(r)𝖿a(r(0+)),r(0+)=av02κ.\gamma x={\sf f}_{a}(r)-{\sf f}_{a}(r(0^{+}))~,~r(0^{+})=\frac{a-v_{0}}{2\kappa}\;. (37)

From (37), (33) and (34) (which still holds for x0x\neq 0) we obtain that ρs(x)\rho_{s}(x) and ρd(x)\rho_{d}(x) exhibit near x=0x=0 an inverse square root divergence

ρs(x)av0κδ(x)+B2|x|,ρd(x)B2|x|sgn(x)\rho_{s}(x)\simeq\frac{a-v_{0}}{\kappa}\delta(x)+\frac{B}{2\sqrt{|x|}}\ ,\ \rho_{d}(x)\simeq\frac{B}{2\sqrt{|x|}}\,{\rm sgn}(x) (38)

with B=12κγv0(v02(aκ)2)B=\frac{1}{2\kappa}\sqrt{\frac{\gamma}{v_{0}}(v_{0}^{2}-(a-\kappa)^{2})}, which involves only ++ particles for x>0x>0 and - particles for x<0x<0.

EsE_{s} and E0E_{0} phases, in the repulsive case, for a<κa<\kappa and a<v0a<v_{0} or a>v0a>v_{0} respectively. In this case a fraction 1aκ1-\frac{a}{\kappa} of particles escape to infinity, while a fraction aκ\frac{a}{\kappa} remain confined. Under the scaling |x|t|x|\sim t one finds at large time in both phases

ρs(x,t)aκδ(x)+12κtθ((κa)t|x|),\rho_{s}(x,t)\simeq\frac{a}{\kappa}\delta(x)+\frac{1}{2\kappa t}\theta((\kappa-a)t-|x|)\;, (39)

hence the expanding part behaves similarly as for a=0a=0, i.e. it is uniform within the support, which now spreads with velocities ±(κa)\pm(\kappa-a). Similarly one finds

ρd(x,t)v04γκt(δ((κa)tx)δ((κa)tx)).\rho_{d}(x,t)\simeq\frac{v_{0}}{4\gamma\kappa t}(\delta((\kappa-a)t-x)-\delta((\kappa-a)t-x))\;. (40)

Both densities are smooth on scales t\sqrt{t}, with similar boundary layers near the edges as for a=0a=0 (see above and SM ).

For the particles which remain bound within x=O(1)x=O(1), r(x)r(x) is smooth away from x=0x=0 given by, for x>0x>0

γx=𝖿~a(κar(x))𝖿~a(κar(0+)),\displaystyle\!\!\!\!\!\!\gamma x=\tilde{\sf f}_{a}\left(\frac{\kappa}{a}r(x)\right)-\tilde{\sf f}_{a}\left(\frac{\kappa}{a}r(0^{+})\right)\;, (41)
𝖿~a(r)=2ar(v02a2(12r)1),r(0+)=av02κθ(av0)\displaystyle\!\!\!\!\!\!\tilde{\sf f}_{a}(r)=2ar(\frac{v_{0}^{2}}{a^{2}(1-2r)}-1)\ ,\ r(0^{+})=\frac{a-v_{0}}{2\kappa}\theta(a-v_{0})

where in the phase E0E_{0}, i.e. a>v0a>v_{0}, there is a cluster at x=0x=0 containing a fraction av0κ\frac{a-v_{0}}{\kappa} of the total number of particles. From (41) one obtains an explicit formula for the densities, see SM . The most notable feature is that they decay as power laws at large distance, in both phases

ρs(x)v022κγx2,ρd(x)v032κγ2x3,\rho_{s}(x)\simeq\frac{v_{0}^{2}}{2\kappa\gamma x^{2}}\quad,\quad\rho_{d}(x)\simeq\frac{v_{0}^{3}}{2\kappa\gamma^{2}x^{3}}\;, (42)

while in the phases IsI_{s} and I0I_{0} the densities decay exponentially at large distance. In that sense in the phases EsE_{s} and E0E_{0} the steady state of the bound particles is always critical. Indeed, the bound particles can be effectively described by an interaction constant κeff=a\kappa_{\rm eff}=a, see SM . Finally, the non-analytic behavior of the densities near x=0x=0 is similar to the one in the phases IsI_{s} and I0I_{0}. In SM we compare these results with the case of passive particles, where only the phases IsI_{s} and EsE_{s} are present.

More general interactions. Consider now a pairwise interaction potential W(x)W(x), such that the force term in (1) is replaced by 1NjW(xixj)-\frac{1}{N}\sum_{j}W^{\prime}(x_{i}-x_{j}), Eq. (1) being recovered for W(x)=κ¯|x|W(x)=\bar{\kappa}|x|. W(x)W^{\prime}(x) being odd, we restrict to the case W(0)=0W^{\prime}(0)=0, i.e. we exclude the possibility that W(x)W^{\prime}(x) diverges at x=0x=0 for which the large NN limit is problematic TouzoDBM2023 . At large NN, from (Non-equilibrium phase transitions in active rank diffusions) with the replacement κ¯sgn(xy)W(xy)\bar{\kappa}\,{\rm sgn}(x-y)\to W^{\prime}(x-y) we obtain at stationarity

0=x[(v0ρdF~(x)ρs]\displaystyle 0=\partial_{x}[(-v_{0}\rho_{d}-\tilde{F}(x)\rho_{s}] (43)
0=x[(v0ρsF~(x)ρd]2γρd\displaystyle 0=\partial_{x}[(-v_{0}\rho_{s}-\tilde{F}(x)\rho_{d}]-2\gamma\rho_{d}

where

F~(x)=V(x)𝑑yW(xy)ρs(y).\tilde{F}(x)=-V^{\prime}(x)-\int dy\,W^{\prime}(x-y)\rho_{s}(y)\;. (44)

Eqs. (43) are identical to those of a single RTP in an effective force field F~(x)\tilde{F}(x) which itself depends on the density. They can be solved formally, leading (in the simplest case, see SM ) to

ρs(x)=Kv02F~(x)2e2γ0x𝑑zF~(z)v02F~(z)2\rho_{s}(x)=\frac{K}{v_{0}^{2}-\tilde{F}(x)^{2}}e^{2\gamma\int_{0}^{x}dz\frac{\tilde{F}(z)}{v_{0}^{2}-\tilde{F}(z)^{2}}} (45)

where KK is determined by the normalization. Eqs. (44) and (45) are self-consistent equations for the density ρs(x)\rho_{s}(x). In SM we show how it recovers the above results for the rank interaction. For a harmonic interaction the density at large NN is found to be the same as for a single particle in a harmonic well. Other examples will be studied in inprep .

Conclusion. We have obtained a description of an active version of rank diffusion, controlled at large NN. We have computed exactly the densities of RTP particles at large time in that limit. In particular, in the attractive case we found a collective dynamical phase transition in the steady state where the support of the density changes from infinite to finite, with shocks (i.e clusters of particles) appearing at the edges. In presence of a linear confining potential we find a rich phase diagram with six phases. In the expanding phases we find that a fraction of the particles remain bound, and in a critical state with power-law decay of the densities at large distance. We also obtained a self-consistent equation for the densities which opens the way to treat more general interaction potentials. Some remaining open questions are finite NN fluctuations and calculation of other observables, such as entropy production.


Acknowledgments: We thank Gregory Schehr for collaborations on related topics. We thank LPTMS (Orsay) and Collège de France for hospitality.

References

  • (1) J. Masoliver and K. Lindenberg, Continuous time persistent random walk: a review and some generalizations, Eur. Phys. J. B 90, 1 (2017).
  • (2) M. Kac, A stochastic model related to the telegrapher’s equation, Rocky Mountain J. Math. 4, 497 (1974).
  • (3) H. C. Berg, E. Coli in Motion, (Springer Verlag, Heidel- berg, Germany) (2004).
  • (4) J. Tailleur, and M. E. Cates, Statistical mechanics of interacting run-and-tumble bacteria, Phys. Rev. Lett. 100, 218103 (2008).
  • (5) M. C. Marchetti, J. F. Joanny, S. Ramaswamy, T. B. Liverpool, J. Prost, M. Rao, and R. Aditi Simha, Hydrodynamics of soft active matter, Rev. Mod. Phys. 85, 1143 (2013).
  • (6) A. B. Slowman, M. R. Evans, and R. A. Blythe, Jamming and attraction of interacting run-and-tumble random walkers, Phys. Rev. Lett. 116, 218101 (2016).
  • (7) A. B. Slowman, M. R. Evans, and R. A. Blythe, Exact solution of two interacting run-and-tumble random walkers with finite tumble duration, J. Phys. A: Math. Theor. 50, 375601 (2017).
  • (8) Y. Fily, and M. C. Marchetti, Athermal phase separation of self-propelled particles with no alignment, Phys. Rev. Lett. 108, 235702 (2012).
  • (9) M. Cates, and J. Tailleur, Motility-induced phase separation, Annu. Rev. Condens. Matter Phys. 6, 219 (2015).
  • (10) C. M. Barriuso Gutiérrez, C. Vanhille-Campos, Francisco Alarcon, I. Pagonabarraga, R. Britoai, and C. Valeriani, Collective motion of run-and-tumble repulsive and attractive particles in one-dimensional systems, Soft Matter 17(46) (2021).
  • (11) R. Soto, R. Golestanian, Run-and-tumble dynamics in a crowded environment: Persistent exclusion process for swimmers, Phys. Review E 89, 012706 (2014).
  • (12) P. Le Doussal, S. N. Majumdar, G. Schehr, Stationary nonequilibrium bound state of a pair of run and tumble particles, Physical Review E 104(4), 044103 (2021).
  • (13) P. Dolai, S. Krekels, and C. Maes, Inducing a bound state between active particles, Phys. Rev. E 105, 044605 (2022).
  • (14) I. Mukherjee, A. Raghu, and P. K. Mohanty, Nonexistence of motility induced phase separation transition in one dimension, SciPost Phys. 14, 165 (2023) .
  • (15) E. Mallmin, R. A. Blythe, and M. R. Evans, Exact spectral solution of two interacting run-and-tumble particles on a ring lattice, J. Stat. Mech., 013204 (2019).
  • (16) A. Das, A. Dhar, and A. Kundu, Gap statistics of two interacting run and tumble particles in one dimension, J. Phys. A: Math. Theor. 53, 345003 (2020).
  • (17) P. Le Doussal, S. N. Majumdar, and G. Schehr, Noncrossing run-and-tumble particles on a line, Phys. Rev. E 100, 012113 (2019).
  • (18) P. Singh, A. Kundu, Crossover behaviours exhibited by fluctuations and correlations in a chain of active particles, J. Phys. A: Math. Theor. 54, 305001 (2021)
  • (19) S. Put, J. Berx, and C. Vanderzande, Non-Gaussian anomalous dynamics in systems of interacting run-and-tumble particles, J. Stat. Mech. 123205 (2019).
  • (20) M. J. Metson, M. R. Evans, and R. A. Blythe, Tuning attraction and repulsion between active particles through persistence, EPL 141, 41001 (2023).
  • (21) M. J. Metson, M. R. Evans, and R. A. Blythe, From a microscopic solution to a continuum description of interacting active particles, Phys. Rev. E 107, 044134 (2023).
  • (22) R. Dandekar, S. Chakraborti, and R. Rajesh, Hard core run and tumble particles on a one-dimensional lattice, Phys. Rev. E 102, 062111 (2020).
  • (23) A. G. Thompson, J. Tailleur, M. E. Cates, R. A. Blythe, Lattice models of nonequilibrium bacterial dynamics, J. Stat. Mech. P02029 (2011).
  • (24) L. Touzo, P. Le Doussal, G. Schehr, Interacting, running and tumbling: the active Dyson Brownian motion, EPL 142, 61004 (2023) and arXiv:2302.02937.
  • (25) A. D. Banner, R. Fernholz, I. Karatzas, Atlas models of equity markets, Ann. Appl. Probab. 15, 2296 (2005)
  • (26) S. Pal, J. Pitman, One-dimensional Brownian particle systems with rank-dependent drifts, Ann. Appl. Probab. 18, 2179 (2008).
  • (27) G. B. Rybicki, Exact statistical mechanics of a one-dimensional self-gravitating system, Astrophys. Space Sci. 14, 56 (1971).
  • (28) P. H. Chavanis, C. Sire, Anomalous diffusion and collapse of self-gravitating Langevin particles in DD dimensions, Phys. Rev. E 69, 016116 (2004).
  • (29) A. Lenard, Exact statistical mechanics of a one-dimensional system with Coulomb forces, J. Math. Phys. 2, 682 (1961).
  • (30) R. J. Baxter, Statistical mechanics of a one-dimensional Coulomb system with a uniform charge background, Proc. Camb. Phil. Soc. 59, 779 (1963)
  • (31) G. Tellez, E. Trizac, Screening like charges in one-dimensional Coulomb systems: Exact results, Phys. Rev. E 92, 042134 (2015).
  • (32) For a recent review, see M. Lewin, Coulomb and Riesz gases: The known and the unknown, J. Math. Phys. 63, 061101 (2022).
  • (33) A. Dhar, A. Kundu, S. N. Majumdar, S. Sabhapandit, G. Schehr, Exact extremal statistics in the classical 1d Coulomb gas, Phys. Rev. Lett. 119, 060601 (2017).
  • (34) A. Flack, S. N. Majumdar, G. Schehr, Gap probability and full counting statistics in the one-dimensional one-component plasma, J. Stat. Mech. 053211 (2022).
  • (35) D. Chafai, D. Garcia-Zelada, P. Jung, At the edge of a one-dimensional jellium, Bernoulli 28, 1784 (2022).
  • (36) P. Le Doussal, Ranked diffusion, delta Bose gas and Burgers equation, Phys. Rev. E 105, L012103 (2022).
  • (37) D. S. Dean, Langevin Equation for the density of a system of interacting Langevin processes, J. Phys. A: Math. Gen. 29, L613 (1996).
  • (38) K. Kawasaki, Microscopic analyses of the dynamical density functional equation of dense fluids, J. Stat. Phys. 93, 527 (1998).
  • (39) A. Flack, P. Le Doussal, S. N. Majumdar, G. Schehr, Out of equilibrium dynamics of repulsive ranked diffusions: the expanding crystal, Phys. Rev. E 107, 064105 (2023).
  • (40) P. Le Doussal and G. Schehr, unpublished.
  • (41) See supplementary material
  • (42) Since f(r) is odd, this is the same as imposing r(0)=0r(0)=0.
  • (43) Note however that the higher order terms, e.g. r1(y)r_{1}(y), depend on the initial condition SM .
  • (44) A. Dhar, A. Kundu, S. N. Majumdar, S. Sabhapandit and G. Schehr, Run-and-tumble particle in one-dimensional confining potentials: Steady-state, relaxation, and first-passage properties, Phys. Rev. E 99, 032132 (2019).
  • (45) Solutions with 0<C<10<C<1, with a shock at x=0x=0 are in principle possible but are excluded by our numerical results.
  • (46) L. Touzo, P. Le Doussal to be published.

