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

\newconstantfamily

csymbol=c \dottedcontentssection[4em]2.9em0.7pc \dottedcontentssubsection[0em]3.3em1pc

SHARP THRESHOLD FOR THE BALLISTICITY OF THE RANDOM WALK ON THE EXCLUSION PROCESS

Abstract

We study a non-reversible random walk advected by the symmetric simple exclusion process, so that the walk has a local drift of opposite sign when sitting atop an occupied or an empty site. We prove that the back-tracking probability of the walk exhibits a sharp transition as the density ρ\rho of particles in the underlying exclusion process varies across a critical density ρc\rho_{c}. Our results imply that the speed v=v(ρ)v=v(\rho) of the walk is a strictly monotone function and that the zero-speed regime is either absent or collapses to a single point, ρc\rho_{c}, thus solving a conjecture of [40]. The proof proceeds by exhibiting a quantitative monotonicity result for the speed of a truncated model, in which the environment is renewed after a finite time horizon LL. The truncation parameter LL is subsequently pitted against the density ρ\rho to carry estimates over to the full model. Our strategy is somewhat reminiscent of certain techniques recently used to prove sharpness results in percolation problems. A key instrument is a combination of renormalisation arguments with refined couplings of environments at slightly different densities, which we develop in this article. Our results hold in fact in greater generality and apply to a class of environments with possibly egregious features, outside perturbative regimes.

Guillaume Conchon-Kerjan1, Daniel Kious2 and Pierre-François Rodriguez3


 

September 2024

1King’s College London

Department of Mathematics

London WC2R 2LS

United Kingdom

guillaume.conchon-kerjan@kcl.ac.uk


2University of Bath

Department of Mathematical Sciences

Bath BA2 7AY

United Kingdom

daniel.kious@bath.ac.uk


3Imperial College London

Department of Mathematics

London SW7 2AZ

United Kingdom

p.rodriguez@imperial.ac.uk

1 Introduction

Transport in random media has been an active field of research for over fifty years but many basic questions remain mathematically very challenging unless the medium satisfies specific (and often rather restrictive) structural assumptions; see for instance [48, 72, 77] and references therein. In this article we consider the problem of ballistic behavior in a benchmark setting that is both i) non-reversible, and ii) non-perturbative (in the parameters of the model). To get a sense of the difficulty these combined features entail, general results in case i) are already hard to come by in perturbative regimes, see, e.g., [24, 75] regarding the problem of diffusive behavior.

Our focus is on a certain random walk in dynamic random environment, which lies outside of the well-studied ‘classes’ and has attracted increasing attention in the last decade, see for instance [40, 43, 44] and references below. The interest in this model stems in no small part from the nature of the environment, which is driven by a particle system (e.g. the exclusion process) that is typically conservative and exhibits slow mixing. Even in the (1+1)-dimensional case, unless one restricts its parameters to perturbative regimes, the features of the model preclude the use of virtually all classical techniques: owing to the dynamics of the environment, the model is genuinely non-reversible, and its properties make the search for an invariant measure of the environment as seen by the walker (in the spirit of [49, 62, 47]) inaccessible by current methods; see Section 1.2 for a more thorough discussion of these and related matters.

The model we study depends on a parameter ρ\rho governing the density of particles in the environment, which in turn affects the walk, whose transition probabilities depend on whether the walk sits on top of a particle (advection occurs) or not. This may (or not) induce ballistic behaviour. Our aim in this article is to prove that ballisticity is a property of the walk undergoing a ‘sharp transition’ as ρ\rho varies, which is an inherently non-perturbative result. Our results answer a number of conjectures of the past fifteen years, and stand in stark contrast with (i.i.d.) static environments [46, 69], where the zero-speed regime is typically extended. The sharpness terminology is borrowed from critical phenomena, and the analogy runs deep, as will become apparent. Drawing inspiration from it is one of the cornerstones of the present work.

1.1. Main results

We present our main results with minimal formalism in the model case where the environment is the (simple) symmetric exclusion process, abbreviated (S)SEP in the sequel, and refer to Sections 2 and 6 for full definitions. We will in fact prove more general versions of these results, see Theorem 3.1, which hold under rather broad assumptions on the environment, satisfied for instance by the SEP. Another environment of interest that fulfils these conditions is discussed in Appendix B, and is built using a Poisson system of particles performing independent simple random walks (PCRW, short for Poisson Cloud of Random Walks).

The SEP is the continuous-time Markov process η=(ηt)t0\eta=(\eta_{t})_{t\geq 0} taking values in {0,1}\{0,1\}^{\mathbb{Z}} and describing the simultaneous evolution of continuous-time simple random walks subject to the exclusion rule. That is, for a given realization of η0\eta_{0}, one puts a particle at those sites xx such that η0(x)=1\eta_{0}(x)=1. Each particle then attempts to move at a given rate ν>0\nu>0, independently of its previous moves and of the other particles, to one of its two neighbours chosen with equal probability. The move happens if and only if the target site is currently unoccupied; see Section 6.1 for precise definitions. For the purposes of this article, one could simply set ν=1\nu=1. We have kept the dependence on ν\nu explicit in anticipation of future applications, for which one may wish to slow down/speed up the environment over time.

On top of the SEP η\eta, a random walk X=(Xn)n0X=(X_{n})_{n\geq 0} starting from X0=0X_{0}=0 moves randomly as follows. Fixing two parameters p,p(0,1)p_{\bullet},p_{\circ}\in(0,1) such that p>pp_{\bullet}>p_{\circ}, the process XX moves at integer times to a neighbouring site, the right site being chosen with probability pp_{\bullet}, resp. pp_{\circ}, depending on whether XX is currently located on an occupied or empty site. Formally, given a realization η\eta of the SEP, and for all n1n\geq 1,

(1.1) Xn+1Xn={+1, with prob. p𝟏ηn(Xn)=1+p𝟏ηn(Xn)=01, with prob. (1p)𝟏ηn(Xn)=1+(1p)𝟏ηn(Xn)=1X_{n+1}-X_{n}=\begin{cases}+1,&\text{ with prob.~$p_{\bullet}\mathbf{1}_{\eta_{n}(X_{n})=1}+p_{\circ}\mathbf{1}_{\eta_{n}(X_{n})=0}$}\\ -1,&\text{ with prob.~$(1-p_{\bullet})\mathbf{1}_{\eta_{n}(X_{n})=1}+(1-p_{\circ})\mathbf{1}_{\eta_{n}(X_{n})=1}$}\end{cases}

(notice that the transition probabilities depend on η\eta since ppp_{\circ}\neq p_{\bullet} by assumption). For ρ(0,1)\rho\in(0,1) we denote by ρ\mathbb{P}^{\rho} the annealed (i.e. averaged over all sources of randomness; see Section 2.3 for details) law of (η,X)(\eta,X) whereby η0\eta_{0} is sampled according to a product Bernoulli measure ((1ρ)δ0+ρδ1)((1-\rho)\delta_{0}+\rho\delta_{1})^{\otimes\mathbb{Z}}, with δx\delta_{x} a Dirac measure at xx, which is an invariant measure for the SEP, see Lemma 6.1. Incidentally, one can check that except for the trivial cases ρ{0,1}\rho\in\{0,1\}, none of the invariant measures for the SEP (cf. [51, Chap. III, Cor. 1.11]) is invariant for the environment viewed from the walker (to see this, one can for instance consider the probability that the origin and one of its neighbours are both occupied). In writing ρ\mathbb{P}^{\rho}, we leave the dependence on ν,p\nu,p_{\bullet} and pp_{\circ} implicit, which are regarded as fixed, and we will focus on the dependence of quantities on ρ\rho, the main parameter of interest. For a mental picture, the reader is invited to think of XX as evolving in a half-plane, with space being horizontal and time running upwards. All figures below will follow this convention.

We now describe our main results, using a language that will make analogies to critical phenomena apparent. Our first main result is simplest to formulate in terms of the fast-tracking probability, defined for n0n\geq 0 and ρ(0,1)\rho\in(0,1) as

(1.2) θn(ρ)=ρ(Hn<H1),\theta_{n}(\rho)=\mathbb{P}^{\rho}(H_{n}<H_{-1}),

where Hk=inf{n0:Xn=k}H_{k}=\inf\{n\geq 0:X_{n}=k\}. In words θn(ρ)\theta_{n}(\rho) is the probability that XX visits nn before 1-1. Analogous results can be proved for a corresponding back-tracking probability, involving the event {Hn<H1}\{H_{-n}<H_{1}\} instead; see the end of Section 1.1 for more on this. The events in (1.2) being decreasing in nn, the following limit

(1.3) θ(ρ)=def.limnθn(ρ)\theta(\rho)\stackrel{{\scriptstyle\text{def.}}}{{=}}\lim_{n}\theta_{n}(\rho)

is well-defined and constitutes an order parameter (in the parlance of statistical physics) for the model. Indeed, it is not difficult to see from (1.1) and monotonicity properties of the environment (see (P.3) in Section 2.2 and Lemma 2.2), that the functions θn()\theta_{n}(\cdot), n0n\geq 0, and hence θ()\theta(\cdot) are non-decreasing, i.e. that θ(ρ)θ(ρ)\theta(\rho)\leq\theta(\rho^{\prime}) whenever ρρ\rho\leq\rho^{\prime}. One thus naturally associates to the function θ()\theta(\cdot) a corresponding critical threshold

(1.4) ρc=def.sup{ρ[0,1]:θ(ρ)=0}.\rho_{c}\stackrel{{\scriptstyle\text{def.}}}{{=}}\sup\{\rho\in[0,1]:\theta(\rho)=0\}.

The transition from the subcritical (ρ<ρc\rho<\rho_{c}) to the supercritical (ρ>ρc\rho>\rho_{c}) regime corresponds to the onset of a phase where the walk has a positive chance to escape to the right. In fact it will do so at linear speed, as will follow from our second main result. To begin with, in view of (1.4) one naturally aims to quantify the behavior of θn(ρ)\theta_{n}(\rho) for large nn. Our first theorem exhibits a sharp transition for its decay.

Theorem 1.1.

For all ρ(0,1)\rho\in(0,1), there exist constants c1,c2(0,)c_{1},c_{2}\in(0,\infty) depending on ρ\rho such that, with ρc\rho_{c} as in (1.4) and for all n1n\geq 1,

  1. (i)

    θn(ρ)c1exp((logn)3/2)\theta_{n}(\rho)\leq c_{1}\exp(-(\log n)^{3/2}), if ρ<ρc\rho<\rho_{c};

  2. (ii)

    θn(ρ)c2\theta_{n}(\rho)\geq c_{2}, if ρ>ρc\rho>\rho_{c}.

The proof of Theorem 1.1 appears at the end of Section 3.2. Since θ=infnθn\theta=\inf_{n}\theta_{n}, item (ii) is an immediate consequence of (1.4). The crux of Theorem 1.1 is therefore to exhibit the rapid decay of item (i) in the full subcritical regime, and not just perturbatively in ρ1\rho\ll 1, which is the status quo, cf. Section 1.2. In the context of critical phenomena, this is the famed question of (subcritical) sharpness, see for instance [53, 1, 36, 76] for sample results of this kind in the context of Ising and Bernoulli percolation models. Plausibly, the true order of decay in item (i) is in fact exponential in nn (for instance, a more intricate renormalisation following [40] should already provide a stretched-exponential bound in nn). We will not delve further into this question in the present article.

Our approach to proving Theorem 1.1 is loosely inspired by the recent sharpness results [31] and [33, 34, 32] concerning percolation of the Gaussian free field and the vacant set of random interlacements, respectively, which both exhibit long-range dependence (somewhat akin to the SEP). Our model is nonetheless very different, and our proof strategy, outlined below in Section 1.3, vastly differs from these works. In particular, it does not rely on differential formulas (in ρ\rho), nor does it involve the OSSS inequality or sharp threshold techniques, which have all proved useful in the context of percolation, see e.g. [35, 38, 17, 23]. One similarity with [33], and to some extent also with [76] (in the present context though, stochastic domination results may well be too much to ask for), is our extensive use of couplings. Developing these couplings represents one of the most challenging technical aspects of our work; we return to this in Section 1.3. It is also interesting to note that, as with statistical physics models, the regime ρρc\rho\approx\rho_{c} near and at the critical density encompasses a host of very natural (and mostly open) questions that seem difficult to answer and point towards interesting phenomena; see Section 1.2 and Corollary 1.3 for more on this.


Our second main result concerns the asymptotic speed of the random walk, and addresses an open problem of [40] which inspired our work. In [40], the authors prove the existence of a deterministic non-decreasing function v:(0,1)v:(0,1)\to\mathbb{R} such that, for all ρ(0,1){ρ,ρ+}\rho\in(0,1)\setminus\{\rho_{-},\rho_{+}\},

(1.5) ρ-a.s., Xnnv(ρ) as n,\mathbb{P}^{\rho}\text{-a.s.,~}\frac{X_{n}}{n}\to v(\rho)\text{ as }n\to\infty,

where

(1.6) ρ=def.sup{ρ:v(ρ)<0},ρ+=def.inf{ρ:v(ρ)>0},\begin{split}&\rho_{-}\stackrel{{\scriptstyle\text{def.}}}{{=}}\sup\{\rho:v(\rho)<0\},\\ &\rho_{+}\stackrel{{\scriptstyle\text{def.}}}{{=}}\inf\{\rho:v(\rho)>0\},\end{split}

leaving open whether ρ+>ρ\rho_{+}>\rho_{-} or not, and with it the possible existence of an extended (critical) zero-speed regime. Our second main result yields the strict monotonicity of v()v(\cdot) and provides the answer to this question. Recall that our results hold for any choice of the parameters 0<p<p<10<p_{\circ}<p_{\bullet}<1 and ν>0\nu>0.

Theorem 1.2.

With ρc\rho_{c} as defined in (1.4), one has

(1.7) ρ=ρ+=ρc.\rho_{-}=\rho_{+}=\rho_{c}.

Moreover, for every ρ,ρ(0,1)\rho,\rho^{\prime}\in(0,1) such that ρ>ρ\rho>\rho^{\prime}, one has

(1.8) v(ρ)>v(ρ).v(\rho)>v(\rho^{\prime}).

The proof of Theorem 1.2 is given in Section 3.2 below. It will follow from a more general result, Theorem 3.1, which applies to a class of environments η\eta satisfying certain natural conditions. These will be shown to hold when η\eta is the SEP.

Let us now briefly relate Theorems 1.1 and 1.2. Loosely speaking, Theorem 1.1 indicates that v(ρ)>0v(\rho)>0 whenever ρ>ρc\rho>\rho_{c}, whence ρ+ρc\rho_{+}\leq\rho_{c} in view of (1.6), which is morally half of Theorem 1.2. One can in fact derive a result akin to Theorem 1.1, but concerning the order parameter θ^(ρ)=limnρ(Hn<H1)\widehat{\theta}(\rho)=\lim_{n}\mathbb{P}^{\rho}(H_{-n}<H_{1}) associated to the back-tracking probability instead. Defining ρ^c\widehat{\rho}_{c} in the same way as (1.4) but with θ^\widehat{\theta} in place of θ\theta, our results imply that ρ^c=ρc\widehat{\rho}_{c}=\rho_{c} and that the transition for the back-tracking probability has similar sharpness features as in Theorem 1.1, but in opposite directions – the sub-critical regime is now for ρ>ρ^c(=ρc)\rho>\widehat{\rho}_{c}(=\rho_{c}). Intuitively, this corresponds to the other inequality ρρc\rho_{-}\geq\rho_{c}, which together with ρ+ρc\rho_{+}\leq\rho_{c}, yields (1.7).

Finally, Theorem 1.2 implies that v(ρ)0v(\rho)\neq 0 whenever ρρc\rho\neq\rho_{c}. Determining whether a law of large number holds at ρ=ρc\rho=\rho_{c} holds or not, let alone whether the limiting speed v(ρc)v(\rho_{c}) vanishes when 0<ρc<10<\rho_{c}<1, is in general a difficult question, to which we hope to return elsewhere; we discuss this and related matters in more detail below in Section 1.2.

One noticeable exception occurs in the presence of additional symmetry, as we now explain. In [40] the authors could prove that, at the ‘self-dual’ point p=1pp_{\bullet}=1-p_{\circ} (for any given value of p(1/2,1)p_{\bullet}\in(1/2,1)), one has v(1/2)=0v(1/2)=0 and the law of large numbers holds with limit speed 0, but they could not prove that vv was non-zero for ρ1/2\rho\neq 1/2, or equivalently that ρ+=ρ=1/2\rho_{+}=\rho_{-}=1/2. Notice that the value p=1pp_{\bullet}=1-p_{\circ} is special, since in this case (cf. (1.1)) XX has the same law under ρ\mathbb{P}^{\rho} as X-X under 1ρ\mathbb{P}^{1-\rho}, and in particular X=lawXX\stackrel{{\scriptstyle\text{law}}}{{=}}-X when ρ=12\rho=\frac{1}{2}. In the result below, which is an easy consequence of Theorem 1.2 and the results of [40], we summarize the situation in the symmetric case. We include the (short) proof here. The following statement is of course reminiscent of a celebrated result of Kesten concerning percolation on the square lattice [45]; see also [23, 17, 58] in related contexts.

Corollary 1.3.

If p=1pp_{\circ}=1-p_{\bullet}, then

(1.9) ρc=12 and θ(12)=0.\rho_{c}=\tfrac{1}{2}\,\text{ and }\,\theta(\tfrac{1}{2})=0.

Moreover, under 1/2\mathbb{P}^{1/2}, XX is recurrent, and the law of large numbers (1.5) holds with vanishing limiting speed v(1/2)=0v(1/2)=0.

Proof.

From the proof of [40, Theorem 2.2] (and display (3.26) therein), one knows that v(1/2)=0v(1/2)=0 and that the law of large numbers holds at ρ=12\rho=\frac{1}{2}. By Theorem 1.2, see (1.6) and (1.7), vv is negative on (0,ρc)(0,\rho_{c}) and positive on (ρc,1)(\rho_{c},1), which implies that ρc=1/2\rho_{c}=1/2. As for the recurrence of XX under 1/2\mathbb{P}^{1/2}, it follows from the ergodic argument of [61] (Corollary 2.2, see also Theorem 3.2 therein; these results are stated for a random walk jumping at exponential times but are easily adapted to our case). Since θ(ρ)ρ(lim infXn0)\theta(\rho)\leq\mathbb{P}^{\rho}(\liminf X_{n}\geq 0) in view of (1.3), recurrence implies that θ(1/2)=0\theta(1/2)=0, and (1.9) follows. ∎

1.2. Discussion

We now place the above results in broader context and contrast our findings with existing results. We then discuss a few open questions in relation with Theorems 1.1 and 1.2.

1.2.1. Related works

Random walks in dynamic random environments have attracted increasing attention in the past two decades, both in statistical physics - as a way to model a particle advected by a fluid [29, 59, 16, 42], and in probability theory - where they provide a counterpart to the more classically studied random walks in static environments [9, 22, 26, 40, 44, 66].

On \mathbb{Z}, the static setting already features a remarkably rich phenomenology, such as transience with zero speed [71], owing to traps that delay the random walk, and anomalous fluctuations [46] - which can be as small as polylogarithmic in the recurrent case [69]. On d\mathbb{Z}^{d} for d2d\geq 2, some fundamental questions are still open in spite of decades of efforts, for instance the conjectures around Sznitman’s Condition (T) and related effective ballisticity criteria [72, 18], and the possibility that, for a uniformly elliptic environment, directional transience is equivalent to ballisticity. If true, these conjectures would have profound structural consequences, in essentially ascertaining that the above trapping phenomena can only be witnessed in dimension one. As with the model studied in this article, a circumstance that seriously hinders progress is the truly non-reversible character of these problems. This severely limits the tools available to tackle them.

New challenges arise when the environment is dynamic, requiring new techniques to handle the fact that correlations between transition probabilities are affected by time. In particular, the trapping mechanisms identified in the static case do not hold anymore and it is difficult to understand if they simply dissolve or if they are replaced by different, possibly more complex, trapping mechanisms. As of today, how the static and dynamic worlds relate is still far from understood. Under some specific conditions however, laws of large numbers (and sometimes central limit theorems) have been proved in dynamic contexts: when the environment has sufficiently good mixing properties [9, 22, 66], a spectral gap [7], or when one can show the existence of an invariant measure for the environment as seen by the walk ([19], again under some specific mixing conditions). One notable instance of such an environment is the supercritical contact process (see [25, 56], or [3] for a recent generalization). We also refer to [5, 57, 20, 6, 27] for recent results in the time-dependent reversible case, for which the assumptions on the environment can be substantially weakened.

The environments we consider are archetypal examples that do not satisfy any of the above conditions. In the model case of the environment driven by the SEP, the mixing time over a closed segment or a circle is super-linear (even slightly super-quadratic [55]), creating a number of difficulties, e.g. barring the option of directly building a renewal structure without having to make further assumptions (see [13] in the simpler non-nestling case). The fact that the systems may be conservative (as is the case for the SEP) further hampers their mixing properties. As such, they have been the subject of much attention in the past decade, see [10, 11, 12, 22, 28, 40, 41, 43, 44, 52] among many others. All of these results are of two types: either they require particular assumptions, or they apply to some perturbative regime of parameters, e.g. high density of particles [21, 26], strong drift, or high/low activity rate of the environment [43, 68].

Recently, building on this series of work, a relatively comprehensive result on the Law of Large Numbers (LLN) in dimension 1+1 has been proved in [40], see (1.5) above, opening an avenue to some fundamental questions that had previously remained out of reach, such as the possible existence of transient regime with zero speed. Indeed, cf. (1.5) and (1.6), the results of [40] leave open the possibility of having an extended ‘critical’ interval of ρ\rho’s for which v(ρ)=0v(\rho)=0, which is now precluded as part of our main results, see (1.7). We seek the opportunity to stress that our strategy, outlined below in Section 1.3, is completely new and rather robust, and we believe a similar approach will lead to progress on related questions for other models.

1.2.2. Open questions


1) Existence and value of v(ρc)v(\rho_{c}). Returning to Theorems 1.1 and 1.2, let us start by mentioning that ρc\rho_{c}, defined by (1.4) and equivalently characterized via (1.6)-(1.7), may in fact be degenerate (i.e. equal to 0 or 11) if v()v(\cdot) stays of constant sign. It is plausible that ρc(0,1)\rho_{c}\in(0,1) if and only if p<1/2<pp_{\circ}<1/2<p_{\bullet}, which corresponds to the (more challenging) nestling case, in which the walk has a drift of opposite signs depending on whether it sits on top of a particle or an empty site. Moreover, the function v()v(\cdot), as given by [40, Theorem 3.4], is well-defined for all values of ρ(0,1)\rho\in(0,1) (this includes a candidate velocity at ρc\rho_{c}), but whether a LLN holds at ρc\rho_{c} with speed v(ρc)v(\rho_{c}) or the value of the latter is unknown in general, except in cases where one knows that v(ρc)=0v(\rho_{c})=0 by other means (e.g. symmetry), in which case a LLN can be proved, see [40, Theorem 3.5]. When ρc\rho_{c} is non-degenerate, given Theorem 1.2, it is natural to expect that v(ρc)=0v(\rho_{c})=0, but this is not obvious as we do not presently know if vv is continuous at ρc\rho_{c}. It is relatively easy to believe that continuity on (0,1){ρc}(0,1)\setminus\{\rho_{c}\}, when the speed is non-zero, can be obtained through an adaptation of the regeneration structure defined in [43], but continuity at ρc\rho_{c} (even in the symmetric case) seems to be more challenging.

Suppose now that one can prove that a law of large numbers holds with vanishing speed at ρ=ρc\rho=\rho_{c}, then one can ask whether XX is recurrent or transient at ρc\rho_{c}. A case in point where one knows the answer is the ‘self-dual’ point p=1pp_{\circ}=1-p_{\bullet} where the critical density equals 1/21/2 and the walk is recurrent. In particular, this result and Theorem 1.2 imply that v(ρ)v(\rho) is zero if and only if ρ=1/2\rho=1/2, and thus there exists no transient regime with zero-speed in the symmetric case. This also answers positively the conjecture in [8] (end of Section 1.4), which states that in the symmetric case, the only density with zero speed is in fact recurrent.

2) Regularity of v()v(\cdot) near ρc\rho_{c} and fluctuations at ρc\rho_{c}. In cases where v(ρc)=0v(\rho_{c})=0 is proved, one may further wonder about the regularity of ρv(ρ)\rho\mapsto v(\rho) around ρc\rho_{c} (and similarly of θ(ρ)\theta(\rho) as ρρc\rho\downarrow\rho_{c}). Is it continuous, and if yes, is it Hölder-continous, or even differentiable? This could be linked to the fluctuations of the random walk when ρ=ρc\rho=\rho_{c}, in the spirit of Einstein’s relation, see for instance [37] in the context of reversible dynamics. Whether the fluctuations of XX under ρc\mathbb{P}^{\rho_{c}} are actually diffusive, super-diffusive or sub-diffusive is a particularly difficult question. It has so far been the subject of various conjectures, both for the RWdRE (e.g. Conjecture 3.5 in [14]) and very closely related models in statistical physics (e.g. [42], [39], [60]). One aspect making predictions especially difficult is that the answer may well depend on all parameters involved, i.e. ρ\rho, pp_{\circ} and pp_{\bullet}. At present, any rigorous upper or lower bound on the fluctuations at ρ=ρc\rho=\rho_{c} would be a significant advance.

3) Comparison with the static setting. In the past decade, there have been many questions as to which features of a static environment (when ν=0\nu=0, so that the particles do not move) are common to the dynamic environment, and which are different. In particular, it is well-known that in the static case, there exists a non-trivial interval of densities for which the random walker has zero speed, due to mesoscopic traps that the random walker has to cross on its way to infinity (see for instance [63, Example 1], and above references). It was conjectured that if ν>0\nu>0 is small enough, there could be such an interval in the dynamic set-up, see [14, (3.8)]. Theorem 1.2 thus disproves this conjecture.

Combined with the CLT from [40, Theorem 2.1], which is valid outside of (ρ,ρ+)(\rho_{-},\rho_{+}), hence for all ρρc\rho\neq\rho_{c} by Theorem 1.2, this also rules out the possibility of non-diffusive fluctuations for any ρρc\rho\neq\rho_{c}, which were conjectured in [14, Section 3.5]. In light of this, one may naturally wonder how much one has to slow down the environment with time (one would have to choose ν=νt\nu=\nu_{t} a suitably decreasing function of the time tt) in order to start seeing effects from the static world. We plan to investigate this in future work.

1.3. Overview of the proof

We give here a relatively thorough overview of the proof, intended to help the reader navigate the upcoming sections. For concreteness, we focus on Theorem 1.2 (see below its statement as to how Theorem 1.1 relates to it), and specifically on the equalities (1.7), in the case of SEP (although our results are more general; see Section 3). The assertion (1.7) is somewhat reminiscent of the chain of equalities u¯=u=u\bar{u}=u_{*}=u_{**} associated to the phase transition of random interlacements that was recently proved in [34, 32, 33], and draws loose inspiration from the interpolation technique of [2], see also [30]. Similarly to [32] in the context of interlacements, and as is seemingly often the case in situations lacking key structural features (in the present case, a notion of reversibility or more generally self-adjointness, cf. [24, 75]), our methods rely extensively on the use of couplings.


The proof essentially consists of two parts, which we detail individually below. In the first part, we compare our model with a finite-range version, in which we fully re-sample the environment at times multiple of some large integer parameter LL (see e.g. [32] for a similar truncation to tame the long-range dependence). One obtains easily that for any LL and any density ρ(0,1)\rho\in(0,1), the random walk in this environment satisfies a strong Law of Large Numbers with some speed vL(ρ)v_{L}(\rho) (Lemma 3.2), and we show that in fact,

(1.10) vL(ρεL)δLv(ρ)vL(ρ+εL)+δL,v_{L}(\rho-\varepsilon_{L})-\delta_{L}\leq v(\rho)\leq v_{L}(\rho+\varepsilon_{L})+\delta_{L},

for some δL,εL\delta_{L},\varepsilon_{L} that are quantitative and satisfy δL,εL=oL(1)\delta_{L},\varepsilon_{L}=o_{L}(1) as LL\to\infty (see Proposition 3.3).

In the second part, we show that for any fixed ρ,ρ+ε(0,1)\rho,\rho+\varepsilon\in(0,1), we have

(1.11) vL(ρ+ε)>vL(ρ)+3δL,v_{L}(\rho+\varepsilon)>v_{L}(\rho)+3\delta_{L},

for LL large enough (Proposition 3.4). Together, (1.10) and (1.11) readily imply that v(ρ+ε)>v(ρ)v(\rho+\varepsilon)>v(\rho), and (1.7) follows. Let us now give more details.

First part: from infinite to finite range. For a density ρ(0,1)\rho\in(0,1) the environment η\eta of the range-LL model is the SEP starting from η0Ber(ρ)\eta_{0}\sim\text{Ber}(\rho)^{\otimes\mathbb{Z}} during the time interval [0,L)[0,L). Then for every integer k1k\geq 1, at time kLkL we sample ηkL\eta_{kL} again as Ber(ρ)\text{Ber}(\rho)^{\otimes\mathbb{Z}}, independently from the past, and let it evolve as the SEP during the interval [kL,(k+1)L)[kL,(k+1)L). The random walk XX on this environment is still defined formally as in (1.1), and we denote ρ,L\mathbb{P}^{\rho,L} the associated annealed probability measure. Clearly, the increments (XkLX(k1)L)k1(X_{kL}-X_{(k-1)L})_{k\geq 1} are i.i.d. (and bounded), so that by the strong LLN, Xn/nvL(ρ):=𝔼ρ[XL/L]X_{n}/n\rightarrow v_{L}(\rho):=\mathbb{E}^{\rho}[X_{L}/L], a.s. as nn\rightarrow\infty.

Now, to relate this to our original ’infinite-range’ (L=L=\infty) model, the main idea is to compare the range-LL with the range-2L2L, and more generally the range-2kL2^{k}L with the range-2k+1L2^{k+1}L model for all k0k\geq 0 via successive couplings. The key is to prove a chain of inequalities of the kind

(1.12) vL(ρ)v2L(ρ+ε1,L)+δ1,Lv2kL(ρ+εk,L)+δk,L,v_{L}(\rho)\leq v_{2L}(\rho+\varepsilon_{1,L})+\delta_{1,L}\leq\ldots\leq v_{2^{k}L}(\rho+\varepsilon_{k,L})+\delta_{k,L}\leq\ldots,

where the sequences (εk,L)k1(\varepsilon_{k,L})_{k\geq 1} of and (δk,L)k1(\delta_{k,L})_{k\geq 1} are increasing, with εL:=limkεk,L=oL(1)\varepsilon_{L}:=\lim_{k\rightarrow\infty}\varepsilon_{k,L}=o_{L}(1) and δL:=limkδk,L=oL(1)\delta_{L}:=\lim_{k\rightarrow\infty}\delta_{k,L}=o_{L}(1). In other words, we manage to pass from a scale 2kL2^{k}L to a larger scale 2k+1L2^{k+1}L, at the expense of losing a bit of speed (δk+1,Lδk,L)(\delta_{k+1,L}-\delta_{k,L}), and using a little sprinkling in the density (εk,Lεk1,L)(\varepsilon_{k,L}-\varepsilon_{k-1,L}), with δ0,L=ε0,L=0\delta_{0,L}=\varepsilon_{0,L}=0. From (1.12) (and a converse inequality proved in a similar way), standard arguments allow us to deduce (1.10). Let us mention that such a renormalization scheme, which trades scaling against sprinkling in one or several parameters, is by now a standard tool in percolation theory, see for instance [73, 67, 30].