.

.

Supplementary Material for

Active rank diffusion


We give the principal details of the calculations described in the main text of the Letter. We display additional analytical and numerical results which support the findings presented in the Letter.


I A few definitions and properties

I.1 Center of mass and density fields for finite NN

The center of mass for any NN is defined by

x¯(t)=1Nixi(t).\bar{x}(t)=\frac{1}{N}\sum_{i}x_{i}(t)\;. (46)

Its equation of motion is given by

ddtx¯=v0Niσi(t)+2Tξi(t)\frac{d}{dt}\bar{x}=\frac{v_{0}}{N}\sum_{i}\sigma_{i}(t)+\sqrt{2T}\xi_{i}(t) (47)

i.e. it evolves as the sum of the positions of NN independent RTP’s with velocities v0/Nv_{0}/N. Since σi(t)σj(t)=δije2γ|tt|\langle\sigma_{i}(t)\sigma_{j}(t^{\prime})\rangle=\delta_{ij}e^{-2\gamma|t-t^{\prime}|}, one obtains that for γ>0\gamma>0 at large times the center of mass has a diffusive behavior x¯2DNt\bar{x}\sim\sqrt{2D_{N}t}, with

DN=1N(T+v022γ).D_{N}=\frac{1}{N}(T+\frac{v_{0}^{2}}{2\gamma})\;. (48)

At finite NN one can define two density fields

ρσ(x,t)=1Niδσ,σi(t)δ(xxi(t))\displaystyle\rho_{\sigma}(x,t)=\frac{1}{N}\sum_{i}\delta_{\sigma,\sigma_{i}(t)}\delta(x-x_{i}(t)) (49)
ρ^σ(x,t)=1Niδσ,σi(t)δ(xxi(t)x¯(t)),x¯(t)=1Njxj(t).\displaystyle\hat{\rho}_{\sigma}(x,t)=\frac{1}{N}\sum_{i}\delta_{\sigma,\sigma_{i}(t)}\delta(x-x_{i}(t)-\bar{x}(t))\quad,\quad\bar{x}(t)=\frac{1}{N}\sum_{j}x_{j}(t)\;. (50)

The first definition is the one used in the main text, for which we derive a Dean equation (valid in principle for any NN, but studied mostly for large NN). The second definition is the one we use for the numerical simulations (except when a confining potential is present, in which case we use the first definition). The difference between the two definitions becomes irrelevant in the large NN limit since DN0D_{N}\to 0 in that limit. We will return to this point below.

I.2 Results for N=2N=2

As an example let us recall the explicit result for the stationary distribution for N=2N=2 obtained in us_bound_state in the attractive case κ¯>0\bar{\kappa}>0. Here we consider the variable z=x1x22z=\frac{x_{1}-x_{2}}{2}, while there we denoted y=x1x2=2zy=x_{1}-x_{2}=2z. We also have the correspondence κ¯=2c¯\bar{\kappa}=2\bar{c} with the notations of us_bound_state ). Taking this into account, the stationary probabilities Pσ1,σ2(z)P_{\sigma_{1},\sigma_{2}}(z) for the variable zz read, for v0>κ¯/2v_{0}>\bar{\kappa}/2,

(P++(z)P+(z)P+(z)P(z))=γκ¯4v02+κ¯2e8γκ¯4v02κ¯2|z|(12v0+κ¯2v0κ¯θ(z)+2v0κ¯2v0+κ¯θ(z)2v0κ¯2v0+κ¯θ(z)+2v0+κ¯2v0κ¯θ(z)1)+12κ¯24v02+κ¯2δ(z)(1001).\displaystyle\left(\begin{array}[]{c}P_{++}(z)\\ P_{+-}(z)\\ P_{-+}(z)\\ P_{--}(z)\\ \end{array}\right)=\frac{\gamma\bar{\kappa}}{4v_{0}^{2}+\bar{\kappa}^{2}}\;e^{-\frac{8\gamma\bar{\kappa}}{4v_{0}^{2}-\bar{\kappa}^{2}}|z|}\left(\begin{array}[]{c}1\\ \frac{2v_{0}+\bar{\kappa}}{2v_{0}-\bar{\kappa}}\theta(z)+\frac{2v_{0}-\bar{\kappa}}{2v_{0}+\bar{\kappa}}\theta(-z)\\ \frac{2v_{0}-\bar{\kappa}}{2v_{0}+\bar{\kappa}}\theta(z)+\frac{2v_{0}+\bar{\kappa}}{2v_{0}-\bar{\kappa}}\theta(-z)\\ 1\\ \end{array}\right)+\frac{1}{2}\frac{\bar{\kappa}^{2}}{4v_{0}^{2}+\bar{\kappa}^{2}}\delta(z)\left(\begin{array}[]{c}1\\ 0\\ 0\\ 1\\ \end{array}\right)\;. (63)

The total probability is P(z)=σ1=±1,σ2=±1Pσ1,σ2(z)P(z)=\sum_{\sigma_{1}=\pm 1,\sigma_{2}=\pm 1}P_{\sigma_{1},\sigma_{2}}(z), which is normalized to +𝑑yP(y)=1\int_{-\infty}^{+\infty}dyP(y)=1. One can check from (63) that one recovers the expression of the P(z)P(z) given in the text in (2). Note the symmetry Pσ1,σ2(z)=Pσ2,σ1(z)P_{\sigma_{1},\sigma_{2}}(z)=P_{\sigma_{2},\sigma_{1}}(-z) and that for each state (σ1,σ2)(\sigma_{1},\sigma_{2}) one has +𝑑yPσ1,σ2(y)=14\int_{-\infty}^{+\infty}dyP_{\sigma_{1},\sigma_{2}}(y)=\frac{1}{4}, i.e. each of the four states is equiprobable in the steady state, as expected.

From this one can compute the second definition of the density, i.e. the density in the reference frame of the center of mass. It is given by

ρ^σ(x)=12δσ1,σδ(x1x¯x)+12δσ2,σδ(x2x¯x)\displaystyle\hat{\rho}_{\sigma}(x)=\frac{1}{2}\langle\delta_{\sigma_{1},\sigma}\delta(x_{1}-\bar{x}-x)\rangle+\frac{1}{2}\langle\delta_{\sigma_{2},\sigma}\delta(x_{2}-\bar{x}-x)\rangle (64)
=12δσ1,σδ(zx)+12δσ2,σδ(zx)\displaystyle=\frac{1}{2}\langle\delta_{\sigma_{1},\sigma}\delta(z-x)\rangle+\frac{1}{2}\langle\delta_{\sigma_{2},\sigma}\delta(-z-x)\rangle (65)
=12σPσ,σ(z)+12σPσ,σ(z)=σPσ,σ(z)\displaystyle=\frac{1}{2}\sum_{\sigma^{\prime}}P_{\sigma,\sigma^{\prime}}(z)+\frac{1}{2}\sum_{\sigma^{\prime}}P_{\sigma^{\prime},\sigma}(-z)=\sum_{\sigma^{\prime}}P_{\sigma,\sigma^{\prime}}(z) (66)

where we used the above mentioned symmetry. Hence we obtain

ρ^s(x)=ρ^+(x)+ρ^(x)=P(x)\displaystyle\hat{\rho}_{s}(x)=\hat{\rho}_{+}(x)+\hat{\rho}_{-}(x)=P(x) (67)

where P(x)P(x) is the total probability given in the text in (2), and

ρ^d(x)=ρ^+(x)ρ^(x)=P++(x)+P+(x)P+(x)P(x)=sgn(x)8γv0κ¯216v04κ¯4e8γκ¯4v02κ¯2|x|.\displaystyle\hat{\rho}_{d}(x)=\hat{\rho}_{+}(x)-\hat{\rho}_{-}(x)=P_{++}(x)+P_{+-}(x)-P_{-+}(x)-P_{--}(x)={\rm sgn}(x)\frac{8\gamma v_{0}\bar{\kappa}^{2}}{16\,v_{0}^{4}-\bar{\kappa}^{4}}\;e^{-\frac{8\gamma\bar{\kappa}}{4v_{0}^{2}-\bar{\kappa}^{2}}|x|}\,. (68)

Note that for v0<κ¯/2v_{0}<\bar{\kappa}/2 one finds us_bound_state that P++(z)=P(z)=12δ(z)P_{++}(z)=P_{--}(z)=\frac{1}{2}\delta(z) and P+(z)=P+(z)=0P_{+-}(z)=P_{-+}(z)=0. Hence the density ρ^s(x)=δ(x)\hat{\rho}_{s}(x)=\delta(x), ρ^d(x)=0\hat{\rho}_{d}(x)=0, and the two particles form a single cluster.

I.3 Center of mass and median point at large NN

In terms of the rank field r(x,t)r(x,t) the position of the center of mass is given for any NN as

x¯(t)=1Nixi(t)=+𝑑xxxr(x,t)=1/2+1/2𝑑rx(r,t).\bar{x}(t)=\frac{1}{N}\sum_{i}x_{i}(t)=\int_{-\infty}^{+\infty}dx\,x\partial_{x}r(x,t)=\int_{-1/2}^{+1/2}dr\,x(r,t)\;. (69)

Consider now the limit N+N\to+\infty. In that limit, the fractions p±(t)=N±(t)/Np_{\pm}(t)=N_{\pm}(t)/N of ±\pm particles follow a deterministic evolution equation, t(p+p)=2γ(p+p)\partial_{t}(p_{+}-p_{-})=-2\gamma(p_{+}-p_{-}). Hence, for γ>0\gamma>0, p+(t)p(t)p_{+}(t)-p_{-}(t) decreases exponentially to zero at large tt.

Consider now the equations (9-10) valid for N=+N=+\infty. Recall that s(+,t)=p+(t)p(t)s(+\infty,t)=p_{+}(t)-p_{-}(t). In the case where the initial condition satisfies p+(0)=p(0)p_{+}(0)=p_{-}(0), which implies s(+,t)=0s(+\infty,t)=0 for all tt, one can check from these equations that the center of mass does not move. Indeed one obtains

ddtx¯(t)=+𝑑xxx2(Txr(x,t)v0s(x,t)+κ¯(r(x,t)214))=0\frac{d}{dt}\bar{x}(t)=\int_{-\infty}^{+\infty}dx\,x\partial^{2}_{x}(T\partial_{x}r(x,t)-v_{0}s(x,t)+\bar{\kappa}(r(x,t)^{2}-\frac{1}{4}))=0 (70)

which vanishes by integration by part since the term inside the second derivative vanishes at x=±x=\pm\infty.

One can also define the median point x0(t)x_{0}(t), which for any NN odd, and for N=+N=+\infty is such that

x0(t)=r(x0(t),t)=0,x0=x(0,t).x_{0}(t)=r(x_{0}(t),t)=0\quad,\quad x_{0}=x(0,t)\;. (71)

This point however may move (depending on the initial condition) if the system has not reached the stationary state (even if p+=pp_{+}=p_{-}).

II Numerical simulation details

The simulations were performed using simple Euler dynamics. To make the computations faster, we use the fact that, as the name suggests, the total force applied on a particle due to the rank interaction only depends on the rank of the particle’s position. We thus keep in memory the updated rank of each particle if they were ordered by position (taking into account the fact that some can have the same position if a cluster forms). The positions of all particles are then updated simultaneously at each time step according to

xi(t+dt)=xi(t)+dt(κ¯N(Niright(t)Nileft(t))V(xi(t))+v0σi(t))x_{i}(t+dt)=x_{i}(t)+dt\left(\frac{\bar{\kappa}}{N}(N_{i}^{\rm right}(t)-N_{i}^{\rm left}(t))-V^{\prime}(x_{i}(t))+v_{0}\sigma_{i}(t)\right) (72)

where Niright(t)N_{i}^{\rm right}(t) (resp. Nileft(t)N_{i}^{\rm left}(t)) is the number of particles strictly at the right (resp. left) of particle ii at time tt, and the σi\sigma_{i}’s switch sign independently with probability γdt\gamma dt at each time step. Strictly speaking, one should keep track of every particles that cross and check if they form a new cluster. In practice, in the phases where no shocks are present in the density at finite NN (i.e. IsI_{s}, FeF_{e}, EsE_{s} and δ0\delta_{0} inside the dotted rectangle), one can simply let the particles cross, in which case they start to perform very small oscillations around each other (if the time step is small enough for the bin size of the histogram, this situation can not be distinguished from a real cluster). In the presence of a linear potential however, we do keep track of clusters forming at x=0x=0 (in the phases I0I_{0} and E0E_{0}, i.e. whenever a>v0a>v_{0}). More precisely when a particle should cross x=0x=0, we place it exactly at x=0x=0 and wait for the next time step to see if it keeps moving. This avoids having a large number of particles oscillating around x=0x=0, which would slow down the simulations.

For the phases where a steady state exists (or when looking at the bound particles in an expanding phase), the plots are obtained using a single run of the dynamics. We run the dynamics long enough to let the system reach stationarity, and then build the histogram of positions (after subtracting the position of the center of mass if no confining potential is present) over a large time window. In this case we use a time step dt=0.001dt=0.001 and run the dynamics for 10710^{7} to 10910^{9} steps, depending on the value of NN. To look at the escaping particles in the expanding phases, we instead run the dynamics 10510^{5} times and build a distinct histogram for each time step. In this case we use dt=0.01dt=0.01.

III More details about the particle clusters

Let us first recall what the phase diagram in Fig. 6 looks like at finite NN. In the repulsive case it is not qualitatively different from the infinite NN case. In particular, the shock at x=0x=0 for a>v0a>v_{0} and the transition to an expanding crystal for κ>a\kappa>a are effects that already exist at finite NN and are independent of NN. In the attractive case the situation is a bit different. Indeed, in the range of parameters corresponding to the phase FeF_{e}, the delta peaks in the densities at x=±xex=\pm x_{e} are replaced by peaks of finite width. Thus the transition from a smooth density to the presence of shocks is an effect of large NN. Additionally, in the phase δ0\delta_{0}, the density ρs\rho_{s} is a delta function even at finite NN only outside the dotted rectangle (a>v0a>v_{0} or κ¯>2v0\bar{\kappa}>2v_{0}) while inside it has a finite width which decreases with NN. Below we will nevertheless use the names of the infinite NN phases, even at finite NN, to refer to the range of parameters corresponding to these phases.

As mentioned in the text, for any finite N2N\geq 2 the particles tend to aggregate into clusters (i.e. groups of particles having all the same position for a finite period of time) due to the discontinuity in the interaction force κ¯Nsgn(xjxi)-\frac{\bar{\kappa}}{N}{\rm sgn}(x_{j}-x_{i}) (in the attractive case), as well as in the external force V(xi)=asgn(xi)-V^{\prime}(x_{i})=-a\,{\rm sgn}(x_{i}) when it is present 111Note that if instead these discontinuities where smooth with a microscopic scale ll, the clusters would still be present but would have a finite size O(l)O(l).. In the phases where the density is smooth, such clusters can only form if two or more particles keep the same value of σ\sigma over an extended period of time, thus they are generally small. For instance in the case N=2N=2 without external potential, for v0>κ¯/2v_{0}>\bar{\kappa}/2, which we recalled in App. I.2, this corresponds to the delta part in P++(z)P_{++}(z) and P(z)P_{--}(z) in (63). More precisely, we expect that the probability to observe a cluster containing any finite fraction of particles decays exponentially with NN in this case. Thus such clusters are not visible in the densities for N+N\to+\infty. The delta peak in the density ρs\rho_{s} at x=0x=0 for finite NN, with weight cNc_{N} (see Fig. 1, which indeed shows the exponential decay in NN), corresponds to the particular case where all the particles form a single cluster, an event occuring with probability cNc_{N} in the steady state. Note that for finite NN, only the clusters which have a fixed position (in the reference frame of the center of mass for the case a=0a=0) are visible as delta peaks in the density. For N3N\geq 3 clusters of all sizes are present, but only clusters of size NN have a fixed position and thus appear in the density.

However, larger clusters tend to form when the attraction between particles or the external potential become too strong compared to the noise. In particular, in the phase FeF_{e}, since a single particle at the edge is always attracted towards the center of mass, clusters containing a finite fraction of particles form near x=±xex=\pm x_{e} (with a high probability which goes to 11 as NN goes to infinity). However, the position of these clusters as well as the number of particles that they contain fluctuate. Thus they appear as peaks in the densities, with a finite width which decreases with NN (see right panel of Fig. 3). These peaks only become delta functions in the limit N+N\to+\infty. This is also the case for the phase δ0\delta_{0} inside the dotted rectangle in Fig. 6.

Finally, in the phases I0I_{0}, E0E_{0} and δ0\delta_{0} (outside the dotted rectangle), we observe the formation of stable clusters, which have a fixed position, in that case x=0x=0 (in the reference frame of the center of mass for the case a=0a=0), and contain a fixed number of particles, corresponding to a finite fraction of the particles, independent of NN. In this case, this leads to a delta peak in the density even at finite NN, with a weight which is independent of NN.

All these statements could be made more quantitative through a numerical study of the distribution of cluster sizes in the different phases (and its dependence on the number of particles NN). This analysis goes beyond the scope of the present paper and is left for future work.

IV Derivation of the main equations

Using the Dean-Kawasaki approach Dean ; Kawa , one can establish, as in PLDRankedDiffusion for the passive case, and as in Sec. III A of the Supp Mat. in TouzoDBM2023 in the active case (replacing in (74) arXiv version W(x)V(x)W^{\prime}(x)\to V^{\prime}(x), V~(x)κ¯Nsgn(x)\tilde{V}^{\prime}(x)\to\frac{\bar{\kappa}}{N}{\rm sgn}(x), ρ~ρ++ρ\tilde{\rho}\to\rho_{+}+\rho_{-}, and TNTT\to NT), a stochastic evolution equation for the density fields, which takes the following form

tρσ(x,t)=Tx2ρσ(x,t)+x[ρσ(x,t)(v0σ+V(x)+κ¯𝑑y(ρ+(y,t)+ρ(y,t))sgn(xy))]\displaystyle\partial_{t}\rho_{\sigma}(x,t)=T\partial_{x}^{2}\rho_{\sigma}(x,t)+\partial_{x}[\rho_{\sigma}(x,t)(-v_{0}\sigma+V^{\prime}(x)+\bar{\kappa}\int dy(\rho_{+}(y,t)+\rho_{-}(y,t)){\rm sgn}(x-y))]
+γρσ(x,t)γρσ(x,t)+1Nxξσ(x,t)+σNζ(x,t).\displaystyle+\gamma\rho_{-\sigma}(x,t)-\gamma\rho_{\sigma}(x,t)+\frac{1}{\sqrt{N}}\partial_{x}\xi_{\sigma}(x,t)+\frac{\sigma}{\sqrt{N}}\zeta(x,t)\;.

Here ξ±(x,t)=2Tρσ(x,t)ησ(x,t)\xi_{\pm}(x,t)=\sqrt{2T\rho_{\sigma}(x,t)}\eta_{\sigma}(x,t) are two independent demographic passive noises, where η±\eta_{\pm} are two independent standard space time white noises, originating from the thermal white noises, and ζ(x,t)\zeta(x,t) is a non-Gaussian active noise, white in time, originating from the telegraphic noises. Some more details about these noises are provided in Sec. III A of the Supp Mat. in TouzoDBM2023 .

It is very important to note that we have used the fact that the interaction force vanishes at x=0x=0, i.e. sgn(0)=0{\rm sgn}(0)=0, in the present model, which allows to establish the DK equation without difficulties (see the discussion in Sections III A and III B of the Supp Mat. in TouzoDBM2023 ). Physically, this is related to the fact that here the particles can cross.

Equation (IV) is formally exact for any NN. However it is most useful at large NN. Indeed both noise terms are typically O(1/N)O(1/\sqrt{N}), hence for most applications one can either neglect them or treat them in perturbation. At this stage the interaction term is still in a non-local form. However, for the Coulomb interaction, it can be brought to a local form, as used in PLDRankedDiffusion , and as we now discuss. Another case of long range interactions where this is possible (and maybe the only other one) is the logarithmic interaction, e.g. for the Dyson Brownian motion, where the resolvant also obeys, in some cases, a local equation (see e.g. TouzoDBM2023 ).

Recalling the definitions ρs(x,t)=ρ+(y,t)+ρ(y,t)\rho_{s}(x,t)=\rho_{+}(y,t)+\rho_{-}(y,t) and ρd(x,t)=ρ+(y,t)ρ(y,t)\rho_{d}(x,t)=\rho_{+}(y,t)-\rho_{-}(y,t), we now study the large NN limit. Neglecting the noise (hence assuming that the density become self-averaging in that limit) we obtain the pair of deterministic dynamical equations

tρs(x,t)=Tx2ρs(x,t)+x[(v0ρd(x,t)+ρs(x,t)(V(x)+κ¯dyρs(y,t)sgn(xy))],\displaystyle\partial_{t}\rho_{s}(x,t)=T\partial_{x}^{2}\rho_{s}(x,t)+\partial_{x}[(-v_{0}\rho_{d}(x,t)+\rho_{s}(x,t)(V^{\prime}(x)+\bar{\kappa}\int dy\rho_{s}(y,t){\rm sgn}(x-y))]\;, (74)
tρd(x,t)=Tx2ρd(x,t)+x[(v0ρs(x,t)+ρd(x,t)(V(x)+κ¯dyρs(y,t)sgn(xy))]2γρd(x,t).\displaystyle\partial_{t}\rho_{d}(x,t)=T\partial_{x}^{2}\rho_{d}(x,t)+\partial_{x}[(-v_{0}\rho_{s}(x,t)+\rho_{d}(x,t)(V^{\prime}(x)+\bar{\kappa}\int dy\rho_{s}(y,t){\rm sgn}(x-y))]-2\gamma\rho_{d}(x,t)\;.

We will search for solutions of these equations with initial conditions ρs(x,t=0)\rho_{s}(x,t=0), ρd(x,t=0)\rho_{d}(x,t=0) vanishing for x=±x=\pm\infty.

As in the main text, and in PLDRankedDiffusion for the passive case, we now define rank fields r(x,t)r(x,t) and s(x,t)s(x,t) (also called height, or counting fields) through

ρs(x,t)=xr(x,t),r(x,t)=xdxρs(x,t)12,\displaystyle\rho_{s}(x,t)=\partial_{x}r(x,t)\quad,\quad r(x,t)=\int^{x}_{-\infty}dx^{\prime}\rho_{s}(x^{\prime},t)-\frac{1}{2}\;, (75)
ρd(x,t)=xs(x,t),s(x,t)=xdyρd(y,t).\displaystyle\rho_{d}(x,t)=\partial_{x}s(x,t)\quad,\quad s(x,t)=\int_{-\infty}^{x}dy\,\rho_{d}(y,t)\;.

Hence r(x,t)r(x,t) increases monotonically from 1/2-1/2 at x=x=-\infty to +1/2+1/2 at x=+x=+\infty, while s(x,t)s(x,t) vanishes for xx\to-\infty. The equations (74) become

txr(x,t)=Tx3r(x,t)+x[v0xs(x,t)+2κ¯r(x,t)xr(x,t)+V(x)xr(x,t)],\displaystyle\partial_{t}\partial_{x}r(x,t)=T\partial_{x}^{3}r(x,t)+\partial_{x}[-v_{0}\partial_{x}s(x,t)+2\bar{\kappa}r(x,t)\partial_{x}r(x,t)+V^{\prime}(x)\partial_{x}r(x,t)]\;, (76)
txs(x,t)=Tx3s(x,t)+x[v0xr(x,t)+(V(x)+2κ¯r(x,t))xs(x,t)]2γxs(x,t),\displaystyle\partial_{t}\partial_{x}s(x,t)=T\partial_{x}^{3}s(x,t)+\partial_{x}[-v_{0}\partial_{x}r(x,t)+(V^{\prime}(x)+2\bar{\kappa}r(x,t))\partial_{x}s(x,t)]-2\gamma\partial_{x}s(x,t)\;,

where we have used that upon integration by part

𝑑yρs(y,t)sgn(xy)\displaystyle\int dy\rho_{s}(y,t){\rm sgn}(x-y) =\displaystyle= 𝑑yyr(y,t)sgn(xy)=2r(x,t)+[r(y,t)sgn(xy)]+\displaystyle\int dy\partial_{y}r(y,t){\rm sgn}(x-y)=2r(x,t)+[r(y,t){\rm sgn}(x-y)]^{+\infty}_{-\infty} (77)
=\displaystyle= 2r(x,t)(r(,t)+r(+,t))=2r(x,t).\displaystyle 2r(x,t)-(r(-\infty,t)+r(+\infty,t))=2r(x,t)\;.

We can now integrate the first equation over xx noting that the integration constant vanishes from the boundary conditions for r(x,t)r(x,t) at x=±x=\pm\infty and since ρs(x)\rho_{s}(x) and ρd(x,t)\rho_{d}(x,t) vanish at infinity. We also make the natural assumption that V(x)ρs(x,t)V^{\prime}(x)\rho_{s}(x,t) vanishes at infinity. We can integrate also the second equation from -\infty to xx, which gives our final set of dynamical equations

tr(x,t)=Tx2r(x,t)v0xs(x,t)+2κ¯r(x,t)xr(x,t)+V(x)xr(x,t),\displaystyle\partial_{t}r(x,t)=T\partial_{x}^{2}r(x,t)-v_{0}\partial_{x}s(x,t)+2\bar{\kappa}r(x,t)\partial_{x}r(x,t)+V^{\prime}(x)\partial_{x}r(x,t)\;, (78)
ts(x,t)=Tx2s(x,t)v0xr(x,t)+2κ¯r(x,t)xs(x,t)+V(x)xs(x,t)2γs(x,t).\displaystyle\partial_{t}s(x,t)=T\partial_{x}^{2}s(x,t)-v_{0}\partial_{x}r(x,t)+2\bar{\kappa}r(x,t)\partial_{x}s(x,t)+V^{\prime}(x)\partial_{x}s(x,t)-2\gamma s(x,t)\;.

These are the equations given in the text, and they are valid both for attractive and repulsive gas. Denoting p±(t)=N±(t)/Np_{\pm}(t)=N_{\pm}(t)/N the fraction of ±\pm particles in the system at time tt, we note that with this choice of definition, s(+,t)=+𝑑xρd(x,t)=p+(t)p(t)s(+\infty,t)=\int_{-\infty}^{+\infty}dx\rho_{d}(x,t)=p_{+}(t)-p_{-}(t) is not necessarily zero at finite time, depending on the initial condition. However, for γ>0\gamma>0 it decays to zero, since t𝑑xρd(x,t)=2γt𝑑xρd(x,t)\partial_{t}\int dx\rho_{d}(x,t)=-2\gamma\partial_{t}\int dx\rho_{d}(x,t). Hence for γ>0\gamma>0 (i) if one starts with p+(t=0)=p(t=0)p_{+}(t=0)=p_{-}(t=0) then p+(t)p(t)=0p_{+}(t)-p_{-}(t)=0 for all times and (ii) in the stationary state, one has p+=pp_{+}=p_{-}. In both cases, the field s(x,t)=s(x)s(x,t)=s(x) should vanish at both x=±x=\pm\infty. The above properties are true of course only because here we have neglected the noise terms at large NN. Note that γ=0\gamma=0 is a special case, with different and conserved values of the ratios p±p_{\pm}, where there can be many stationary states called fixed points.

Remark. In the diffusive limit v0,γ+v_{0},\gamma\to+\infty with Ta=v022γT_{a}=\frac{v_{0}^{2}}{2\gamma} fixed, one sees that it is natural to assume that in the second equation in (78) all terms but two are negligible and one gets

v0xr(x,t)2γs(x,t)=0.-v_{0}\partial_{x}r(x,t)-2\gamma s(x,t)=0\;. (79)

Inserting in the first equation in (78) one sees that the term v0xs(x,t)-v_{0}\partial_{x}s(x,t) tends to a thermal term Tax2r(x,t)T_{a}\partial_{x}^{2}r(x,t), and one recovers the equation for r(x,t)r(x,t) of the passive case, with a thermal noise at temperature TaT_{a}.

V Large time solution in the presence of a linear potential

In this section we consider the equations (9-10) which describe the system at large NN in the presence of a linear potential V(x)=a|x|V(x)=a|x| (a0a\geq 0), both in the attractive and repulsive case setting T=0T=0. This leads to the phase diagram in Fig. 6 for which we provide here a derivation. This also provides a derivation of the first part of the Letter by simply setting a=0a=0. We compute the rank fields r(x,t)r(x,t) and s(x,t)s(x,t) at large times. Whenever there is a steady state, one can deduce the densities ρs(x)\rho_{s}(x) and ρd(x)\rho_{d}(x) using the parametric representation (35) and its extension in all phases.