We now describe how we couple the range-LL and the range-2L2L models in order to obtain the first inequality, the other couplings being identical up to a scaling factor. We do this in detail in Lemma 4.1, where for technical purposes, we need in fact more refined estimates than only the first moment, but they follow from this same coupling. The coupling works roughly as follows (cf. Fig. 1). We consider two random walks X(1)=lawρ,LX^{(1)}\stackrel{{\scriptstyle\text{law}}}{{=}}\mathbb{P}^{\rho,L} and X(2)=lawρ+εL,2LX^{(2)}\stackrel{{\scriptstyle\text{law}}}{{=}}\mathbb{P}^{\rho+\varepsilon_{L},2L} coupled via their respective environments η(1)\eta^{(1)} and η(2)\eta^{(2)} such that we retain good control (in a sense explained below) on the relative positions of their respective particles during the time interval [0,2L][0,2L], except during a short time interval after time LL after renewing the particles of η(1)\eta^{(1)}. In some sense, we want to dominate η(1)\eta^{(1)} by η(2)\eta^{(2)} ‘as much as possible’, and use it in combination with the following monotonicity property of the walk: since p>pp_{\bullet}>p_{\circ}, X(2)X^{(2)} cannot be overtaken by X(1)X^{(1)} if η(2)\eta^{(2)} covers η(1)\eta^{(1)} (cf. (1.1); see also Lemma 2.2 for a precise formulation). Roughly speaking, we will use this strategy from time 0 to LL, then lose control for short time before recovering a domination and using the same argument but with a lateral shift (in space), as illustrated in Figure 1. The key for recovering a domination in as little time as possible is a property of the following flavour.

(1.13) { Let L,t,1 be such that Lt24, and let η0,η0{0,1}Z be such that over [-3L,3L], η0 (resp. η0) has empirical density +ρε (resp. ρ) on segments of length . Then one can couple the time evolutions of η,η as SEPs during [0,t] such that P(ηt(x)ηt(x),x[+-Lct,-Lct])-1ctLexp(-c′′ε2t/14), for some constants >c,c,c′′0. \begin{cases}&\text{\begin{minipage}{364.00139pt} Let $L,t,\ell\geq 1$ be such that $L\gg t^{2}\gg\ell^{4}$, and let $\eta_{0},\eta^{\prime}_{0}\in\{0,1\}^{\mathbb{Z}}$ be such that over $[-3L,3L]$, $\eta_{0}$ (resp.~$\eta^{\prime}_{0}$) has empirical density $\rho+\varepsilon$ (resp.~$\rho$) on segments of length $\ell$. Then one can couple the time evolutions of $\eta,\eta^{\prime}$ as SEPs during $[0,t]$ such that $$\mathbb{P}\big{(}\eta_{t}(x)\geq\eta_{t}^{\prime}(x),\,\forall x\in[-L+ct,L-ct]\big{)}\geq 1-c^{\prime}tL\exp(-c^{\prime\prime}\varepsilon^{2}t^{1/4}),$$ for some constants $c,c^{\prime},c^{\prime\prime}>0$. \end{minipage}}\end{cases}

Roughly, for tt poly-logarithmic in LL the right-hand side of (1.13) will be as close to 1 as we need. We do not give precise requirements on the scales ,t,L\ell,t,L (nor a definition of empirical density), except that they must satisfy some minimal power ratio due to the diffusivity of the particles. We will formalize this statement in Section 3.1, see in particular condition (C.2.2), with exponents and constants that are likely not optimal but sufficient for our purposes. Even if the environment were to mix in super-quadratic (but still polynomial) time, our methods would still apply, up to changing the exponents in our conditions (C.1)-(C.3).

The only statistic we control is the empirical density, i.e. the number of particles over intervals of a given length. The relevant control is formalised in Condition (C.1). In particular, our environment couplings have to hold with a quenched initial data, and previous annealed couplings in the literature (see e.g. [15]) are not sufficient for this purpose (even after applying the usual annealed-to-quenched tricks), essentially due to the difficulty of controlling the environment around the walker in the second part of the proof, as we explain in Remark 5.5; cf. also [64, 4, 65] for related issues in other contexts.

Refer to caption
Figure 1: Coupling of X(1)=lawρ,LX^{(1)}\stackrel{{\scriptstyle\text{law}}}{{=}}\mathbb{P}^{\rho,L} and X(2)=lawρ+εL,2LX^{(2)}\stackrel{{\scriptstyle\text{law}}}{{=}}\mathbb{P}^{\rho+\varepsilon_{L},2L}. The particles of η(1)\eta^{(1)} and their trajectories are pictured in black, those of η(2)\eta^{(2)} in red. In blue, from time 0 to LL, is the trajectory of X(1)X^{(1)}, which is a lower bound for that of X(2)X^{(2)}. From time LL to 2L2L, we split it between the rightmost (resp. leftmost) possible realization of X(1)X^{(1)} (resp. X(2)X^{(2)}). These correspond to worst-case scenarios. Owing to the efficiency of the coupling (grey band), the discrepancy incurred at time 2L2L remains controlled with high probability, but there is a barrier to how small this gap can be made, which is super-linear in logL\log L.

Let us now track the coupling a bit more precisely to witness the quantitative speed loss δL\delta_{L} incurred by (1.13). We start by sampling η0(1)\eta^{(1)}_{0} and η0(2)\eta^{(2)}_{0} such that a.s., η0(1)(x)η0(2)(x)\eta^{(1)}_{0}(x)\leq\eta^{(2)}_{0}(x) for all xx\in\mathbb{Z}, which is possible by stochastic domination. In plain terms, wherever there is a particle of η(1)\eta^{(1)}, there is one of η(2)\eta^{(2)}. Then, we can let η(1)\eta^{(1)} and η(2)\eta^{(2)} evolve during [0,L][0,L] such that this domination is deterministically preserved over time (this feature is common to numerous particle systems, including both SSEP and the PCRW). As mentioned above, conditionally on such a realization of the environments, one can then couple X(1)X^{(1)} and X(2)X^{(2)} such that Xs(2)Xs(1)X^{(2)}_{s}\geq X^{(1)}_{s} for all s[0,L]s\in[0,L]. The intuition for this is that ’at worst’, X(1)X^{(1)} and X(2)X^{(2)} sit on the same spot, where X(2)X^{(2)} might see a particle and X(1)X^{(1)} an empty site (hence giving a chance for X(2)X^{(2)} to jump to the right, and for X(1)X^{(1)} to the left), but not the other way around (recall to this effect that p>pp_{\bullet}>p_{\circ}).

At time LL, we have to renew the particles of η(1)\eta^{(1)}, hence losing track of the domination of η(1)\eta^{(1)} by η(2)\eta^{(2)}. We intend to recover this domination within a time t:=(logL)100t:=(\log L)^{100} by coupling the particles of ηL(1)\eta^{(1)}_{L} with those of ηL(2)\eta^{(2)}_{L} using (1.13) (at least over a space interval [3L,3L][-3L,3L], that neither X(1)X^{(1)} nor X(2)X^{(2)} can leave during [0,2L][0,2L]). Since during that time, X(1)X^{(1)} could at worst drift of tt steps to the right (and X(2)X^{(2)} make tt steps to the left), we couple de facto the shifted configurations η(1)()\eta^{(1)}(\cdot) with η(2)(2t)\eta^{(2)}(\cdot-2t). Hence, we obtain that ηL+t(1)(x)ηL+t(2)(x2t)\eta^{(1)}_{L+t}(x)\leq\eta^{(2)}_{L+t}(x-2t) for all x[3L+ct,3L+ct]x\in[-3L+ct,3L+ct] with high probability as given by (1.13). As a result, all the particles of ηL+t(1)\eta^{(1)}_{L+t} are covered by particles of ηL+t(2)\eta^{(2)}_{L+t}, when observed from the worst-case positions of XL+t(1)X^{(1)}_{L+t} and XL+t(2),X^{(2)}_{L+t}, respectively.

Finally, during [L+t,2L][L+t,2L] we proceed similarly as we did during [0,L][0,L], coupling η(1)\eta^{(1)} with η(2)(2t)\eta^{(2)}(\cdot-2t) so as to preserve the domination, and two random walks starting respectively from the rightmost possible position for XL+t(1)X^{(1)}_{L+t}, and the leftmost possible one for XL+t(2)X^{(2)}_{L+t}. The gap between X(1)X^{(1)} and X(2)X^{(2)} thus cannot increase, and we get that X2L(1)X2L(2)+2tX^{(1)}_{2L}\leq X^{(2)}_{2L}+2t, except on a set of very small probability where (1.13) fails. Dividing by 2L2L and taking expectations, we get

(1.14) vL(ρ)v2L(ρ+ε)+δ, where δ=O((logL)100L),v_{L}(\rho)\leq v_{2L}(\rho+\varepsilon)+\delta\text{, where }\delta=O\big{(}\tfrac{(\log L)^{100}}{L}\big{)},

for suitable ε>0\varepsilon>0. The exponent 100100 is somewhat arbitrary, but importantly, owing to the coupling time of the two processes (and the diffusivity of the SEP particles), this method could not work with a value of δ\delta smaller than (log2L)/L(\log^{2}L)/L. In particular, the exponent of the logarithm must be larger than one, and this prevents potentially simpler solutions for the second part.

Second part: quantitative speed increase at finite range. We now sketch the proof of (1.11), which is a quantitative estimate on the monotonicity of vLv_{L}, equivalently stating that for ρ(0,1)\rho\in(0,1) and ε(0,1ρ)\varepsilon\in(0,1-\rho),

(1.15) 𝔼ρ+ε[XL]>𝔼ρ[XL]+3LδL,\mathbb{E}^{\rho+\varepsilon}[X_{L}]>\mathbb{E}^{\rho}[X_{L}]+3L\delta_{L},

with δL\delta_{L} explicit and supplied by the first part. This is the most difficult part of the proof, as we have to show that a denser environment actually yields a positive gain for the displacement of the random walk, and not just to limit the loss as in the first part. The main issue when adding an extra density ε\varepsilon of particles is that whenever XX is on top of one such particle, and supposedly makes a step to the right instead of one to the left, it could soon after be on top of an empty site that would drive it to the left, and cancel its previous gain.

We design a strategy to get around this and preserve gaps, illustrated in Figure 2 (cf. also Figure 5 for the full picture, which is more involved). When coupling two walks Xρ=lawρX^{\rho}\stackrel{{\scriptstyle\text{law}}}{{=}}\mathbb{P}^{\rho}, Xρ+ε=lawρ+εX^{\rho+\varepsilon}\stackrel{{\scriptstyle\text{law}}}{{=}}\mathbb{P}^{\rho+\varepsilon} with respective environments ηρ,ηρ+ε\eta^{\rho},\eta^{\rho+\varepsilon}, our strategy is to spot times when i) Xρ+εX^{\rho+\varepsilon} sees an extra particle (that we call sprinkler) and jumps to the right, while XρX^{\rho} sees an empty site and jumps to the left, and ii) this gap is extended with XρX^{\rho} then drifting to the left and Xρ+εX^{\rho+\varepsilon} drifting to the right, for some time tt that must necessarily satisfy tlogLt\ll\log L – so that this has a chance of happening often during the time interval [0,L][0,L].

Once such a gap is created, on a time interval say [s1,s2][s_{1},s_{2}] with s2s1=ts_{2}-s_{1}=t, we attempt to recouple the environments ηρ,ηρ+ε\eta^{\rho},\eta^{\rho+\varepsilon} around the respective positions of their walker, so that the local environment seen from Xρ+εX^{\rho+\varepsilon} again dominates the environment seen from XρX^{\rho}, which then allows us to preserve the gap previously created, up to some time TT which is a polynomial in logL\log L. As we will explain shortly, this gap arises with not too small a probability, so that repeating the same procedure between times iTiT and (i+1)T(i+1)T (for iL/T1i\leq L/T-1) will provide us with the discrepancy 3δL3\delta_{L} needed.

The re-coupling of the environments mentioned in the previous paragraph must happen in time less than O(logL)O(\log L) (in fact less than tt), in order to preserve the gap previously created. This leads to two difficulties:

  • in such a short time, we cannot possibly recouple the two environments over a space interval of size comparable to LL (\approx the space horizon the walk can explore during [0,L][0,L]), and

  • the coupling will have a probability L1\gg L^{-1} to fail, hence with high probability, this will actually happen during [0,L][0,L]! In this case, we lose track of the domination of the environments completely, hence it could even be possible that XρX^{\rho} overtakes Xρ+εX^{\rho+\varepsilon}.

We handle the first of these difficulties with a two-step surgery coupling:

  1. 1)

    (Small coupling). We perform a first coupling of ηρ\eta^{\rho} and ηρ+ε\eta^{\rho+\varepsilon} on an interval of stretched exponential width (think exp(t1/2)\exp(t^{1/2}) for instance) during a time s3s2=t/2s_{3}-s_{2}=t/2 (cf. the orange region in Fig. 2), so as to preserve at least half of the gap created. We call this coupling “small” because tt is small (compared to LL). If that coupling is successful, Xρ+εX^{\rho+\varepsilon} now sees an environment that strictly covers the one seen by XρX^{\rho}, on a spatial interval II of width exp(t1/2)t\asymp\exp(t^{1/2})\gg t. Due to the length of II, this domination extends in time, on an interval [s3,T][s_{3},T] of duration say t200t^{200} (it could even be extended to timescales exp(t1/2)\asymp\exp(t^{1/2})): it could only be broken by SEP particles of ηs3ρ\eta^{\rho}_{s_{3}} lying outside of II, who travel all the distance to meet the two walkers, but there is a large deviation control on the drift of SEP particles. The space-time zone in which the environment as seen from Xρ+εX^{\rho+\varepsilon} dominates that from XρX^{\rho} is the green trapezoid depicted in Figure 2.

  2. 2)

    (Surgery coupling). When the coupling in 1) is successful, we use the time interval [s3,T][s_{3},T] to recouple the environments ηρ\eta^{\rho} and ηρ+ε\eta^{\rho+\varepsilon} on the two outer sides of the trapezoid, on a width of length 10L10L say, so that ηTρ+ε(+t/2)\eta^{\rho+\varepsilon}_{T}(\cdot+t/2) dominates ηTρ(t/2)\eta^{\rho}_{T}(\cdot-t/2) on the entire width, all the while preserving the fact that ηρ+ε(+t/2)\eta^{\rho+\varepsilon}(\cdot+t/2) dominates ηρ(t/2)\eta^{\rho}(\cdot-t/2) inside the trapezoid at all times (for simplicity and by monotonicity, we can suppose assume that Xs3ρ+εXs3ρX^{\rho+\varepsilon}_{s_{3}}-X^{\rho}_{s_{3}} is exactly equal to tt, as pictured in Figure 2). This step is quite technical, and formulated precisely as Condition (C.2) in Section 3.1. It requires that different couplings can be performed on disjoint contiguous intervals and glued in a ‘coherent’ fashion, whence the name surgery coupling. Again, we will prove this specifically for the SEP in a dedicated section, see Lemma 6.9. Since this second coupling operates on a much longer time interval, we can now ensure its success with overwhelming probability, say 1O(L100)1-O(L^{-100}).

Refer to caption
Figure 2: Coupling of Xρ=lawρX^{\rho}\stackrel{{\scriptstyle\text{law}}}{{=}}\mathbb{P}^{\rho} and Xρ+ε=lawρ+εX^{\rho+\varepsilon}\stackrel{{\scriptstyle\text{law}}}{{=}}\mathbb{P}^{\rho+\varepsilon} during [0,T][0,T]. The black curve(s) represent the minimal gap created between XρX^{\rho} and Xρ+εX^{\rho+\varepsilon} when all couplings succeed. The dashed blue lines represent the worst-case trajectories of the two walks when the first coupling (between s2s_{2} and s3s_{3}) fails. In this case item 3), the parachute coupling (not depicted) comes into effect.

We still need to address the difficulty described in the second bullet point above, which corresponds to the situation where the small coupling described at 1) fails. As argued, will happen a number of times within [0,L] because tt is small. If this occurs, we lose control of the domination of the environments seen by the two walkers. This leads to:

  1. 3)

    (Parachute coupling). If the coupling in 1) is unsuccessful, instead of the surgery coupling, we perform a (parachute) coupling of ηρ\eta^{\rho} and ηρ+ε\eta^{\rho+\varepsilon} in [s3,T][s_{3},T] so that ηρ+ε((Ts3))\eta^{\rho+\varepsilon}(\cdot-(T-s_{3})) covers ηρ(+(Ts3))\eta^{\rho}(\cdot+(T-s_{3})), as we did in the first part of the proof. This again happens with extremely good probability 1O(L100)1-O(L^{-100}). The price to pay is that on this event, the two walks may have drifted linearly in the wrong direction during [s3,T][s_{3},T] (see the dashed blue trajectories in Figure 2), but we ensure at least that XTρ+εXTρ2(Ts3)X^{\rho+\varepsilon}_{T}-X^{\rho}_{T}\geq-2(T-s_{3}), and we have now re-coupled the environments seen from the walkers, hence we are ready to start a new such step during the interval [T,2T][T,2T], etc. This last point is crucial, as we need to iterate to ensure an expected gain 𝔼ρ+ε[XL]𝔼ρ[XL]\mathbb{E}^{\rho+\varepsilon}[X_{L}]-\mathbb{E}^{\rho}[X_{L}] that is large enough, cf. (1.15).

We now sketch a back-of-the-envelope calculation to argue that this scheme indeed generates the necessary discrepancy between 𝔼ρ+ε[XL]\mathbb{E}^{\rho+\varepsilon}[X_{L}] and 𝔼ρ[XL]\mathbb{E}^{\rho}[X_{L}]. During an interval of the form [iT,(i+1)T][iT,(i+1)T] for i=0,1,,L/T1i=0,1,\ldots,\lfloor L/T\rfloor-1 (we take i=0i=0 in the subsequent discussion for simplicity), with probability 1ect\simeq 1-e^{-ct}, we do not have the first separating event (around time s1s_{1}) between Xρ+εX^{\rho+\varepsilon} and XρX^{\rho}, and we simply end up with XTρ+εXTρ0X^{\rho+\varepsilon}_{T}-X^{\rho}_{T}\geq 0, which is a nonnegative expected gain.

Else, with probability ect\simeq e^{-ct} we have the first separation happening during [s1,s2][s_{1},s_{2}]. On this event, the coupling in 1) (and the subsequent impermeability of the green trapezoid) succeeds with probability 1et1/100\geq 1-e^{-t^{1/100}}, and we end up with XTρ+εXTρtX^{\rho+\varepsilon}_{T}-X^{\rho}_{T}\geq t, resulting in an expected gain of order t\simeq t. If the first coupling fails, and we resort to the parachute coupling, we have a (negative) expected gain 2Tet1/100\simeq-2Te^{-t^{1/100}}. Multiplying by the probability of the first separation, we obtain a net expected gain ect(t2Tet1/100)\simeq e^{-ct}(t-2Te^{-t^{1/100}}).

Finally, we evaluate the possibility that the surgery coupling (in step 2) above) or the parachute coupling 3) fails, preventing us to restore the domination of ηρ(+Xρ)\eta^{\rho}(\cdot+X^{\rho}) by ηρ+ε(+Xρ+ε)\eta^{\rho+\varepsilon}(\cdot+X^{\rho+\varepsilon}) at time TT, and we do not control anymore the coupling of the walks. This has probability O(L100)O(L^{-100}), and in the worst case we have deterministically XLρ+εXLρ=2LX^{\rho+\varepsilon}_{L}-X^{\rho}_{L}=-2L, so the expected loss is O(L99)O(L^{-99}).

Putting it all together, we get, for LL large enough by choosing e.g. t=logLt=\sqrt{\log L} and T=(logL)100T=(\log L)^{100} (it turns out that our couplings work with this choice of scales), that

(1.16) 𝔼ρ+ε[XTρ+ε]𝔼ρ[XTρ]ect(t2Tet1/100)+O(L99)Lo(1)\mathbb{E}^{\rho+\varepsilon}[X^{\rho+\varepsilon}_{T}]-\mathbb{E}^{\rho}[X^{\rho}_{T}]\geq e^{-ct}(t-2Te^{-t^{1/100}})+O(L^{-99})\geq L^{-o(1)}

As long as the surgery coupling or the parachute succeed, we can repeat this coupling for up to L/T\simeq L/T times during [0,L][0,L], thus obtaining

(1.17) 𝔼ρ+ε[XLρ+ε]𝔼ρ[XLρ]L1o(1)/TL1o(1),\mathbb{E}^{\rho+\varepsilon}[X^{\rho+\varepsilon}_{L}]-\mathbb{E}^{\rho}[X^{\rho}_{L}]\geq L^{1-o(1)}/T\geq L^{1-o(1)},

which establishes (1.15), with a sizeable margin.

1.4. Organization of the paper

In Section 2, we give rigorous definitions of general environments with a minimal set of properties (P.1)-(2.2), and of the random walk XX. In Section 3, we impose the mild coupling conditions (C.1)-(C.3) on the environments, state our general result, Theorem 3.1, and deduce Theorems 1.1 and 1.2 from it. We then define the finite-range models and give a skeleton of the proof. In Section 4, we proceed to the first part of the proof, and in Section 5, we proceed to the second part (cf. Section 1.3 reagrding the two parts). In Section 6, we define rigorously the SEP and show that it satisfies all the conditions mentioned above. The (short) Appendix A contains a few tail estimates in use throughout this article. In Appendix B, we introduce the PCRW environment and show that it equally satisfies the above conditions, which entails that our results also apply to this environment. Throughout this article all quantities may implicitly depend on the two parameters p,p(0,1)p_{\bullet},p_{\circ}\in(0,1) that are assumed to satisfy p>pp_{\bullet}>p_{\circ}, cf. above (1.1) (and also Section 2.3).

2 Setup and useful facts

In this section, after introducing a small amount of notation (Section 2.1), we proceed to define in Section 2.2 a class of (dynamic) random environments of interest, driven by a Markov process η\eta characterized by properties (P.1)–(2.2) below. We will primarily be interested in environments driven by the exclusion process, which is introduced in Section 6 and shown to satisfy these properties. The framework developed in Section 2.2 and further in Section 3 will allow our results to apply directly to a second environment of interest, considered in Appendix B. We conclude by introducing in Section 2.3 the relevant walk in random environment along with its associated quenched and annealed laws and collect a few basic features of this setup.

2.1. Notation

We write +\mathbb{Z}_{+}, resp. +\mathbb{R}_{+}, for the set of nonnegative integers, resp. real numbers. We use the letter z×+z\in\mathbb{R}\times\mathbb{R}_{+} exclusively to denote space-time points z=(x,t)z=(x,t), and typically x,y,x,yx,y,x^{\prime},y^{\prime}\dots for spatial coordinates and s,t,s,t,s,t,s^{\prime},t^{\prime},\dots for time coordinates. We usually use m,n,m,n,\dots for non-negative integers. With a slight abuse of notation, for two integers aba\leq b, we will denote by [a,b][a,b] the set of integers {a,,b}\{a,\ldots,b\} and declare that the length of [a,b][a,b] is |{a,,b}|=ba+1|\{a,\ldots,b\}|=b-a+1. Throughout, c,c,c,c^{\prime},\dots and C,C,C,C^{\prime},\dots denote generic constants in (0,)(0,\infty) which are purely numerical and can change from place to place. Numbered constants are fixed upon first appearance.

2.2. A class of dynamic random environments

We will consider stationary Markov (jump) processes with values in Σ=(+)\Sigma=(\mathbb{Z}_{+})^{\mathbb{Z}}. The state space Σ\Sigma carries a natural partial order: for two configurations η,ηΣ\eta,\eta^{\prime}\in\Sigma, we write ηη\eta\preccurlyeq\eta^{\prime} (or ηη\eta^{\prime}\succcurlyeq\eta) if, for all xx\in\mathbb{Z}, η(x)η(x)\eta(x)\leq\eta^{\prime}(x). More generally, for II\subset\mathbb{Z}, η|Iη|I\eta|_{I}\preccurlyeq\eta^{\prime}|_{I} (or η|Iη|I\eta^{\prime}|_{I}\succcurlyeq\eta|_{I}) means that η(x)η(x)\eta(x)\leq\eta^{\prime}(x) for all xIx\in I. For a finite subset II\subset\mathbb{Z}, we use the notation η(I)=xIη(x)\eta(I)=\sum_{x\in I}\eta(x). We will denote st.\geq_{\text{st.}} and st.\leq_{\text{st.}} the usual stochastic dominations for probability measures. Let JJ be a (fixed) non-empty open interval of +\mathbb{R}^{+}. An environment is specified in terms of two families of probability measures (𝐏η0:η0Σ)(\mathbf{P}^{\eta_{0}}:\eta_{0}\in\Sigma) governing the process (ηt)t0(\eta_{t})_{t\geq 0} and (μρ:ρJ)({\mu_{\rho}}:\rho\in J), where μρ\mu_{\rho} is a measure on Σ\Sigma, which are required satisfy the following conditions:

(P.1) (Markov property and invariance). For every η0Σ\eta_{0}\in\Sigma, the process (ηt)t0(\eta_{t})_{t\geq 0} defined under 𝐏η0\mathbf{P}^{\eta_{0}} is a time-homogeneous Markov process, such that for all s0s\geq 0 and xx\in\mathbb{Z}, the process (ηs+t(x+))t0(\eta_{s+t}(x+\cdot))_{t\geq 0} has law 𝐏ηs(x+)\mathbf{P}^{\eta_{s}(x+\cdot)}; moreover, the process (ηt)t0(\eta_{t})_{t\geq 0} exhibits ‘axial symmetry’ in the sense that (ηt(x))t0(\eta_{t}(x-\cdot))_{t\geq 0} has law 𝐏η0(x)\mathbf{P}^{\eta_{0}(x-\cdot)}.
(P.2) (Stationary measure). The initial distribution μρ{\mu}_{\rho} is a stationary distribution for the Markov process 𝐏η0\mathbf{P}^{\eta_{0}}. More precisely, letting 𝐏ρ=μρ(dη0)𝐏η0,ρJ,{\mathbf{P}}^{\rho}=\int{\mu}_{\rho}(d\eta_{0})\mathbf{P}^{\eta_{0}},\quad\rho\in J, the 𝐏ρ{\mathbf{P}}^{\rho}-law of (ηt+s)t0(\eta_{t+s})_{t\geq 0} is identical to the 𝐏ρ{\mathbf{P}}^{\rho}-law of (ηt)t0(\eta_{t})_{t\geq 0} for all s0s\geq 0 and ρJ\rho\in J. In particular, the marginal law of ηt\eta_{t} under 𝐏ρ{\mathbf{P}}^{\rho} is μρ{\mu}_{\rho} for all t0t\geq 0.
(P.3) (Monotonicity). The following stochastic dominations hold: i) (quenched) For all η0η0\eta^{\prime}_{0}\preccurlyeq\eta_{0}, one has that 𝐏η0st.𝐏η0\mathbf{P}^{\eta^{\prime}_{0}}\leq_{\text{st.}}\mathbf{P}^{\eta_{0}} , i.e. there exists a coupling of (ηt)t0(\eta^{\prime}_{t})_{t\geq 0} and (ηt)t0(\eta_{t})_{t\geq 0} such that ηtηt\eta^{\prime}_{t}\preccurlyeq\eta_{t} for all t0t\geq 0. ii) (annealed) For all ρρ\rho^{\prime}\leq\rho, one has that μρst.μρ{\mu}_{\rho^{\prime}}\leq_{\text{st.}}{\mu}_{\rho}. Together with i), this implies that 𝐏ρst.𝐏ρ{\mathbf{P}}^{\rho^{\prime}}\leq_{\text{st.}}{\mathbf{P}}^{\rho}.

A prime example satisfying the above conditions is the simple exclusion process, introduced and discussed further in Section 6; see in particular Lemma 6.1 regarding the validity of properties (P.1)–(2.2). We refer to Appendix B for another example.

2.3. Random walk

We now introduce the random walk in dynamic environment (RWdRE) that will be the main object of interest in this article. To this effect, we fix two constants p,p(0,1)p_{\bullet},p_{\circ}\in(0,1) such that p>pp_{\bullet}>p_{\circ} and an environment configuration η=(ηt)t0\eta=(\eta_{t})_{t\geq 0} with ηtΣ\eta_{t}\in\Sigma (=+=\mathbb{Z}_{+}^{\mathbb{Z}}, see Section 2.2). Given this data, the random walk evolving on top of the environment η\eta is conveniently defined in terms of a family (Un)n0(U_{n})_{n\geq 0} of i.i.d. uniform random variables on [0,1][0,1], as follows. For an initial space-time position z=(x,m)×+z=(x,m)\in\mathbb{Z}\times\mathbb{Z}_{+}, let PzηP^{\eta}_{z} be the law of the discrete-time Markov chain X=(Xn)n0X=(X_{n})_{n\geq 0} such that X0=xX_{0}=x and for all integer n0n\geq 0,

(2.1) Xn+1=Xn+2×𝟙{Un(pp)1{ηn+m(Xn)=0}+p}1.X_{n+1}=X_{n}+2\times{\mathds{1}}\big{\{}U_{n}\leq(p_{\circ}-p_{\bullet})1\{\eta_{n+m}(X_{n})=0\}+p_{\bullet}\big{\}}-1.

Let us call xx\in\mathbb{Z} an occupied site of ηtΣ\eta_{t}\in\Sigma if ηt(x)>0\eta_{t}(x)>0, and empty otherwise. With this terminology, (2.1) implies for instance that when XnX_{n}, started at a point z=(x,t=0)z=(x,t=0), is on an occupied (resp. empty) site of ηn\eta_{n}, it jumps to its right neighbour with probability pp_{\bullet} (resp. pp_{\circ}) and to its left neighbour with probability 1p1-p_{\bullet} (resp. 1p)1-p_{\circ}). We call PzηP^{\eta}_{z} the quenched law of the walk started at zz and abbreviate Pη=P(0,0)ηP^{\eta}=P^{\eta}_{(0,0)}; here quenched refers to the fact that the environment η\eta is deterministic.

We now discuss annealed measures, i.e. including averages over the dynamics of the environment η\eta. We assume from here on that η=(ηt)t0\eta=(\eta_{t})_{t\geq 0} satisfies the assumptions of Section 2.2. Recall that 𝐏η0\mathbf{P}^{\eta_{0}} denotes the law of η\eta starting from the configuration η0Σ\eta_{0}\in\Sigma and that 𝐏ρ\mathbf{P}^{\rho} is declared in (P.2). Correspondingly, one introduces the following two annealed measures for the walk

(2.2) zρ[]=𝐏ρ(dη)Pzη[],ρJ,\displaystyle\mathbb{P}^{\rho}_{z}[\,\cdot\,]=\int\mathbf{P}^{\rho}(d\eta)P^{\eta}_{z}[\,\cdot\,],\quad\rho\in J,
(2.3) zη0[]=𝐏η0(dη)Pzη[],η0Σ,\displaystyle\mathbb{P}^{\eta_{0}}_{z}[\,\cdot\,]=\int\mathbf{P}^{\eta_{0}}(d\eta)P^{\eta}_{z}[\,\cdot\,],\quad\eta_{0}\in\Sigma,

for arbitrary z×+z\in\mathbb{Z}\times\mathbb{Z}_{+}. Whereas the latter averages over the dynamics of η\eta for a fixed initial configuration η0\eta_{0}, the former includes η0\eta_{0}, which is sampled from the stationary distribution μρ\mu_{\rho} for η\eta. Observe that zρ=μρ(dη0)zη0\mathbb{P}^{\rho}_{z}=\int{\mu}_{\rho}(d\eta_{0})\mathbb{P}^{\eta_{0}}_{z}. For {η,ρ}\star\in\{\eta,\rho\}, write x\mathbb{P}^{\star}_{x} for (x,0)\mathbb{P}^{\star}_{(x,0)}, for all xx\in\mathbb{Z}, and write simply \mathbb{P}^{\star} for 0\mathbb{P}^{\star}_{0}.