To test our predictions for large NN we have also performed numerical simulations by solving the Langevin equation and measuring the densities in the steady state for some finite values of NN. The results are displayed in Fig 7 and Fig 8 and show excellent agreement with the predictions.

V.1 Attractive case

We start with the attractive case. We first assume that there exists a steady state and solve the T=0T=0 stationarity equations

v0s\displaystyle v_{0}s^{\prime} =\displaystyle= 2κ¯rr+asgn(x)r,\displaystyle 2\bar{\kappa}rr^{\prime}+a\,{\rm sgn}(x)\,r^{\prime}\;, (80)
v0r\displaystyle v_{0}r^{\prime} =\displaystyle= 2κ¯rs+asgn(x)s2γs.\displaystyle 2\bar{\kappa}rs^{\prime}+a\,{\rm sgn}(x)\,s^{\prime}-2\gamma s\;. (81)

For a>0a>0, the center of mass is attracted to the origin. It is thus natural to search for a stationary solution such that r(x)r(x) is odd and s(x)s(x) is even in xx. For a=0a=0 we use coordinates such that the center of mass is at zero, and thus it is also reasonable to assume such symmetry (note however that in the main text we did not make this assumption). Thus we can restrict ourselves to x0x\geq 0 in the following, and impose the condition r(0)=0r(0)=0. We proceed as in the text for a=0a=0, by integrating the first equation with r(+)=1/2r(+\infty)=1/2 and s(+)=0s(+\infty)=0, to obtain for x0x\geq 0

s(x)=κ¯v0(r(x)214)+av0(r(x)12).s(x)=\frac{\bar{\kappa}}{v_{0}}\big{(}r(x)^{2}-\frac{1}{4}\big{)}+\frac{a}{v_{0}}\big{(}r(x)-\frac{1}{2}\big{)}\;. (82)

Note that by symmetry this equation holds for any xx\in\mathbb{R} if r(x)r(x) in the last term is replaced by |r(x)||r(x)|.

Substituting in the second equation one obtains

(v02(2κ¯r+a)2)r=γ2(12r)(2a+κ¯+2κ¯r).(v_{0}^{2}-(2\bar{\kappa}r+a)^{2})r^{\prime}=\frac{\gamma}{2}(1-2r)(2a+\bar{\kappa}+2\bar{\kappa}r)\;. (83)

Let us now analyze this equation, starting with the case κ¯=0\bar{\kappa}=0.

Non-interacting case κ¯=0\bar{\kappa}=0. Let us start with the non interacting case κ¯=0\bar{\kappa}=0. Equation (83) becomes

(v02a2)r=aγ(12r),(v_{0}^{2}-a^{2})r^{\prime}=a\gamma(1-2r)\;, (84)

a linear equation whose general solution for x0x\geq 0 is

12r(x)=Ce2aγxv02a2.1-2r(x)=Ce^{-\frac{2a\gamma x}{v_{0}^{2}-a^{2}}}\;. (85)

Since r(+)=12r(+\infty)=\frac{1}{2} and r(0)=0r(0)=0, there are two phases:

(i) av0a\geq v_{0}, in which case C=0C=0, leading to r(x)=12r(x)=\frac{1}{2} for x>0x>0 and, when extending to xx\in\mathbb{R}

r(x)=12sgn(x),ρs(x)=δ(x)r(x)=\frac{1}{2}{\rm sgn}(x)\quad,\quad\rho_{s}(x)=\delta(x) (86)

leading to s(x)=0s(x)=0.

(ii) a<v0a<v_{0}, in which case C=1C=1 footnoteC , leading to a smooth solution which reads, upon extending to xx\in\mathbb{R}

r(x)=12sgn(x)(1e2aγxv02a2),ρs(x)=γav02a2exp(2γav02a2|x|)r(x)=\frac{1}{2}\operatorname{sgn}(x)(1-e^{-\frac{2a\gamma x}{v_{0}^{2}-a^{2}}})\quad,\quad\rho_{s}(x)=\frac{\gamma a}{v_{0}^{2}-a^{2}}\,\exp\left(-\frac{2\gamma a}{v_{0}^{2}-a^{2}}|x|\right) (87)

where s(x)s(x) is given for x0x\geq 0 by (82). We thus recover the known result of DKM19 .

Interacting case κ¯>0\bar{\kappa}>0. Let us now focus on the attractive case κ¯>0\bar{\kappa}>0. The phase diagram in the space (κ¯/v0,a/v0)(\bar{\kappa}/v_{0},a/v_{0}) is similar to the case a=0a=0 (the phases are the same, but the phase boundaries depend on a/v0a/v_{0}).

Phase IsI_{s}. For v0>a+κ¯v_{0}>a+\bar{\kappa}, the prefactor on the l.h.s of (83) is strictly positive. Since the r.h.s. is positive and bounded, we get that r(x)r^{\prime}(x) is bounded on +\mathbb{R}^{+}. Hence the stationary density is smooth with infinite support, and r(x)r(x) for x0x\geq 0 is obtained by inversion of the equation

γx=𝖿a(r)=2κ¯r+v02(κ¯+a)22(κ¯+a)log(κ¯+2a+2κ¯r(κ¯+2a)(12r))\gamma x={\sf f}_{a}(r)=2\bar{\kappa}r+\frac{v_{0}^{2}-(\bar{\kappa}+a)^{2}}{2(\bar{\kappa}+a)}\log\left(\frac{\bar{\kappa}+2a+2\bar{\kappa}r}{(\bar{\kappa}+2a)(1-2r)}\right) (88)

and s(x)s(x) is then given for x0x\geq 0 by (82). Note that for convenience we normalized so that 𝖿a=0(r)=κ¯f(r){\sf f}_{a=0}(r)=\bar{\kappa}f(r), and that 𝖿a(0)=0{\sf f}_{a}(0)=0, 𝖿a(0)=2(v02a2)/(2a+κ¯){\sf f}^{\prime}_{a}(0)=2(v_{0}^{2}-a^{2})/(2a+\bar{\kappa}) and 𝖿a′′(0)=8a(v02(a+κ¯)2)/(2a+κ¯)2{\sf f}^{\prime\prime}_{a}(0)=8a(v_{0}^{2}-(a+\bar{\kappa})^{2})/(2a+\bar{\kappa})^{2}. This is the phase denoted IsI_{s} in Fig. 6. The densities ρs\rho_{s} and ρd\rho_{d} have an infinite support. They are smooth on their support except at x=0x=0 for a>0a>0. Let us now examine their behavior near x=0x=0.

Let us assume, as is natural here, that r(0+)=0r(0^{+})=0 (no shock at zero). Then Eq. (83) gives r(0+)=γ(2a+κ¯)2(v02a2)r^{\prime}(0^{+})=\frac{\gamma(2a+\bar{\kappa})}{2(v_{0}^{2}-a^{2})}. Since s(0+)=av0r(0+)s^{\prime}(0^{+})=\frac{a}{v_{0}}r^{\prime}(0^{+}) from (82), we obtain that for xx near zero

ρd(x)av0γ(2a+κ¯)2(v02a2)sgn(x).\rho_{d}(x)\simeq\frac{a}{v_{0}}\frac{\gamma(2a+\bar{\kappa})}{2(v_{0}^{2}-a^{2})}{\rm sgn}(x)\;. (89)

Hence ρd(x)\rho_{d}(x) has a step discontinuity at x=0x=0. It means that there is an excess of ++ particles for x>0x>0 and of - particles for x<0x<0. The result (89) is valid with the phases IsI_{s} and FeF_{e}. It is also true for the non-interacting case κ¯=0\bar{\kappa}=0. Note that the discontinuity diverges as v0av_{0}\to a^{-}.

Now, it is easy to see that for a>0a>0, ρs(x)\rho_{s}(x) has also a non-analyticity at x=0x=0. Indeed inverting (88) near x=0x=0 one finds (taking into account that r(x)r(x) must be odd)

r(x)=γx𝖿a(0)𝖿a′′(0)2𝖿a(0)3γ2x2sgn(x)+O(x3)r(x)=\frac{\gamma x}{{\sf f}^{\prime}_{a}(0)}-\frac{{\sf f}^{\prime\prime}_{a}(0)}{2{\sf f}^{\prime}_{a}(0)^{3}}\gamma^{2}x^{2}{\rm sgn}(x)+O(x^{3}) (90)

which leads to

r(x)=2a+κ¯2(v02a2)γxa(2a+κ¯)(v02(a+κ¯)2)2(v02a2)3γ2x2sgn(x)+O(x3)r(x)=\frac{2a+\bar{\kappa}}{2(v_{0}^{2}-a^{2})}\gamma x-\frac{a(2a+\bar{\kappa})(v_{0}^{2}-(a+\bar{\kappa})^{2})}{2(v_{0}^{2}-a^{2})^{3}}\gamma^{2}x^{2}{\rm sgn}(x)+O(x^{3}) (91)

and finally one obtains for xx near 0

ρs(x)=2a+κ¯2(v02a2)γa(2a+κ¯)(v02(a+κ¯)2)(v02a2)3γ2|x|+O(x2)\rho_{s}(x)=\frac{2a+\bar{\kappa}}{2(v_{0}^{2}-a^{2})}\gamma-\frac{a(2a+\bar{\kappa})(v_{0}^{2}-(a+\bar{\kappa})^{2})}{(v_{0}^{2}-a^{2})^{3}}\gamma^{2}|x|+O(x^{2}) (92)

which shows that ρs(x)\rho_{s}(x) is continuous but with a derivative discontinuity at x=0x=0. In this phase the density is maximum at the origin.

In the critical case v0=κ¯+av_{0}=\bar{\kappa}+a, i.e. at the phase boundary IsI_{s}-FeF_{e}, one has simply 𝖿a(r)=2κ¯r{\sf f}_{a}(r)=2\bar{\kappa}r and (88) leads to exactly the same linear solution (19) for r(x)r(x) than for a=0a=0, hence one obtains

r(x)=12xxe,s(x)=κ¯4v0((xxe)21)+a2v0(|x|xe1),|x|xe=κ¯γr(x)=\frac{1}{2}\frac{x}{x_{e}^{*}}\quad,\quad s(x)=\frac{\bar{\kappa}}{4v_{0}}\big{(}(\frac{x}{x^{*}_{e}})^{2}-1\big{)}+\frac{a}{2v_{0}}(\frac{|x|}{x_{e}^{*}}-1)\quad,\quad|x|\leq x_{e}^{*}=\frac{\bar{\kappa}}{\gamma} (93)

which lead to the densities

ρs(x)=γ2κ¯,ρd(x)=γ2x2v0κ¯+aγ2v0κ¯sgn(x),|x|xe=κ¯γ\rho_{s}(x)=\frac{\gamma}{2\bar{\kappa}}\quad,\quad\rho_{d}(x)=\frac{\gamma^{2}x}{2v_{0}\bar{\kappa}}+\frac{a\gamma}{2v_{0}\bar{\kappa}}{\rm sgn}(x)\quad,\quad|x|\leq x_{e}^{*}=\frac{\bar{\kappa}}{\gamma} (94)

which vanish for |x|>xe|x|>x_{e}^{*}. They have step discontinuities at the two edges, and in addition ρd(x)\rho_{d}(x) has a step discontinuity at the origin, as discussed above.

Phase FeF_{e}. For v0<κ¯+av_{0}<\bar{\kappa}+a, 𝖿a(r){\sf f}_{a}(r) becomes non-invertible (recall that 𝖿0(r){\sf f}_{0}(r) is plotted in Fig. 3). The density then has a finite support with shocks at the edges ±xe\pm x_{e}. As in the case a=0a=0, integrating equations (80-81) around xe>0x_{e}>0 with the correct interpretation of the force term as discussed in the text leads to

v0Δs=(κ¯(r(xe+)+r(xe))+a)Δr,v0Δr=(κ¯(r(xe+)+r(xe))+a))Δsv_{0}\Delta s=\left(\bar{\kappa}(r(x_{e}^{+})+r(x_{e}^{-}))+a\right)\Delta r~,~v_{0}\Delta r=\left(\bar{\kappa}(r(x_{e}^{+})+r(x_{e}^{-}))+a)\right)\Delta s (95)

with Δr=r(xe+)r(xe)\Delta r=r(x_{e}^{+})-r(x_{e}^{-}) and Δs=s(xe+)s(xe)\Delta s=s(x_{e}^{+})-s(x_{e}^{-}). A non-zero Δr\Delta r then implies r(xe+)+r(xe)=(v0a)/κ¯r(x_{e}^{+})+r(x_{e}^{-})=(v_{0}-a)/\bar{\kappa}, leading to

r(xe)=v0aκ¯12r(x_{e}^{-})=\frac{v_{0}-a}{\bar{\kappa}}-\frac{1}{2} (96)

since r(xe+)=1/2r(x_{e}^{+})=1/2. Inside the support, r(x)r(x) is again given by (88), where the function 𝖿a(r){\sf f}_{a}(r) is invertible on the interval [0,r(xe)][0,r(x_{e}^{-})]. Indeed the point where 𝖿a(r)=0{\sf f}^{\prime}_{a}(r)=0 is r=r=(v0a)/(2κ)>r(xe)r=r^{*}=(v_{0}-a)/(2\kappa)>r(x_{e}^{-}). This is the phase denoted FeF_{e} in Fig. 6. The densities ρs\rho_{s} and ρd\rho_{d} have a bounded support [xe,xe][-x_{e},x_{e}] with delta singularities at the two edges, i.e one has for xx\in\mathbb{R}

ρs(x)=Δr(δ(xxe)+δ(x+xe))+ρ~s(x),ρd(x)=Δs(δ(xxe)δ(x+xe))+ρ~d(x),Δr=Δs=1v0aκ¯\rho_{s}(x)=\Delta r(\delta(x-x_{e})+\delta(x+x_{e}))+\tilde{\rho}_{s}(x)~,~\rho_{d}(x)=\Delta s(\delta(x-x_{e})-\delta(x+x_{e}))+\tilde{\rho}_{d}(x)~,~\Delta r=\Delta s=1-\frac{v_{0}-a}{\bar{\kappa}} (97)

where ρ~s\tilde{\rho}_{s} and ρ~d\tilde{\rho}_{d} are smooth functions on ]xe,0[]-x_{e},0[ and ]0,xe[]0,x_{e}[ (since 𝖿a(r){\sf f}_{a}(r) is generically analytic near r=r(xe)r=r(x_{e}^{-}) it implies that ρ~s(xe)\tilde{\rho}_{s}(x_{e}^{-}) and ρ~d(xe)\tilde{\rho}_{d}(x_{e}^{-}) are finite, together with all their left derivatives at x=xex=x_{e}^{-}). This corresponds to a cluster of ++ particles at x=xex=x_{e} and a cluster of - particles at x=xex=-x_{e}. This is the phase denoted FeF_{e} in Fig. 6. Since inside the support the formulae for r(x),s(x),ρs(x)r(x),s(x),\rho_{s}(x) and ρd(x)\rho_{d}(x) are the same as for the phase IsI_{s} above, one finds that their behavior around x=0x=0 is again given by (89) and (92). Hence in this phase ρd(x)\rho_{d}(x) has also a jump at x=0x=0 and ρs(x)\rho_{s}(x) has a linear cusp, with however positive amplitude, i.e. the density ρs(x)\rho_{s}(x) is minimum at x=0x=0.

Phase δ0\delta_{0}. For v0κ¯/2+av_{0}\leq\bar{\kappa}/2+a all particles are in the same cluster. Indeed from (96) we see that as v0κ¯/2+av_{0}\to\bar{\kappa}/2+a, r(xe)0r(x_{e}^{-})\to 0 implying xe0x_{e}\to 0. This is the phase denoted δ0\delta_{0} in Fig. 6. In that phase r(x)=12sgn(x)r(x)=\frac{1}{2}{\rm sgn}(x) and s(x)=0s(x)=0, leading to ρs(x)=δ(x)\rho_{s}(x)=\delta(x) and ρd(x)=0\rho_{d}(x)=0.

Remark: stability of a single cluster at finite NN. Let us recall that for a=0a=0 and for v0<κ¯/2v_{0}<\bar{\kappa}/2 there is a stability argument (given in the text) for all particles to be in a single cluster for arbitrary NN. Let us now extend this argument for a>0a>0. Denote x+x_{+} (resp. xx_{-}) the position of one of the rightmost (resp. leftmost) particles and n+n_{+} (resp. nn_{-}) the number of particles at the same location. One has

dx+dtv0κ¯N(Nn+)asgn(x+)\displaystyle\frac{dx_{+}}{dt}\leq v_{0}-\frac{\bar{\kappa}}{N}(N-n_{+})-a\operatorname{sgn}(x_{+}) (98)
dxdtv0+κ¯N(Nn)asgn(x)\displaystyle\frac{dx_{-}}{dt}\geq-v_{0}+\frac{\bar{\kappa}}{N}(N-n_{-})-a\operatorname{sgn}(x_{-}) (99)

which leads to

d(x+x)dt2v0κ¯(2n++nN)a(sgn(x+)sgn(x))2v0κ¯\frac{d(x_{+}-x_{-})}{dt}\leq 2v_{0}-\bar{\kappa}\left(2-\frac{n_{+}+n_{-}}{N}\right)-a(\operatorname{sgn}(x_{+})-\operatorname{sgn}(x_{-}))\leq 2v_{0}-\bar{\kappa} (100)

since one has sgn(x+)sgn(x)0\operatorname{sgn}(x_{+})-\operatorname{sgn}(x_{-})\geq 0. This implies that for v0<κ¯/2v_{0}<\bar{\kappa}/2 all particles will end up in a single cluster. Note however that for finite NN the δ\delta peak in the density is not necessarily at x=0x=0 in this case. On the other hand if v0<av_{0}<a, then as long as x+>0x_{+}>0, dx+dt<v0a<0\frac{dx_{+}}{dt}<v_{0}-a<0, and symmetrically for xx_{-}. This also implies, for any NN, that all particles will end up in a single cluster, which in this case must be at x=0x=0. Thus, a delta peak in the density ρs(x)\rho_{s}(x) containing all the particles is stable for any NN for v0<min(a,κ¯/2)v_{0}<\min(a,\bar{\kappa}/2). However in the regime min(a,κ¯/2)<v0<κ¯/2+a\min(a,\bar{\kappa}/2)<v_{0}<\bar{\kappa}/2+a, it is stable only in the limit N+N\to+\infty, while for finite NN there can be deviations from this state. Consider for example that all NN particles are at x=0x=0, all in the ++ state. Then they will have a total velocity v0a>0v_{0}-a>0 and will be able to escape from x=0x=0. Imagine now that at some point half of the particles switch sign. Then the cluster will break since v0>κ¯/2v_{0}>\bar{\kappa}/2. As N+N\to+\infty, the probability of such events decreases exponentially, leading to the δ0\delta_{0} phase.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 7: Density ρs(x)\rho_{s}(x) in the steady state in the different phases, in the presence of a potential V(x)=a|x|V(x)=a|x|, for different values of NN. For all figures v0=1v_{0}=1 and γ=1\gamma=1. For finite NN it is obtained by numerical simulations (note that here we have not made the shift to the reference frame of the center of mass contrary to the case a=0a=0 discussed in the main text). The black curves correspond to the limit N+N\to+\infty and are computed using the parametric representation (35) extended to all phases. The delta peaks in the density are represented by dotted vertical lines. Note that in all phases except FeF_{e} the data for N=100N=100 is already hardly distinguishable from the prediction at N=+N=+\infty.

V.2 Repulsive case

We now turn to the repulsive case, κ=κ¯>0\kappa=-\bar{\kappa}>0. While for a=0a=0, there is a single, expanding phase, the addition of a linear potential makes the phenomenology much richer: for a>0a>0, 5 different phases can be distinguished. Two of these phases (IsI_{s} and δ0\delta_{0}) are continuations of phases which exist for κ¯>0\bar{\kappa}>0.

Phase δ0\delta_{0}. First of all, when av0+κa\geq v_{0}+\kappa, the density trivially converges to ρs(x)=δ(x)\rho_{s}(x)=\delta(x) independently of NN (as can be seen e.g. from (98) setting κ¯=κ\bar{\kappa}=-\kappa). This is again the phase δ0\delta_{0} in Fig. 6, but on the side κ¯<0\bar{\kappa}<0.

For a<v0+κa<v_{0}+\kappa, one needs to distinguish between a stationary case for κa<v0+κ\kappa\leq a<v_{0}+\kappa, and an expanding case for a<κa<\kappa. Each of these cases will give rise to two distinct phases, separated by the line a=v0a=v_{0}, leading to four phases in total. Let us start with the stationary case, where we again solve Eqs. (80-81), but this time for κ¯=κ<0\bar{\kappa}=-\kappa<0. We can again look for solutions with r(x)r(x) odd in xx and s(x)s(x) even.

Phase IsI_{s}. For κ=κ¯>0\kappa=-\bar{\kappa}>0, (83) becomes (for x0x\geq 0)

(v02(a2κr)2)r=γ2(12r)(2aκ2κr).(v_{0}^{2}-(a-2\kappa r)^{2})r^{\prime}=\frac{\gamma}{2}(1-2r)(2a-\kappa-2\kappa r)\;. (101)

When a>κa>\kappa, the right hand side is always positive. If in addition a<v0a<v_{0}, then the factor in the left-hand side is strictly positive and r(x)r(x) is correctly given by equation (88) for any x0x\geq 0. The function 𝖿a(r){\sf f}_{a}(r) is invertible, and this is the same phase IsI_{s} as discussed above: both densities ρs(x)\rho_{s}(x) and ρd(x)\rho_{d}(x) are smooth with infinite support.

Phase I0I_{0}. However, if a>v0a>v_{0} (still assuming κa<v0+κ\kappa\leq a<v_{0}+\kappa) the left-hand side of (101) is positive only above a certain value of rr. Thus, for r(x)r^{\prime}(x) to be positive everywhere there must be a jump at x=0x=0, such that

r(0+)=av02κ.r(0^{+})=\frac{a-v_{0}}{2\kappa}\;. (102)

and r(x)r(x) for x>0x>0 is now given by

γx=𝖿a(r)𝖿a(r(0+))\gamma x={\sf f}_{a}(r)-{\sf f}_{a}(r(0^{+})) (103)

where 𝖿a(r){\sf f}_{a}(r) is still given by (88). This the phase I0I_{0} in Fig. 6, which is stationary, has infinite support and exhibits a shock at x=0x=0. Note also that ρs(x)=r(x)\rho_{s}(x)=r^{\prime}(x) diverges near x=0x=0. More precisely near x=0x=0 one has, by linearizing (101) for rr near r(0+)r(0^{+}) and integrating w.r.t. x>0x>0,

2κv0(r(x)r(0+))2γ2κ(v02(aκ)2)x.2\kappa v_{0}(r(x)-r(0^{+}))^{2}\sim\frac{\gamma}{2\kappa}(v_{0}^{2}-(a-\kappa)^{2})\,x\;. (104)

Hence there is generically an inverse square root divergence of the density ρs(x)\rho_{s}(x) near the shock at x=0x=0, i.e. for xx\in\mathbb{R} and xx near 0 one has

r(x)(r(0+)+B|x|)sgn(x),ρs(x)av0κδ(x)+B2|x|withB=12κγv0(v02(aκ)2).r(x)\simeq\left(r(0^{+})+B\sqrt{|x|}\right){\rm sgn}(x)\quad,\quad\rho_{s}(x)\simeq\frac{a-v_{0}}{\kappa}\delta(x)+\frac{B}{2\sqrt{|x|}}\quad{\rm with}\quad B=\frac{1}{2\kappa}\sqrt{\frac{\gamma}{v_{0}}(v_{0}^{2}-(a-\kappa)^{2})}\;. (105)

Using (82) and the value of r(0+)r(0^{+}) in (102), we find that

s(x)(aκ)2v024κv0+B|x|s(x)\simeq\frac{(a-\kappa)^{2}-v_{0}^{2}}{4\kappa v_{0}}+B\sqrt{|x|} (106)

and since s(x)s(x) is even, s(x)s(x) is continuous at x=0x=0 which implies that the cluster of particles there has equal fractions of ±\pm species. However, Eq. (106) leads to an inverse square root divergence of the density ρd(x)\rho_{d}(x)

ρd(x)B2|x|sgn(x).\rho_{d}(x)\simeq\frac{B}{2\sqrt{|x|}}\,{\rm sgn}(x)\;. (107)

This result shows that the inverse square root divergence involves only ++ particles for x>0x>0 and minus particles for x<0x<0.

The presence of the cluster of particles at x=0x=0 can easily be understood via a qualitative argument. When v0<av_{0}<a, non-interacting particles would remain stuck at x=0x=0, and it is only the repulsion that allows some particles to escape from x=0x=0. Therefore there must be a cluster of particles at x=0x=0, which contains a fraction r(0+)r(0)=2r(0+)r(0^{+})-r(0^{-})=2r(0^{+}) of particles. The particle immediately at the right of this cluster is able to escape if the total force resulting from the potential and the interactions a+2κr(0+)-a+2\kappa r(0^{+}) is larger than v0-v_{0}. Balancing these two forces fixes the value of r(0+)r(0^{+}) in agreement with (102). The presence of the square root divergence in the density near the cluster comes from the fact that the velocity of particles decreases to zero as one approaches the cluster.

Note that as one gets closer to the phase boundary I0δ0I_{0}-\delta_{0}, i.e. aκv0a-\kappa\to v_{0}, the weight of the δ\delta peak in ρs(x)\rho_{s}(x) tends to unity and B0B\to 0.

Large distance behavior in IsI_{s} and I0I_{0} phases.

Let us first note that the behaviors at large |x||x| of ρs(x)\rho_{s}(x) and ρd(x)\rho_{d}(x) are related in a simple way. Indeed, Eq. (82) implies that whenever the support of the densities in infinite one has as |x|+|x|\to+\infty

ρd(x)a+κ¯v0ρs(x)sgn(x).\rho_{d}(x)\simeq\frac{a+\bar{\kappa}}{v_{0}}\rho_{s}(x){\rm sgn}(x)\;. (108)

One can check that in the phases IsI_{s} and I0I_{0} the prefactor is smaller than unity, as it should. One sees that as v0v_{0} increases the proportion of - on the positive side increases and vice-versa, i.e. the mixing of ++ and - particles becomes stronger.

Next, let us determine this large distance behavior. Expanding Eq. (88) for rr near 1/21/2 and x+x\to+\infty one finds

r(x)=12a+κ¯2a+κ¯e2(a+κ¯)v02(a+κ¯)2(γxκ¯)r(x)=\frac{1}{2}-\frac{a+\bar{\kappa}}{2a+\bar{\kappa}}e^{-\frac{2(a+\bar{\kappa})}{v_{0}^{2}-(a+\bar{\kappa})^{2}}(\gamma x-\bar{\kappa})} (109)

which recovers (87) for κ¯=0\bar{\kappa}=0. This leads to

ρs(x)Ase|x|xeξ,ρd(x)Adsgn(x)e|x|xeξ\displaystyle\rho_{s}(x)\simeq A_{s}\,e^{-\frac{|x|-x_{e}^{*}}{\xi_{\infty}}}~,~\rho_{d}(x)\simeq A_{d}\,{\rm sgn}(x)e^{-\frac{|x|-x_{e}^{*}}{\xi_{\infty}}} (110)
ξ=v02(κ¯+a)22γ(κ¯+a),xe=κ¯γ,As=2γ(κ¯+a)2(v02(κ¯+a)2)(κ¯+2a),Ad=a+κ¯v0As.\displaystyle\xi_{\infty}=\frac{v_{0}^{2}-(\bar{\kappa}+a)^{2}}{2\gamma(\bar{\kappa}+a)}~,~x_{e}^{*}=\frac{\bar{\kappa}}{\gamma}~,~A_{s}=\frac{2\gamma(\bar{\kappa}+a)^{2}}{(v_{0}^{2}-(\bar{\kappa}+a)^{2})(\bar{\kappa}+2a)}~,~A_{d}=\frac{a+\bar{\kappa}}{v_{0}}A_{s}\;.

which for a=0a=0 reproduces the result given in the text. This is valid within the phases IsI_{s} and I0I_{0} both in attractive and repulsive case with κ¯=κ\bar{\kappa}=-\kappa.