We introduce a joint construction for the walk XX when started at different space-time points. This involves a graphical representation using arrows similar to that used in [40, Section 3], but simpler, which will be sufficient for our purposes. For a point w=(x,n)×+w=(x,n)\in\mathbb{Z}\times\mathbb{Z}_{+}, we let π1(w)=x\pi_{1}(w)=x and π2(w)=n\pi_{2}(w)=n denote the projection onto the first (spatial) and second (temporal) coordinate. We consider the discrete lattice

(2.4) 𝕃=(2×2+)((1,1)+(2×2+)).\mathbb{L}=(2\mathbb{Z}\times 2\mathbb{Z}_{+})\cup\big{(}(1,1)+(2\mathbb{Z}\times 2\mathbb{Z}_{+})\big{)}.

Note that the process (Xn,n)n0(X_{n},n)_{n\geq 0} evolves on the lattice 𝕃(×+)\mathbb{L}\subset(\mathbb{Z}\times\mathbb{Z}_{+}) defined by (2.4) when X=(Xn)n0X=(X_{n})_{n\geq 0} is started at z=(0,0)z=(0,0) under any of PzηP_{z}^{\eta} and the measures in (2.2)-(2.3).

We proceed to define a family of processes (Xw=(Xnw)n0:w𝕃)(X^{w}=(X^{w}_{n})_{n\geq 0}:w\in\mathbb{L}), such that X0w=π1(w)X^{w}_{0}=\pi_{1}(w) almost surely and (Xn(0,0))n0(X^{(0,0)}_{n})_{n\geq 0} has the same law as XX under P(0,0)ηP_{(0,0)}^{\eta}. Furthermore, XwX^{w^{\prime}} and XwX^{w} have the property that they coalesce whenever they intersect, that is if Xmw=XnwX^{w^{\prime}}_{m}=X^{w}_{n} for some w,w𝕃w,w^{\prime}\in{\mathbb{L}} and n,m0n,m\geq 0, then Xm+kw=Xn+kwX^{w^{\prime}}_{m+k}=X^{w}_{n+k} for all k0k\geq 0.

Let U=(Uw)w𝕃U=\left(U_{w}\right)_{w\in\mathbb{L}} be a collection of i.i.d. uniform random variables on [0,1][0,1]. Given the environment η\eta, we define a field A=(Aw)w𝕃{1,1}𝕃A=(A_{w})_{w\in\mathbb{L}}\in\{-1,1\}^{\mathbb{L}} of arrows (see Figure 3), measurably in (η,U)(\eta,U) as follows:

(2.5) Aw=A(ηn(x),Uw)=2×𝟙{Uw(pp)1{ηn(x)=0}+p}1,w=(x,n)𝕃.\displaystyle A_{w}=A(\eta_{n}(x),U_{w})=2\times{\mathds{1}}\big{\{}U_{w}\leq(p_{\circ}-p_{\bullet})1\{\eta_{n}(x)=0\}+p_{\bullet}\big{\}}-1,\quad w=(x,n)\in\mathbb{L}.

For any w=(x,n)𝕃w=(x,n)\in\mathbb{L}, we then set X0w=xX^{w}_{0}=x and, for all integer k0k\geq 0, we define recursively

(2.6) Xk+1w=Xkw+A(Xkw,n+k).\displaystyle X^{w}_{k+1}=X^{w}_{k}+A_{(X^{w}_{k},\,n+k)}.
Refer to caption
Figure 3: A representation of 𝕃\mathbb{L} (in dashed black) with the arrows in blue, the space \mathbb{Z} being horizontal and time going upwards. The trajectory of X(0,0)X^{(0,0)} is in red. Trajectories simply follow the arrows. Notice that two coalescing trajectories stay merged forever.

This defines the coupled family ((Xkw)k0:w𝕃)((X^{w}_{k})_{k\geq 0}:w\in\mathbb{L}), and we note that trajectories the process (Xkw,k+n)k0(X_{k}^{w},k+n)_{k\geq 0}, where π2(w)=n\pi_{2}(w)=n, are embedded in (i.e. subsets of) 𝕃\mathbb{L}. In view of (2.1) and (2.5)-(2.6), it follows plainly that the law of X(0,0)=(Xn(0,0))n0X^{(0,0)}=(X^{(0,0)}_{n})_{n\geq 0} when averaging over UU while keeping η\eta fixed is the same as that of XX under P(0,0)η=PηP^{\eta}_{(0,0)}=P^{\eta}.

We now discuss a variation of the construction specified around (2.5)-(2.6), which will be practical in Section 5.2.

Lemma 2.1.

Let η0Σ\eta_{0}\in\Sigma and suppose that (η,U)(\eta,U) are coupled under the probability measure QQ, with η\eta having marginal law 𝐏η0\mathbf{P}^{\eta_{0}} and (Uw)w𝕃\left(U_{w}\right)_{w\in\mathbb{L}} a collection of i.i.d. uniform random variables on [0,1][0,1], in such a way that under QQ, for all integers t0t\geq 0,

(2.7) (Uw:π2(w)=t) is independent from t,\text{$(U_{w}:\pi_{2}(w)=t)$ is independent from $\mathcal{F}_{t}$},

where

(2.8) t:=σ((ηs)0st,(Uw:π2(w)t1)).\mathcal{F}_{t}:=\sigma((\eta_{s})_{0\leq s\leq t},(U_{w}:\pi_{2}(w)\leq t-1)).

Then, defining Xw=Xw(η,U)X^{w}=X^{w}(\eta,U) as in (2.5)-(2.6), it follows that X(0,0)X^{(0,0)} has law η0\mathbb{P}^{\eta_{0}} (cf. (2.3)) under QQ.

Proof.

We show by induction over k1k\geq 1 integer that for all such kk, the pair ((ηt)0tk1,(Xt)0tk)((\eta_{t})_{0\leq t\leq k-1},(X_{t})_{0\leq t\leq k}) has the same law under both QQ and η0\mathbb{P}^{\eta_{0}} (which, for the duration of the proof, we extend here to denote the joint distribution of (X,η)(X,\eta), by a slight abuse of notation). The initialisation at k=1k=1 is immediate, since Q(X1=1)=p𝟙{η0(0)=1}+p𝟙{η0(0)=0}Q(X_{1}=1)=p_{\bullet}{\mathds{1}}\big{\{}\eta_{0}(0)=1\big{\}}+p_{\circ}{\mathds{1}}\big{\{}\eta_{0}(0)=0\big{\}} and η0\eta_{0} is deterministic.

As for the induction step, assume the induction hypothesis for some k1k\geq 1, and condition on the σ\sigma- algebra Σk:=σ((ηt)0tk1,(Xt)0tk)\Sigma_{k}:=\sigma((\eta_{t})_{0\leq t\leq k-1},(X_{t})_{0\leq t\leq k}). We first let η\eta evolve between times k1k-1 and kk. Under both QQ and η0\mathbb{P}^{\eta_{0}} and conditionally on Σk\Sigma_{k}, we have that (ηt)k1tk(\eta_{t})_{k-1\leq t\leq k} is distributed as (ηt)0t1(\eta_{t})_{0\leq t\leq 1} under 𝐏ηk1\mathbf{P}^{\eta_{k-1}}, given the marginal distribution of η\eta and by the Markov property (P.1). Therefore, ((ηt)0tk,(Xt)0tk)((\eta_{t})_{0\leq t\leq k},(X_{t})_{0\leq t\leq k}) has the same law under both QQ and η0\mathbb{P}^{\eta_{0}}. Now, conditioning on k\mathcal{F}_{k} as defined in (2.8), and noticing that (ηt)0tk(\eta_{t})_{0\leq t\leq k}, (Xt)0tk)(X_{t})_{0\leq t\leq k}) (hence also ηk(Xk)\eta_{k}(X_{k})) are all k\mathcal{F}_{k}-measurable (for (Xt)0tk(X_{t})_{0\leq t\leq k} this follows from (2.5)-(2.6)), we obtain that

(2.11) Q(Xk+1Xk=1|k)=(2.5),(2.6)(U(Xk,k)p|k)𝟙{ηk(Xk)=1}+(U(Xk,k)p|k)𝟙{ηk(Xk)=0}.Q(X_{k+1}-X_{k}=1\,|\,\mathcal{F}_{k})\\ \stackrel{{\scriptstyle\eqref{def:A},\eqref{eq:defX}}}{{=}}\mathbb{Q}(U_{(X_{k},k)}\leq p_{\bullet}\,|\,\mathcal{F}_{k})\mathds{1}\big{\{}\eta_{k}(X_{k})=1\big{\}}+\mathbb{Q}(U_{(X_{k},k)}\leq p_{\circ}\,|\,\mathcal{F}_{k})\mathds{1}\big{\{}\eta_{k}(X_{k})=0\big{\}}.

Now, by (2.8), and using that XkX_{k} is k\mathcal{F}_{k}-measurable, we can evaluate the conditional probabilities in (2.11) to find that

(2.14) Q(Xk+1Xk=1|k)=p𝟙{ηk(Xk)=1}+p𝟙{ηk(Xk)=0}=η0(Xk+1Xk=1|k),Q(X_{k+1}-X_{k}=1\,|\,\mathcal{F}_{k})=p_{\bullet}\mathds{1}\big{\{}\eta_{k}(X_{k})=1\big{\}}+p_{\circ}\mathds{1}\big{\{}\eta_{k}(X_{k})=0\big{\}}\\ =\mathbb{P}^{\eta_{0}}(X_{k+1}-X_{k}=1\,|\,\mathcal{F}_{k}),

where the second equality comes from the independence of k\mathcal{F}_{k} and UkU_{k} under η0\mathbb{P}^{\eta_{0}}, see (2.1) and (2.3). Integrating both sides of (2.14) under QQ and η0\mathbb{P}^{\eta_{0}}, respectively, against a suitable (k\mathcal{F}_{k}-measurable) test function of ((ηt)0tk,(Xt)0tk)((\eta_{t})_{0\leq t\leq k},(X_{t})_{0\leq t\leq k}), it follows that ((ηt)0tk,(Xt)0tk+1)((\eta_{t})_{0\leq t\leq k},(X_{t})_{0\leq t\leq k+1}) have the same distribution under QQ and η0\mathbb{P}^{\eta_{0}}. This concludes the proof of the induction step. ∎

We conclude this section by collecting a useful monotonicity property for the collection of random walks defined above, similar to [40, Proposition 3.1]. Its proof hinges on the fact that the trajectories considered cannot cross without first meeting at a vertex, after which they merge. In the sequel, for a given environment configuration η~=(η~t)t0\widetilde{\eta}=(\widetilde{\eta}_{t})_{t\geq 0}, we refer to X~w=(X~nw)n0\widetilde{X}^{w}=(\widetilde{X}_{n}^{w})_{n\geq 0} the process defined as in (2.6) but with η~\widetilde{\eta} in place of η\eta entering the definition of the arrows in (2.5). We will be interested in the case where η\eta and η~\widetilde{\eta} are such that

(2.15) ηn(x)η~n(x), for all (x,n)𝕃K,\displaystyle\eta_{n}(x)\leq\widetilde{\eta}_{n}(x)\text{, for all }(x,n)\in\mathbb{L}\cap K,

for some K×+K\subseteq\mathbb{Z}\times\mathbb{Z}_{+}. The following result is already interesting for the choice η=η~\eta=\widetilde{\eta}, in which case Xw=X~wX^{w}=\widetilde{X}^{w} below.

Lemma 2.2.

If η,η~Σ\eta,\widetilde{\eta}\in\Sigma and K×+K\subseteq\mathbb{Z}\times\mathbb{Z}_{+} are such that (2.15) holds, then for every w,wKw,w^{\prime}\in K with π1(w)π1(w)\pi_{1}(w^{\prime})\leq\pi_{1}(w) and π2(w)=π2(w)\pi_{2}(w)=\pi_{2}(w^{\prime}),and for every n0n\geq 0 such that [π1(w)k,π1(w)+k]×[π2(w),π2(w)+k]K[\pi_{1}(w)-k,\pi_{1}(w^{\prime})+k]\times[\pi_{2}(w),\pi_{2}(w)+k]\subseteq K for all 0kn0\leq k\leq n, one has that

(2.16) XnwX~nw.X^{w^{\prime}}_{n}\leq\widetilde{X}^{w}_{n}.
Proof.

We only treat the case K=×+K=\mathbb{Z}\times\mathbb{Z}_{+} to lighten the argument. The adaptation to a general KK is straightforward, as all possible trajectories considered for XX and X~\widetilde{X} lie in KK by assumption. The proof proceeds by a straightforward induction argument. Indeed, since X0w=π1(w)X_{0}^{w^{\prime}}=\pi_{1}(w^{\prime}) and X~0w=π1(w)\widetilde{X}_{0}^{w}=\pi_{1}(w), one has that X~0wX0w0\widetilde{X}^{w}_{0}-X^{w}_{0}\geq 0 by assumption. To carry out the induction step, one notes that X~nwXnw\widetilde{X}^{w}_{n}-X^{w^{\prime}}_{n} is even for any n0n\geq 0 with increments ranging in 2-2, 0 or +2+2, and combines this with the following observation: if X~nwXnw=0\widetilde{X}^{w}_{n}-X^{w^{\prime}}_{n}=0, then

X~n+1wXn+1w=(2.6)A(X~nw,π2(w)+n)A(Xnw,π2(w)+n)=(2.5)A(η~n(Xnw),Uw)A(ηn(Xnw),Uw)0,\widetilde{X}^{w}_{n+1}-X^{w^{\prime}}_{n+1}\stackrel{{\scriptstyle\eqref{eq:defX}}}{{=}}A_{(\widetilde{X}_{n}^{w},\pi_{2}(w)+n)}-A_{({X}_{n}^{w^{\prime}},\pi_{2}(w^{\prime})+n)}\stackrel{{\scriptstyle\eqref{def:A}}}{{=}}A(\widetilde{\eta}_{n}(X^{w}_{n}),U_{w})-A({\eta}_{n}(X^{w}_{n}),U_{w})\geq 0,

where the second equality follows using that π2(w)=π2(w)\pi_{2}(w)=\pi_{2}(w^{\prime}) (along with X~nw=Xnw\widetilde{X}^{w}_{n}=X^{w^{\prime}}_{n}) and the inequality is due to (2.15) and the fact that A(,ξ)A(\cdot,\xi) is increasing for any ξ[0,1]\xi\in[0,1], which is straightforward from (2.5). ∎

For later reference, we also record that for any w𝕃w\in\mathbb{L} and nm0n\geq m\geq 0,

(2.18) |XnwXmw|nm,\left|X^{w}_{n}-X^{w}_{m}\right|\leq n-m,

as follows clearly from (2.6).

3 Main results

In this section, we start by formulating in Section 3.1 precise coupling conditions that we require from the environments, and which are of independent interest. These are given by conditions (C.1)-(C.3) below (a flavor of the second of these was given in (1.13) in the introduction). Our main result, Theorem 3.1, appears in Section 3.2. It concerns the generic random walk in random environment defined in Section 2.2-2.3, subject to the conditions of Section 3.1. Our standing assumptions will thus be that all of properties (P.1)-(2.2) and the conditions (C.1)-(C.3) hold. We will verify separately in Section 6 that all of these conditions hold for SEP. From Theorem 3.1 we then readily deduce Theorems 1.1and 1.2 at the end of Section 3.2.

Towards the proof of Theorem 3.1, and following the outline of Section 1.3, we proceed to introduce in Section 3.3 a finite-range approximation of the model, in which the environment is renewed after LL time steps and gather its essential features that will be useful for us. We then state two key intermediate results, Propositions 3.3 and 3.4, which correspond to the first and second parts from the discussion in Section 1.3; cf. also (1.10) and (1.11). From these, we deduce Theorem 3.1 in Section 3.4. The proofs of the two propositions appear in forthcoming sections.

3.1. Coupling conditions on the environment

Recall the framework of Section 2.2-2.3, which we now amend with three further conditions. These involve an additional parameter ν>0\nu>0 which quantifies the activity of the environment (in practice it will correspond to the rate parameter appearing in (6.1) in the case of SEP). We keep the dependence on ν\nu explicit in the following conditions for possible future applications, for which one may wish to tamper with the speed of the environment (for instance, by slowing it down). For the purposes of the present article however, one could simply set ν=1\nu=1 in what follows.

The first two conditions (C.1) and (C.2) below regard the environment (ηt)t0(\eta_{t})_{t\geq 0} alone, which, following the setup of Section 2.2, is assumed to be specified in terms of the measures (𝐏η0:η0Σ)(\mathbf{P}^{\eta_{0}}:\eta_{0}\in\Sigma) and (μρ:ρJ)({\mu_{\rho}}:\rho\in J) that satisfy (P.1)-(2.2). The first condition, (C.1), concerns the empirical density of the environment. Roughly speaking, it gives a quantitative control on how (2.2) is conserved over time. The second condition, (C.2), is more technical. In a nutshell, it states that if one environment η\eta covers another environment η\eta^{\prime} on a finite interval II at a given time, and if η\eta has a larger empirical density than η\eta^{\prime} outside II at the same time, the evolutions of η\eta and η\eta^{\prime} can be coupled in a way that η\eta covers η\eta^{\prime} on a larger interval after some time.

We proceed to formalize these two properties:

  1. (C.1)

    (Conservation of density). There exists constants \Cldensitystable,\Cl[c]densitystableexpo(0,)\Cl{densitystable},\Cl[c]{densitystableexpo}\in(0,\infty) such that for all ρJ\rho\in J and ε(0,1)\varepsilon\in(0,1) (with ρ+εJ\rho+\varepsilon\in J), and for all ,,H,t1\ell,\ell^{\prime},H,t\geq 1 satisfying H>4νt>\Crdensitystable2ε2(1+|log3(νt)|)H>4{\nu}t>\Cr{densitystable}\ell^{2}\varepsilon^{-2}(1+{|\log^{3}(\nu t)}|) and t\ell^{\prime}\leq\sqrt{t}, the following two inequalities hold. Let η0\eta_{0} be such that on every interval II of length \ell included [H,H][-H,H], one has η0(I)(ρ+ε)\eta_{0}(I)\leq(\rho+\varepsilon)\ell (resp. η0(I)(ρε)\eta_{0}(I)\geq(\rho-\varepsilon)\ell). Then

    𝐏η0(for all intervals I of length  included in[H+2νt,H2νt]ηt(I)(ρ+3ε)(resp. ηt(I)(ρ3ε)))14Hexp(\Crdensitystableexpoε2).{\mathbf{P}}^{\eta_{0}}\left(\begin{array}[]{c}\text{for all intervals $I^{\prime}$ of length $\ell^{\prime}$ included in}\\ \text{$[-H+2{\nu}t,H-2{\nu}t]$: $\eta_{t}(I^{\prime})\leq(\rho+3\varepsilon)\ell^{\prime}$}\\ \text{(resp.~$\eta_{t}(I^{\prime})\geq(\rho-3\varepsilon)\ell^{\prime}$)}\end{array}\right)\geq 1-4H\exp(-\Cr{densitystableexpo}\varepsilon^{2}\ell^{\prime}).
  2. (C.2)

    (Couplings). There exists \Clcompatible,\ClSEPcoupling2(0,)\Cl{compatible},\Cl{SEPcoupling2}\in(0,\infty) such that the following holds. Let ρJ\rho\in J, ε(0,1)\varepsilon\in(0,1) (with ρ+εJ\rho+\varepsilon\in J), and H1,H2,t,1H_{1},H_{2},t,\ell\geq 1 be integers such that min{H1,H2H11}>10νt>4ν100>\Crcompatible\min\{H_{1},H_{2}-H_{1}-1\}>10{\nu}t>4\nu\ell^{100}>\Cr{compatible}, ν>\Crcompatibleε2(1+|log3(ν4)|)\nu\ell>\Cr{compatible}\varepsilon^{-2}(1+|\log^{3}(\nu\ell^{4})|) and >80νε1+ν2\ell>80\nu\varepsilon^{-1}+\nu^{-2}. Let η0,η0Σ\eta_{0},\eta^{\prime}_{0}\in\Sigma be such that η0|[H1,H1]η0|[H1,H1]\eta_{0}|_{[-H_{1},H_{1}]}\succcurlyeq\eta_{\color[rgb]{1,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{1,0,0}0\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}}^{\prime}|_{[-H_{1},H_{1}]} and such that for every interval I[H2,H2]I\subseteq[-H_{2},H_{2}] of length /2|I|\lfloor\ell/2\rfloor\leq|I|\leq\ell , we have η0(I)(ρ+3ε/4)|I|\eta_{0}(I)\geq(\rho+3\varepsilon/4)|I| and η0(I)(ρ+ε/4)|I|\eta^{\prime}_{0}(I)\leq(\rho+\varepsilon/4)|I|. Then there exists a coupling \mathbb{Q} of two environments η,η\eta,\eta^{\prime} with respective marginals 𝐏η0,𝐏η0\mathbf{P}^{\eta_{0}},\mathbf{P}^{\eta^{\prime}_{0}} such that

    (3.4) (s[0,t],ηs|[H1+4νt,H14νt]ηs|[H1+4νt,H14νt])120texp(νt/4)\mathbb{Q}\big{(}\forall s\in[0,t],\,\eta_{s}|_{[-H_{1}+4{\nu}t,H_{1}-4{\nu}t]}\succcurlyeq\eta^{\prime}_{s}|_{[-H_{1}+4{\nu}t,H_{1}-4{\nu}t]}\big{)}\geq 1-20t\exp(-{\nu}t/4)

    and

    (3.5) (ηt|[H2+6νt,H26νt]ηt|[H2+6νt,H26νt])15\CrSEPcoupling24H2exp(\CrSEPcoupling21νν+1ε2).\mathbb{Q}\big{(}\eta_{t}|_{[-H_{2}+6{\nu}t,H_{2}-6{\nu}t]}\succcurlyeq\eta^{\prime}_{t}|_{[-H_{2}+6{\nu}t,H_{2}-6{\nu}t]}\big{)}\geq 1-5\Cr{SEPcoupling2}\ell^{4}H_{2}\exp\left(-\Cr{SEPcoupling2}^{-1}\textstyle\frac{\nu}{\nu+1}\varepsilon^{2}\ell\right).

For later reference, we record the following two particular instances of (C.2), which correspond to the cases where H2=0H_{2}=0 and H1=0H_{1}=0, respectively. For convenience, we state them with better constants and exponents than in (C.2). In fact, when verifying condition (C.2) for the SEP in Section 6.2, we will first prove that these two conditions hold, and use them to prove (C.2).

  1. (C.2.1)

    (No particle drifting in from the side). Let t,H0t,H\geq 0 and k1k\geq 1 be integers, and let η0,η0Σ\eta_{0},\eta^{\prime}_{0}\in\Sigma be such that η0|[H,H]η0|[H,H]\eta_{0}|_{[-H,H]}\succcurlyeq\eta_{0}^{\prime}|_{[-H,H]}. There exists a coupling \mathbb{Q} of environments η,η\eta,\eta^{\prime} with respective marginals 𝐏η0\mathbf{P}^{\eta_{0}} and 𝐏η0\mathbf{P}^{\eta^{\prime}_{0}} such that

    (3.6) (s[0,t],ηs|[H+2νkt,H2νkt]ηs|[H+2νkt,H2νkt])120exp(kνt/4).\mathbb{Q}\big{(}\forall s\in[0,t],\,\eta_{s}|_{[-H+2{\nu}kt,H-2{\nu}kt]}\succcurlyeq\eta^{\prime}_{s}|_{[-H+2{\nu}kt,H-2{\nu}kt]}\big{)}\geq 1-20\exp(-k{\nu}t/4).

    Informally, with high probability no particle of η\eta^{\prime} outside of [H,H][-H,H] can drift into [H+2νt,H2νt][-H+2{\nu}t,H-2{\nu}t] (when k=1k=1 for instance) before time tt to perturb the domination of η\eta^{\prime} by η\eta.

  2. (C.2.2)

    (Covering η\eta^{\prime} by η\eta). There exists \ClSEPcoupling>0\Cl{SEPcoupling}>0 such that for all ρJ\rho\in J and ε(0,1)\varepsilon\in(0,1) (with ρ+εJ\rho+\varepsilon\in J), the following holds. If H,t1H,t\geq 1 satisfy H>4νtH>4{\nu}t, ν8t>1\nu^{8}t>1 and νt1/4>\CrSEPcouplingε2(1+|log3(νt)|)\nu t^{1/4}>\Cr{SEPcoupling}\varepsilon^{-2}(1+|\log^{3}(\nu t)|), then for all η0,η0Σ\eta_{0},\eta^{\prime}_{0}\in\Sigma such that on each interval I[H,H]I\subset[-H,H] of length /2|I|\lfloor\ell/2\rfloor\leq|I|\leq\ell, where :=t1/4\ell:=\lfloor t^{1/4}\rfloor, η0(I)(ρ+3ε/4)|I|\eta_{0}(I)\geq(\rho+3\varepsilon/4)|I| and η0(I)(ρ+ε/4)|I|\eta^{\prime}_{0}(I)\leq(\rho+\varepsilon/4)|I|, there exists a coupling \mathbb{Q} of η,η\eta,\eta^{\prime} with marginals 𝐏η0,𝐏η0\mathbf{P}^{\eta_{0}},\mathbf{P}^{\eta^{\prime}_{0}} so that

    (3.7) (ηt|[H+4νt,H4νt]ηt|[H+4νt,H4νt])1\CrSEPcoupling2tHexp((\CrSEPcoupling2(1+ν1))1ε2t1/4).\mathbb{Q}(\eta_{t}|_{[-H+4{\nu}t,H-4{\nu}t]}\succcurlyeq\eta_{t}^{\prime}|_{[-H+4{\nu}t,H-4{\nu}t]})\geq 1-\Cr{SEPcoupling2}tH\exp\big{(}-(\Cr{SEPcoupling2}(1+\nu^{-1}))^{-1}\varepsilon^{2}t^{1/4}\big{)}.

    In words, the coupling \mathbb{Q} achieves order between η\eta and η\eta^{\prime} at time tt under suitable regularity assumptions on the empirical density of their initial configurations η0\eta_{0} and η0\eta_{0}^{\prime}.

The third and last property ensures that if an environment η\eta covers another environment η\eta^{\prime} and has at least one extra particle at distance \ell of the origin (which typically happens if η\eta has higher density than η\eta^{\prime}), then with probability at least exponentially small in \ell, by time \ell, this particle can reach the origin which will be empty for η\eta^{\prime}, while preserving the domination of η\eta^{\prime} by η\eta. We will combine this property with the uniform ellipticity of the random walk on top of the environment to show that the walker has at least an exponentially small probability to reach a position where η\eta has a particle but not η\eta^{\prime}, which in turn yields a probability bounded away from zero that a walker on η\eta steps to the right while a walker on η\eta^{\prime} steps to the left (under the coupling mentioned in Section 1.3), hence creating the desired initial gap that we will then exploit in our constructions.

  1. (C.3)

    (Sprinkler). For all ρJ\rho\in J, for all integers H,,k1H,\ell,k\geq 1 with H2νkH\geq 2\nu\ell k and k48ν1(ν+log(40)log(p(1p)ν/2))k\geq 48\nu^{-1}(\nu+\log(40)-\log(p_{\circ}(1-p_{\bullet})\nu/2)), the following holds. If η0,η0Σ\eta_{0},\eta^{\prime}_{0}\in\Sigma satisfy η0(x)η0(x)\eta_{0}(x)\geq\eta^{\prime}_{0}(x) for all x[H,H]x\in[-H,H], η0([0,])η0([0,])+1\eta_{0}([0,\ell])\geq\eta^{\prime}_{0}([0,\ell])+1 and η0([3+1,3])6(ρ+1)\eta^{\prime}_{0}([-3\ell+1,3\ell])\leq 6(\rho+1)\ell, then there is a coupling \mathbb{Q} of η\eta^{\prime} under 𝐏η0{\mathbf{P}}^{\eta_{0}^{\prime}} and η\eta under 𝐏η0{\mathbf{P}}^{\eta_{0}} such that, with δ=(ν/2eν)6(ρ+1),\delta=(\nu/2e^{{\nu}})^{6(\rho+1)\ell},

    (3.8) (η(x)>0,η(x)=0)2δ\begin{split}&\mathbb{Q}(\eta_{\ell}(x)>0,\,\eta^{\prime}_{\ell}(x)=0)\geq 2\delta\end{split}

    for x{0,1},x\in\{0,1\}, and with δ=δ(p(1p))6(ρ+1)\delta^{\prime}=\delta(p_{\circ}(1-p_{\bullet}))^{6(\rho+1)\ell},

    (3.9) ({s[0,],ηs|[H+2νk,H2νk]ηs|[H+2νk,H2νk]}c)20ekν/4δ.\begin{split}\mathbb{Q}(\{\forall s\in[0,\ell],\,\eta_{s}|_{[-H+2{\nu}k\ell,H-2{\nu}k\ell]}\succcurlyeq\eta^{\prime}_{s}|_{[-H+2{\nu}k\ell,H-2{\nu}k\ell]}\}^{c})&\leq 20e^{-k\nu\ell/4}\leq\delta^{\prime}.\end{split}

3.2. Main result

Following is our main theorem.

Theorem 3.1 (Sharpness of v()v(\cdot)).

Let η\eta be an environment as in Section 2.2 satisfying (C.1)-(C.3). Assume that for all ρJ\rho\in J, there exists v(ρ)v(\rho) such that

(3.10) ρ-a.s., limnn1Xn=v(ρ).\mathbb{P}^{\rho}\text{-a.s., }\lim_{n}n^{-1}{X_{n}}=v(\rho).

Then, for all ρ,ρJ\rho,\rho^{\prime}\in J such that ρ>ρ\rho>\rho^{\prime}, one has that

(3.11) v(ρ)>v(ρ).\displaystyle v(\rho)>v(\rho^{\prime}).

With Theorem 3.1 at hand, we first give the proofs of Theorems 1.1 and 1.2. We start with the latter.

Proof of Theorem 1.2.

Theorem 1.2 concerns the particular case where ρ\mathbb{P}^{\rho} refers to the walk of Section 2.3 evolving on top of the exclusion process η\eta started from product Bernoulli(ρ)(\rho) distribution, for ρJ(0,1)\rho\in J\subset(0,1). The properties (P.1)-(2.2) and (C.1)-(C.3) are indeed all satisfied in this case, as is proved separately in Section 6 below, see Lemma 6.1 and Proposition 6.3.