Critical case a=κa=\kappa. This corresponds to the boundaries IsEsI_{s}-E_{s} and I0E0I_{0}-E_{0}. The reasoning above still holds in the marginal case a=κa=\kappa, both for av0a\leq v_{0} (boundary IsEsI_{s}-E_{s}) and for a>v0a>v_{0} (boundary I0E0I_{0}-E_{0}). Eq. (88) takes a much simpler form in this case,

γx=𝖿~a(r)=2ar(v02a2(12r)1),a<v0,x0\displaystyle\gamma x=\tilde{\sf f}_{a}(r)=2ar\left(\frac{v_{0}^{2}}{a^{2}(1-2r)}-1\right)\quad,\quad a<v_{0}\quad,\quad x\geq 0 (111)
γx=𝖿~a(r)𝖿~a(r(0+))=(v0a+2ar)2a(12r),a>v0,x>0.\displaystyle\gamma x=\tilde{\sf f}_{a}(r)-\tilde{\sf f}_{a}(r(0^{+}))=\frac{(v_{0}-a+2ar)^{2}}{a(1-2r)}\quad,\quad a>v_{0}\quad,\quad x>0\;. (112)

This leads to

r(x)=14a2((aγx+(av0)2)(aγx+(a+v0)2)+a2v02aγx),a<v0,x0\displaystyle r(x)=\frac{1}{4a^{2}}\left(\sqrt{(a\gamma x+(a-v_{0})^{2})(a\gamma x+(a+v_{0})^{2})}+a^{2}-v_{0}^{2}-a\gamma x\right)\quad,\quad a<v_{0}\quad,\quad x\geq 0 (113)
r(x)=14a(γx(4v0+γx)+2(av0)γx),a>v0,x>0\displaystyle r(x)=\frac{1}{4a}\left(\sqrt{\gamma x(4v_{0}+\gamma x)}+2(a-v_{0})-\gamma x\right)\quad,\quad a>v_{0}\quad,\quad x>0 (114)

and, for xx\in\mathbb{R}

ρs(x)=γ4a(a2+v02+aγ|x|(aγ|x|+(av0)2)(aγ|x|+(a+v0)2)1),a<v0\displaystyle\rho_{s}(x)=\frac{\gamma}{4a}\left(\frac{a^{2}+v_{0}^{2}+a\gamma|x|}{\sqrt{(a\gamma|x|+(a-v_{0})^{2})(a\gamma|x|+(a+v_{0})^{2})}}-1\right)\quad,\quad a<v_{0} (115)
ρs(x)=γ4a(2v0+γ|x|γ|x|(4v0+γ|x|)1)+(1v0a)δ(x),a>v0.\displaystyle\rho_{s}(x)=\frac{\gamma}{4a}\left(\frac{2v_{0}+\gamma|x|}{\sqrt{\gamma|x|(4v_{0}+\gamma|x|)}}-1\right)+\left(1-\frac{v_{0}}{a}\right)\delta(x)\quad,\quad a>v_{0}\;. (116)

We note that for a>v0a>v_{0} the behavior of the density near the shock at x=0x=0 is compatible with the more general result (105) with an inverse square root divergence of amplitude B=γv0/(2a)B=\sqrt{\gamma v_{0}}/(2a). Finally, an unusual feature is the algebraic decay of the density at infinity. This can be anticipated since the bound state decay length ξ\xi_{\infty} in (110) diverges as κa\kappa\to a^{-}. Indeed, one finds at large xx in both cases (a<v0a<v_{0} and a>v0a>v_{0})

ρs(x)v022aγx2\rho_{s}(x)\simeq\frac{v_{0}^{2}}{2a\gamma x^{2}} (117)

where only the term O(1/|x|3)O(1/|x|^{3}) differs for a<v0a<v_{0} and a>v0a>v_{0}. Note that for a=κa=\kappa one has

s(x)=a4v0(12r(x))2.s(x)=-\frac{a}{4v_{0}}(1-2r(x))^{2}\;. (118)

This implies that at large x>0x>0 one has s(x)14v03aγ2x2s(x)\simeq-\frac{1}{4}\frac{v_{0}^{3}}{a\gamma^{2}x^{2}}, leading to

ρd(x)v032aγ2x3.\rho_{d}(x)\simeq\frac{v_{0}^{3}}{2a\gamma^{2}x^{3}}\;. (119)
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 8: Density ρd(x)\rho_{d}(x) in the steady state in the different phases, in the presence of a potential V(x)=a|x|V(x)=a|x|, for different values of NN. For finite NN it is obtained by numerical simulations (note that here we have not made the shift to the reference frame of the center of mass contrary to the case a=0a=0 discussed in the main text). For all figures v0=1v_{0}=1 and γ=1\gamma=1. The black curves correspond to the limit N+N\to+\infty and are computed using the parametric representation (35) and its extension to all phases. The delta peaks in the density are represented by dotted vertical lines.

Expanding phases EsE_{s} and E0E_{0}. Let us now turn to the expanding case κ>a\kappa>a, which contains two phases separated by the line v0=av_{0}=a. In this case the repulsion is strong enough to overcome the potential and a fraction of the particles is sent to infinity. Thus at large time, there are two regions of interest in space.

(i) First, the particles which escape to infinity will form an expanding gas on scales |x|t|x|\sim t, with two edges moving at a velocity ±(κa)\pm(\kappa-a), each with an associated boundary layer of spatial width t\sim\sqrt{t}. Below we determine the large time densities which we find identical for the two phases EsE_{s} and E0E_{0}.

(ii) Next the particles which remain inside the potential well will exhibit a stationary density on distances of order O(1)O(1), which we will compute. This density will either exhibit a shock at x=0x=0 (phase E0E_{0}) or will be smooth (phase EsE_{s}), in both cases with infinite support.

(i) Escaping particles. Let us start with the escaping particles. As for a=0a=0, we start from the equations

tr\displaystyle\partial_{t}r =\displaystyle= Tx2rv0xs2κrxr+asgn(x)xr,\displaystyle T\partial_{x}^{2}r-v_{0}\partial_{x}s-2\kappa r\partial_{x}r+a\,{\rm sgn}(x)\partial_{x}r\;, (120)
ts\displaystyle\partial_{t}s =\displaystyle= Tx2sv0xr2κrxs+asgn(x)xs2γs,\displaystyle T\partial_{x}^{2}s-v_{0}\partial_{x}r-2\kappa r\partial_{x}s+a\,{\rm sgn}(x)\partial_{x}s-2\gamma s\;, (121)

where we have reintroduced the temperature since it does not make the derivation more difficult in this case. We define y=x/ty=x/t and search for a solution of the form

r(x,t)=r0(y)+r1(y)t+,s(x,t)=s1(y)t+r(x,t)=r_{0}(y)+\frac{r_{1}(y)}{t}+\dots\quad,\quad s(x,t)=\frac{s_{1}(y)}{t}+\dots (122)

The first equation (120) gives, at leading order in 1/t1/t

(2κr0yasgn(y))r0=0.(2\kappa r_{0}-y-a\,\operatorname{sgn}(y))r_{0}^{\prime}=0\;. (123)

Thus for a given yy either r0(y)=0r_{0}^{\prime}(y)=0 or r0(y)=(y+asgn(y))/(2κ)r_{0}(y)=(y+a\,\operatorname{sgn}(y))/(2\kappa). Since |r0(y)|1/2|r_{0}(y)|\leq 1/2 we obtain the leading behavior of the solution for x>0x>0 in the scaling region xtx\sim t

r(x,t)min(a2κ+x2κt,12).r(x,t)\simeq\min\left(\frac{a}{2\kappa}+\frac{x}{2\kappa t},\frac{1}{2}\right)\;. (124)

By symmetry this gives the following density for xx\in\mathbb{R} at large times in the scaling region |x|t|x|\sim t

ρs(x,t)aκδ(x)+12κtθ((κa)t|x|),\rho_{s}(x,t)\simeq\frac{a}{\kappa}\delta(x)+\frac{1}{2\kappa t}\theta((\kappa-a)t-|x|)\;, (125)

i.e. uniform on the support [(κa)t,(κa)t][-(\kappa-a)t,(\kappa-a)t] with a ”cluster” at xt=y=0\frac{x}{t}=y=0 containing a fraction

2r0(y=0+)=aκ2r_{0}(y=0^{+})=\frac{a}{\kappa} (126)

of particles. Our solution only shows that these particles have zero velocity on average, and have remained in the vicinity of the potential well. They do not however necessarily form a real cluster in the sense encountered above. Below we will study the structure of their density on scales |x|t|x|\ll t.

The second equation (121) gives, at leading order in 1/t1/t the same relation as for a=0a=0, namely

s1=v02γr0,s_{1}=-\frac{v_{0}}{2\gamma}r_{0}^{\prime}\;, (127)

hence, using the solution for r0(y)r_{0}(y) one obtains, in the scaling region xtx\sim t

s(x,t)v04γκtθ((κa)t|x|)s(x,t)\simeq-\frac{v_{0}}{4\gamma\kappa t}\theta((\kappa-a)t-|x|) (128)

and thus

ρd(x,t)v04γκt(δ((κa)tx)δ((κa)tx)).\rho_{d}(x,t)\simeq\frac{v_{0}}{4\gamma\kappa t}(\delta((\kappa-a)t-x)-\delta((\kappa-a)t-x))\;. (129)

As for a=0a=0, (127) shows that at large times, the noise plays the same role as a Brownian noise with temperature Ta=v022γT_{a}=\frac{v_{0}^{2}}{2\gamma}, and one can for instance compute the boundary layer in exactly the same way (see below).

The present expansion is formally correct inside the ”bulk” of the expanding gas, i.e. for |y|=|x|t<κa|y|=\frac{|x|}{t}<\kappa-a. However examining the equations (120-121) to the next orders in 1/t1/t, one finds an indeterminacy, e.g. r1(y)r_{1}(y) cannot be determined. The reason for that is that the higher order terms, starting with r1(y)r_{1}(y) may depend on the details of the initial condition. One can see how this happens in the simpler case of the Burger’s equation with a=0a=0, which amounts to set v0=0v_{0}=0 in (120) and T=0T=0. In that case the solution r(x,t)r(x,t) starting from the initial condition r(x,0)=R(x)r(x,0)=R(x) is given by the solution of the equation

r=R(x2κrt)r=r(x,t)r=R(x-2\kappa rt)\quad\Leftrightarrow\quad r=r(x,t) (130)

which, introducing the inverse function x0(r)x_{0}(r) such that R(x0(r))=rR(x_{0}(r))=r, can also be written as x=x0(r)+2κrtx=x_{0}(r)+2\kappa rt. Dividing by tt one gets, with y=x/ty=x/t

r=y2κ12κtx0(r)r=\frac{y}{2\kappa}-\frac{1}{2\kappa t}x_{0}(r) (131)

This leads to r0(y)=y2κr_{0}(y)=\frac{y}{2\kappa} and

r1(y)=12κx0(y2κ)r_{1}(y)=-\frac{1}{2\kappa}x_{0}\left(\frac{y}{2\kappa}\right) (132)

which shows explicit dependence in the initial condition. Although in the full problem there is also thermal and active noise, their role in the bulk of the expanding gas may be subdominant.

Boundary layer for the expanding phases. Near the edges of the expanding gas the above 1/t1/t expansion fails and one needs to look for a different scaling form, as discussed in the text, within a boundary layer of width t\sim\sqrt{t}.

In the boundary layer we can look for a solution of (120-121) under the form

r(x,t)=121tf(z),s(x,t)=1tg(z),z=x(κa)tt.r(x,t)=\frac{1}{2}-\frac{1}{\sqrt{t}}\,f(z)\quad,\quad s(x,t)=\frac{1}{t}\,g(z)\quad,\quad z=\frac{x-(\kappa-a)t}{\sqrt{t}}\;. (133)

Inserting these forms into (121) gives at leading order in 1t\frac{1}{t}

g(z)=v02γf(z).g(z)=\frac{v_{0}}{2\gamma}f^{\prime}(z)\;. (134)

Replacing g(z)g(z) in (120), leads to

12f+z2f+Tefff′′+2κff=0,Teff=T+v022γ.\frac{1}{2}f+\frac{z}{2}f^{\prime}+T_{eff}f^{\prime\prime}+2\kappa ff^{\prime}=0\quad,\quad T_{eff}=T+\frac{v_{0}^{2}}{2\gamma}\;. (135)

This is exactly the same boundary layer equation as in the passive case with an effective temperature TeffT_{eff}. This equation can be made dimensionless by writing

f~(z~)=κTefff(z),z~=zTeff\tilde{f}(\tilde{z})=\frac{\kappa}{\sqrt{T_{eff}}}f(z)\quad,\quad\tilde{z}=\frac{z}{\sqrt{T_{eff}}} (136)

yielding the equation

12f~+z~2f~+f~′′+2f~f~=0\frac{1}{2}\tilde{f}+\frac{\tilde{z}}{2}\tilde{f}^{\prime}+\tilde{f}^{\prime\prime}+2\tilde{f}\tilde{f}^{\prime}=0 (137)

which is solved by

f~(z~)=ez~2/4πerfc(z~/2).\tilde{f}(\tilde{z})=\frac{e^{-\tilde{z}^{2}/4}}{\sqrt{\pi}\,{\rm erfc}(-\tilde{z}/2)}\;. (138)

We can thus write the full solution as a function of f~(z~)\tilde{f}(\tilde{z})

r(x,t)=121κTefftf~(z~),s(x,t)=v02γκtf~(z~),z~=x(κa)tTefftr(x,t)=\frac{1}{2}-\frac{1}{\kappa}\sqrt{\frac{T_{eff}}{t}}\tilde{f}(\tilde{z})\quad,\quad s(x,t)=\frac{v_{0}}{2\gamma\kappa t}\tilde{f}^{\prime}(\tilde{z})\quad,\quad\tilde{z}=\frac{x-(\kappa-a)t}{\sqrt{T_{eff}t}} (139)

and

ρs(x,t)=1κtf~(z~),ρd(x,t)=v02γκTefft3/2f~′′(z~)\rho_{s}(x,t)=-\frac{1}{\kappa t}\tilde{f}^{\prime}(\tilde{z})\quad,\quad\rho_{d}(x,t)=\frac{v_{0}}{2\gamma\kappa\sqrt{T_{eff}}\,t^{3/2}}\tilde{f}^{\prime\prime}(\tilde{z}) (140)

which is the result that we displayed in the text in the case a=0a=0.

As a final remark let us stress that there is strictly no difference in the above results between the EsE_{s} and E0E_{0} phases, which can only be distinguished by looking at spatial scales O(1)O(1) to which we now turn.

(ii) Stationary densities for the bound particles. Contrary to the case a=0a=0, we see that a finite fraction aκ\frac{a}{\kappa} of particles remains close to x=0x=0. However, because of the rescaling performed above (xtx\sim t) we do not yet know if these particles form a large cluster at x=0x=0 or if they simply have a position which is of order o(t)o(t). To answer this question we now focus on particles which have a position x=O(1)x=O(1) and assume that they reach a steady state. We then need to solve the same system of equations for r(x)r(x) and s(x)s(x) as in (80-81), but with new boundary conditions r(±)=a2κr(\pm\infty)=\frac{a}{2\kappa} and s(±)=0s(\pm\infty)=0. Integrating (80) we obtain for x>0x>0

s(x)=κv0(r(x)2a24κ2)+av0(r(x)a2κ)=κv0(ra2κ)2.s(x)=-\frac{\kappa}{v_{0}}\big{(}r(x)^{2}-\frac{a^{2}}{4\kappa^{2}}\big{)}+\frac{a}{v_{0}}\big{(}r(x)-\frac{a}{2\kappa}\big{)}=-\frac{\kappa}{v_{0}}(r-\frac{a}{2\kappa})^{2}\;. (141)

Substituting in (81) one obtains

(v02(a2κr)2)r=γκ2(aκ2r)2.(v_{0}^{2}-(a-2\kappa r)^{2})r^{\prime}=\frac{\gamma\kappa}{2}(\frac{a}{\kappa}-2r)^{2}\;. (142)

Let us define r^(x)=κar(x)\hat{r}(x)=\frac{\kappa}{a}r(x) and inject in Eq. (142). One finds that r^(x)\hat{r}(x) satisfies the equation for x>0x>0

(v02a2(12r^)2)r^=γa2(12r^)2(v_{0}^{2}-a^{2}(1-2\hat{r})^{2})\hat{r}^{\prime}=\frac{\gamma a}{2}(1-2\hat{r})^{2} (143)

with r^(±)=1/2\hat{r}(\pm\infty)=1/2. This is exactly the same equation as the one obtained for r(x)r(x) in the marginal case, i.e. Eq. (101) with a=κa=\kappa. This can be understood as follows. Since they are equally split between the two sides, the particles which escape to infinity can be ignored when studying the particles which remain confined in the potential (the force that they exert on a confined particles sums to zero). One then only has to consider Nbound=aκNN_{\rm bound}=\frac{a}{\kappa}N particles with a repulsive interaction strength κN=aNbound\frac{\kappa}{N}=\frac{a}{N_{\rm bound}}, i.e. the bound particles behave as if the effective interaction constant was κeff=a\kappa_{\rm eff}=a, corresponding to the critical case described above.

Thus, using the results above, for x=O(1)x=O(1) and x>0x>0, r(x)r(x) is given by

γx=𝖿~a(κar(x)),a<v0\displaystyle\gamma x=\tilde{\sf f}_{a}\left(\frac{\kappa}{a}r(x)\right)\quad,\quad a<v_{0} (144)
γx=𝖿~a(κar(x))𝖿~a(κar(0+)),r(0+)=av02κ,a>v0\displaystyle\gamma x=\tilde{\sf f}_{a}\left(\frac{\kappa}{a}r(x)\right)-\tilde{\sf f}_{a}\left(\frac{\kappa}{a}r(0^{+})\right)\quad,\quad r(0^{+})=\frac{a-v_{0}}{2\kappa}\quad,\quad a>v_{0} (145)

which correspond respectively to phases EsE_{s} and E0E_{0}, where 𝖿~a\tilde{\sf f}_{a} is given in (111). Thus, there is indeed a cluster of particles at x=0x=0 only in the case a>v0a>v_{0} (phase E0E_{0}).

From this observation one obtains the density ρs(x)\rho_{s}(x) explicitly in the phases EsE_{s} and E0E_{0}. It is simply given by the same formulae (115) and (116) by multiplying the density by the factor aκ\frac{a}{\kappa}. The densities have infinite support and are smooth except at x=0x=0. One obtains the following behaviors at large |x||x|

ρs(x)v022κγx2,ρd(x)v032κγ2x3\rho_{s}(x)\simeq\frac{v_{0}^{2}}{2\kappa\gamma x^{2}}\quad,\quad\rho_{d}(x)\simeq\frac{v_{0}^{3}}{2\kappa\gamma^{2}x^{3}} (146)

so, in that sense, in the phases EsE_{s} and E0E_{0} the bound part of the gas is always critical (i.e. the decay of the densities are power laws rather than exponentials as it is in the phases IsI_{s} and I0I_{0}).

Let us now discuss the behavior of the densities around x=0x=0. In the phase EsE_{s}, applying the results obtained above in the phase IsI_{s}, Eqs. (91) and (92), to r^(x)\hat{r}(x) and performing the rescaling r(x)=aκr^(x)r(x)=\frac{a}{\kappa}\hat{r}(x) one finds

ρs(x)=a22κ(v02a2)γa3v022κ(v02a2)3γ2|x|+O(x2).\rho_{s}(x)=\frac{a^{2}}{2\kappa(v_{0}^{2}-a^{2})}\gamma-\frac{a^{3}v_{0}^{2}}{2\kappa(v_{0}^{2}-a^{2})^{3}}\gamma^{2}|x|+O(x^{2})\;. (147)

From (141) one finds again that s(0+)=av0r(0+)=av0ρs(0)s^{\prime}(0^{+})=\frac{a}{v_{0}}r^{\prime}(0^{+})=\frac{a}{v_{0}}\rho_{s}(0), since r(0+)=0r(0^{+})=0 in this phase, hence we obtain near x=0x=0

ρd(x)a3γ2v0κ(v02a2)sgn(x).\rho_{d}(x)\simeq\frac{a^{3}\gamma}{2v_{0}\kappa(v_{0}^{2}-a^{2})}{\rm sgn}(x)\;. (148)

In the phase E0E_{0} one finds, applying the results obtained above in the phase I0I_{0} to r^(x)\hat{r}(x) and performing the rescaling r(x)=aκr^(x)r(x)=\frac{a}{\kappa}\hat{r}(x) that near x=0x=0

r(x)(r(0+)+B|x|)sgn(x),ρs(x)av0κδ(x)+B2|x|withB=12κγv0,r(x)\simeq\left(r(0^{+})+B^{\prime}\sqrt{|x|}\right){\rm sgn}(x)\quad,\quad\rho_{s}(x)\simeq\frac{a-v_{0}}{\kappa}\delta(x)+\frac{B^{\prime}}{2\sqrt{|x|}}\quad{\rm with}\quad B^{\prime}=\frac{1}{2\kappa}\sqrt{\gamma v_{0}}\;, (149)

as well as (using the relation (141))

s(x)v04κ+B|x|,ρd(x)B2|x|sgn(x).s(x)\simeq-\frac{v_{0}}{4\kappa}+B^{\prime}\sqrt{|x|}\quad,\quad\rho_{d}(x)\simeq\frac{B^{\prime}}{2\sqrt{|x|}}\,{\rm sgn}(x)\;. (150)

V.3 Passive limit

In this section we focus on the phase IsI_{s} and we consider the passive limit v0+v_{0}\to+\infty, γ+\gamma\to+\infty with Ta=v02/(2γ)T_{a}=v_{0}^{2}/(2\gamma) fixed. Taking this limit in (88), one obtains for r>0r>0

xTa=1κ¯+alog(κ¯+2a+2κ¯r(κ¯+2a)(12r)),\frac{x}{T_{a}}=\frac{1}{\bar{\kappa}+a}\log\left(\frac{\bar{\kappa}+2a+2\bar{\kappa}r}{(\bar{\kappa}+2a)(1-2r)}\right)\;, (151)

which leads to, for any xx (using that r(x)r(x) is odd),

r(x)=121e(κ¯+a)|x|/Ta1+κ¯κ¯+2ae(κ¯+a)|x|/Tasgn(x)r(x)=\frac{1}{2}\frac{1-e^{-(\bar{\kappa}+a)|x|/T_{a}}}{1+\frac{\bar{\kappa}}{\bar{\kappa}+2a}e^{-(\bar{\kappa}+a)|x|/T_{a}}}{\rm sgn}(x) (152)

and thus

ρ(x)=(κ¯+a)2T(κ¯+2a)e(κ¯+a)xT(1+κ¯κ¯+2ae(κ¯+a)xT)2.\rho(x)=\frac{(\bar{\kappa}+a)^{2}}{T(\bar{\kappa}+2a)}\frac{e^{-\frac{(\bar{\kappa}+a)x}{T}}}{(1+\frac{\bar{\kappa}}{\bar{\kappa}+2a}e^{-\frac{(\bar{\kappa}+a)x}{T}})^{2}}\;. (153)

This result can also be obtained by solving directly the Dean-Kawasaki equation for the passive case in the large NN limit, i.e. solving

0=Tr′′(x)+(2κ¯r(x)+asgn(x))r(x).0=Tr^{\prime\prime}(x)+(2\bar{\kappa}r(x)+a\,{\rm sgn}(x))r^{\prime}(x)\;. (154)

This solution is valid as long as κ¯>a\bar{\kappa}>-a, i.e. κ<a\kappa<a. It is always continuous with infinite support, which is consistent with the phase IsI_{s}.

The density for the passive system can also be obtained using the Coulomb gas. The NN particle equilibrium Gibbs measure at temperature TT is

exp(κ¯NTi<j|xixj|aTi|xi|)\exp(-\frac{\bar{\kappa}}{NT}\sum_{i<j}|x_{i}-x_{j}|-\frac{a}{T}\sum_{i}|x_{i}|) (155)

which is normalizable if N1Nκ¯>a\frac{N-1}{N}\bar{\kappa}>-a. For large NN one can rewrite it in terms of the density, in a Coulomb gas description as

exp(κ¯N2T𝑑xρ(x)ρ(x)|xx|NaT𝑑xρ(x)|x|N𝑑xρ(x)logρ(x))\exp(-\frac{\bar{\kappa}N}{2T}\int dx\rho(x)\rho(x^{\prime})|x-x^{\prime}|-\frac{Na}{T}\int dx\rho(x)|x|-N\int dx\rho(x)\log\rho(x)) (156)

with the constraint 𝑑xρ(x)=1\int dx\rho(x)=1, and we have included the entropy term which, given the way we scaled the interaction, is of the same order as the other terms. For T=O(1)T=O(1) and large NN the Gibbs measure is concentrated around the optimal density, which is obtained as the solution of a self-consistent equation

ρ(x)=Keκ¯T𝑑xρ(x)|xx|aT|x|.\rho(x)=Ke^{-\frac{\bar{\kappa}}{T}\int dx^{\prime}\rho(x^{\prime})|x-x^{\prime}|-\frac{a}{T}|x|}\;. (157)

One can check that the solution (153) satisfies this self-consistent equation. Indeed, for x>0x>0 one has

𝑑xρ(x)|xx|=0𝑑x(r(x)+12)+0x𝑑x(r(x)+12)x+𝑑x(r(x)12)\displaystyle\int dx^{\prime}\rho(x^{\prime})|x-x^{\prime}|=\int_{-\infty}^{0}dx^{\prime}(r(x^{\prime})+\frac{1}{2})+\int_{0}^{x}dx^{\prime}(r(x^{\prime})+\frac{1}{2})-\int_{x}^{+\infty}dx^{\prime}(r(x^{\prime})-\frac{1}{2}) (158)
=(κ¯+2a)xκ¯+2Tκ¯log(κ¯κ¯+2a+e(κ¯+a)xT)\displaystyle=-\frac{(\bar{\kappa}+2a)x}{\bar{\kappa}}+2\frac{T}{\bar{\kappa}}\log(\frac{\bar{\kappa}}{\bar{\kappa}+2a}+e^{\frac{(\bar{\kappa}+a)x}{T}}) (159)

which shows that (157) is indeed obeyed with K=(κ¯+a)2T(κ¯+2a)K=\frac{(\bar{\kappa}+a)^{2}}{T(\bar{\kappa}+2a)}. Note that the equivalence of the DK approach and the Coulomb gas in the passive case was discussed in PLDRankedDiffusion .

If κ¯<a\bar{\kappa}<-a, we obtain instead an expanding phase similar to the active case, but where for x=O(1)x=O(1) the density never has a delta peak at x=0x=0. Thus in the passive case the phase diagram is much simpler and only exhibits 2 phases, corresponding to IsI_{s} and EsE_{s} respectively.

V.4 Moments in the steady states

For the stationary measures obtained here,here one can always define the reciprocal function x(r)x(r) of r(x)r(x), which has a plateau whenever r(x)r(x) has a shock. The moments of |x||x| for k1k\geq 1 can then be expressed as

|x|k=20+𝑑xρs(x)xk=20+1/2𝑑rx(r)k.\langle|x|^{k}\rangle=2\int_{0}^{+\infty}dx\rho_{s}(x)x^{k}=2\int_{0^{+}}^{1/2^{-}}dr\,x(r)^{k}\;. (160)

In the part of the support where the density is smooth one can use the relation x(r)=1γ𝖿a(r)x(r)=\frac{1}{\gamma}{\sf f}_{a}(r) from (88). Hence in the phases IsI_{s} and I0I_{0} one has

|x|k={2γkr(0+)1/2dr𝖿a(r)k,in phasesIsandI02γk0r(xe)dr𝖿a(r)k+2pexek,in phaseFe\langle|x|^{k}\rangle=\begin{cases}\frac{2}{\gamma^{k}}\int_{r(0^{+})}^{1/2}dr\,{\sf f}_{a}(r)^{k}\quad,\quad\text{in phases}\ I_{s}\ \text{and}\ I_{0}\\ \frac{2}{\gamma^{k}}\int_{0}^{r(x_{e}^{-})}dr\,{\sf f}_{a}(r)^{k}+2p_{e}x_{e}^{k}\quad,\quad\text{in phase}\,F_{e}\end{cases} (161)

where we recall that r(xe)=v0aκ¯12r(x_{e}^{-})=\frac{v_{0}-a}{\bar{\kappa}}-\frac{1}{2} and pe=κ¯+av0κ¯p_{e}=\frac{\bar{\kappa}+a-v_{0}}{\bar{\kappa}} in the phase FeF_{e}, as well as r(0+)=av02κr(0^{+})=\frac{a-v_{0}}{2\kappa} in I0I_{0} and 0 in IsI_{s}.

Let us give some explicit formulas in the phase IsI_{s}. One finds

|x|=κ¯2γ+v02(κ¯+a)2κ¯γlog(2(κ¯+a)κ¯+2a).\langle|x|\rangle=\frac{\bar{\kappa}}{2\gamma}+\frac{v_{0}^{2}-(\bar{\kappa}+a)^{2}}{\bar{\kappa}\gamma}\log(\frac{2(\bar{\kappa}+a)}{\bar{\kappa}+2a})\;. (162)

Specializing now to the case a=0a=0 one obtains

|x|=κ¯2γ(1+(v02κ¯21)log4),\langle|x|\rangle=\frac{\bar{\kappa}}{2\gamma}(1+(\frac{v_{0}^{2}}{\bar{\kappa}^{2}}-1)\log 4)\;, (163)

as well as

x2=κ¯212γ2(12v02κ¯2+π2(v02κ¯21)28).\langle x^{2}\rangle=\frac{\bar{\kappa}^{2}}{12\gamma^{2}}\left(12\frac{v_{0}^{2}}{\bar{\kappa}^{2}}+\pi^{2}\left(\frac{v_{0}^{2}}{\bar{\kappa}^{2}}-1\right)^{2}-8\right)\;. (164)

Clearly the moments of any order are simple polynomials in v0v_{0}.