In the notation of (1.6), we now separately consider the intervals J{J,J0,J+}J\in\{J_{-},J_{0},J_{+}\}, where J=(0,ρ)J_{-}=(0,\rho_{-}), J0=(ρ,ρ+)J_{0}=(\rho_{-},\rho_{+}) and J+=(ρ+,1)J_{+}=(\rho_{+},1). The fact that the law of large numbers (3.10) holds for any choice of JJ is the content of [40, Theorem 1.1]. Thus Theorem 3.1 is in force, and vv is strictly increasing on JJ for any J{J,J0,J+}J\in\{J_{-},J_{0},J_{+}\}. Since by direct application of (P.3) and Lemma 2.2, vv is already non-decreasing on JJ0J+J_{-}\cup J_{0}\cup J_{+}, it must be (strictly) increasing on JJ0J+J_{-}\cup J_{0}\cup J_{+} altogether, and (1.8) follows. In view of (1.6), the first equality in (1.7) is an immediate consequence of (3.11) with J=J0J=J_{0}.

As to the second equality in (1.7), recalling the definition of ρc\rho_{c} from (1.4), we first argue that ρρc\rho_{-}\leq\rho_{c}. If ρ<ρ\rho<\rho_{-}, then by (1.5)-(1.6), Xn/nv(ρ)<0X_{n}/n\to v(\rho)<0 ρ\mathbb{P}^{\rho}-a.s. and thus in particular ρ(lim supXn<0)=1\mathbb{P}^{\rho}(\limsup X_{n}<0)=1 (in fact it equals -\infty but we will not need this). On the other hand, {Hn<H1}{H1>n}{Xk0,kn}\{H_{n}<H_{-1}\}\subset\{H_{-1}>n\}\subset\{X_{k}\geq 0,\,\forall k\leq n\}, and the latter has probability tending to 0 as nn\to\infty under ρ\mathbb{P}^{\rho}. It follows that θ(ρ)=0\theta(\rho)=0 in view of (1.3), whence ρρc\rho\leq\rho_{c}, and thus ρρc\rho_{-}\leq\rho_{c} upon letting ρρ\rho\uparrow\rho_{-}.

Since ρ=ρ+\rho_{-}=\rho_{+}, in order to complete the proof it is enough to show that ρ+ρc\rho_{+}\geq\rho_{c}. Let ρ>ρ+\rho>\rho_{+}. We aim to show that θ(ρ)>0\theta(\rho)>0. Using the fact that v(ρ)>0v(\rho)>0 and (1.5), one first picks n0=n0(ρ)1n_{0}=n_{0}(\rho)\geq 1 such that zρ(Xn>0,nn0)1/2\mathbb{P}^{\rho}_{z}(X_{n}>0,\,\forall n\geq n_{0})\geq 1/2 for any z=(m,m)z=(m,m) with m0m\geq 0 (the worst case is m=0m=0, the other cases follow from the case m=0m=0 using invariance under suitable space-time translations). Now, observe that for all n0n\geq 0, under ρ\mathbb{P}^{\rho},

(3.14) {Hn<H1}({XkXk1=+1,1kn0}{X2n0+k>0,kn0}{lim supnXn=+});\{H_{n}<H_{-1}\}\supset\big{(}\{X_{k}-X_{k-1}=+1,\,\forall 1\leq k\leq n_{0}\}\\ \cap\{X_{2n_{0}+k^{\prime}}>0,\,\forall k^{\prime}\geq n_{0}\}\cap\{\limsup_{n\rightarrow\infty}X_{n}=+\infty\}\big{)};

for, on the event on the right-hand side, one has that Xn0=n0X_{n_{0}}=n_{0} and the walk can in the worst case travel n0n_{0} steps to the left during the time interval (n0,2n0](n_{0},2n_{0}], whence in fact Xk0X_{k}\geq 0 for all k0k\geq 0, and Hn<H_{n}<\infty since lim supnXn=+\limsup_{n\rightarrow\infty}X_{n}=+\infty. Combining (3.14), the fact that the last event on the RHS has full probability due to (1.5)-(1.6), the fact that P(0,0)η(XkXk1=+1,1kn0)(pp)n0P^{\eta}_{(0,0)}(X_{k}-X_{k-1}=+1,\,\forall 1\leq k\leq n_{0})\geq(p_{\bullet}\wedge p_{\circ})^{n_{0}} on account of (2.5)-(2.6) and the Markov property of the quenched law at time n0n_{0}, one finds that

θn(ρ)(1.2)(pp)n0(n0,n0)ρ(Xn0+k>0,kn0)21(pp)n0>0,\theta_{n}(\rho)\stackrel{{\scriptstyle\eqref{eq:Bn}}}{{\geq}}(p_{\bullet}\wedge p_{\circ})^{n_{0}}\cdot\mathbb{P}_{(n_{0},n_{0})}^{\rho}(X_{n_{0}+k^{\prime}}>0,\,\forall k^{\prime}\geq n_{0})\geq 2^{-1}(p_{\bullet}\wedge p_{\circ})^{n_{0}}>0,

where the second inequality follows by choice of n0n_{0}. Thus, θ(ρ)>0\theta(\rho)>0 (see (1.3)), i.e. ρρc\rho\geq\rho_{c}. Letting ρρ+\rho\downarrow\rho_{+} one deduces that ρ+ρc\rho_{+}\geq\rho_{c}, and this completes the verification of (1.7). ∎

Proof of Theorem 1.1.

Only item (i) requires an explanation. This is an easy consequence of [40, Proposition 3.6], e.g. with the choice ε=(ρcρ)/4\varepsilon=(\rho_{c}-\rho)/4 for a given ρ<ρc\rho<\rho_{c}, and Theorem 1.2 (see (1.7)), as we now explain. Indeed one has that {Hn<H1}{Xnv(ρcε)n}\{H_{n}<H_{-1}\}\subset\{X_{n}\geq v(\rho_{c}-\varepsilon)n\} under ρ\mathbb{P}^{\rho} as soon as nn is large enough (depending on ρ\rho); to see the inclusion of events recall that ρc=ρ\rho_{c}=\rho_{-} on account of (1.7), which has already been proved, and therefore v(ρcε)<0v(\rho_{c}-\varepsilon)<0. The conclusion of item (i) now readily follows using the second estimate in [40, (3.27)]. ∎

3.3. The finite-range model 𝐏ρ,L\mathbf{P}^{\rho,L}

Following the strategy outlined in Section 1.3, we will aim at comparing the random walk in dynamic random environment, which has infinite-range correlations, to a finite-range model, which enjoys regeneration properties, and which we now introduce.

For a density ρJ\rho\in J (recall that JJ is an open interval of +\mathbb{R}_{+}) and an integer L1L\geq 1, we define a finite-range version of the environment, that is, a probability measure 𝐏ρ,L\mathbf{P}^{\rho,L} on Ση=(ηt(x):x,t+)\Sigma\ni\eta=(\eta_{t}(x):\,x\in\mathbb{Z},\,t\in\mathbb{R}_{+}) (see Section 2.2 for notation) such that the following holds. At every time tt multiple of LL, ηt\eta_{t} is sampled under 𝐏ρ,L\mathbf{P}^{\rho,L} according to μρ{\mu}_{\rho} (recall (P.2)), independently of (ηs)0s<t(\eta_{s})_{0\leq s<t} and, given ηt\eta_{t}, the process (ηt+s)0s<L(\eta_{t+s})_{0\leq s<L} has the same distribution under 𝐏ρ,L\mathbf{P}^{\rho,L} as (ηs)0s<L(\eta_{s})_{0\leq s<L} under 𝐏ηt{\bf P}^{\eta_{t}}, cf. Section 2.2 regarding the latter. We denote 𝐄ρ,L\mathbf{E}^{\rho,L} the expectation corresponding to 𝐏ρ,L\mathbf{P}^{\rho,L}. It readily follows that η\eta is a homogenous Markov process under 𝐏ρ,L\mathbf{P}^{\rho,L}, and that 𝐏ρ,L\mathbf{P}^{\rho,L} inherits all of Properties (P.1)-(2.2) from 𝐏ρ\mathbf{P}^{\rho}. In particular μρ{\mu}_{\rho} is still an invariant measure for the time-evolution of this environment. Note that 𝐏ρ,\mathbf{P}^{\rho,\infty} is well-defined and 𝐏ρ,=𝐏ρ\mathbf{P}^{\rho,\infty}=\mathbf{P}^{\rho}.

Recalling the quenched law PzηP_{z}^{\eta} of the walk XX in environment η\eta started at zz\in\mathbb{Z} from Section 2.3, we extend the annealed law of the walk from (2.2) by setting zρ,L[]=𝐏ρ,L(dη)Pzη[]\mathbb{P}_{z}^{\rho,L}[\cdot]=\int\mathbf{P}^{\rho,L}(d\eta)P_{z}^{\eta}[\cdot] so that zρ,=zρ\mathbb{P}^{\rho,\infty}_{z}=\mathbb{P}^{\rho}_{z} corresponds to the annealed law defined in (2.2). We also abbreviate ρ,L=0ρ,L\mathbb{P}^{\rho,L}=\mathbb{P}^{\rho,L}_{0}.

We now collect the key properties of finite-range models that will be used in the sequel. A straightforward consequence of the above definitions is that

(3.17) under ρ,L,{(XkL+sXkL)0sL:k} is an i.i.d. family,with common distribution identical to the ρ-law of (Xs)0sL.\begin{array}[]{l}\text{under }\mathbb{P}^{\rho,L},\left\{(X_{kL+s}-X_{kL})_{0\leq s\leq L}:k\in\mathbb{N}\right\}\text{ is an i.i.d.~family,}\\ \text{with common distribution identical to the }\mathbb{P}^{\rho}\text{-law of }(X_{s})_{0\leq s\leq L}.\end{array}

Moreover, by direct inspection one sees that 𝐏ρ,L\mathbf{P}^{\rho,L} inherits the properties listed in (3.1) from 𝐏ρ\mathbf{P}^{\rho}; that is, whenever 𝐏ρ\mathbf{P}^{\rho} does,

(3.18) 𝐏ρ,L\mathbf{P}^{\rho,L} satisfies (C.1)(C.2)(C.2.1)(C.2.2) (all for LtL\geq t) and (C.3) (for LL\geq\ell).

(more precisely, all of these conditions hold with 𝐏η,L\mathbf{P}^{\eta,L} in place of 𝐏η\mathbf{P}^{\eta} everywhere, where 𝐏η,L\mathbf{P}^{\eta,L} refers to the evolution under 𝐏ρ,L\mathbf{P}^{\rho,L} with initial condition η0=η\eta_{0}=\eta). The next result provides a well-defined monotonic speed vL()v_{L}(\cdot) for the finite-range model. This is an easy fact to check. A much more refined quantitative monotonicity result will follow shortly in Proposition 3.3 below (implying in particular strict monotonicity of vL()v_{L}(\cdot)).

Lemma 3.2 (Existence of the finite-range speed vLv_{L}).

For ρJ\rho\in J and an integer L1L\geq 1, let

(3.19) vL(ρ)=def.𝔼ρ,L[XL/L]=𝔼ρ[XL/L].v_{L}(\rho)\stackrel{{\scriptstyle\textnormal{def.}}}{{=}}\mathbb{E}^{\rho,L}\left[{X_{L}}/{L}\right]=\mathbb{E}^{\rho}\left[{X_{L}}/{L}\right].

Then

(3.20) ρ,L-a.s.limn+n1Xn=vL(ρ).\mathbb{P}^{\rho,L}\textnormal{-a.s.}\lim_{n\rightarrow+\infty}n^{-1}{X_{n}}=v_{L}(\rho).

Moreover, for any fixed LL, we have that

(3.21) the map ρvL(ρ)\rho\mapsto v_{L}(\rho) is non-decreasing on JJ.
Proof.

The second equality in (3.19) is justified by (3.17). The limit in (3.20) is an easy consequence of (3.17), (2.18), the definition (3.19) of vLv_{L} and the law of large numbers. The monotonicity (3.21) is obtained by combining (P.3) and Lemma 2.2. ∎

3.4. Key propositions and proof of Theorem 3.1

In this section, we provide two key intermediate results, stated as Propositions 3.3 and 3.4, which roughly correspond to the two parts of the proof outline in Section 1.3. Theorem 3.1 will readily follow from these results, and the proof appears at the end of this section. As explained in the introduction, the general strategy draws inspiration from recent interpolation techniques used in the context of sharpness results for percolation models with slow correlation decay, see in particular [31, 33], but the proofs of the two results stated below are vastly different.

Our first result is proved in Section 4 and allows us to compare the limiting speed of the full-range model to the speed of the easier finite-range model with a slightly different density. Note that, even though both v()v(\cdot) and vL()v_{L}(\cdot) are monotonic, there is a priori no clear link between vv and vLv_{L}.

Proposition 3.3 (Approximation of vv by vLv_{L}).

Under the assumptions of Theorem 3.1, there exists \ClC:approx=\CrC:approx(ν)(0,)\Cl{C:approx}=\Cr{C:approx}(\nu)\in(0,\infty) such that for all L3L\geq 3 and ρ\rho such that [ρ1logL,ρ+1logL]J[\rho-\frac{1}{\log L},\rho+\frac{1}{\log L}]\subseteq J,

(3.22) vL(ρ1logL)\CrC:approx(logL)100Lv(ρ)vL(ρ+1logL)+\CrC:approx(logL)100L.v_{L}\Big{(}\rho-\frac{1}{\log L}\Big{)}-\frac{\Cr{C:approx}(\log L)^{100}}{L}\leq v(\rho)\leq v_{L}\Big{(}\rho+\frac{1}{\log L}\Big{)}+\frac{\Cr{C:approx}(\log L)^{100}}{L}.

With Proposition 3.3 above, we now have a chance to deduce the strict monotonicity of v()v(\cdot) from that of vL()v_{L}(\cdot). Proposition 3.4 below, proved in Section 5, provides a quantitative strict monotonicity for the finite-range speed vL()v_{L}(\cdot). Note however that it is not easy to obtain such a statement, even for the finite-range model: indeed, from the definition of vL()v_{L}(\cdot) in (3.19), one can see that trying to directly compute the expectation in (3.19) for any LL boils down to working with the difficult full range model. One of the main difficulties is that the environment mixes slowly and creates strong space-time correlations. Nonetheless, using sprinkling methods, it turns out that one can speed-up the mixing dramatically by increasing the density of the environment. This is the main tool we use in order to obtain the result below.

Proposition 3.4 (Quantitative monotonicity of vLv_{L}).

Assume (3.18) holds and let ρJ\rho\in J. For all ϵ>0\epsilon>0 such that ρ+ϵJ\rho+\epsilon\in J, there exists L1=L1(ρ,ϵ,ν)1L_{1}=L_{1}(\rho,\epsilon,\color[rgb]{1,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{1,0,0}\nu\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0})\geq 1 such that, for all LL1L\geq L_{1},

(3.23) vL(ρ+ϵ)vL(ρ)3\CrC:approx(logL)100L.v_{L}(\rho+\epsilon)-v_{L}(\rho)\geq\frac{3\Cr{C:approx}\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}(\log L)^{100}}{L}.

Propositions 3.3 and 3.4 imply Theorem 3.1, as we now show.

Proof of Theorem 3.1..

Let ρ,ρJ\rho,\rho^{\prime}\in J with ρ>ρ\rho>\rho^{\prime}, and define ε=ρρ\varepsilon=\rho-\rho^{\prime}. Consider L3L1(ρ+ε/3,ε/3)L\geq 3\vee L_{1}(\rho+\varepsilon/3,\varepsilon/3), so that the conclusions of Propositions 3.3 and 3.4 both hold. By choosing LL sufficiently large in a manner depending on ρ\rho and ρ\rho^{\prime}, we can further ensure that (logL)1<ε/3(\log L)^{-1}<\varepsilon/3 and [ρ(logL)1,ρ+(logL)1][ρ(logL)1,ρ+(logL)1]J[\rho-(\log L)^{-1},\rho+(\log L)^{-1}]\cup[\rho^{\prime}-(\log L)^{-1},\rho^{\prime}+(\log L)^{-1}]\subseteq J. Abbreviating αL=L1(logL)100\alpha_{L}=L^{-1}{(\log L)^{100}}, it follows that

v(ρ)(3.22)vL(ρ(logL)1)\CrC:approxαL(3.21)vL(ρε3)\CrC:approxαL(3.21)vL(ρ+2ε3)\CrC:approxαL(3.23)vL(ρ+ε3)+2\CrC:approxαL(3.21)vL(ρ+(logL)1)+2\CrC:approxαL(3.22)v(ρ)+\CrC:approxαL>v(ρ),v(\rho)\stackrel{{\scriptstyle\eqref{eq:vLapprox}}}{{\geq}}v_{L}(\rho-(\log L)^{-1})-\Cr{C:approx}\alpha_{L}\stackrel{{\scriptstyle\eqref{eq:vLmonotonic}}}{{\geq}}v_{L}(\rho-\textstyle\frac{\varepsilon}{3})-\Cr{C:approx}\alpha_{L}\stackrel{{\scriptstyle\eqref{eq:vLmonotonic}}}{{\geq}}v_{L}(\rho^{\prime}+\frac{2\varepsilon}{3})-\Cr{C:approx}\alpha_{L}\\ \stackrel{{\scriptstyle\eqref{eq:vLincrease}}}{{\geq}}\textstyle v_{L}(\rho^{\prime}+\frac{\varepsilon}{3})+2\Cr{C:approx}\alpha_{L}\stackrel{{\scriptstyle\eqref{eq:vLmonotonic}}}{{\geq}}v_{L}(\rho^{\prime}+(\log L)^{-1})+2\Cr{C:approx}\alpha_{L}\stackrel{{\scriptstyle\eqref{eq:vLapprox}}}{{\geq}}v(\rho^{\prime})+\Cr{C:approx}\alpha_{L}>v(\rho^{\prime}),

yielding (3.11). ∎

4 Finite-range approximation of ρ\mathbb{P}^{\rho}

In this section, we prove Proposition 3.3, which allows us to compare the speed v()v(\cdot) of the full-range model to the speed vL()v_{L}(\cdot) (see (3.19)) of the finite-range model introduced in §3.3. The proof uses a dyadic renormalisation scheme. By virtue of the law(s) of large numbers, see (3.10) and (3.20), and for L0L_{0} and ρ\rho fixed, the speed v(ρ)v(\rho) ought to be close to the speed v2KL0(ρ)v_{2^{K}L_{0}}(\rho), for some large integer KK. Hence, if one manages to control the discrepancies between v2k+1L0(ρ)v_{2^{k+1}L_{0}}(\rho) and v2kL0(ρ)v_{2^{k}L_{0}}(\rho) for all 0kK10\leq k\leq K-1 and prove that their sum is small, then the desired proximity between v(ρ)v(\rho) and vL0(ρ)v_{L_{0}}(\rho) follows. This is roughly the strategy we follow except that at each step, we slightly increase or decrease the density ρ\rho (depending on which bound we want to prove), in order to weaken the (strong) correlations in the model. This is why we only compare v()v(\cdot) and vL()v_{L}(\cdot) for slightly different densities at the end. This decorrelation method is usually referred to as sprinkling; see e.g. [67, 74] for similar ideas in other contexts.

As a first step towards proving Proposition 3.3, we establish in the following lemma a one-step version of the renormalization, with a flexible scaling of the sprinkling (f(L)f(L) below) in anticipation of possible future applications. For xx\in\mathbb{R}, let x=max(x,0)x_{-}=\max(-x,0) denote the negative part of xx. Throughout the remainder of this section, we are always tacitly working under the assumptions of Theorem 3.1.

Lemma 4.1.

There exists L0=L0(ν)1L_{0}=L_{0}(\nu)\geq 1 such that for all LL0L\geq L_{0}, all f(L)[(logL)90,L1/10]f(L)\in[(\log L)^{90},L^{1/10}] and all ρ\rho such that (ρf(L)1/40,ρ+f(L)1/40)J(\rho-f(L)^{-1/40},\rho+f(L)^{-1/40})\subseteq J, the following holds: there exists a coupling L\mathbb{Q}_{L} of (Xs(i))0s2L(X^{(i)}_{s})_{0\leq s\leq 2L}, i=1,2i=1,2, such that X(1)ρ,LX^{(1)}\sim\mathbb{P}^{\rho,L}, X(2)ρ+ε,2LX^{(2)}\sim\mathbb{P}^{\rho+\varepsilon,2L} with ε=f(L)1/40\varepsilon=f(L)^{-1/40}, and

(4.1) L(min0s2L(Xs(2)Xs(1))f(L))ef(L)1/40.\mathbb{Q}_{L}\Big{(}\min_{0\leq s\leq 2L}\big{(}X^{(2)}_{s}-X^{(1)}_{s}\big{)}\leq-f(L)\Big{)}\leq e^{-f(L)^{1/40}}.

Consequently,

(4.2) 𝔼L[max0s2L(Xs(2)Xs(1))]2f(L),\mathbb{E}^{\mathbb{Q}_{L}}\Big{[}\max_{0\leq s\leq 2L}\big{(}X^{(2)}_{s}-X^{(1)}_{s}\big{)}_{-}\Big{]}\leq 2f(L),

and

(4.3) VarL(max0s2L(Xs(2)Xs(1)))2f(L)2.\emph{Var}^{\mathbb{Q}_{L}}\Big{(}\max_{0\leq s\leq 2L}\big{(}X^{(2)}_{s}-X^{(1)}_{s}\big{)}_{-}\Big{)}\leq 2f(L)^{2}.

The same conclusions hold with marginals X(1)ρ,2LX^{(1)}\sim\mathbb{P}^{\rho,2L} and X(2)ρ+ε,LX^{(2)}\sim\mathbb{P}^{\rho+\varepsilon,L} instead.

Proof.

Towards showing (4.1), let us first assume that

(4.4) there exists L0>3 such that for all LL0(ρ+log2L)J and there exists acoupling L of η(1)𝐏ρ,L and η(2)𝐏ρ+ε,2L s.t. L[G]1exp(f(L)1/40),\begin{split}&\text{there exists $L_{0}>3$ such that for all $L\geq L_{0}$, $(\rho+\log^{-2}L)\in J$ and there exists a}\\ &\text{coupling $\mathbb{Q}_{L}$ of $\eta^{(1)}\sim\mathbf{P}^{\rho,L}$ and $\eta^{(2)}\sim\mathbf{P}^{\rho+\varepsilon,2L}$ s.t. }\mathbb{Q}_{L}[G]\geq 1-\exp(-f(L)^{1/40}),\end{split}

where, setting t=f(L)2t=\lfloor\frac{f(L)}{2}\rfloor, the ‘good’ event GG is defined as

(4.7) G={ηs(1)(x)ηs(2)(x),(x,s)[3L,3L]×[0,L)}{ηs(1)(x)ηs(2)(x2t),(x,s)[3L,3L]×[L+t,2L)}.G=\big{\{}\eta^{(1)}_{s}(x)\leq\eta^{(2)}_{s}(x),\ \forall(x,s)\in[-3L,3L]\times[0,L)\big{\}}\\ \cap\big{\{}\eta^{(1)}_{s}(x)\leq\eta^{(2)}_{s}(x-2t),\ \forall(x,s)\in[-3L,3L]\times[L+t,2L)\big{\}}.

Given the above, we now extend the coupling L\mathbb{Q}_{L} to the random walks X(1)Pη(1)X^{(1)}\sim P^{\eta^{(1)}} and X(2)Pη(2)X^{(2)}\sim P^{\eta^{(2)}}, defined as in Section 2.3, up to time 2L2L. To do that, we only need to specify how we couple the collections of independent uniform random variables (Uw(1))w𝕃(U^{(1)}_{w})_{w\in\mathbb{L}} and (Uw(2))w𝕃(U^{(2)}_{w})_{w\in\mathbb{L}} (recall that 𝕃\mathbb{L} denotes space-time, see (2.4)) used to determine the steps of each random walk; the walks X(i)X^{(i)}, i=1,2i=1,2, up to time 2L2L are then specified in terms of (U(i),η(i))(U^{(i)},\eta^{(i)}) as in (2.5)-(2.6). Under L\mathbb{Q}_{L}, we let (Uw(1))w𝕃(U^{(1)}_{w})_{w\in\mathbb{L}} be i.i.d. uniform random variable on [0,1][0,1] and define, for (x,s)𝕃(x,s)\in\mathbb{L},

U(x,s)(2)={U(x,s)(1),if 0s<LU(x2t,s)(1),if Ls<2L\displaystyle U^{(2)}_{(x,s)}=\begin{cases}U^{(1)}_{(x,s)},&\text{if }0\leq s<L\\ U^{(1)}_{(x-2t,s)},&\text{if }L\leq s<2L\end{cases}

(for definiteness let U(x,s)(2)=U(x,s)(1)U^{(2)}_{(x,s)}=U^{(1)}_{(x,s)} when s2Ls\geq 2L). Clearly (Uw(2))w𝕃(U^{(2)}_{w})_{w\in\mathbb{L}} are i.i.d. uniform variables, hence X(2)X^{(2)} also has the desired marginal law. Now, we will explain why, on this coupling, (4.1) holds, and refer to Figure 4 for illustration. We will only consider what happens on the event GG defined in (4.7).

Refer to caption
Figure 4: The trajectory of X(1)X^{(1)} (resp. X(2)X^{(2)}) is pictured in blue (resp. red). The trajectory of XωX^{\omega} coincides with that of X(1)X^{(1)} during [L+t,2L][L+t,2L]. The trajectory of X~ω\widetilde{X}^{\omega} is in black (and shifted by 2t-2t in dashed black). On the event GG, the points X2L(i)X_{2L}^{(i)} for i=1,2i=1,2 both lie in the interval [X~Ltω2t,X~Ltω][\widetilde{X}^{\omega}_{L-t}-2t,\widetilde{X}^{\omega}_{L-t}], where w=(x1,L+t)w=(x_{1},L+t).

On GG, by Lemma 2.2 with K=[3L,3L]×[0,L]K=[-3L,3L]\times[0,L], we have that Xs(2)Xs(1)X^{(2)}_{s}\geq X^{(1)}_{s} for all 0sL0\leq s\leq L. This implies that XL+t(2)XL+t(1)2tX^{(2)}_{L+t}\geq X^{(1)}_{L+t}-2t owing to (2.18). Let η=η(1)\eta=\eta^{(1)}, define η~\widetilde{\eta} by η~n(x)=ηn(2)(x2t)\widetilde{\eta}_{n}(x)=\eta^{(2)}_{n}(x-2t) and w=(XL+t(1),L+t)w=(X^{(1)}_{L+t},L+t). Let XwX^{w} and X~w\widetilde{X}^{w} be random walks evolving on top of η\eta and η~\widetilde{\eta} respectively, and both using the collection of uniform random variables U(1)U^{(1)}. Then, conditionally on ηL+t(1)\eta^{(1)}_{L+t}, ηL+t(2)\eta^{(2)}_{L+t} and XL+t(1)X^{(1)}_{L+t}, and on the event {XL+t(1)=x1}\{X^{(1)}_{L+t}=x_{1}\}, XLtwX^{w}_{L-t} has the law of X2L(1)X^{(1)}_{2L} and X~Ltw2t\widetilde{X}^{w}_{L-t}-2t has the law of the position at time 2L2L of a random walk started at (x12t,L+t)(x_{1}-2t,L+t) and evolving on top of η(2)\eta^{(2)}, using the collection U(2)U^{(2)}. In particular, it evolves on the same environment as X(2)X^{(2)} and starts on the left of XL+t(2)X^{(2)}_{L+t}, so that X~LtwX2L(2)\widetilde{X}^{w}_{L-t}\leq X^{(2)}_{2L} (by Lemma 2.2 applied with K=×+K=\mathbb{Z}\times\mathbb{Z}^{+} and η=η~\eta=\widetilde{\eta}). On GG, using the second event in the intersection on the right-hand side of (4.7), we can again apply Lemma 2.2, now at time L+tL+t, with XX and X~\widetilde{X} as described and K=[3L+2t,3L2t]×[L+t,2L]K=[-3L+2t,3L-2t]\times[L+t,2L]. We obtain that XstwX~stwX^{w}_{s-t}\leq\widetilde{X}^{w}_{s-t} for all s[t,L]s\in[t,L]. Thus, subtracting 2t2t on both sides and using the previous facts, we have on GG that Xs(1)2tXs(2)X^{(1)}_{s}-2t\leq X^{(2)}_{s}, for all 0s2L0\leq s\leq 2L. All in all, we have proved that

G{max0s2L(Xs(2)Xs(1))2t},\displaystyle G\subseteq\big{\{}\max_{0\leq s\leq 2L}\big{(}X^{(2)}_{s}-X^{(1)}_{s}\big{)}_{-}\leq 2t\big{\}},

and therefore (4.1) follows from our assumption (4.4). The fact that (4.2) and (4.3) hold is a simple consequence of (4.1) and a straightforward computation using the deterministic inequality (Xs(2)Xs(1))4L\big{(}X^{(2)}_{s}-X^{(1)}_{s}\big{)}_{-}\leq 4L valid for all 0s2L0\leq s\leq 2L and the fact that LL0>3L\geq L_{0}>3 (note that for LL large enough, we have (2L)2ef(L)1/40f(L)(2L)^{2}e^{-f(L)^{1/40}}\leq f(L) for all f(L)[log90L,L1/10]f(L)\in[\log^{90}L,L^{1/10}]).

It remains to show that (4.4) holds, which brings into play several of the properties gathered in §3.1 and which hold by assumption. Let L0>3L_{0}>3 be such that for every LL0L\geq L_{0} and any choice of f(L)log90Lf(L)\geq\log^{90}L, with ε=f(L)1/40\varepsilon=f(L)^{-1/40} we have that ρ+εJ\rho+\varepsilon\in J and

(4.8) L>t>10100ν1+ν8+ν7\CrSEPcoupling8ε16,L>t>10^{100}\nu^{-1}+\nu^{-8}+\nu^{-7}\Cr{SEPcoupling}^{8}\varepsilon^{-16},

where t=f(L)/2t=\lfloor f(L)/2\rfloor as defined above (4.7).

We will provide a step-by-step construction of L\mathbb{Q}_{L}, in four steps. First, note that by (P.3)-ii), there exists a coupling L\mathbb{Q}_{L} of environments (ηt(1)(x):x,t[0,L))(\eta^{(1)}_{t}(x):x\in\mathbb{Z},t\in[0,L)) and (ηt(2)(x):x,t[0,L])(\eta^{(2)}_{t}(x):x\in\mathbb{Z},t\in[0,L]) such that under L\mathbb{Q}_{L}, η(1)𝐏ρ,L\eta^{(1)}\sim\mathbf{P}^{\rho,L} on the time interval [0,L)[0,L), η(2)𝐏ρ+ε,2L\eta^{(2)}\sim\mathbf{P}^{\rho+\varepsilon,2L} on the time interval [0,L][0,L] and such that L\mathbb{Q}_{L}-a.s., ηt(1)(x)ηt(2)(x)\eta^{(1)}_{t}(x)\leq\eta^{(2)}_{t}(x) for all xx\in\mathbb{Z} and t[0,L)t\in[0,L). We then extend L\mathbb{Q}_{L} to time LL for η(1)\eta^{(1)} by sampling ηL(1)μρ\eta^{(1)}_{L}\sim{\mu}_{\rho}, independently of (ηt(1))0t<L(\eta^{(1)}_{t})_{0\leq t<L} and (ηt(2))0tL(\eta^{(2)}_{t})_{0\leq t\leq L}. In particular the above already yields that the inequalities required as part of the event GG in (4.7) which involve (ηt(i))0t<L(\eta^{(i)}_{t})_{0\leq t<L} actually hold L\mathbb{Q}_{L}-a.s.

Second, letting =def.t1/4\ell\stackrel{{\scriptstyle\text{def.}}}{{=}}\lfloor t^{1/4}\rfloor, we define G1G_{1} to be the event that for every interval I[10L(1+ν),10L(1+ν)]I\subseteq[-10L(1+{\nu}),10L(1+{\nu})] of length /2|I|\lfloor\ell/2\rfloor\leq|I|\leq\ell, the inequalities ηL(1)(I)(ρ+ε/4)\eta^{(1)}_{L}(I)\leq(\rho+\varepsilon/4)\ell and ηL(2)(I)(ρ+3ε/4)\eta^{(2)}_{L}(I)\geq(\rho+3\varepsilon/4)\ell hold (see the beginning of §2.2 for notation). By (2.2) and a union bound over the choices of such intervals II, we get that

(4.9) L(G1c)40L(1+ν)exp(\Crdensitydevε2/32).\mathbb{Q}_{L}(G_{1}^{c})\leq 40\ell L(1+{\nu})\exp(-\Cr{densitydev}\varepsilon^{2}\ell/32).

At time LL and on G1cG_{1}^{c}, extend L\mathbb{Q}_{L} to the time interval [L,2L][L,2L] by letting η(1)\eta^{(1)} and η(2)\eta^{(2)} follow their dynamics PηL(1)P^{\eta^{(1)}_{L}} and PηL(2)P^{\eta^{(2)}_{L}} independently of each other (using (P.1)).

Third, we continue the construction of the coupling on G1G_{1} by applying (C.2.2) with H=10L(1+ν)H=10L(1+\nu), η0=ηL(2)(2t)\eta_{0}=\eta^{(2)}_{L}(\cdot-2t) and η0=ηL(1)\eta^{\prime}_{0}=\eta^{(1)}_{L}, checking the assumptions using (4.8) (note indeed that we have 1+|log3(νt)|(νt)1/81+|\log^{3}(\nu t)|\leq(\nu t)^{1/8} since νt>10100\nu t>10^{100}, and then that \CrSEPcouplingε2(νt)1/8νt1/4\Cr{SEPcoupling}\varepsilon^{-2}(\nu t)^{1/8}\leq\nu t^{1/4}). This implies that, given ηL(1)\eta^{(1)}_{L} and ηL(2)\eta^{(2)}_{L}, on G1G_{1}, there exists an extension of the coupling L\mathbb{Q}_{L} on the time interval [L,L+t][L,L+t] such that, defining

(4.10) G2={x[8L6νL,8L+6νL],ηL+t(1)(x)ηL+t(2)(x2t)},G_{2}=\big{\{}\forall x\in[-8L-6{\nu}L,8L+6{\nu}L],\,\eta^{(1)}_{L+t}(x)\leq\eta^{(2)}_{L+t}(x-2t)\big{\}},

we have

(4.11) L(G2c)10\CrSEPcoupling2tL(1+ν)exp(\CrSEPcoupling21νν+1ε2t1/4).\mathbb{Q}_{L}(G_{2}^{c})\leq 10\Cr{SEPcoupling2}tL(1+{\nu})\exp\big{(}\textstyle-\Cr{SEPcoupling2}^{-1}\frac{{\nu}}{{\nu}+1}\varepsilon^{2}t^{1/4}\big{)}.

At time L+tL+t and on G2cG_{2}^{c}, we extend L\mathbb{Q}_{L} on the time interval [L+t,2L][L+t,2L] by letting η(1)\eta^{(1)} and η(2)\eta^{(2)} follow their dynamics PηL+t(1)P^{\eta^{(1)}_{L+t}} and PηL+t(2)P^{\eta^{(2)}_{L+t}} independently of each other.

Fourth, we continue the construction of the coupling on G2G_{2} by applying (C.2.1) with η0=ηL+t(2)(2t)\eta_{0}=\eta^{(2)}_{L+t}(\cdot-2t) and η0=ηL+t(1)\eta^{\prime}_{0}=\eta^{(1)}_{L+t}, H=8L+6νLH=8L+6{\nu}L, k=1k=1 and tt in (C.2.1) equal to LtL-t. This implies that, given ηL+t(1)\eta^{(1)}_{L+t} and ηL+t(2)\eta^{(2)}_{L+t}, on G2G_{2}, there exists an extension of the coupling L\mathbb{Q}_{L} on the time interval [L+t,2L][L+t,2L] such that, defining

(4.12) G3={x[8L,8L],s[L+t,2L],ηs(1)(x)ηs(2)(x2t)},G_{3}=\big{\{}\forall x\in[-8L,8L],\,s\in[L+t,2L],\,\eta^{(1)}_{s}(x)\leq\eta^{(2)}_{s}(x-2t)\big{\}},

we have

(4.13) L(G3c)20exp(ν(Lt)/4).\mathbb{Q}_{L}(G_{3}^{c})\leq 20\exp(-{\nu}({L-t})/{4}).

Finally, note that G1G2G3GG_{1}\cap G_{2}\cap G_{3}\subseteq G, so that (4.4) is a straightforward consequence of  (4.9), (4.11) and (4.13) provided that L0L_{0} is chosen large enough, depending only on ν\nu (as well as \CrSEPcoupling2\Cr{SEPcoupling2}, \Crdensitydev\Cr{densitydev} and \CrSEPcoupling\Cr{SEPcoupling}). Note in particular that with our choices of f(L)f(L), ε\varepsilon above (4.8) and (cf(L)1/4)\ell(\geq cf(L)^{1/4}), we have that min(Lt,ε2t1/4,ε2)f(L)1/20\min(L-t,\varepsilon^{2}t^{1/4},\varepsilon^{2}\ell)\geq f(L)^{1/20}. All in all (4.4) follows. The case X(1)𝐏ρ,2LX^{(1)}\sim\mathbf{P}^{\rho,2L} and X(2)𝐏ρ+ε,LX^{(2)}\sim\mathbf{P}^{\rho+\varepsilon,L} can be treated in the same way, by means of an obvious analogue of (4.4). The remainder of the coupling (once the environments are coupled) remains the same. ∎

We are now ready to prove Proposition 3.3. The proof combines the law of large numbers for the speed together with Lemma 4.1 applied inductively over increasing scales. The rough strategy is as follows. We aim to compare the speed of the full-range model v(ρ)v(\rho) to the speed of the finite-range model vL(ρ)v_{L}(\rho) for some possibly large, but finite LL, and prove that these two are close. For δ<vL(ρ)\delta<v_{L}(\rho), we know that the probability for the finite-range model to go slower than speed δ\delta goes to 0 on account of Lemma 3.2. Thus, in a large box of size 2KL02^{K}L_{0}, the L0L_{0}-range model will most likely be faster than δ\delta as soon as KK is large enough. Lemma 4.1 is used over dyadic scales to control the discrepancies between the 2kL02^{k}L_{0}-range model and the 2k+1L02^{k+1}L_{0}-range model for all kk from 0 to K1K-1, and to prove that they are small. It will be seen to imply that with high probability, in a box of size 2KL02^{K}L_{0}, the 2KL02^{K}L_{0}-range model will be faster than δ\delta. Now, we only need to observe that when observed in a box of size 2KL02^{K}L_{0}, the 2KL02^{K}L_{0}-range model is equivalent to the full-range model. As Lemma 4.1 already hints at, this is but a simplified picture and the actual argument entails additional complications. This is because each increase in the range (obtained by application of Lemma 4.1) comes not only at the cost of slightly ‘losing speed,’ but also requires a compensation in the form of a slight increase in the density ρ\rho, and so the accumulation of these various effects have to be tracked and controlled jointly.

Proof of Proposition 3.3.

We only show the first inequality in Proposition 3.3, i.e. for LC(ν)L\geq C(\nu) and ρ\rho such that [ρ(logL)1,ρ+(logL)1]J[\rho-(\log L)^{-1},\rho+(\log L)^{-1}]\subseteq J, abbreviating αL=(logL)100L\alpha_{L}=\frac{(\log L)^{100}}{L}, one has

(4.14) vL(ρ(logL)1)αLv(ρ).v_{L}(\rho-(\log L)^{-1})-\alpha_{L}\leq v(\rho).

The first inequality in (3.22) then follows for all L3L\geq 3 by suitably choosing the constant \CrC:approx\Cr{C:approx} since v,vL[1,1]v,v_{L}\in[-1,1]. The second inequality of (3.22) is obtained by straightforward adaptation of the arguments below, using the last sentence of Lemma 4.1.

For L1L\geq 1, define Lk=2kLL_{k}=2^{k}L for k0k\geq 0. As we now explain, the conclusion (4.14) holds as soon as for LC(ν)L\geq C(\nu) and ρ\rho as above, we show that

(4.15) ρ((XLK/LK)vL(ρ(logL)1)αL)0, as K.\mathbb{P}^{\rho}\big{(}({X_{L_{K}}}/L_{K})\leq v_{L}(\rho-(\log L)^{-1})-\alpha_{L}\big{)}\to 0\text{, as }K\to\infty.

Indeed under the assumptions of Proposition 3.3, the law of large numbers (1.5) holds, and therefore in particular, for all δ>v(ρ)\delta>v(\rho), we have that ρ(XLKδLK)\mathbb{P}^{\rho}\left({X_{L_{K}}}\leq\delta{L_{K}}\right) tends to 11 as KK\to\infty. Together with (4.15) this is readily seen to imply (4.14).

We will prove (4.15) for LC(ν)L\geq C(\nu), where the latter is chosen such that the conclusions of Lemma 4.1 hold for LL, and moreover such that

(4.16) (logL)9/4+100(logL)5/41logLlogL5 and (32)k(klog2logL+1)991, for all k0.(\log L)^{-9/4}+\frac{100}{(\log L)^{5/4}}\leq\frac{1}{\log L}\text{, }\log L\geq 5\text{ and }\textstyle\left(\frac{3}{2}\right)^{-k}\big{(}k\frac{\log 2}{\log L}+1\big{)}^{99}\leq 1\text{, for all }k\geq 0.

For such LL we define, for all integer K,k0K,k\geq 0, all ρ>0\rho>0 and all δ\delta\in\mathbb{R}, recalling the finite-range annealed measures ρ,L\mathbb{P}^{\rho,L} from §3.3,

(4.17) pρ,K,δ(k)=ρ,Lk(XLKδLK),pρ,K,δ()=ρ(XLKδLK)\begin{split}&p_{\rho,K,\delta}^{(k)}=\mathbb{P}^{\rho,L_{k}}(X_{L_{K}}\leq\delta{L_{K}}),\\ &p_{\rho,K,\delta}^{(\infty)}=\mathbb{P}^{\rho}(X_{L_{K}}\leq\delta{L_{K}})\end{split}

(observe that the notation is consistent with §3.3, i.e. ρ,=ρ\mathbb{P}^{\rho,\infty}=\mathbb{P}^{\rho}). In this language (4.15) requires that pρ,K,δ()p_{\rho,K,\delta}^{(\infty)} vanishes in the limit KK\to\infty for a certain value of δ\delta. We start by gathering a few properties of the quantities in (4.17). For all ρJ\rho^{\prime}\in J, the following hold:

(4.18) limK+pρ,K,δ(0)=0 for all δ<vL(ρ) (by (3.20)),\displaystyle\lim_{K\rightarrow+\infty}p_{\rho^{\prime},K,\delta}^{(0)}=0\text{ for all }\delta<v_{L}(\rho^{\prime})\text{ (by \eqref{eq:LLNvL}),}
(4.19) pρ,K,δ()=pρ,K,δ(k), for all δ[1,1],K0 and kK,\displaystyle p_{\rho^{\prime},K,\delta}^{(\infty)}=p_{\rho^{\prime},K,\delta}^{(k)}\text{, for all }\delta\in[-1,1],\ K\geq 0\text{ and }k\geq K,
(4.20) pρ,K,δ() and pρ,K,δ(k) are non-increasing in ρ and non-decreasing in δ.\displaystyle p_{\rho^{\prime},K,\delta}^{(\infty)}\text{ and }p_{\rho^{\prime},K,\delta}^{(k)}\text{ are non-increasing in $\rho^{\prime}$ and non-decreasing in $\delta$.}

As explained atop the start of the proof, owing to the form of Lemma 4.1 we will need to simultaneously sprinkle the density and the speed we consider in order to be able to compare the range-Lk+1L_{k+1} model to the range-LkL_{k} model. To this effect, let

(4.21) ρ0=ρ(logL)1 and ρk+1=ρk+(logLk)9/4, for all k0,\rho_{0}=\rho-(\log L)^{-1}\text{ and }\rho_{k+1}=\rho_{k}+{(\log L_{k})^{-9/4}}\text{, for all }k\geq 0,

as well as

(4.22) δ0=vL(ρ0)L2 and δk+1=δklog99(Lk)/Lk, for all k0.\delta_{0}=v_{L}(\rho_{0})-{L^{-2}}\text{ and }\delta_{k+1}=\delta_{k}-{\log^{99}(L_{k})}/L_{k}\text{, for all }k\geq 0.

A straightforward computation, bounding the sum below for k1k\geq 1 by the integral 0𝑑x/(xlog2+logL)9/4\int_{0}^{\infty}dx/(x\log 2+\log L)^{9/4}, yields that

(4.23) limkρk=ρ0+k0(logLk)9/4ρ0+(logL)9/4+100(logL)5/4(4.16)ρ0+1logL(4.21)ρ.\lim_{k}\rho_{k}=\rho_{0}+\sum_{k\geq 0}(\log L_{k})^{-9/4}\leq\rho_{0}+(\log L)^{-9/4}+\frac{100}{(\log L)^{5/4}}\stackrel{{\scriptstyle\eqref{elodie}}}{{\leq}}\rho_{0}+\frac{1}{\log L}\stackrel{{\scriptstyle\eqref{eq:rho_k}}}{{\leq}}\rho.

Another straightforward computation yields that

(4.26) k0(logLk)99Lk=k0(32)k(klog2logL+1)99×(logL)99L(34)k(4.16)(logL)99Lk0(34)k(4.16)(logL)100L1L2.\sum_{k\geq 0}\frac{(\log L_{k})^{99}}{L_{k}}=\sum_{k\geq 0}\left(\frac{3}{2}\right)^{-k}\left(k\frac{\log 2}{\log L}+1\right)^{99}\times\frac{(\log L)^{99}}{L}\left(\frac{3}{4}\right)^{k}\\ \stackrel{{\scriptstyle\eqref{elodie}}}{{\leq}}\frac{(\log L)^{99}}{L}\sum_{k\geq 0}\left(\frac{3}{4}\right)^{k}\stackrel{{\scriptstyle\eqref{elodie}}}{{\leq}}\frac{(\log L)^{100}}{L}-\frac{1}{L^{2}}.

In particular, since (ρk)(\rho_{k}) is increasing in kk, (4.23) implies that for all K0K\geq 0, we have ρρK\rho\geq\rho_{K}, and (4.26) yields in view of (4.22) that δK>vL(ρ0)αL\delta_{K}>v_{L}(\rho_{0})-\alpha_{L}, with αL(=L1(logL)100)\alpha_{L}(=L^{-1}{(\log L)^{100}}) as above (4.14). Using this, it follows that, for all K0K\geq 0,

(4.29) pρ,K,vL(ρ0)αL()=(4.19)pρ,K,vL(ρ0)αL(K)(4.20)pρK,K,δK(K)=pρ0,K,δ0(0)+0k<K(pρk+1,K,δk+1(k+1)pρk,K,δk(k)).p^{(\infty)}_{\rho,K,v_{L}(\rho_{0})-\alpha_{L}}\stackrel{{\scriptstyle\eqref{propofp2}}}{{=}}p_{\rho,K,v_{L}(\rho_{0})-\alpha_{L}}^{(K)}\\ \stackrel{{\scriptstyle\eqref{propofp3}}}{{\leq}}p_{\rho_{K},K,\delta_{K}}^{(K)}=p_{\rho_{0},K,\delta_{0}}^{(0)}+\sum_{0\leq k<K}\left(p_{\rho_{k+1},K,\delta_{k+1}}^{(k+1)}-p_{\rho_{k},K,\delta_{k}}^{(k)}\right).

As the left-hand side in (4.29) is precisely equal to the probability appearing in (4.15), it is enough to argue that the right-hand side of (4.29) tends to 0 as KK\to\infty in order to conclude the proof. By recalling that δ0<vL(ρ0)\delta_{0}<v_{L}(\rho_{0}) from (4.22) and using (4.18), we see that limKpρ0,K,δ0(0)=0\lim_{K\to\infty}p_{\rho_{0},K,\delta_{0}}^{(0)}=0, which takes care of the first term on the right of (4.29).

We now aim to show that the sum over kk in (4.29) vanishes in the limit KK\to\infty, which will conclude the proof. Lemma 4.1 now comes into play. Indeed recalling the definition (4.17), the difference for fixed value of kk involves walks with range LkL_{k} and Lk+1L_{k+1}, and Lemma 4.1 supplies a coupling allowing good control on the negative part of this difference (when expressed under the coupling). Specifically, for a given K1K\geq 1 and 0kK10\leq k\leq K-1, let X(1)ρk+1,Lk+1X^{(1)}\sim\mathbb{P}^{\rho_{k+1},L_{k+1}} and X(2)ρk,LkX^{(2)}\sim\mathbb{P}^{\rho_{k},L_{k}}. Note that for i{1,2}i\in\{1,2\}, one has the rewrite

XLK(i)=0<2Kk1(X(+1)Lk+1(i)XLk+1(i)).X^{(i)}_{L_{K}}=\sum_{0\leq\ell<2^{K-k-1}}\left(X^{(i)}_{(\ell+1)L_{k+1}}-X^{(i)}_{\ell L_{k+1}}\right).

Therefore, due to the regenerative structure of the finite-range model, explicated in (3.17), it follows that, for i{1,2}i\in\{1,2\}, under ρk+2i,Lk+2i\mathbb{P}^{\rho_{k+2-i},L_{k+2-i}},

(4.30) XLK(i)=law0<2Kk1XLk+1(i,),X^{(i)}_{L_{K}}\stackrel{{\scriptstyle\text{law}}}{{=}}\sum_{0\leq\ell<2^{K-k-1}}X^{(i,\ell)}_{L_{k+1}},

where XLk+1(i,)X^{(i,\ell)}_{L_{k+1}}, 0\ell\geq 0, is a collection of independent copies of XLk+1(i)X^{(i)}_{L_{k+1}} under ρk+2i,Lk+2i\mathbb{P}^{\rho_{k+2-i},L_{{k+2-i}}}. Now recall the coupling measure L\mathbb{Q}_{L} provided by Lemma 4.1 with f(L)=(logL)90f(L)=(\log L)^{90} and let us denote by \mathbb{Q} the product measure induced by this couplingfor the choices L=LkL=L_{k} and ε=ρk+1ρk\varepsilon=\rho_{k+1}-\rho_{k} in Lemma 4.1, so that \mathbb{Q} supports the i.i.d. family of pairs (XLk+1(1,),XLk+1(2,))(X^{(1,\ell)}_{L_{k+1}},X^{(2,\ell)}_{L_{k+1}}), 0\ell\geq 0, each sampled under Lk\mathbb{Q}_{L_{k}}. In particular, under \mathbb{Q}, for all 0\ell\geq 0, XLk+1(1,)X^{(1,\ell)}_{L_{k+1}} and XLk+1(2,)X^{(2,\ell)}_{L_{k+1}} have law ρk+1,Lk+1\mathbb{P}^{\rho_{k+1},L_{k+1}} and ρk,Lk\mathbb{P}^{\rho_{k},L_{k}}, respectively. Now, one can write, for any K1K\geq 1 and 0kK10\leq k\leq K-1, with the sum over \ell ranging over 0<2Kk10\leq\ell<2^{K-k-1} below, that

pρk+1,K,δk+1(k+1)pρk,K,δk(k)=(4.17)(XLk+1(1,)LKδk+1)(XLk+1(2,)LKδk)(XLk+1(1,)LKδk+1,XLk+1(2,)>LKδk)(4.22)((XLk+1(1,)XLk+1(2,))2Kklog99(Lk)).p_{\rho_{k+1},K,\delta_{k+1}}^{(k+1)}-p_{\rho_{k},K,\delta_{k}}^{(k)}\stackrel{{\scriptstyle\eqref{def:prhok}}}{{=}}\mathbb{Q}\Big{(}\sum_{\ell}X^{(1,\ell)}_{L_{k+1}}\leq L_{K}\delta_{k+1}\Big{)}-\mathbb{Q}\Big{(}\sum_{\ell}X^{(2,\ell)}_{L_{k+1}}\leq L_{K}\delta_{k}\Big{)}\\ \leq\mathbb{Q}\Big{(}\sum_{\ell}X^{(1,\ell)}_{L_{k+1}}\leq L_{K}\delta_{k+1},\ \sum_{\ell}X^{(2,\ell)}_{L_{k+1}}>L_{K}\delta_{k}\Big{)}\stackrel{{\scriptstyle\eqref{def:delta0delta}}}{{\leq}}\mathbb{Q}\Big{(}\sum_{\ell}(X^{(1,\ell)}_{L_{k+1}}-X^{(2,\ell)}_{L_{k+1}})\leq-2^{K-k}\log^{99}(L_{k})\Big{)}.

Now, using Chebyshev’s inequality together with Lemma 4.1 (recall that f(L)=log90Lf(L)=\log^{90}L), it follows that for K1K\geq 1 and 0kK10\leq k\leq K-1,

(4.33) pρk+1,K,δk+1(k+1)pρk,K,δk(k)2Kk(logLk)190(2Kk(logLk)992Kk(logLk)90)272Kk(k+5)82K2+224K8,\begin{split}p_{\rho_{k+1},K,\delta_{k+1}}^{(k+1)}-p_{\rho_{k},K,\delta_{k}}^{(k)}&\leq\frac{2^{K-k}(\log L_{k})^{190}}{\big{(}2^{K-k}(\log L_{k})^{99}-2^{K-k}(\log L_{k})^{90}\big{)}^{2}}\\ &\leq\frac{7}{2^{K-k}(k+5)^{8}}\leq 2^{-\tfrac{K}{2}}+224K^{-8},\end{split}

where the second line is obtained by considering the cases when k<K/2k<K/2 or kK/2k\geq K/2 separately, together with straightforward computations. The bound (4.33) implies in turn that

k=0K1(pρk+1,K,δk+1(k+1)pρk,K,δk(k))K2K2+224K7K0,\sum_{k=0}^{K-1}\left(p_{\rho_{k+1},K,\delta_{k+1}}^{(k+1)}-p_{\rho_{k},K,\delta_{k}}^{(k)}\right)\leq K2^{-\tfrac{K}{2}}+224K^{-7}\stackrel{{\scriptstyle K\to\infty}}{{\longrightarrow}}0,

which concludes the proof. ∎

5 Quantitative monotonicity for the finite-range model

The goal of this section is to prove Proposition 3.4. For this purpose, recall that JJ is an open interval, and fix ρJ\rho\in J and ϵ>0\epsilon>0 such that (ρ+ϵ)J(\rho+\epsilon)\in J. Morever, in view of (3.21), we can assume that ϵ<1/100\epsilon<1/100. The dependence of quantities on ρ\rho and ϵ\epsilon will be explicit in our notation. As explained above Proposition 3.4, even if we are dealing with the finite-range model, the current question is about estimating the expectation in (3.19), which is actually equivalent to working on the full-range model. Thus, we retain much of the difficulty, including the fact that the environment mixes slowly. The upshot is that the speed gain to be achieved is quantified, and rather small, cf. the right-hand side of (3.23).

The general idea of the proof is as follows. First, recall that we want to prove that at time LL, the expected position of Xρ+ϵρ+ϵX^{\rho+\epsilon}\sim\mathbb{P}^{\rho+\epsilon} (where \sim denotes equality in law in the sequel) is larger than the expected position of XρρX^{\rho}\sim\mathbb{P}^{\rho} by 3(logL)1003(\log L)^{100}. The main conceptual input is to couple XρX^{\rho} and Xρ+ϵX^{\rho+\epsilon} in such a way that, after a well-chosen time TLT\ll L, we create a positive discrepancy between them with a not-so-small probability, and that this discrepancy is negative with negligible probability, allowing us to control its expectation. We will also couple these walks in order to make sure that the environment seen from XTρX^{\rho}_{T} is dominated by the environment seen from XTρ+ϵX^{\rho+\epsilon}_{T}, even if these walks are not in the same position, and we further aim for these environments to be ‘typical’. The last two items will allow us to repeat the coupling argument several times in a row and obtain a sizeable gap between XLρX^{\rho}_{L} and XLρ+ϵX^{\rho+\epsilon}_{L}. In more quantitative terms, we choose below T=5(logL)1000T=5(\log L)^{1000} and create an expected discrepancy of exp((logL)1/20)\exp(-(\log L)^{1/20}) at time TT. Repeating this procedure L/TL/T times provides us with an expected discrepancy at time LL larger than 3(logL)1003(\log L)^{100} (and in fact larger than L1+o(1)L^{1+o(1)}). We refer to the second part of Section 1.3 for a more extensive discussion of how the expected gap size comes about.

We split the proof of Proposition 3.4 into two parts. The main part (Section 5.1) consists of constructing a coupling along the above lines. In Section 5.3, we prove Proposition 3.4.

5.1. The trajectories Y±Y^{\pm}

In this section, we define two discrete-time processes YtY_{t}^{-} and Yt+Y^{+}_{t}, t[0,T]t\in[0,T], for some time horizon TT, see (5.2) below, which are functions of two deterministic environments η\eta^{-} and η+\eta^{+} and an array U=(Uw)w𝕃U=(U_{w})_{w\in\mathbb{L}} (see Section 2 for notation) of numbers in [0,1][0,1]. This construction will lead to a deterministic estimate of the difference Yt+YtY^{+}_{t}-Y^{-}_{t}, stated in Lemma 5.1. In the next section, see Lemma 5.2, we will prove that there exists a measure \mathbb{Q} on (η±,U)(\eta^{\pm},U) such that under \mathbb{Q}, YY^{-} dominates stochastically the law of a random walk XρρX^{\rho}\sim\mathbb{P}^{\rho} (see below (2.3) for notation), Y+Y^{+} is stochastically dominated by the law of a random walk Xρ+ϵρ+ϵX^{\rho+\epsilon}\sim\mathbb{P}^{\rho+\epsilon}, and we have a lower bound on 𝔼[YT+YT]\mathbb{E}^{\mathbb{Q}}[Y^{+}_{T}-Y^{-}_{T}], thus yielding a lower bound on 𝔼ρ+ϵ[Xρ+ϵ]𝔼ρ[Xρ]\mathbb{E}^{\rho+\epsilon}[X^{\rho+\epsilon}]-\mathbb{E}^{\rho}[X^{\rho}].


The construction of the two processes Y±Y^{\pm} will depend on whether some events are realized for η±\eta^{\pm} and UU. We will denote E1,E2,E_{1},E_{2},\ldots these events, which will occur (or not) successively in time. We introduce the convenient notation

(5.1) Eij=def.EiEi+1Ej, for all j>i1,E_{i-j}\stackrel{{\scriptstyle\text{def.}}}{{=}}E_{i}\cap E_{i+1}\cap\ldots\cap E_{j},\text{ for all $j>i\geq 1$,}

and write EijcE_{i-j}^{c} for the complement of EijE_{i-j}. We refer to Figure 5 (which is a refined version of Figure 2) for visual aid for the following construction of Y±Y^{\pm}, and to the discussion in Section 1.3 for intuition. Let us give a brief outlook on what follows. The outcome of the construction will depend on whether a sequence of events E1E_{1}-E9E_{9} happens or not: when all of these events occur, which we call a success, then YT+Y_{T}^{+} and YTY^{-}_{T} have a positive discrepancy and we retain a good control on the environments at time TT, namely η+\eta^{+} dominates η\eta^{-} (or, more precisely, η\eta^{-} shifted by YT+YTY_{T}^{+}-Y^{-}_{T}). When at least one of these nine events does not happen, this can either result in a neutral event, where the trajectories end up at the same position and we still have control on the environments, or it could result in a bad event, where we can have YT+<YTY_{T}^{+}<Y^{-}_{T} (but we will show in Lemma 5.2 that we keep control on the environments with overwhelming probability, owing to the parachute coupling evoked at the end of Section 1.3). Whenever we observe a neutral or a bad event, we will exit the construction and define the trajectories Y±Y^{\pm} at once from the time of observation all the way up to time TT. Finally, we note that the walks Y+Y^{+} and YY^{-} are actually not Markovian. This is however not a problem as we use them later to have bounds on the expected displacements of the actual walks XρX^{\rho} and Xρ+εX^{\rho+\varepsilon}.

Refer to caption
Figure 5: Trajectories of Y±Y^{\pm} during a time-interval of length TT. For readability, we only picture the scenario when the events EiE_{i} are all realized (except during [s3,T][s_{3},T]). The picture is not up to scale (all linear stretches on the trajectories of Y±Y^{\pm} have slope 1, those of the grey trapezoid have slope 2ν2\nu). The red disk represents a sprinkler (extra particle of η+η\eta^{+}\setminus\eta^{-}). The arrows on the left indicate which of the coupling conditions (C.1)-(C.3) from Section 3.1 is being used to perform the coupling in the given time interval. This is at the core of the proof of Lemma 5.2. The various couplings involved in the time interval [s3,T][s_{3},T] meet the necessity of having to restore domination of environments (possibly with a relative spatial shift) on a full space interval of length MM comparable to LL with high probability, so as to be able to repeat proceedings in the next step.

We now proceed to make this precise. For L3L\geq 3, define

(5.2) G=(logL)1000\ell_{G}=\lfloor(\log L)^{1000}\rfloor, g=(logL)1/1000\ell_{g}=\lfloor(\log L)^{1/1000}\rfloor and T=5GT=5\ell_{G},

as well as

(5.3) s0=G,s1=s0+g,s2=s1+2g20 and s3=s2+g20.s_{0}=\ell_{G},\,s_{1}=s_{0}+\ell_{g},\,s_{2}=s_{1}+2\ell_{g}^{20}\text{ and }s_{3}=s_{2}+\ell_{g}^{20}.

Let M10(ν+1)LM\geq 10(\nu+1)L, which will parametrize the spatial length of a space-time box in which the entire construction takes place, and let Y0±=0Y_{0}^{\pm}=0. The definition of Y±Y^{\pm} depends on the values of MM and LL, but we choose not to emphasize it in the notation. Given processes

(5.4) γ=(η+,η,U)\gamma=(\eta^{+},\eta^{-},U)

on a state space Ω\Omega such that the first two coordinates take values in (+)[0,T]×(\mathbb{Z}_{+})^{[0,T]\times\mathbb{Z}} and the last one in [0,1]𝕃[0,1]^{\mathbb{L}}, we will define the events Ei=Ei(γ)E_{i}=E_{i}(\gamma) below measurably in γ\gamma and similarly Y±=(Yt±(γ))0tTY^{\pm}=(Y^{\pm}_{t}(\gamma))_{0\leq t\leq T} with values in the set of discrete-time trajectories starting at 0. Unlike the sample paths of Xρ,Xρ+εX^{\rho},X^{\rho+\varepsilon}, the trajectories of Y±Y^{\pm} may perform jumps that are not to nearest neighbors. Probability will not enter the picture until Lemma 5.2 below, see in particular (5.51), which specifies the law of γ\gamma. We now properly define the aforementioned three scenarii of success, neutral events and bad events, which will be mutually exclusive, and we specify Y±Y^{\pm} in all cases.