VI More general interaction and self-consistent equation

In this section we consider more general interactions and derive a self-consistent equation for the steady state density.

Consider NN particles in 1D with positions xix_{i} and total energy

E(x)=1Ni<jW(xixj)+iV(xi)E(\vec{x})=\frac{1}{N}\sum_{i<j}W(x_{i}-x_{j})+\sum_{i}V(x_{i}) (165)

where V(x)V(x) is the external potential and where W(x)W(x) is the pairwise interaction energy, the function W(x)W(x) being even. The equation of motion reads (here we assume no passive noise)

dxidt=1NjiW(xixj)V(xi)+v0σi(t).\frac{dx_{i}}{dt}=-\frac{1}{N}\sum_{j\neq i}W^{\prime}(x_{i}-x_{j})-V^{\prime}(x_{i})+v_{0}\sigma_{i}(t)\;. (166)

Here W(x)W^{\prime}(x) is thus an odd function. Hence either one has W(0)=0W^{\prime}(0)=0 or W(x)W^{\prime}(x) diverges at x=0x=0. In the second case, for divergent repulsive interactions with W(0+)=+W^{\prime}(0^{+})=+\infty, the particles cannot cross and we have found in TouzoDBM2023 that the Dean-Kawasaki (DK) method is problematic in the active case. Hence we will restrict here to the case W(0)=0W^{\prime}(0)=0 where the particles can cross freely. In that case, the large NN limit can be taken more straightforwardly, and the method should work, as confirmed by our numerics in the particular case of the rank diffusion, i.e. W(x)=sgn(x)W^{\prime}(x)={\rm sgn}(x).

Then, the DK method gives, in the limit of large NN, the evolution equation for the densities

tρσ(x,t)=x[ρσ(x,t)(v0σ+V(x)+𝑑yW(xy)(ρ+(y,t)+ρ(y,t)))]+γρσ(x,t)γρσ(x,t).\partial_{t}\rho_{\sigma}(x,t)=\partial_{x}\left[\rho_{\sigma}(x,t)\left(-v_{0}\sigma+V^{\prime}(x)+\int dyW^{\prime}(x-y)(\rho_{+}(y,t)+\rho_{-}(y,t))\right)\right]+\gamma\rho_{-\sigma}(x,t)-\gamma\rho_{\sigma}(x,t)\;. (167)

In terms of the densities ρs\rho_{s} and ρd\rho_{d} this leads to

tρs=x[(v0ρdF~(x,t)ρs]\displaystyle\partial_{t}\rho_{s}=\partial_{x}[(-v_{0}\rho_{d}-\tilde{F}(x,t)\rho_{s}] (168)
tρd=x[(v0ρsF~(x,t)ρd]2γρd\displaystyle\partial_{t}\rho_{d}=\partial_{x}[(-v_{0}\rho_{s}-\tilde{F}(x,t)\rho_{d}]-2\gamma\rho_{d}

where

F~(x,t)=V(x)𝑑yW(xy)ρs(y,t).\tilde{F}(x,t)=-V^{\prime}(x)-\int dyW^{\prime}(x-y)\rho_{s}(y,t)\;. (169)

Note that the equations (168) are identical to the equations of a single RTP, N=1N=1, in a effective time dependent force field F~(x,t)\tilde{F}(x,t) which depends itself on the time dependent density. One can thus use the known results for that simpler problem, whenever these are available.

Consider now the stationary measure (which we assume here to exist). The steady state densities must be solution of

0=x[(v0ρdF~(x)ρs]\displaystyle 0=\partial_{x}[(-v_{0}\rho_{d}-\tilde{F}(x)\rho_{s}] (170)
0=x[(v0ρsF~(x)ρd]2γρd\displaystyle 0=\partial_{x}[(-v_{0}\rho_{s}-\tilde{F}(x)\rho_{d}]-2\gamma\rho_{d}

where the effective force field is now static

F~(x)=V(x)𝑑yW(xy)ρs(y)\tilde{F}(x)=-V^{\prime}(x)-\int dy\,W^{\prime}(x-y)\rho_{s}(y) (171)

and depends on the steady state density.

These equations can be solved formally since the stationary measure for a single RTP in an arbitrary force field is known. Let us recall the main steps. One can integrate the first equation on the real line, leading to

ρd(x)=F~(x)v0ρs(x)+Cd\rho_{d}(x)=-\frac{\tilde{F}(x)}{v_{0}}\rho_{s}(x)+C_{d} (172)

where CdC_{d} is an integration constant. Assuming now that F~(x)ρs(x)\tilde{F}(x)\rho_{s}(x) as well as ρd(x)\rho_{d}(x) vanish at infinity leads to Cd=0C_{d}=0. Inserting in the first equation we obtain

2γF~(x)ρs(x)=x((v02F~(x)2)ρs(x)).2\gamma\tilde{F}(x)\rho_{s}(x)=\partial_{x}((v_{0}^{2}-\tilde{F}(x)^{2})\rho_{s}(x))\;. (173)

Denoting ρs(x)=f(x)/(v02F~(x)2)\rho_{s}(x)=f(x)/(v_{0}^{2}-\tilde{F}(x)^{2}), one has f(x)=2γF~(x)/(v02F~(x)2)f(x)f^{\prime}(x)=2\gamma\tilde{F}(x)/(v_{0}^{2}-\tilde{F}(x)^{2})f(x). This equation can be integrated. There are then several cases depending on how many roots the equation F~(x)=v02\tilde{F}(x)=v_{0}^{2}. If we assume here for simplicity the case of a connected support of the density (a single interval which can be infinite) one obtains

ρs(x)=Kv02F~(x)2e2γ0x𝑑zF~(z)v02F~(z)2,F~(z)=F(z)dyW(zy)ρs(y)\rho_{s}(x)=\frac{K}{v_{0}^{2}-\tilde{F}(x)^{2}}e^{2\gamma\int_{0}^{x}dz\frac{\tilde{F}(z)}{v_{0}^{2}-\tilde{F}(z)^{2}}}\quad,\quad\tilde{F}(z)=F(z)-\int dy\,W^{\prime}(z-y)\rho_{s}(y) (174)

where KK is a constant determined by normalization and we have assumed that x=0x=0 is within the support of the density. The equation (174) must be understood as a self-consistent equation for the steady state density ρs(x)\rho_{s}(x). Once ρs(x)\rho_{s}(x) is known, ρd(x)\rho_{d}(x) is obtained from (172). We defer its general study to future work and mention here two simple applications.

Recovering active rank diffusion. Let us show how one recovers the solution for the active rank diffusion problem obtained in our paper. In that case one has W(x)=κ¯|x|W(x)=\bar{\kappa}|x|, W(x)=κ¯sgn(x)W^{\prime}(x)=\bar{\kappa}\,{\rm sgn}(x), with sgn(0)=0{\rm sgn}(0)=0, and κ¯𝑑yρs(y)sgn(xy)=2κ¯r(x)\bar{\kappa}\int dy\,\rho_{s}(y){\rm sgn}(x-y)=2\bar{\kappa}r(x). Hence the effective force field is F~(x)=2κ¯r(x)\tilde{F}(x)=-2\bar{\kappa}r(x). Considering here for simplicity only the case of the phase IsI_{s} where the support of the densities is infinite, the self-consistent equation (174) then reads

r(x)=Kv024κ¯2r(x)2e4γκ¯0x𝑑zr(z)v024κ¯2r(z)2.r^{\prime}(x)=\frac{K}{v_{0}^{2}-4\bar{\kappa}^{2}r(x)^{2}}e^{-4\gamma\bar{\kappa}\int_{0}^{x}dz\frac{r(z)}{v_{0}^{2}-4\bar{\kappa}^{2}r(z)^{2}}}\;. (175)

It is not so obvious to solve that equation directly. One can guess that there exists a function g(r)g(r) such that

4γκ¯0x𝑑zr(z)v024κ¯2r(z)2=g(r(x)).4\gamma\bar{\kappa}\int_{0}^{x}dz\frac{r(z)}{v_{0}^{2}-4\bar{\kappa}^{2}r(z)^{2}}=g(r(x))\;. (176)

Note that since r(x)r^{\prime}(x) must vanish at |x|=+|x|=+\infty, the function g(r)g(r) must diverge to ++\infty at r=±1/2r=\pm 1/2. Taking a derivative w.r.t. xx one obtains

g(r(x))r(x)=4γκ¯r(x)v024κ¯2r(x)2g^{\prime}(r(x))r^{\prime}(x)=4\gamma\bar{\kappa}\frac{r(x)}{v_{0}^{2}-4\bar{\kappa}^{2}r(x)^{2}} (177)

and using (175) one can close the equation and obtain an equation for g(r)g(r)

g(r)eg(r)=4γκ¯Kr.g^{\prime}(r)e^{-g(r)}=\frac{4\gamma\bar{\kappa}}{K}r\;. (178)

Hence

eg=2γκ¯K(14r2)e^{-g}=\frac{2\gamma\bar{\kappa}}{K}(\frac{1}{4}-r^{2}) (179)

leading, using again (175) to

r(x)=2γκ¯v024κ¯2r(x)2(14r2)r^{\prime}(x)=\frac{2\gamma\bar{\kappa}}{v_{0}^{2}-4\bar{\kappa}^{2}r(x)^{2}}(\frac{1}{4}-r^{2}) (180)

which is identical to the equation (16) of the text.

Harmonic interaction. In the case of an harmonic interaction W(x)=μ2x2W(x)=\frac{\mu}{2}x^{2} (μ>0\mu>0), we have

F~(z)=μ𝑑y(zy)ρs(y).\tilde{F}(z)=-\mu\int dy(z-y)\rho_{s}(y)\;. (181)

If we assume that ρs(x)\rho_{s}(x) is an even function, this becomes

F~(z)=μz.\tilde{F}(z)=-\mu z\;. (182)

Therefore the stationary density of particles at large NN is exactly the same as for independent particles in a harmonic potential of strength μ\mu centered at x=0x=0 (see e.g. DKM19 )

ρs(x)=K~(1(μxv0)2)γμ1,K~=24γ/μB(γ/μ,γ/μ)μv0,|x|<v0/μ.\rho_{s}(x)=\tilde{K}\left(1-\left(\frac{\mu x}{v_{0}}\right)^{2}\right)^{\frac{\gamma}{\mu}-1}\quad,\quad\tilde{K}=\frac{2}{4^{\gamma/\mu}B(\gamma/\mu,\gamma/\mu)}\frac{\mu}{v_{0}}\quad,\quad|x|<v_{0}/\mu\;. (183)

This was to be expected since the sum of the forces of two particles at ±x0\pm x_{0} on a third particle at xx is an attractive harmonic force towards x=0x=0. More precisely the equation of motion reads for any finite NN

x˙i=μNj(xjxi)+v0σi=μxi+μx¯+v0σi,x¯=1Njxj\dot{x}_{i}=\frac{\mu}{N}\sum_{j}(x_{j}-x_{i})+v_{0}\sigma_{i}=-\mu x_{i}+\mu\bar{x}+v_{0}\sigma_{i}\quad,\quad\bar{x}=\frac{1}{N}\sum_{j}x_{j} (184)

and in the large NN limit the position of the center of mass x¯\bar{x} tends to zero, leading to an effective decoupling.

For finite NN, it is more convenient to go to the reference frame of the center of mass (which, as in the rank interaction case follows a simple diffusion with DN=1N(T+v022γ)D_{N}=\frac{1}{N}(T+\frac{v_{0}^{2}}{2\gamma}) and define x~i=xix¯\tilde{x}_{i}=x_{i}-\bar{x}. Using that ddtx¯=v0Niσi(t)\frac{d}{dt}\bar{x}=\frac{v_{0}}{N}\sum_{i}\sigma_{i}(t) it satisfies

dx~idt=dxidtdx¯dt=μx~i+(11N)v0σi(t)v0Nj(i)σj(t).\frac{d\tilde{x}_{i}}{dt}=\frac{dx_{i}}{dt}-\frac{d\bar{x}}{dt}=-\mu\tilde{x}_{i}+\big{(}1-\frac{1}{N}\big{)}v_{0}\sigma_{i}(t)-\frac{v_{0}}{N}\sum_{j(\neq i)}\sigma_{j}(t)\;. (185)

For any value of NN, the equation (185) can be solved explicitly, and from it one can compute the first and second moments. One finds x~i(t)=x~i(0)eμt\langle\tilde{x}_{i}(t)\rangle=\tilde{x}_{i}(0)e^{-\mu t} and

x~i(t)2x~i(t)2\displaystyle\langle\tilde{x}_{i}(t)^{2}\rangle-\langle\tilde{x}_{i}(t)\rangle^{2} =\displaystyle= (11N)v02(12γμ+μ2+2e(μ+2γ)t4γ2μ2+e2μtμ(μ2γ))t+(11N)v02μ(2γ+μ)forμ2γ\displaystyle\big{(}1-\frac{1}{N}\big{)}v_{0}^{2}\left(\frac{1}{2\gamma\mu+\mu^{2}}+\frac{2e^{-(\mu+2\gamma)t}}{4\gamma^{2}-\mu^{2}}+\frac{e^{-2\mu t}}{\mu(\mu-2\gamma)}\right)\quad\xrightarrow[t\to+\infty]{}(1-\frac{1}{N})\frac{v_{0}^{2}}{\mu(2\gamma+\mu)}\quad{\rm for}\,\mu\neq 2\gamma (186)
=\displaystyle= (11N)v028γ2(1e4γt4γte4γt)t+(11N)v028γ2forμ=2γ\displaystyle\big{(}1-\frac{1}{N}\big{)}\frac{v_{0}^{2}}{8\gamma^{2}}\left(1-e^{-4\gamma t}-4\gamma te^{-4\gamma t}\right)\xrightarrow[t\to+\infty]{}(1-\frac{1}{N})\frac{v_{0}^{2}}{8\gamma^{2}}\quad{\rm for}\,\mu=2\gamma

where the average includes an average over the initial σi(0)=±1\sigma_{i}(0)=\pm 1, and we have used that σi(t)σj(t)=e2γ|tt|δij\langle\sigma_{i}(t)\sigma_{j}(t^{\prime})\rangle=e^{-2\gamma|t-t^{\prime}|}\delta_{ij}. Hence, apart from the factor 11N1-\frac{1}{N}, it is identical to the moment for a single RTP in a quadratic well DKM19 .