Below ss and tt always denote integer times. Define the event (see Section 2.2 regarding the notation \succcurlyeq)

(5.5) E1=\displaystyle E_{1}= {s[0,s0],ηs+|[M+2νs0,M2νs0]ηs|[M+2νs0,M2νs0)]}.\displaystyle\{\forall s\in[0,s_{0}],\,\eta^{+}_{s}|_{[-M+2{\nu}s_{0},M-2{\nu}s_{0}]}\succcurlyeq\eta^{-}_{s}|_{[-M+2{\nu}s_{0},M-2{\nu}s_{0})]}\}.

The event E1E_{1} guarantees the domination of the environment η\eta^{-} by η+\eta^{+} and is necessary for a success, while E1cE_{1}^{c} will be part of the bad event. On E1E_{1}, for all 1ts01\leq t\leq s_{0}, let recursively

(5.6) Yt=Yt+=Yt1+A(ηt1(Yt1),U(Yt1,t1))Y^{-}_{t}=Y^{+}_{t}=Y^{-}_{t-1}+A(\eta^{-}_{t-1}(Y^{-}_{t-1}),U_{(Y^{-}_{t-1},t-1)})

where AA was defined in (2.5), so that we let the walks evolve together up to time s0s_{0}. On the bad event E1cE_{1}^{c}, we exit the construction by defining

(5.7) Yt=t and Yt+=t, for all 1tT.Y^{-}_{t}=t\text{ and }Y^{+}_{t}=-t\text{, for all }1\leq t\leq T.

Next, define the events

(5.8) E2=\displaystyle E_{2}= {ηs0+([Ys0,Ys0+g])>ηs0([Ys0,Ys0+g])}\displaystyle\{\eta^{+}_{s_{0}}([Y^{-}_{s_{0}},Y^{-}_{s_{0}}+\ell_{g}])>\eta^{-}_{s_{0}}([Y^{-}_{s_{0}},Y^{-}_{s_{0}}+\ell_{g}])\}
{ηs0([Ys03g+1,Ys0+3g])(ρ+1)6g},\displaystyle\cap\{\eta^{-}_{s_{0}}([Y^{-}_{s_{0}}-3\ell_{g}+1,Y^{-}_{s_{0}}+3\ell_{g}])\leq(\rho+1)\cdot 6\ell_{g}\},
(5.9) E3=\displaystyle E_{3}= {s[s0,s1],ηs+|[M+(4ν+1)s1,M(4ν+1)s1]ηs|[M+(4ν+1)s1,M(4ν+1)s1]}.\displaystyle\{\forall s\in[s_{0},s_{1}],\,\eta^{+}_{s}|_{[-M+(4\nu+1)s_{1},M-(4\nu+1)s_{1}]}\succcurlyeq\eta^{-}_{s}|_{[-M+(4\nu+1)s_{1},M-(4\nu+1)s_{1}]}\}.

The event E3E_{3} again ensures the domination of η\eta^{-} by η+\eta^{+} on a suitable spatial interval while E2E_{2} creates favourable conditions at time s0s_{0} to possibly see a sprinkler at time s1s_{1}. On the event E13E_{1-3} (recall (5.1) for notation), for all s0+1ts1s_{0}+1\leq t\leq s_{1}, let recursively

(5.10) Yt=Yt+=Yt1+A(ηt1(Yt1),U(Yt1,t1)),Y^{-}_{t}=Y^{+}_{t}=Y^{-}_{t-1}+A(\eta^{-}_{t-1}(Y^{-}_{t-1}),U_{(Y^{-}_{t-1},t-1)}),

so that the walks, from time s0s_{0}, continue to evolve together up to time s1s_{1} and, in particular, we have that

(5.11) on E13,Ys1=Ys1+.\text{on }E_{1-3},\quad Y^{-}_{s_{1}}=Y^{+}_{s_{1}}.

In order to deal with the case where E2E_{2} or E3E_{3} fail, we will distinguish two mutually exclusive cases, that will later contribute to an overall neutral and bad event, respectively; cf. (5.42) and (5.46). To this end, we introduce

(5.12) E2,bis={s[s0,T],ηs+|[M+2νT,M2νT]ηs|[M+2νT,M2νT]}.E_{2,\textnormal{bis}}=\{\forall s\in[s_{0},T],\,\eta^{+}_{s}|_{[-M+2{\nu}T,M-2{\nu}T]}\succcurlyeq\eta^{-}_{s}|_{[-M+2{\nu}T,M-2{\nu}T]}\}.

On E1E2cE2,bisE_{1}\cap E_{2}^{c}\cap E_{2,\textnormal{bis}}, which will be a neutral event, we exit the construction by letting the walk evolve together up to time TT, that is, we define

(5.13) Yt=Yt+=Yt1+A(ηt1(Yt1),U(Yt1,t1)), for all s0+1tT.Y^{-}_{t}=Y^{+}_{t}=Y^{-}_{t-1}+A(\eta^{-}_{t-1}(Y^{-}_{t-1}),U_{(Y^{-}_{t-1},t-1)})\text{, for all }s_{0}+1\leq t\leq T.

On the bad event (E1E2cE2,bisc)(E12E3c)\big{(}E_{1}\cap E_{2}^{c}\cap E_{2,\textnormal{bis}}^{c}\big{)}\cup\left(E_{1-2}\cap E_{3}^{c}\right), we exit the construction by defining

(5.14) Yt=t and Yt+=t, for all s0+1tT.Y^{-}_{t}=t\text{ and }Y^{+}_{t}=-t\text{, for all }s_{0}+1\leq t\leq T.

In the above, Y+Y^{+} and YY^{-} may take a non-nearest neighbour jump at time s0s_{0}, which is fine for our purpose. Next, define

(5.15) E4=\displaystyle E_{4}= {ηs1+(Ys1+)1,ηs1(Ys1)=0},\displaystyle\{\eta^{+}_{s_{1}}(Y_{s_{1}}^{+\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}})\geq 1,\eta^{-}_{s_{1}}(Y_{s_{1}}^{-\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}})=0\},
(5.16) E5=\displaystyle E_{5}= {s[s1,s2],ηs+|[M+(6ν+1)s2,M(6ν+1)s2]ηs|[M+(6ν+1)s2,M(6ν+1)s2]},\displaystyle\{\forall s\in[s_{1},s_{2}],\,\eta^{+}_{s}|_{[-M+(6\nu+1)s_{2},M-(6\nu+1)s_{2}]}\succcurlyeq\eta^{-}_{s}|_{[-M+(6\nu+1)s_{2},M-(6\nu+1)s_{2}]}\},

where E4E_{4} states that the walkers see a sprinkler at time s1s_{1} and E5E_{5} guarantees domination of the environments from time s1s_{1} to s2s_{2}. We will first define what happens on the neutral and the bad events. For this purpose, define

(5.17) E4,bis={s[s1,T],ηs+|[M+(4ν+1)T,M(4ν+1)T]ηs|[M+(4ν+1)T,M(4ν+1)T]}.E_{4,\textnormal{bis}}=\{\forall s\in[s_{1},T],\,\eta^{+}_{s}|_{[-M+(4{\nu}+1)T,M-(4{\nu}+1)T]}\succcurlyeq\eta^{-}_{s}|_{[-M+(4{\nu}+1)T,M-(4{\nu}+1)T]}\}.

Recall that on E13E_{1-3}, we have defined Y±Y^{\pm} up to time s1s_{1}. On the neutral event E13E45cE4,bisE_{1-3}\cap E_{4-5}^{c}\cap E_{4,\textnormal{bis}}, we define

(5.18) Yt=Yt+=Yt1+A(ηt1(Yt1),U(Yt1,t1)), for all s1+1tT.Y^{-}_{t}=Y^{+}_{t}=Y^{-}_{t-1}+A(\eta^{-}_{t-1}(Y^{-}_{t-1}),U_{(Y^{-}_{t-1},t-1)})\text{, for all }s_{1}+1\leq t\leq T.

On the bad event E13E45cE4,biscE_{1-3}\cap E_{4-5}^{c}\cap E_{4,\textnormal{bis}}^{c}, we exit the construction by defining

(5.19) Yt=t and Yt+=t, for all s1+1tT.Y^{-}_{t}=t\text{ and }Y^{+}_{t}=-t\text{, for all }s_{1}+1\leq t\leq T.

On the event E15E_{1-5}, because of the sprinkler, the walkers have a chance to split apart hence, for t[s1+1,s2]t\in[s_{1}+1,s_{2}], we let

(5.20) Yt=Yt1+A(ηt1(Yt1),U(Yt1,t1)) and Yt+=Yt1++A(ηt1+(Yt1+),U(Yt1+,t1)).Y^{-}_{t}=Y^{-}_{t-1}+A(\eta^{-}_{t-1}(Y^{-}_{t-1}),U_{(Y^{-}_{t-1},t-1)})\text{ and }Y^{+}_{t}=Y^{+}_{t-1}+A(\eta^{+}_{t-1}(Y^{+}_{t-1}),U_{(Y^{+}_{t-1},t-1)}).

Above, Y+Y^{+} and YY^{-} evolve on top of their respective environments η+\eta^{+} and η\eta^{-} from time s1+1s_{1}+1 to time s2s_{2}. To be able to continue the construction from time s2s_{2} to s3s_{3}, we define two events E6E_{6} and E7E_{7} concerning the environments from times s1s_{1} to s2s_{2}. If we are on E16E_{1-6\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}}, between times s1+1s_{1}+1 and s2s_{2} we will require that Y+Y^{+} and YY^{-} drift away, regardless of the states of η+\eta^{+} and η\eta^{-}, using only the information provided by UU. For this purpose, define

(5.21) E6=s1ts21E6t,E_{6}=\bigcap_{s_{1}\leq t\leq s_{2}-1}E_{6}^{t},

where we set

(5.22) E6s1={U(Ys1,s1)(p,p)} (recall that p>p) andE6t={U(Ys1(ts1),t)>p}{U(Ys1+ts1,t)<p} for s1+1ts21.\begin{split}&E^{s_{1}}_{6}=\big{\{}U_{(Y^{-}_{s_{1}},s_{1})}\in(p_{\circ},p_{\bullet})\big{\}}\text{ (recall that $p_{\bullet}>p_{\circ}$) and}\\ &E^{t}_{6}=\big{\{}U_{(Y_{s_{1}}^{-}-(t-s_{1})\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0},t)}>p_{\bullet}\big{\}}\cap\big{\{}U_{(Y^{-}_{s_{1}}+t-s_{1}\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0},t)}<p_{\circ}\big{\}}\text{ for }s_{1}+1\leq t\leq s_{2}-1.\end{split}

On E4E6s1E_{4}\cap E_{6}^{s_{1}}, YY^{-} steps to the left and Y+Y^{+} steps to the right from their common position, thus creating a gap at time s1+1s_{1}+1. Then for all s1<t<s2s_{1}<t<s_{2}, as long as E4E_{4} and E6s1E_{6}^{s_{1}} to E6t1E_{6}^{t-1} happen, YtY^{-}_{t} and Yt+Y^{+}_{t} are at distinct positions and E6tE_{6}^{t} allows YY^{-} to take one more step to the left and Y+Y^{+} one more step to the right. Hence, using what we have constructed so far on E15E_{1-5}, we have that

(5.23) on E16,Ys2+=Ys2+2(s2s1),\text{on }E_{1-6},\quad Y^{+}_{s_{2}}=Y^{-}_{s_{2}}+2(s_{2}-s_{1}),

and we still have the domination of η\eta^{-} by η+\eta^{+} at time s2s_{2}, cf. (5.16).

Since Y+Y^{+} and YY^{-} are no longer at the same position, we are going to momentarily allow to lose this domination in order to recreate it at time s3s_{3} but in a suitably shifted manner, namely, achieve that η+(Ys3++)|Iη(Ys3+)|I\eta^{+}(Y^{+}_{s_{3}}+\cdot)|_{I}\succcurlyeq\eta^{-}(Y^{-}_{s_{3}}+\cdot)|_{I} for a suitable interval II, see (5.33). To do so, we first need favourable conditions at time s2s_{2} encapsulated by the event

(5.26) E7={for all intervals I[Ys13(ν+1)T,Ys1+3(ν+1)T] with g2/2|I|g2,ηs2+(I)(ρ+3ϵ/4)|I| and ηs2(I)(ρ+ϵ/4)|I|}E_{7}=\left\{\begin{array}[]{c}\text{for all intervals }I\subseteq[Y^{-}_{s_{1}}-3(\nu+1)T,\,Y^{-}_{s_{1}}+3(\nu+1)T]\text{ with }\\ \lfloor\ell_{g}^{2}/2\rfloor\leq|I|\leq\ell_{g}^{2},\eta^{+}_{s_{2}}(I)\geq(\rho+3\epsilon/4)|I|\text{ and }\eta^{-}_{s_{2}}(I)\leq(\rho+\epsilon/4)|I|\end{array}\right\}

The above requires good empirical densities at time s2s_{2} on an interval of length of order TT centred around the common position of the walkers Y±Y^{\pm} at time s1s_{1}. On E17E_{1-7}, we do not precisely control the position of the walkers from time s2s_{2} to s3s_{3} and define, for all s2<ts3s_{2}<t\leq s_{3},

(5.27) YtYt1=1,Yt+Yt1+=1,Y_{t}^{-}-Y_{t-1}^{-}=1,\quad Y_{t}^{+}-Y_{t-1}^{+}=-1,

which corresponds to the worst case scenario assuming nearest-neighbour jumps. In particular, using (5.27), (5.23) and (5.3), we have that

(5.28) on E17,Ys3+Ys3=2g20.\text{on }E_{1-7},\quad Y^{+}_{s_{3}}-Y^{-}_{s_{3}}=2\ell_{g}^{20}.

We now need to consider the case where E67E_{6-7} fails, and we will again distinguish two types of failure. For this purpose, define

(5.29) E6,bis={s[s2,T],ηs+|[M+(6ν+1)T,M(6ν+1)T]ηs|[M+(6ν+1)T,M(6ν+1)T]},E_{6,\textnormal{bis}}=\{\forall s\in[s_{2},T],\,\eta^{+}_{s}|_{[-M+(6\nu+1)T,M-(6\nu+1)T]}\succcurlyeq\eta^{-}_{s}|_{[-M+(6\nu+1)T,M-(6\nu+1)T]}\},

and, on the neutral event E15E6,7cE6,bisE_{1-5}\cap E_{6,7}^{c}\cap E_{6,\textnormal{bis}}, we merge Y+Y^{+} with YY^{-} at time s2+1s_{2}+1 and then let them walk together, by defining

(5.30) Ys2+1+\displaystyle Y_{s_{2}+1}^{+} =Ys2+1=Ys2+A(ηs2(Ys2),U(Ys2,s2)), and\displaystyle=Y_{s_{2}+1}^{-}=Y_{s_{2}}^{-}+A\big{(}\eta_{s_{2}}^{-}(Y^{-}_{s_{2}}),U_{(Y^{-}_{s_{2}},s_{2})}\big{)},\text{ and }
(5.31) Yt\displaystyle Y^{-}_{t} =Yt+=Yt1+A(ηt1(Yt1),U(Yt1,t1)), for all s2+2tT\displaystyle=Y^{+}_{t}=Y^{-}_{t-1}+A(\eta^{-}_{t-1}(Y^{-}_{t-1}),U_{(Y^{-}_{t-1},t-1)})\text{, for all }s_{2}+2\leq t\leq T

(above one line would be sufficient but we single out (5.30) because the merging will typically occasion a jump for Y+Y^{+}). On the remaining bad event E15E67cE6,biscE_{1-5}\cap E_{6-7}^{c}\cap E_{6,\textnormal{bis}}^{c}, we exit the construction in the now usual way by defining

(5.32) Yt=t and Yt+=t, for all s2+1tT.Y^{-}_{t}=t\text{ and }Y^{+}_{t}=-t\text{, for all }s_{2}+1\leq t\leq T.

It remains to define Y±Y^{\pm} from time s3s_{3} to TT on the event E17E_{1-7}. The good event will require

(5.33) E8=def.{ηs3+(+g20)|[Ys12(ν+1)T,Ys1+2(ν+1)T]ηs3(g20)|[Ys12(ν+1)T,Ys1+2(ν+1)T]},E_{8}\stackrel{{\scriptstyle\text{def.}}}{{=}}\{\eta^{+}_{s_{3}}(\cdot+\ell_{g}^{20})|_{[Y^{-}_{s_{1}}-2(\nu+1)T,Y^{-}_{s_{1}}+2(\nu+1)T]}\succcurlyeq\eta^{-}_{s_{3}}(\cdot-\ell_{g}^{20})|_{[Y^{-}_{s_{1}}-2(\nu+1)T,Y^{-}_{s_{1}}+2(\nu+1)T]}\},

that is, we want that the environment η+\eta^{+} seen from Ys3+Y^{+}_{s_{3}} covers η\eta^{-} seen from Ys3Y^{-}_{s_{3}} on an interval of length of order TT, which will enable us to let the walkers move in parallel (i.e. taking the same steps at the same time), even if they are at different positions. Moreover, we want this domination to persist from time s3s_{3} to time TT, hence we require

(5.34) E9=def.{s[s3,T],ηs+(+g20)|[Ys12T,Ys1+2T]ηs(g20)|[Ys12T,Ys1+2T]}.E_{9}\stackrel{{\scriptstyle\text{def.}}}{{=}}\{\forall s\in[s_{3},T],\,\eta^{+}_{s}(\cdot+\ell_{g}^{20})|_{[Y^{-}_{s_{1}}-2T,Y^{-}_{s_{1}}+2T]}\succcurlyeq\eta^{-}_{s}(\cdot-\ell_{g}^{20})|_{[Y^{-}_{s_{1}}-2T,Y^{-}_{s_{1}}+2T]}\}.

On the good event E19E_{1-9}, we let, for t[s3+1,T]t\in[s_{3}+1,T],

(5.35) Yt=Yt1+A(ηt1(Yt1),U(Yt1,t1)) and Yt+=Yt1++A(ηt1+(Yt1+),U(Yt1+2g20,t1)).Y^{-}_{t}=Y^{-}_{t-1}+A(\eta^{-}_{t-1}(Y^{-}_{t-1}),U_{(Y^{-}_{t-1},t-1)})\text{ and }Y^{+}_{t}=Y^{+}_{t-1}+A(\eta^{+}_{t-1}(Y^{+}_{t-1}),U_{(Y^{+}_{t-1}-2\ell_{g}^{20},t-1)}).

Note that above, we choose to shift spatially the collection (Uw)(U_{w}) by 2g202\ell_{g}^{20} spatially for Y+Y^{+}, which corresponds to the difference between YY^{-} and Y+Y^{+}. Therefore, both walks are using the same UwU_{w} to determine their next step. Using (5.35), (5.28) and (5.34), one thus proves recursively that

(5.36) on E19,Yt+Yt=2g20, for all s3tT.\text{on }E_{1-9},\,Y^{+}_{t}-Y^{-}_{t}=2\ell_{g}^{20}\text{, for all }s_{3}\leq t\leq T.

Finally, on the bad event E17cE89cE_{1-7}^{c}\cap E_{8-9}^{c}, we finish the construction by defining

(5.37) Yt=Yt1+1,Yt+=Yt1+1, for all s3+1tT1, and YT=T,YT+=T.Y^{-}_{t}=Y^{-}_{t-1}+1,\,Y^{+}_{t}=Y^{+}_{t-1}-1\text{, for all }s_{3}+1\leq t\leq T-1,\text{ and }Y^{-}_{T}=T,Y^{+}_{T}=-T.

This ends the definition of the trajectories Y+Y^{+} and YY^{-}. Let us emphasize once more that these trajectories are not nearest-neighbour and not Markovian w.r.t. the canonical filtration associated to (η+,η,U)(\eta^{+},\eta^{-},U), but that this will not prevent us from obtaining the desired bounds.

Below, we proceed to define our three key events and summarise in Lemma 5.1 some of the important (deterministic) properties we will use.

  • Scenario I: Good event. We define

    (5.38) Egood=def.E19.E_{\textnormal{good}}\stackrel{{\scriptstyle\text{def.}}}{{=}}E_{1-9}.

    Notice that by (5.36), on the event EgoodE_{\textnormal{good}} we have that

    (5.39) YT+YT=2g20>0,Y^{+}_{T}-Y^{-}_{T}=2\ell_{g}^{20}>0,

    and also that

    (5.40) s[0,s2],ηs+|[T,T]ηs|[T,T],\forall s\in[0,s_{2}],\,\eta^{+}_{s}|_{[-T,T]}\succcurlyeq\eta^{-}_{s}|_{[-T,T]},

    which follows by (5.16), (5.9) and (5.5), provided that M(6ν+1)s2TM-(6\nu+1)s_{2}\geq T. Finally, (5.34) and the fact that on EgoodE_{\textnormal{good}}, |Ys1|s1|Y^{-}_{s_{1}}|\leq s_{1} (by (5.6) and (5.10)) imply that

    (5.41) s[s3,T],ηs+(+2g20)|[T2g20,T2g20]ηs|[T,T].\forall s\in[s_{3},T],\,\eta^{+}_{s}(\cdot+2\ell_{g}^{20})|_{[-T-2\ell_{g}^{20},T-2\ell_{g}^{20}]}\succcurlyeq\eta^{-}_{s}|_{[-T,T]}.
  • Scenario II: Neutral event. We define

    (5.42) Eneutral=def.(E1E2cE2,bis)(E13E45cE4,bis)(E15E67cE6,bis).E_{\textnormal{neutral}}\stackrel{{\scriptstyle\text{def.}}}{{=}}\left(E_{1}\cap E_{2}^{c}\cap E_{2,\textnormal{bis}}\right)\cup\left(E_{1-3}\cap E_{4-5}^{c}\cap E_{4,\textnormal{bis}}\right)\cup\left(E_{1-5}\cap E_{6-7}^{c}\cap E_{6,\textnormal{bis}}\right).

    Note that on EneutralE_{\textnormal{neutral}}, (5.30)-(5.31), (5.18) and (5.13) imply that

    (5.43) YT+=YT\begin{split}&Y^{+}_{T}=Y^{-}_{T}\end{split}

    and (5.31), (5.30), (5.20), (5.18), (5.13), (5.10) and (5.6) imply that

    (5.44) Yt=Yt1+A(ηt1(Yt1),U(Yt1,t1)), for all 1tT.Y^{-}_{t}=Y^{-}_{t-1}+A(\eta^{-}_{t-1}(Y^{-}_{t-1}),U_{(Y^{-}_{t-1},t-1)})\text{, for all }1\leq t\leq T.

    From (5.29), (5.17), (5.16), (5.12), (5.9) and (5.5), we also have

    (5.45) s[0,T],ηs+|[T,T]ηs|[T,T],\forall s\in[0,T],\,\eta^{+}_{s}|_{[-T,T]}\succcurlyeq\eta^{-}_{s}|_{[-T,T]},

    provided that M(6ν+1)TTM-(6\nu+1)T\geq T.

  • Scenario III: Bad event. We finally define the event

    (5.46) Ebad=def.(EgoodEneutral)c=E1c(E1E2cE2,bisc)(E12E3c)(E13E45cE4,bisc)(E15E67cE6,bisc)(E17E89c).\begin{split}E_{\textnormal{bad}}&\stackrel{{\scriptstyle\text{def.}}}{{=}}\left(E_{\textnormal{good}}\cup E_{\textnormal{neutral}}\right)^{c}=E_{1}^{c}\cup\left(E_{1}\cap E_{2}^{c}\cap E_{2,\textnormal{bis}}^{c}\right)\cup\left(E_{1-2}\cap E_{3}^{c}\right)\cup\\ &\left(E_{1-3}\cap E_{4-5}^{c}\cap E_{4,\textnormal{bis}}^{c}\right)\cup\left(E_{1-5}\cap E_{6-7}^{c}\cap E_{6,\textnormal{bis}}^{c}\right)\cup\left(E_{1-7}\cap E_{8-9}^{c}\right).\end{split}

    In particular, (5.37), (5.32), (5.19), (5.14), and (5.7) yield

    (5.47) YT+=YT2T=T on the event Ebad.Y^{+}_{T}=Y^{-}_{T}-2T=-T\text{ on the event }E_{\textnormal{bad}}.

The next deterministic lemma is a restatement of (5.39), (5.43) and (5.47).

Lemma 5.1.

The event EgoodE_{\textnormal{good}}, EneutralE_{\textnormal{neutral}} and EbadE_{\textnormal{bad}} defined in (5.36), (5.42) and (5.46), respectively, form a partition of Ω\Omega such that

(5.48) YT+\displaystyle{Y}_{T}^{+} =YT+2g20 on Egood;\displaystyle=Y_{T}^{-}+2\ell_{g}^{20}\text{ on }E_{\textnormal{good}};
(5.49) YT+\displaystyle{Y}^{+}_{T} =YT on Eneutral;\displaystyle=Y_{T}^{-}\text{ on }E_{\textnormal{neutral}};
(5.50) YT+\displaystyle{Y}_{T}^{+} =YT2T on Ebad.\displaystyle=Y^{-}_{T}-2T\text{ on }E_{\textnormal{bad}}.

5.2. The coupling \mathbb{Q}

We now aim to compare Y±Y^{\pm} to XρX^{\rho} and Xρ+εX^{\rho+\varepsilon}, and to integrate over the dynamics of the environments when started from a typical initial configuration. We will derive from this a bound on the expected discrepancy between Y+Y^{+} and YY^{-}, and thus on the one between XρX^{\rho} and Xρ+εX^{\rho+\varepsilon}. The main result of this section is Lemma 5.2, which entails a coupling \mathbb{Q} with these features. Lemma 5.2 is the key ingredient in the proof of Proposition 3.4, which appears in the next subsection.

We will require the environments η+\eta^{+} and η\eta^{-} to have a typical initial configuration under 𝐏ρ+ϵ\mathbf{P}^{\rho+\epsilon} and 𝐏ρ\mathbf{P}^{\rho} in the following sense. Recall that ρJ\rho\in J and ϵ(0,1/100)\epsilon\in(0,1/100) have been fixed at the start of this section, and that (ρ+ϵ)J(\rho+\epsilon)\in J. For M,LM,L\in\mathbb{N}, we say that (η0+,η0)Σ2(\eta^{+}_{0},\eta^{-}_{0})\in\Sigma^{2} (recall that Σ=(+)\Sigma=(\mathbb{Z}_{+})^{\mathbb{Z}} from Section 2.2) is (M,L)(M,L)-balanced if all of the following occur:

  1. (i)

    η0+(x)η0(x)\eta^{+}_{0}(x)\geq\eta^{-}_{0}(x) for all x[M,M]x\in[-M,M],

  2. (ii)

    η0+([x,x+(logL)21])(ρ+99ϵ100)(logL)2\eta^{+}_{0}([x,x+\lfloor(\log L)^{2}\rfloor-1])\geq(\rho+\tfrac{99\epsilon}{100})\lfloor(\log L)^{2}\rfloor for all x[M,M(logL)2+1]x\in[-M,M-\lfloor(\log L)^{2}\rfloor+1],

  3. (iii)

    η0([x,x+(logL)21])(ρ+ϵ100)(logL)2\eta^{-}_{0}([x,x+\lfloor(\log L)^{2}\rfloor-1])\leq(\rho+\tfrac{\epsilon}{100})\lfloor(\log L)^{2}\rfloor for all x[M,M(logL)2+1]x\in[-M,M-\lfloor(\log L)^{2}\rfloor+1].

Lemma 5.2.

There exists L2=L2(ρ,ϵ,ν)1L_{2}=L_{2}(\rho,\epsilon,\nu\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0})\geq 1 such that for all LL2L\geq L_{2}, the following holds. For all M[10(ν+1)L,20(ν+1)L]M\in[10(\nu+1)L,20(\nu+1)L] and for every (M,L)(M,L)-balanced choice of (η0+,η0)(\eta^{+}_{0},\eta^{-}_{0}), there exists a coupling =(η0+,η0)\mathbb{Q}=\mathbb{Q}_{(\eta^{+}_{0},\eta^{-}_{0})} of (η+,η,U)(\eta^{+},\eta^{-},U) with the following properties:

(5.51) η𝐏η0,η+𝐏η0+\eta^{-}\sim\mathbf{P}^{\eta^{-}_{0}},\ \eta^{+}\sim\mathbf{P}^{\eta^{+}_{0}}, and (Uw)w𝕃\left(U_{w}\right)_{w\in\mathbb{L}} are i.i.d. uniform variables on [0,1][0,1];
(5.52) There exist X±=X±(η±,U)X^{\pm}=X^{\pm}(\eta^{\pm},U) such that X±η0±X^{\pm}\sim\mathbb{P}^{\eta^{\pm}_{0}} and \mathbb{Q}-a.s., the inequalities XTYTX^{-}_{T}\leq Y^{-}_{T} and YT+XT+Y^{+}_{T}\leq X^{+}_{T} hold, with Y±=Y±(η+,η,U)Y^{\pm}=Y^{\pm}(\eta^{+},\eta^{-},U) as in Section 5.1;
(5.53) 𝔼[YT+YT]exp((logL)1/20);\displaystyle\mathbb{E}^{\mathbb{Q}}\left[{Y}^{+}_{T}-Y^{-}_{T}\right]\geq\exp\big{(}-\left(\log L\right)^{1/20}\big{)};
(5.54) (Erestart)1L100, where Erestart=def.{x[(M(8ν+3)T),M(8ν+3)T],ηT+(x+YT+)ηT(x+YT)}.\displaystyle\begin{split}&\mathbb{Q}\left(E_{\textnormal{restart}}\right)\geq 1-L^{-100},\text{ where }\\ &E_{\textnormal{restart}}\stackrel{{\scriptstyle\textnormal{def.}}}{{=}}\{\forall x\in[-(M-(8\nu+3)T),M-(8\nu+3)T],\,\eta^{+}_{T}(x+{Y}^{+}_{T})\geq\eta^{-}_{T}(x+Y^{-}_{T})\}.\end{split}
Proof.

Let L21L_{2}\geq 1 to be chosen later, and LL2L\geq L_{2}. Fix M[10(ν+1)L,20(ν+1)L]M\in[10(\nu+1)L,20(\nu+1)L], and let (η0+,η0)(\eta_{0}^{+},\eta_{0}^{-}) be (M,L)(M,L)-balanced. We construct the coupling \mathbb{Q} below. We will then define XX^{-} and X+X^{+} such that (5.52) is satisfied. To prove the main estimate (5.53), we will control the probabilities of the events EgoodE_{\textnormal{good}}, EneutralE_{\textnormal{neutral}} and EbadE_{\textnormal{bad}} emerging from the construction of Y±Y^{\pm}, and use Lemma 5.1. Finally, proving (5.54) will require to bound the probability of losing the synchronisation of η+\eta^{+} and η\eta^{-} by time TT.

We split the proof into five parts: the construction of \mathbb{Q} and the proofs of (5.51)-(5.54).

Part I: Construction of \mathbb{Q}.

We will denote the natural filtration generated by the triplet (η+,η,U)(\eta^{+},\eta^{-},U) by

(5.55) t=σ{(ηt+)0tt,(ηt)0tt,(Uw)w𝕃,π2(w)t1}, for t0.\mathcal{F}_{t}=\sigma\{(\eta^{+}_{t^{\prime}})_{0\leq t^{\prime}\leq t},(\eta^{-}_{t^{\prime}})_{0\leq t^{\prime}\leq t},(U_{w})_{w\in\mathbb{L},\pi_{2}(w)\leq t-1}\}\text{, for }t\geq 0.

Under \mathbb{Q}, we first let U=(Uw)w𝕃U=(U_{w})_{w\in\mathbb{L}} be a collection of i.i.d. uniform random variables on [0,1][0,1]. We now define the coupling of η\eta^{-} and η+\eta^{+} under \mathbb{Q} in such a way that given t\mathcal{F}_{t}, the evolution of (ηs±)tst+1(\eta^{\pm}_{s})_{t\leq s\leq t+1} has the right marginal 𝐏ηt±\mathbf{P}^{\eta^{\pm}_{t}} (up to time one), cf. (P.1). Moreover, throughout the construction of \mathbb{Q}, i.e. from (5.57) to (5.73) below, we will in fact ensure that for all integer t[0,T]t\in[0,T], under \mathbb{Q},

(5.56) (Uw:π2(w)=t) is independent from t.\text{$(U_{w}:\pi_{2}(w)=t)$ is independent from $\mathcal{F}_{t}$}.

We now proceed to specify η±\eta^{\pm} under \mathbb{Q}, and refer again to Figure 5 for visual aid. On the time-interval [0,s0][0,s_{0}], we

(5.57) couple (η+,η)(\eta^{+},\eta^{-}) as (η,η)(\eta,\eta^{\prime}) in (C.2.1) with η0=η0+\eta_{0}=\eta^{+}_{0}, η0=η0\eta^{\prime}_{0}=\eta^{-}_{0}, H=MH=M, t=s0t=s_{0} and k=1k=1;

in particular, the domination η0|[H,H]η0|[H,H]\eta_{0}|_{[-H,H]}\succcurlyeq\eta_{0}^{\prime}|_{[-H,H]} required by (C.2.1) is ensured by the fact that (η0+,η0)(\eta^{+}_{0},\eta^{-}_{0}) is (M,L)(M,L)-balanced; see item (i) in the corresponding definition above Lemma 5.2. Since E1s0E_{1}\in\mathcal{F}_{s_{0}} by (5.5) and (5.55), we can observe at time s0s_{0} whether E1E_{1} occurred. Conditionally on s0\mathcal{F}_{s_{0}} and on E1cE_{1}^{c}, during [s0,T][s_{0},T], we

(5.58) let η\eta^{-} and η+\eta^{+} evolve independently with respective marginals 𝐏ηs0\mathbf{P}^{\eta^{-}_{s_{0}}} and 𝐏ηs0+\mathbf{P}^{\eta^{+}_{s_{0}}}.

Recalling the definition (5.8), we have that E2s0E_{2}\in\mathcal{F}_{s_{0}}. Conditionally on s0\mathcal{F}_{s_{0}} and on E1E2cE_{1}\cap E_{2}^{c}, we couple (η+,η)(\eta^{+},\eta^{-}) during the interval [s0,T][s_{0},T] as

(5.59) (η,η)(\eta,\eta^{\prime}) in (C.2.1) with η0=ηs0+\eta_{0}=\eta^{+}_{s_{0}}, η0=ηs0\eta^{\prime}_{0}=\eta^{-}_{s_{0}}, H=M2νs0H=M-2\nu s_{0}, t=Ts0t=T-s_{0} and k=1k=1.

On E12E_{1-2}, note that |Ys0|s0<s1|Y^{-}_{s_{0}}|\leq s_{0}<s_{1}. Hence, as we now explain, during the time-interval [s0,s1][s_{0},s_{1}], conditionally on s0\mathcal{F}_{s_{0}} and on E12E_{1-2}, we can

(5.60)  apply the coupling of (C.3) to ηt()=ηs0+t+(+Ys0) and ηt()=ηs0+t(+Ys0)t0, with =gH=M(2ν+1)s1k=s1/g and x= mod 2\begin{split}&\text{ apply the coupling of~\ref{pe:nacelle} to $\eta_{t}(\cdot)=\eta^{+}_{s_{0}+t\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}}(\cdot+Y^{-}_{s_{0}})$ and $\eta_{t}^{\prime}(\cdot)=\eta^{-}_{s_{0}+t\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}}(\cdot+Y^{-}_{s_{0}})$, $t\geq 0$,}\\ &\text{ with $\ell=\ell_{g}$, $H=M-(2\nu+1)s_{1}$, $k=\lfloor s_{1}/\ell_{g}\rfloor$ and $x=\ell\text{ mod }2$. }\end{split}

Note indeed that by (5.3) and (5.2), kg>48ν1(ν+log(40)p(1p)ν/2)k\geq\ell_{g}>48\nu^{-1}(\nu+\log(40)-p_{\circ}(1-p_{\bullet})\nu/2) for LL large enough and H=M(2ν+1)s12νs12νkg=2νkH=M-(2\nu+1)s_{1}\geq 2\nu s_{1}\geq 2\nu k\ell_{g}=2\nu\ell k for all L3L\geq 3, as required in (C.3). Together with the definition of E2E_{2} in (5.8), this allows us to apply the coupling of (C.3).

So far we have specified η±\eta^{\pm} for all of [0,T][0,T] on E12cE_{1-2}^{c} and up to time s1s_{1} on E12E_{1-2}. As we now briefly elaborate, it is also plain from the construction above that the independence property postulated in (5.56) holds for all ts1t\leq s_{1}. This is a trivial matter for t<s0t<s_{0} in view of (5.57), which does not involve UU at all. For s0ts1s_{0}\leq t\leq s_{1}, the only dependence on UU arises through Ys0Y^{-}_{s_{0}} via E2E_{2} and (5.60). However, on account of (5.6), (5.7), (5.10), (5.13) and (5.14), Ys0Y^{-}_{s_{0}} only relies on variables in UU with time label at most s01s_{0}-1, whence the claim. In the sequel (more precisely, up to (5.73)), considerations along similar lines allow to extend (5.56) to larger times tt. These will not be made explicit.

Returning to the construction of \mathbb{Q}, it remains to specify η±\eta^{\pm} after time s1s_{1} on the event E12E_{1-2}. Recall that by (5.9) and (5.15), both E3E_{3} and E4E_{4} are in s1\mathcal{F}_{s_{1}}. Over the time interval [s1,T][s_{1},T], conditionally on s1\mathcal{F}_{s_{1}} and on E12E3cE_{1-2}\cap E_{3}^{c}, we choose to

(5.61) let η\eta^{-} and η+\eta^{+} evolve independently with respective marginals 𝐏ηs1\mathbf{P}^{\eta^{-}_{s_{1}}} and 𝐏ηs1+\mathbf{P}^{\eta^{+}_{s_{1}}}.

On the other hand, conditionally on s1\mathcal{F}_{s_{1}} and on E13E4cE_{1-3}\cap E_{4}^{c}, during [s1,T][s_{1},T], we

(5.62) couple (η+,η) as (η,η) in (C.2.1),with H=M(4ν+1)s1,t=t=Ts1,k=1.\begin{split}&\quad\text{couple }(\eta^{+},\eta^{-})\text{ as }(\eta,\eta^{\prime})\text{ in~\ref{pe:drift}},\\ \text{with }&H=M-(4\nu+1)s_{1},\quad t=t=T-s_{1},\quad k=1.\end{split}

Conditionally on s1\mathcal{F}_{s_{1}} and on E14E_{1-4}, during the time-interval [s1,s2][s_{1},s_{2}], we

(5.63) couple (η+,η) as (η,η) in (C.2.1),with H=M(4ν+1)s1,t=s2s1,k=s1/(s2s1).\begin{split}&\quad\text{couple }(\eta^{+},\eta^{-})\text{ as }(\eta,\eta^{\prime})\text{ in~\ref{pe:drift}},\\ \text{with }&H=M-(4\nu+1)s_{1},\quad t=s_{2}-s_{1},\quad k=\lfloor s_{1}/(s_{2}-s_{1})\rfloor.\end{split}

Note that k1k\geq 1 in (5.63) by (5.3) and since LL2L\geq L_{2} by assumption (upon possibly enlarging L2L_{2}). Moreover, note that the conditions on the environments at time s1s_{1} needed for (C.2.1) to apply in both (5.62) and (5.63) are met owing to the occurrence of E3E_{3}, see (5.9).

We still have to specify η±\eta^{\pm} from time s2s_{2} onwards on the event E14E_{1-4}. By definition (5.16), we have that E5s2E_{5}\in\mathcal{F}_{s_{2}}. Hence, conditionally on s2\mathcal{F}_{s_{2}} and on E14E5cE_{1-4}\cap E_{5}^{c}, during [s2,T][s_{2},T], we

(5.64) let η\eta^{-} and η+\eta^{+} evolve independently with respective marginals 𝐏ηs2\mathbf{P}^{\eta^{-}_{s_{2}}} and 𝐏ηs2+\mathbf{P}^{\eta^{+}_{s_{2}}}.

Using the definitions (5.21) and (5.26), we have that both E6E_{6} and E7E_{7} are in s2\mathcal{F}_{s_{2}}. During [s2,T][s_{2},T], conditionally on s2\mathcal{F}_{s_{2}} and on E15E67cE_{1-5}\cap E_{6-7}^{c}, observing that the defining features of E5E_{5} (see (5.16)) allow to apply (C.2.1), we

(5.65) couple (η+,η) as (η,η) in (C.2.1) with H=M(6ν+1)s2t=Ts2 and k=1.\text{couple $(\eta^{+},\eta^{-})$ as $(\eta,\eta^{\prime})$ in~\ref{pe:drift} with $H=M-(6\nu+1)s_{2\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}}$, $t=T-s_{2}$ and $k=1$}.

Conditionally on s2\mathcal{F}_{s_{2}} and on E17E_{1-7}, during [s2,s3][s_{2},s_{3}], as we now explain, we

(5.66) couple (η+(+Ys2+g20),η(+Ys2g20)) as (η,η) in (C.2.2)with H=3(ν+1)Tt=s3s2 and (ρ,ε)=(ρ,ϵ).\begin{split}&\text{couple $(\eta^{+}(\cdot+Y^{-}_{s_{2}}+\ell_{g}^{20}),\eta^{-}(\cdot+Y^{-}_{s_{2}}-\ell_{g}^{20}))$ as $(\eta,\eta^{\prime})$ in~\ref{pe:couplings}}\\ &\text{with $H=3(\nu+1)T$, $t=s_{3}-s_{2}$ and $(\rho,\varepsilon)=(\rho,\epsilon)$}.\end{split}

Indeed, one checks that the conditions needed for (C.2.2) to apply are all satisfied on the event E7E_{7} whenever L2(ρ,ϵ,ν)L_{2}(\rho,\epsilon,\nu) is large enough. First, by choice of L2L_{2} and for all LL2L\geq L_{2}, one readily ensures (recall (5.2)-(5.3)) that all of 3(ν+1)T>4ν(s3s2)3(\nu+1)T>4\nu(s_{3}-s_{2}), ν8(s3s2)>1\nu^{8}(s_{3}-s_{2})>1 and ν(s3s2)1/4>\CrSEPcouplingϵ2(1+|log3(ν(s3s2))|)\nu(s_{3}-s_{2})^{1/4}>\Cr{SEPcoupling}\epsilon^{-2}(1+|\log^{3}(\nu(s_{3}-s_{2}))|) hold. Moreover, the conditions on the empirical densities of η0\eta_{0} and η0\eta_{0}^{\prime} appearing above (3.7) hold by definition of E7E_{7} at (5.26), since (s3s2)1/4g2\lfloor(s_{3}-s_{2})^{1/4}\rfloor\geq\ell_{g}^{2} by (5.3). Note that the relevant intervals II in the context of (C.2.2), which have length /2|I|\lfloor\ell/2\rfloor\leq|I|\leq\ell with =(s3s2)1/4\ell=\lfloor(s_{3}-s_{2})^{1/4}\rfloor, may in practice be much larger than those appearing in the definition of E7E_{7}, but they can be paved by disjoint contiguous intervals as entering E7E_{7}. The required controls on the corresponding empirical densities η0(I)\eta_{0}(I) and η0(I)\eta_{0}^{\prime}(I) are thus inherited from those defining E7E_{7}. Similar considerations also apply below (whenever either of (C.2.2) or (C.2) are used).

It remains to specify η±\eta^{\pm} on the event E17E_{1-7} for the time interval [s3,T][s_{3},T]. We now define an additional event whose realisation will allow us to couple η±\eta^{\pm} on the whole window of width order MM at time TT, regardless of whether the coupling at (5.66) succeeds or not; here, by “succeed” we mean that the (high-probability) event appearing in (3.7) is realized. The following will play a key role when establishing (5.54). We set

(5.69) E10=def.{I[M+2νs3,M2νs3] with (logL)2/2|I|(logL)2,ηs3(I)(ρ+ϵ/4)|I| and ηs3+(I)(ρ+3ϵ/4)|I|}.E_{10}\stackrel{{\scriptstyle\text{def.}}}{{=}}\left\{\begin{array}[]{c}\forall I\subseteq[-M+2\nu s_{3},M-2\nu s_{3}]\text{ with }\lfloor\lfloor(\log L)^{2}\rfloor/2\rfloor\leq|I|\leq\lfloor(\log L)^{2}\rfloor,\\ \eta^{-}_{s_{3}}(I)\leq(\rho+\epsilon/4)|I|\text{ and }\eta^{+}_{s_{3}}(I)\geq(\rho+3\epsilon/4)|I|\end{array}\right\}.

The event E10E_{10} above and E8E_{8} defined in (5.33) are both s3\mathcal{F}_{s_{3}}-measurable. Conditionally on s3\mathcal{F}_{s_{3}} and on E18E10E_{1-8}\cap E_{10}, during [s3,T][s_{3},T], we couple

(5.70) (η+(+Ys1+g20),η(+Ys1g20)) as in (C.2) with H1=2(ν+1)T,H2=M2νs3s1,t=Ts3,=G1/200 and (ρ,ε)=(ρ,ϵ).\begin{split}&(\eta^{+}(\cdot+Y^{-}_{s_{1}}+\ell_{g}^{20}),\eta^{-}(\cdot+Y^{-}_{s_{1}}-\ell_{g}^{20}))\text{ as in~\ref{pe:compatible} with }H_{1}=2(\nu+1)T,\\ &H_{2}=M-2\nu s_{3}-s_{1},\ t=T-s_{3},\ \ell=\ell_{G}^{1/200}\text{ and }(\rho,\varepsilon)=(\rho,\epsilon).\end{split}

To this effect, we verify that for LL large enough, by (5.3), (5.2) and since M10(ν+1)LM\geq 10(\nu+1)L, we indeed have that min{H1,H2H11}>10νt>4ν100>\Crcompatible\min\{H_{1},H_{2}-H_{1}-1\}>10{\nu}t>4\nu\ell^{100}>\Cr{compatible}, ν>\Crcompatibleϵ2(1+|log3(ν4)|)\nu\ell>\Cr{compatible}\epsilon^{-2}(1+|\log^{3}(\nu\ell^{4})|) and >80νϵ1+ν2.\ell>80\nu\epsilon^{-1}+\nu^{-2}. Also, the domination on [H1,H1][-H_{1},H_{1}] and the empirical density condition are respectively guaranteed by definition of E8E_{8} at (5.33) and E10E_{10} at (5.69).

If E8E_{8} does not occur (hence we temporarily lose the domination of η\eta^{-} by η+\eta^{+}) but E10E_{10} does, we proceed to what was referred to as the parachute coupling in Section 1.3, in order to recover this domination by time TT. Precisely, conditionally on s3\mathcal{F}_{s_{3}} and on E17E8cE10E_{1-7}\cap E_{8}^{c}\cap E_{10}, we

(5.71) couple (η+(T),η(+T)) as (η,η) in (C.2.2)with H=MT2νs3t=Ts3 and (ρ,ε)=(ρ,ϵ).\begin{split}&\text{couple $(\eta^{+}(\cdot-T),\eta^{-}(\cdot+T))$ as $(\eta,\eta^{\prime})$ in~\ref{pe:couplings}}\\ &\text{with $H=M-T-2\nu s_{3}$, $t=T-s_{3}$ and $(\rho,\varepsilon)=(\rho,\epsilon)$}.\end{split}

One checks indeed that the conditions of (C.2.2) hold, since for LL large enough (recalling (5.3) and (5.2)) we have that H>4νtH>4\nu t,ν8t>1\nu^{8}t>1, νt1/4>\CrSEPcouplingϵ2(1+|log3(νt)|)\nu t^{1/4}>\Cr{SEPcoupling}\epsilon^{-2}(1+|\log^{3}(\nu t)|) and t1/4>(logL)2t^{1/4}>\lfloor(\log L)^{2}\rfloor, so that on E10E_{10} the condition on the empirical density holds.

The remaining cases are straightforward to specify. Conditionally on s3\mathcal{F}_{s_{3}}, on E18E10cE_{1-8}\cap E_{10}^{c}, we

(5.72) couple (η+(+Ys1+g20),η(+Ys1g20)) as in (C.2.1) with H=2(ν+1)Tg20t=Ts3, and k=1.\begin{split}&\text{couple $(\eta^{+}(\cdot+Y^{-}_{s_{1}}+\ell_{g}^{20}),\eta^{-}(\cdot+Y^{-}_{s_{1}}-\ell_{g}^{20}))$ as in~\ref{pe:drift} }\\ &\text{with $H=2(\nu+1)T-\ell_{g}^{20}$, $t=T-s_{3}$, and $k=1$.}\end{split}

Finally, conditionally on s3\mathcal{F}_{s_{3}} and on E17E8cE10cE_{1-7}\cap E_{8}^{c}\cap E_{10}^{c}, during [s3,T][s_{3},T], we

(5.73) let η\eta^{-} and η+\eta^{+} evolve independently with respective marginals 𝐏ηs3\mathbf{P}^{\eta^{-}_{s_{3}}} and 𝐏ηs3+\mathbf{P}^{\eta^{+}_{s_{3}}}.

We have now fully defined the measure \mathbb{Q} and with it the triplet (η+,η,U)(\eta^{+},\eta^{-},U) (in the time interval [0,T][0,T]). The task is now to verify that with these choices, all of (5.51)-(5.54) hold. For concreteness we extend all three processes (η+,η,U)(\eta^{+},\eta^{-},U) independently at times t>Tt>T, using the Markov property at time TT for η±\eta^{\pm}. These extensions will de facto play no role because all of (5.52)-(5.54) only concern matters up to time T.T.


Part II: Proof of (5.51).

The fact that UU has the desired law is immediate, see below (5.55). We proceed to show that η𝐏η0\eta^{-}\sim\mathbf{P}^{\eta_{0}^{-}}. Since the construction of η\eta^{-} only consists of successive couplings at times 0,s0,s1,s20,s_{0},s_{1},s_{2} and s3s_{3} where the marginals have the desired distributions (recall (C.2)(C.2.1)(C.2.2) and (C.3)), by virtue of the Markov property (P.1), it is enough to check that the events deciding which coupling to apply at time sis_{i} are si\mathcal{F}_{s_{i}}-measurable for i{0,1,2,3}i\in\{0,1,2,3\}, which we already did at (5.58)-(5.66), (5.70), (5.71), (5.72) and (5.73). Therefore, η𝐏η0\eta^{-}\sim\mathbf{P}^{\eta_{0}^{-}}. In the same way we deduce that η+𝐏η0+\eta^{+}\sim\mathbf{P}^{\eta_{0}^{+}}.


Part III: Proof of (5.52).

We now use (η+,η,U)(\eta^{+},\eta^{-},U) to construct explicit functions X+=X+(η+,U)X^{+}=X^{+}(\eta^{+},U) and X=X(η,U)X^{-}=X^{-}(\eta^{-},U) with the correct marginal laws X±η0±X^{\pm}\sim\mathbb{P}^{\eta_{0}^{\pm}}. The case of XX^{-} is easily dispensed with: we define XX^{-} as in (2.5) and (2.6) with (η,U)(\eta^{-},U) instead of (η,U)(\eta,U). Since we have already established that η𝐏η0\eta^{-}\sim\mathbf{P}^{\eta_{0}^{-}} as part of (5.51), it follows using (5.56) and Lemma 2.1 that Xη0X^{-}\sim\mathbb{P}^{\eta_{0}^{-}}.

As for X+X^{+}, we also define it via (2.5) and (2.6) using the construction specified around (2.7), replacing (η,U)(\eta,U) by (η+,U+)(\eta^{+},U^{+}) where U+U^{+} is defined as follows: for all w𝕃w\in\mathbb{L} with π2(w)s31\pi_{2}(w)\leq s_{3}-1, Uw+=UwU^{+}_{w}=U_{w}, and for all w𝕃w\in\mathbb{L} with π2(w)s3\pi_{2}(w)\geq s_{3},

(5.74) Uw+={Uw(2g20,0) on E18,Uw on E18c.U^{+}_{w}=\begin{cases}U_{w-(2\ell_{g}^{20},0)}\text{ on }E_{1-8},\\ U_{w}\text{ on }E_{1-8}^{c}.\end{cases}

Recalling that E18s3E_{1-8}\in\mathcal{F}_{s_{3}}, and hence is independent of (Uw;w𝕃,π2(w)s3)(U_{w};w\in\mathbb{L},\pi_{2}(w)\geq s_{3}), it follows that X+η0+X^{+}\sim\mathbb{P}^{\eta_{0}^{+}} again by combining the established fact that η+𝐏η0+\eta^{+}\sim\mathbf{P}^{\eta_{0}^{+}} and (5.56). The rationale behind (5.74) will become clear momentarily.

We show that X±X^{\pm} as defined above satisfy YTXTY^{-}_{T}\geq X^{-}_{T} and YT+XT+Y^{+}_{T}\geq X^{+}_{T} \mathbb{Q}-a.s. To this end, note first that on EbadE_{\textnormal{bad}} and by (5.47), YT=TY^{-}_{T}=T and YT+=TY^{+}_{T}=-T, hence \mathbb{Q}-a.s. on EbadE_{\textnormal{bad}} we have using the trivial bounds XT±[T,T]X^{\pm}_{T}\in[-T,T] that XTYTX^{-}_{T}\leq Y^{-}_{T} and YT+XT+Y^{+}_{T}\leq X^{+}_{T}.

Second, on EneutralE_{\textnormal{neutral}} and by (5.44), under \mathbb{Q}, the process (Yt:0tT)(Y^{-}_{t}:0\leq t\leq T) simply follows the environment (η,U)(\eta^{-},U) as per (2.5) and (2.6), and so does (Xt:0tT)(X_{t}^{-}:0\leq t\leq T) by above definition of XX^{-}. Hence YT=XTY^{-}_{T}=X^{-}_{T}. Moreover, since EneutralE18cE_{\textnormal{neutral}}\subset E_{1-8}^{c} by (5.42), the process (Xt+:0tT)(X^{+}_{t}:0\leq t\leq T) follows the environment (η+,U)(\eta^{+},U) by (5.74). By Lemma 2.2 applied with (X,X~)=(Y,X+)(X,\widetilde{X})=(Y^{-},X^{+}), (η,η~)=(η,η+)(\eta,\widetilde{\eta})=(\eta^{-},\eta^{+}) and K=[T,T]×[0,T]K=[-T,T]\times[0,T], recalling (5.45), we get that XT+YTX^{+}_{T}\geq Y^{-}_{T}. Using (5.43), we finally obtain XT+YT+=YT=XTX^{+}_{T}\geq Y^{+}_{T}=Y^{-}_{T}=X^{-}_{T} \mathbb{Q}-a.s. on EneutralE_{\textnormal{neutral}}.

Third, on EgoodE16E_{\textnormal{good}}\subset E_{1-6}, the process (Yt:0ts2)(Y^{-}_{t}:0\leq t\leq s_{2}) follows the environment (η,U)(\eta^{-},U) by (5.6), (5.10) and (5.20), as is the case of XX^{-}, so that Xs2=Ys2X^{-}_{s_{2}}=Y^{-}_{s_{2}}. Then by (5.27) we have that Ys3=Ys2+(s3s2)Y^{-}_{s_{3}}=Y^{-}_{s_{2}}+(s_{3}-s_{2}) which guarantees that Xs3Ys3X^{-}_{s_{3}}\leq Y^{-}_{s_{3}}. From time s3s_{3} to time TT, XX^{-} and YY^{-} follow the same environment (η,U)(\eta^{-},U) by (5.35) and thus, by Lemma 2.2 with (X,X~)=(Y,X)(X,\widetilde{X})=(Y^{-},X^{-}), (η,η~)=(η,η)(\eta,\widetilde{\eta})=(\eta^{-},\eta^{-}) and K=×[s3,T]K=\mathbb{Z}\times[s_{3},T], we have that XTYTX^{-}_{T}\leq Y^{-}_{T}.

Similarly for X+X^{+} and Y+Y^{+}, on EgoodE13E_{\textnormal{good}}\subset E_{1-3}, during [0,s1][0,s_{1}] the process X+X^{+} follows the environment (η+,U)(\eta^{+},U) while Y+Y^{+} follows (η,U)(\eta^{-},U) by (5.6), (5.10). By (5.40) we can apply Lemma 2.2 with (X,X~)=(Y+,X+)(X,\widetilde{X})=(Y^{+},X^{+}), (η,η~)=(η,η+)(\eta,\widetilde{\eta})=(\eta^{-},\eta^{+}) and K=[T,T]×[0,s1]K=[-T,T]\times[0,s_{1}] to deduce that Ys1+Xs1+Y^{+}_{s_{1}}\leq X^{+}_{s_{1}}. Then during [s1,s2][s_{1},s_{2}], both Y+Y^{+} and X+X^{+} follow (η+,U)(\eta^{+},U) (recall (5.20), and that EgoodE15E_{\textnormal{good}}\subset E_{1-5}). Thus Lemma 2.2 with (X,X~)=(Y+,X+)(X,\widetilde{X})=(Y^{+},X^{+}), (η,η~)=(η+,η+)(\eta,\widetilde{\eta})=(\eta^{+},\eta^{+}) and K=×[s1,s2]K=\mathbb{Z}\times[s_{1},s_{2}] ensures that Ys2+Xs2+Y^{+}_{s_{2}}\leq X^{+}_{s_{2}} \mathbb{Q}-a.s. on EgoodE_{\textnormal{good}}. Next, (5.27), which holds on E17EgoodE_{1-7}\supset E_{\textnormal{good}}, implies that Ys3+Xs3+Y^{+}_{s_{3}}\leq X^{+}_{s_{3}}, given that X+X^{+} only takes nearest-neighbour steps. Finally, since EgoodE19E_{\textnormal{good}}\subset E_{1-9}, from time s3s_{3} to TT, both Y+Y^{+} and X+X^{+} follow the environment (η+,U+)(\eta^{+},U^{+}), where U+U^{+} is as in (5.74) (recall (5.35); this explains the choice in (5.74)). Applying Lemma 2.2 with (X,X~)=(Y+,X+)(X,\widetilde{X})=(Y^{+},X^{+}), (η,η~)=(η+,η+)(\eta,\widetilde{\eta})=(\eta^{+},\eta^{+}) and K=×[s3,T]K=\mathbb{Z}\times[s_{3},T] therefore yields YT+XT+Y^{+}_{T}\leq X^{+}_{T} on EgoodE_{\textnormal{good}}. This concludes the proof of (5.52).

Part IV: Proof of (5.53).

By Lemma 5.1, we have that

(5.75) 𝔼[YT+YT]=2g20(Egood)2T(Ebad).\mathbb{E}^{\mathbb{Q}}[Y^{+}_{T}-Y^{-}_{T}]=2\ell_{g}^{20}\mathbb{Q}(E_{\textnormal{good}})-2T\mathbb{Q}(E_{\textnormal{bad}}).

Recalling the definitions of EgoodE_{\textnormal{good}} and EbadE_{\textnormal{bad}} from (5.38) and (5.46), as well as (5.2) and (5.3), it is thus enough to show that

(5.76) ((Egood)=)(E19)exp(g22)(\mathbb{Q}(E_{\textnormal{good}})=)\ \mathbb{Q}(E_{1-9})\geq\exp(-\ell_{g}^{22})

and

(5.79) S=def.(E1c)+(E1E2cE2,bisc)+(E12E3c)+(E13E4,5cE4,bisc)+(E15E67cE6,bisc)+(E17E89c)(E19)/(10T).S\stackrel{{\scriptstyle\text{def.}}}{{=}}\mathbb{Q}(E_{1}^{c})+\mathbb{Q}(E_{1}\cap E_{2}^{c}\cap E_{2,\textnormal{bis}}^{c})+\mathbb{Q}(E_{1-2}\cap E_{3}^{c})+\mathbb{Q}(E_{1-3}\cap E_{4,5}^{c}\cap E_{4,\textnormal{bis}}^{c})\\ +\mathbb{Q}(E_{1-5}\cap E_{6-7}^{c}\cap E_{6,\textnormal{bis}}^{c})+\mathbb{Q}(E_{1-7}\cap E_{8-9}^{c})\leq\mathbb{Q}(E_{1-9})/(10T).

Proof of (5.76). Recall E1E_{1} from (5.5). By (5.57) and (3.6), which is in force, we get

(5.80) (E1c)20exp(νs0/4)exp(g105),\mathbb{Q}(E_{1}^{c})\leq 20\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\exp(-\nu s_{0}/4)\leq\exp(-\ell_{g}^{10^{5}}),

the last inequality being true for LL large enough by (5.3) and (5.2). Next, recalling (5.8), that 4s0>s0+3g4s_{0}>s_{0}+3\ell_{g} by (5.3) and that ϵ<1/100\epsilon<1/100, and using that |Ys0|s0|Y^{-}_{s_{0}}|\leq s_{0}, we note that

(5.81) E2cx[4s0+1,4s0],g/2g{η([x,x+1])(ρ+ϵ/2)<η+([x,x+1])(ρ+2ϵ)}c.\begin{split}E_{2}^{c}\subseteq&\bigcup_{\begin{subarray}{c}x\in[-4s_{0}+1,4s_{0}],\\ \lfloor\ell_{g}/2\rfloor\leq\ell^{\prime}\leq\ell_{g}\end{subarray}}\{\eta^{-}([x,x+\ell^{\prime}-1])\leq(\rho+\epsilon/2)\ell^{\prime}<\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\eta^{+}([x,x+\ell^{\prime}-1])\leq(\rho+2\epsilon)\ell^{\prime}\}^{c}.\end{split}

We note in passing that we cannot locate precisely Ys0Y^{-}_{s_{0}} without (possibly heavily) conditioning the evolution of η\eta^{-} on [0,s0][0,s_{0}]. By (C.1) applied with either (η0,ρ,ε)=(η0,ρ,ϵ/20)(\eta_{0},\rho,\varepsilon)=(\eta^{-}_{0},\rho,\epsilon/20) or (η0,ρ,ε)=(η0,ρ+ϵ,ϵ/20)(\eta_{0},\rho,\varepsilon)=(\eta^{-}_{0},\rho+\epsilon,\epsilon/20), both times with (,H,t)=(log2L,(4ν+8)s0,s0)(\ell,H,t)=(\lfloor\log^{2}L\rfloor,(4\nu+8)s_{0},s_{0}) and \ell^{\prime} ranging from g/2\lfloor\ell_{g}/2\rfloor to g\ell_{g} we have by a union bound on xx and \ell^{\prime}:

(5.82) (E2c)(2×8s0×g)4(4ν+8)s0exp(\Crdensitystableexpoϵ2g/2/400)1/4,\mathbb{Q}(E_{2}^{c})\leq(2\times 8s_{0}\times\ell_{g})4(4\nu+8)s_{0}\exp(-\Cr{densitystableexpo}\epsilon^{2}\lfloor\ell_{g}/2\rfloor/400)\leq 1/4,

the last inequality holding for LL large enough by (5.2) and (5.3). Note that we can indeed apply (C.1) since (η0+,η0)(\eta^{+}_{0},\eta^{-}_{0}) is (M,L)(M,L)-balanced (see items (ii) and (iii) above Lemma 5.2), and M>(4ν+8)s0>4νs0>\Crdensitystableϵ2(logL)22(1+|log3(νs0)|)M>(4\nu+8)s_{0}>4\nu s_{0}>\Cr{densitystable}\epsilon^{-2}\lfloor(\log L)^{2}\rfloor^{2}(1+|\log^{3}(\nu s_{0})|) and gs0\ell_{g}\leq\sqrt{s_{0}} for LL large enough.

Next, we aim to derive a suitable deterministic lower bound on (E34|s0)\mathbb{Q}(E_{3-4}\,|\mathcal{F}_{s_{0}}), on the event E12E_{1-2}, which is s0\mathcal{F}_{s_{0}}-measurable. We aim to apply (C.3), cf. (5.60), and dealing with E3E_{3} is straightforward, see (3.9), but E4E_{4} requires a small amount of work, cf. (3.8) and (5.15). To this effect, we first observe that under \mathbb{Q}, with η,η\eta,\eta^{\prime} as defined in (5.60), one has the inclusions

(5.85) E14(5.11),(5.15){ηs1+(Ys1)1,ηs1(Ys1)=0}E13{η(x)1,η(x)=0}{Ys1Ys0=x}E13,E_{1-4}\stackrel{{\scriptstyle\eqref{Ypmats1},\eqref{eq:E4deff}}}{{\supset}}\{\eta^{+}_{s_{1}}(Y_{s_{1}}^{-})\geq 1,\eta^{-}_{s_{1}}(Y_{s_{1}}^{-})=0\}\cap E_{1-3}\\ \supset\{\eta_{\ell}(x)\geq 1,\eta^{\prime}_{\ell}(x)=0\}\cap\{Y_{s_{1}}^{-}-Y_{s_{0}}^{-}=x\}\cap E_{1-3},

where =g=s1s0\ell=\ell_{g}=s_{1}-s_{0} and x= mod 2x=\ell\text{ mod }2 on account of (5.60) and (5.3). Now recalling that the evolution of YY^{-} on the time interval [s0,s1][s_{0},s_{1}] follows (5.10) on E13E_{1-3}, and in view of (2.5), one readily deduces that the event {Ys1Ys0=x}\{Y_{s_{1}}^{-}-Y_{s_{0}}^{-}=x\} is implied by the event E13FE_{1-3}\cap F, where FF refers to the joint occurrence of {U(Ys0,s0+n)<p}\{U_{(Y^{-}_{s_{0}},s_{0}+n)}<p_{\circ}\} for even integer nn satisfying 0n10\leq n\leq\ell-1 and {U(Ys0,s0+n)>p}\{U_{(Y^{-}_{s_{0}},s_{0}+n)}>p_{\bullet}\} for odd integer nn satisfying 0n10\leq n\leq\ell-1. (Observe indeed that, if \ell is odd, whence x=1x=1, there will be one more step to the right than to the left in the resulting trajectory for Y+Y^{+}). Feeding this into (5.85), applying a union bound, using (3.9) and (3.8) (which are in force on account of (5.60)), it follows that on E12E_{1-2},

(E34|s0)({η(x)1,η(x)=0}F|s0)δ2δ(p(1p))δδ\mathbb{Q}(E_{3-4}\,|\mathcal{F}_{s_{0}})\geq\mathbb{Q}(\{\eta_{\ell}(x)\geq 1,\eta^{\prime}_{\ell}(x)=0\}\cap F\,|\mathcal{F}_{s_{0}})-\delta^{\prime}\geq 2\delta(p_{\circ}(1-p_{\bullet}))^{\ell}-\delta^{\prime}\geq\delta^{\prime}

(see above (C.3) regarding δ\delta and δ\delta^{\prime}); in the penultimate step above, we have also used that, conditionally on 0\mathcal{F}_{0}, the events {η(x)1,η(x)=0}\{\eta_{\ell}(x)\geq 1,\eta^{\prime}_{\ell}(x)=0\} and FF are independent. Combining this with (5.80) and (5.82), we deduce that for LL large enough, using (5.2),

(5.86) (E14)3δ/420eg105δ/2.\mathbb{Q}(E_{1-4})\geq 3\delta^{\prime}/4-20\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}e^{-\ell_{g}^{10^{5}}}\geq\delta^{\prime}/2.

Next, recalling the definition (5.16) of E5E_{5} and that E14s1E_{1-4}\in\mathcal{F}_{s_{1}}, by (5.63) and the bound given in (C.2.1), we have that, \mathbb{Q}-a.s. on E14E_{1-4},

(5.87) (E5c|s1)20exp(s1/(s2s1)(s2s1)ν4)exp(ν4G)exp(ν8g106),\mathbb{Q}(E_{5}^{c}\,|\mathcal{F}_{s_{1}})\leq 20\exp(-\lfloor s_{1}/(s_{2}-s_{1})\rfloor(s_{2}-s_{1})\tfrac{\nu}{4})\leq\exp(-\tfrac{\nu}{4}\ell_{G})\leq\exp\big{(}-\tfrac{\nu}{8}\ell_{g}^{10^{6}}\big{)},

for all LL large enough by (5.3).

We will return to E6E_{6} momentarily and first consider E7E_{7}. To compute the probability of E7cE_{7}^{c}, we take a union bound over \ell^{\prime} such that g2/2g2\lfloor\ell_{g}^{2}/2\rfloor\leq\ell^{\prime}\leq\ell_{g}^{2}, use the definition (5.26) and the fact that (η0,η0+)(\eta_{0}^{-},\eta_{0}^{+}) is (M,L)(M,L)-balanced in order to apply (C.1) twice, once for η\eta^{-} and once for η+\eta^{+}, choosing in both cases (,H,t)=((logL)2,38(ν+1)G,s2)(\ell,H,t)=(\lfloor(\log L)^{2}\rfloor,38(\nu+1)\ell_{G},s_{2}) and, for η\eta^{-}, with (η0,ρ,ε)=(η0,ρ,ϵ/100)(\eta_{0},\rho,\varepsilon)=(\eta^{-}_{0},\rho,\epsilon/100) and, for η+\eta^{+}, with (η0,ρ,ε)=(η0+,ρ+ϵ,ϵ/100)(\eta_{0},\rho,\varepsilon)=(\eta^{+}_{0},\rho+\epsilon,\epsilon/100). Note that for LL large enough by (5.3) and (5.2), we have indeed M>H>4νt>\Crdensitystable2ε2(1+|log3(νt)|)M>H>4\nu t>\Cr{densitystable}\ell^{2}\varepsilon^{-2}(1+|\log^{3}(\nu t)|) and g2t\ell^{\prime}\leq\ell_{g}^{2}\leq\sqrt{t}, as required in (C.1). We thus obtain that

(5.88) (E7c)400(ν+1)g2Gexp(\Crdensitystableexpoϵ2g22/104)exp(\Crdensitystableexpoϵ2105g2),\mathbb{Q}(E_{7}^{c})\leq 400(\nu+1)\ell_{g}^{2}\ell_{G}\exp(-\Cr{densitystableexpo}\epsilon^{2}\lfloor\tfrac{\ell_{g}^{2}}{2}\rfloor/10^{4})\leq\exp\big{(}-\tfrac{\Cr{densitystableexpo}\epsilon^{2}}{10^{5}}\ell_{g}^{2}\big{)},

where the last inequality holds for LL large enough (depending on ν\nu, ρ\rho and ϵ\epsilon). Putting (5.86), (5.87) and (5.88) together and taking LL (hence g\ell_{g}) large enough, it follows that

(5.89) (E15E7)(E15)(E7c)δ/4.\mathbb{Q}(E_{1-5}\cap E_{7})\geq\mathbb{Q}(E_{1-5})-\mathbb{Q}(E_{7}^{c})\geq\delta^{\prime}/4.

Concerning E6E_{6}, we first observe that under \mathbb{Q}, the field (Uw:s1π2(w)s2)(U_{w}:s_{1}\leq\pi_{2}(w)\leq s_{2}) is independent from the σ\sigma-algebra generated by s1\mathcal{F}_{s_{1}} and (ηt±)s1ts2(\eta^{\pm}_{t})_{s_{1}\leq t\leq s_{2}}; this can be seen by direct inspection of the coupling construction until time s2s_{2}, paying particular attention (with regards to the evolution of η±\eta^{\pm} during [s1,s2][s_{1},s_{2}]) to (5.61), (5.62), (5.63) (and (5.58)), which all involve only either i) trivial couplings or ii) couplings relying on (C.2.1). In particular these couplings do not involve UU at all. Using the previous observation and recalling (5.21)-(5.22), it follows that \mathbb{Q}-a.s.,

(5.90) (E6|s1,(ηt±)s1ts2)=(pp)(p(1p))s2s11(pp)(p(1p))2g20.\mathbb{Q}\left(\left.E_{6}\right|\mathcal{F}_{s_{1}},(\eta^{\pm}_{t})_{s_{1}\leq t\leq s_{2}}\right)=(p_{\bullet}-p_{\circ})\left(p_{\circ}(1-p_{\bullet})\right)^{s_{2}-s_{1}-1}\geq(p_{\bullet}-p_{\circ})\left(p_{\circ}(1-p_{\bullet})\right)^{2\ell_{g}^{20}}.

Next, observe that E15E_{1-5} and E7E_{7} are measurable with respect to the σ\sigma -algebra generated by s1\mathcal{F}_{s_{1}} and (ηt±)s1ts2(\eta^{\pm}_{t})_{s_{1}\leq t\leq s_{2}} owing to (5.5), (5.8),  (5.9), (5.15),  (5.16) and (5.26). Combining this with (5.89), and recalling the value of δ\delta^{\prime} from (C.3), we obtain that

(5.91) (E17)(p(1p))2g20δ/4(p(1p))3g20exp(g21),\begin{split}\mathbb{Q}(E_{1-7})&\geq\left(p_{\circ}(1-p_{\bullet})\right)^{2\ell_{g}^{20}}\delta^{\prime}/4\geq\left(p_{\circ}(1-p_{\bullet})\right)^{3\ell_{g}^{20}}\geq\exp(-\ell_{g}^{21}),\end{split}

where, in the last two inequalities, we took LL (hence g\ell_{g}) large enough.

Recalling E8E_{8} from (5.33) along with the coupling defined in (5.66) and using that E17s2E_{1-7}\in\mathcal{F}_{s_{2}}, we can apply (C.2.2), and obtain that, \mathbb{Q}-a.s. on E17E_{1-7}, for large enough LL (owing to (5.3) and (5.2)),

(E8c|s2)\CrSEPcoupling2×3(ν+1)T(s3s2)exp((\CrSEPcoupling2(1+ν1))1ϵ2(s3s2)1/4)exp(g4),\mathbb{Q}\left(\left.E_{8}^{c}\right|\mathcal{F}_{s_{2}}\right)\leq\Cr{SEPcoupling2}\times 3(\nu+1)T(s_{3}-s_{2})\exp\big{(}-(\Cr{SEPcoupling2}(1+\nu^{-1}))^{-1}\epsilon^{2}(s_{3}-s_{2})^{1/4}\big{)}\leq\exp\left(-\ell_{g}^{4}\right),

which implies that

(5.92) (E8c|E17)exp(g4).\mathbb{Q}\left(\left.E_{8}^{c}\right|E_{1-7}\right)\leq\exp\left(-\ell_{g}^{4}\right).

Next, we control the probability of E10E_{10} defined at (5.69). To do so, we take a union bound over [(logL)2/2,(logL)2]\ell^{\prime}\in[\lfloor\lfloor(\log L)^{2}\rfloor/2\rfloor,\lfloor(\log L)^{2}\rfloor] and, using that (η0,η0+)(\eta_{0}^{-},\eta_{0}^{+}) is (M,L)(M,L)-balanced, we apply (C.1) twice for any fixed \ell^{\prime} in this interval: once with (η0,ρ,ε)=(η0,ρ,ϵ/20)(\eta_{0},\rho,\varepsilon)=(\eta^{-}_{0},\rho,\epsilon/20), once with (η0,ρ,ε)=(η0+,ρ+ϵ,ϵ/20)(\eta_{0},\rho,\varepsilon)=(\eta^{+}_{0},\rho+\epsilon,\epsilon/20) and both times with (,H,t)=((logL)2,M,s3)(\ell,H,t)=(\lfloor(\log L)^{2}\rfloor,M,s_{3}). Note once again that for LL large enough by (5.3) and (5.2), we have M>H>4νt>\Crdensitystable2ε2(1+|log3(νt)|)M>H>4\nu t>\Cr{densitystable}\ell^{2}\varepsilon^{-2}(1+|\log^{3}(\nu t)|) and g2t\ell^{\prime}\leq\ell_{g}^{2}\leq\sqrt{t}, as required for (C.1) to apply. This yields, applying a union bound over the values of \ell^{\prime}, that for large enough LL (recalling that M20(ν+1)LM\leq 20(\nu+1)L),

(5.93) (E10c)8M(logL)2exp(\Crdensitystableexpoϵ2log2L/400)exp(g1500).\mathbb{Q}(E_{10}^{c})\leq 8M\lfloor(\log L)^{2}\rfloor\exp\left(-\Cr{densitystableexpo}\epsilon^{2}\lfloor\log^{2}L\rfloor/400\right)\leq\exp(-\ell_{g}^{1500}).

Finally, recalling the coupling (5.70) (together with (5.33), (5.34) and (5.69)), we have by (C.2), more precisely by (3.4), for large enough LL,

(5.94) (E9c|E18E10)20Texp(ν(Ts3)/4)exp(g105).\mathbb{Q}(E_{9}^{c}\,|\,E_{1-8}\cap E_{10})\leq 20T\exp(-\nu(T-s_{3})/4)\leq\exp(-\ell_{g}^{10^{5}}).

Putting together (5.91), (5.92), (5.93) and (5.94), we obtain (5.76), as desired.

Proof of (5.79). We bound individually the six terms comprising SS on the left-hand side of (5.79). The first of these is already controlled by (5.80). Recall the coupling defined in (5.59) together with the definitions of E1E_{1}, E2E_{2} and E2,bisE_{2,\textnormal{bis}} in (5.5), (5.8) and (5.12). Observe that by (5.59) and (C.2.1), we have \mathbb{Q}-a.s. on E1E2cE_{1}\cap E_{2}^{c} that (E2,bisc|s0)20exp(ν(Ts0)/4).\mathbb{Q}(E_{2,\textnormal{bis}}^{c}|\mathcal{F}_{s_{0}})\leq 20\exp(-\nu(T-s_{0})/4). Hence for large enough LL, by (5.3) and (5.2), this yields

(5.95) (E1E2cE2,bisc)exp(g105).\mathbb{Q}(E_{1}\cap E_{2}^{c}\cap E_{2,\textnormal{bis}}^{c})\leq\exp(-\ell_{g}^{10^{5}}).

Recall now the coupling (5.60) and the definition (5.9) of E3E_{3}. By (3.9) which is in force, we have

(5.96) (E12E3c)20exp(νgs1/g/4)exp(g105),\mathbb{Q}(E_{1-2}\cap E_{3}^{c})\leq 20\exp\left(-\nu\ell_{g}\lfloor s_{1}/\ell_{g}\rfloor/4\right)\leq\exp(-\ell_{g}^{10^{5}}),

for LL large enough due to (5.3) and (5.2). Next, remark that

(5.97) E13E45cE4,bisc{E14E5c}{E13E4cE4,bisc}.E_{1-3}\cap E_{4-5}^{c}\cap E_{4,\textnormal{bis}}^{c}\subseteq\{E_{1-4}\cap E_{5}^{c}\}\cup\{E_{1-3}\cap E_{4}^{c}\cap E_{4,\textnormal{bis}}^{c}\}.

On one hand, by (5.63) and (C.2.1) (recalling (5.16)) we have

(5.98) (E14E5c)20exp(ν(s2s1)s1/(s2s1)/4).\mathbb{Q}(E_{1-4}\cap E_{5}^{c})\leq 20\exp\left(-\nu(s_{2}-s_{1})\lfloor s_{1}/(s_{2}-s_{1})\rfloor/4\right).

On the other hand, by (5.62) and (C.2.1) (recalling (5.17)), we have

(5.99) (E13E4cE4,bisc)20exp(ν(Ts1)/4)\mathbb{Q}(E_{1-3}\cap E_{4}^{c}\cap E_{4,\textnormal{bis}}^{c})\leq 20\exp\left(-\nu(T-s_{1})/4\right)

Together, (5.97), (5.98) and (5.99) yield that

(5.100) (E13E45cE4,bisc)exp(g105),\mathbb{Q}(E_{1-3}\cap E_{4-5}^{c}\cap E_{4,\textnormal{bis}}^{c})\leq\exp(-\ell_{g}^{10^{5}}),

for LL large enough, again via (5.3) and (5.2). Similarly, by (5.65), (C.2.1) and (5.29), for LL large enough we have that

(5.101) (E15E67cE6,bisc)exp(g105).\mathbb{Q}(E_{1-5}\cap E_{6-7}^{c}\cap E_{6,\textnormal{bis}}^{c})\leq\exp(-\ell_{g}^{10^{5}}).

In view of (5.79), putting together (5.80), (5.95), (5.96), (5.100) and (5.101) yields that

(5.102) S(E17E89c)+5exp(g105).S\leq\mathbb{Q}(E_{1-7}\cap E_{8-9}^{c})+5\exp(-\ell_{g}^{10^{5}}).

By (5.76), which has been already established, we know that 5exp(g105)(E19)/(20T)5\exp(-\ell_{g}^{10^{5}})\leq\mathbb{Q}(E_{1-9})/(20T) for LL large enough by (5.2). Hence, in order to conclude the proof, it is enough to show that (E17E89c)(E19)/(20T)\mathbb{Q}(E_{1-7}\cap E_{8-9}^{c})\leq\mathbb{Q}(E_{1-9})/(20T), or equivalently that

(5.103) (E89|E17)1(20T)1.\mathbb{Q}(E_{8-9}\,|\,E_{1-7})\geq 1-({20T})^{-1}.

Indeed,

(E89|E17)=(E8|E17)(E9|E18)=(E8|E17)(E19|E18E10)(E18E10)(E18)(E8|E17)(E19|E18E10)(1(E10c)(E19))(1exp(g4))(1exp(g105))(1exp(g1500)exp(g22)),\mathbb{Q}(E_{8-9}\,|\,E_{1-7})=\mathbb{Q}(E_{8}\,|\,E_{1-7})\mathbb{Q}(E_{9}\,|\,E_{1-8})=\mathbb{Q}(E_{8}\,|\,E_{1-7})\mathbb{Q}(E_{1-9}\,|\,E_{1-8}\cap E_{10})\frac{\mathbb{Q}(E_{1-8}\cap E_{10})}{\mathbb{Q}(E_{1-8})}\\ \geq\mathbb{Q}(E_{8}\,|\,E_{1-7})\mathbb{Q}(E_{1-9}\,|\,E_{1-8}\cap E_{10})\left(1-\frac{\mathbb{Q}(E_{10}^{c})}{\mathbb{Q}(E_{1-9})}\right)\\ \geq\left(1-\exp(-\ell_{g}^{4})\right)\left(1-\exp(-\ell_{g}^{10^{5}})\right)\left(1-\exp(-\ell_{g}^{1500})\exp(\ell_{g}^{22})\right),

by virtue of (5.92), (5.94), (5.93) and (5.76) in the last line. For large enough LL (recalling (5.3) and (5.2)), this readily yields (5.103) and concludes the proof of (5.79).

Part V: Proof of (5.54).

Recall the event ErestartE_{\textnormal{restart}} from (5.54). The proof of (5.54) is relatively straightforward at this point, we just need to keep careful track of those events in the above construction that can force us out of the event ErestartE_{\textnormal{restart}}; this is key to get the rapid decay in (5.54). First note that EneutralE2,bisE4,bisE6,bisE_{\textnormal{neutral}}\subseteq E_{2,\textnormal{bis}}\cup E_{4,\textnormal{bis}}\cup E_{6,\textnormal{bis}} by (5.42), so that by (5.12), (5.17), (5.29) and (5.49) (and using that |YT±|T|Y_{T}^{\pm}|\leq T along with (5.2)), ErestartE_{\textnormal{restart}} holds on EneutralE_{\textnormal{neutral}} for large enough LL (tacitly assumed in the sequel), hence ErestartcE_{\textnormal{restart}}^{c} can only happen on EgoodEbadE_{\textnormal{good}}\cup E_{\textnormal{bad}}. As we now explain, looking at the decomposition of EgoodE_{\textnormal{good}} and EbadE_{\textnormal{bad}} at (5.38) and (5.46), and inspecting closely the construction η±\eta^{\pm} (and Y±Y^{\pm}), especially in Part I of the proof, starting with the paragraph of (5.69) until (5.73), one notices that ErestartcE_{\textnormal{restart}}^{c} can in fact only happen:

  • (i)

    on all but the last event (i.e. all but E17E89cE_{1-7}\cap E_{8-9}^{c}) defining EbadE_{\textnormal{bad}} at (5.46), or

  • (ii)

    on E17E10cE_{1-7}\cap E_{10}^{c} (see (5.69)), or

  • (iii)

    on E17E8cE10E_{1-7}\cap E_{8}^{c}\cap E_{10} if the coupling (C.2.2) applied at (5.71) fails, i.e. if the event

    (5.107) E11=def.{ηT+(T)|[M+T+4νT,MT4νT]ηT(+T)|[M+T+4νT,MT4νT]}E_{11}\stackrel{{\scriptstyle\text{def.}}}{{=}}\big{\{}\eta^{+}_{T\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}}(\cdot-T)|_{[-M+T+4\nu T,M-T-4\nu T]}\succcurlyeq\eta^{-}_{T\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}}(\cdot+T)|_{[-M+T+4\nu T,M-T-4\nu T]}\big{\}}

    (cf. (3.7)) does not occur, or

  • (iv)

    on E18E9cE10E_{1-8}\cap E_{9}^{c}\cap E_{10}, or

  • (v)

    on E110E_{1-10} if the coupling (C.2) (more precisely (3.5)) applied at (5.70) fails, i.e. if the event

    (5.108) E12=def.{ηT+(+Ys1+g20)|IηT(+Ys1g20)|I}E_{12}\stackrel{{\scriptstyle\text{def.}}}{{=}}\big{\{}\eta^{+}_{T\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}}(\cdot+Y^{-}_{s_{1}}+\ell_{g}^{20})|_{I}\succcurlyeq\eta^{-}_{T\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}}(\cdot+Y^{-}_{s_{1}}-\ell_{g}^{20})|_{I}\big{\}}

    with I=[M+T+6νT,MT6νT]I=[-M+T+6\nu T,M-T-6\nu T] does not occur.

We now detail how the cases (i)-(v) arise. First note that, since Erestart(EbadEgood)E_{\textnormal{restart}}\subset(E_{\textnormal{bad}}\cup E_{\textnormal{good}}) as established above, and since Egood=E19E_{\textnormal{good}}=E_{1-9} by definition (see (5.38)), after discarding item (i) from the above list it only remains to investigate matters on the event E17E_{1-7}, and the cases considered in items (ii)-(v) indeed form a partition of this event, save for the additional specifications (“if the coupling…”) in items (iii) and (v), which we now discuss. For item (iii), note indeed that if E17E8cE10EbadE_{1-7}\cap E_{8}^{c}\cap E_{10}\subseteq E_{\textnormal{bad}} holds (see (5.46)) then YT+=YT2T{Y}^{+}_{T}=Y^{-}_{T}-2T by Lemma 5.1, and that since |YT|T|Y^{-}_{T}|\leq T, if in addition E11E_{11} holds then so does ErestartE_{\textnormal{restart}} in view of (5.54). Similarly, regarding item (v), we have by (5.38) that E110EgoodE_{1-10}\subseteq E_{\textnormal{good}} so that if E110E_{1-10} holds, Y~T+=YT+2g20\widetilde{Y}^{+}_{T}=Y^{-}_{T}+2\ell_{g}^{20} by Lemma 5.1. Since |YT(Ys1g20)|Ts1+g20T|Y^{-}_{T}-(Y^{-}_{s_{1}}-\ell_{g}^{20})|\leq T-s_{1}+\ell_{g}^{20}\leq T, if in addition E12E_{12} occurs then ErestartE_{\textnormal{restart}} occurs as well.

Combining items (i)-(v) above and recalling (5.46) in the context of item (i), by a union bound we have that

(Erestartc)(E1c)+(E1E2cE2,bisc)+(E12E3c)+(E13E4,5cE4,bisc)+(E15E67cE6,bisc)+(E10c)+(E18E9cE10)+(E17E8cE10E11c)+(E18E10E12c).\mathbb{Q}(E_{\textnormal{restart}}^{c})\leq\mathbb{Q}(E_{1}^{c})+\mathbb{Q}(E_{1}\cap E_{2}^{c}\cap E_{2,\textnormal{bis}}^{c})\\ +\mathbb{Q}(E_{1-2}\cap E_{3}^{c})+\mathbb{Q}(E_{1-3}\cap E_{4,5}^{c}\cap E_{4,\textnormal{bis}}^{c})+\mathbb{Q}(E_{1-5}\cap E_{6-7}^{c}\cap E_{6,\textnormal{bis}}^{c})+\mathbb{Q}(E_{10}^{c})\\ +\mathbb{Q}(E_{1-8}\cap E_{9}^{c}\cap E_{10})+\mathbb{Q}(E_{1-7}\cap E_{8}^{c}\cap E_{10}\cap E_{11}^{c})+\mathbb{Q}(E_{1-8}\cap E_{10}\cap E_{12}^{c}).

By (5.80), (5.93), (5.94), (5.95), (5.96),  (5.100) and (5.101), for large enough LL we obtain that

(5.112) (Erestartc)7exp(g1500)+Q(E17E10E8cE11c)+(E18E10E12c).\mathbb{Q}(E_{\textnormal{restart}}^{c})\leq 7\exp(-\ell_{g}^{1500})+Q(E_{1-7}\cap E_{10}\cap E_{8}^{c}\cap E_{11}^{c})+\mathbb{Q}(E_{1-8}\cap E_{10}\cap E_{12}^{c}).

As to the last two terms in (5.112), using the coupling defined in (5.71) and (C.2.2), we get

(5.115) Q(E11c|E17E8cE10)\CrSEPcoupling2TMexp((\CrSEPcoupling2(1+ν1))1ϵ2(Ts3)1/4)C32g10620(ν+1)Lexp((\CrSEPcoupling2(1+ν1))1ϵ2g2105)exp(g105),Q(E_{11}^{c}\,|\,E_{1-7}\cap E_{8}^{c}\cap E_{10})\leq\Cr{SEPcoupling2}TM\exp\big{(}-(\Cr{SEPcoupling2}(1+\nu^{-1}))^{-1}\epsilon^{2}(T-s_{3})^{1/4}\big{)}\\ \leq C_{3}2\ell_{g}^{10^{6}}\cdot 20(\nu+1)L\cdot\exp\big{(}-(\Cr{SEPcoupling2}(1+\nu^{-1}))^{-1}\epsilon^{2}\ell_{g}^{2\cdot 10^{5}}\big{)}\leq\exp\big{(}-\ell_{g}^{10^{5}}\big{)},

where we used (5.2)-(5.3), the fact that M20(ν+1)LM\leq 20(\nu+1)L and took LL large enough. Similarly, using the coupling defined in (5.70) and (C.2) (more precisely (3.5)), we obtain that

(5.116) Q(E12c|E18E10)5\CrSEPcoupling2G1/20Mexp((\CrSEPcoupling2(1+ν1))1ϵ2G1/100)exp(g2103).Q(E_{12}^{c}\,|\,E_{1-8}\cap E_{10})\leq 5\Cr{SEPcoupling2}\ell_{G}^{1/20}M\exp\big{(}-(\Cr{SEPcoupling2}(1+\nu^{-1}))^{-1}\epsilon^{2}\ell_{G}^{1/100}\big{)}\leq\exp(-\ell_{g}^{2\cdot 10^{3}}).

Finally, putting (5.112), (5.115) and (5.116) together, and using (5.2) with LL large enough, leads to (Erestartc)L100\mathbb{Q}(E_{\textnormal{restart}}^{c})\leq L^{-100}. This concludes the proof of (5.54) and thus of Lemma 5.2, taking L2L_{2} large enough so that (5.60)-(5.116) (which represent a finite number of constraints) hold. ∎

Remark 5.3.

In the coupling \mathbb{Q} constructed in Lemma 5.2, UU is a priori not independent from (η+,η)(\eta^{+},\eta^{-}); for instance a synchronous evolution of η+(+2g20)\eta^{+}(\cdot+2\ell_{g}^{20}) and η\eta^{-} after s3s_{3} could indicate that E6E_{6}, which depends on UU during [s1,s2][s_{1},s_{2}], has happened.

5.3. Proof of Proposition 3.4

Let L1L\geq 1 be an integer satisfying LL2(ρ,ϵ,ν)L\geq L_{2}(\rho,\epsilon,\nu), where L2L_{2} is given by Lemma 5.2. Let k:=L/Tk:=\lfloor L/T\rfloor, with TT as defined in (5.2). We start with a brief overview of the proof. To deduce (3.23), we will couple two walks X^+ρ+ϵ\widehat{X}^{+}\sim\mathbb{P}^{\rho+\epsilon} and X^ρ\widehat{X}^{-}\sim\mathbb{P}^{\rho} on the time interval [0,iT][0,iT], 1ik1\leq i\leq k, recursively in ii. The processes X^±\widehat{X}^{\pm} will be specified in terms of associated environments η^+𝐏ρ+ϵ\widehat{\eta}^{+}\sim\mathbf{P}^{\rho+\epsilon}, η^𝐏ρ\widehat{\eta}^{-}\sim\mathbf{P}^{\rho}, and an i.i.d. array U^=(U^w)w𝕃\widehat{U}=(\widehat{U}_{w})_{w\in\mathbb{L}} using Lemma 5.2 repeatedly, cf. (5.51)-(5.52). We will denote ^\widehat{\mathbb{Q}} the associated coupling measure defined below, which will also comprise associated auxiliary walks Y^±\widehat{Y}^{\pm} that will be defined using the construction of Y±Y^{\pm} in Section 5.1, iterated over ii and allow to keep control on the gap between X^+\widehat{X}^{+} and X^\widehat{X}^{-} via a combination of (5.52)-(5.53). The very possibility of iteration is guaranteed by the high-probability event ErestartE_{\textnormal{restart}} in (5.54).

We now proceed to make the above precise. For later reference we set Mi=20(ν+1)L(8ν+3)iM_{i}=20(\nu+1)L-(8\nu+3)i, for i{0,,k1}i\in\{0,\dots,k-1\}. We further recall the filtration t\mathcal{F}_{t} defined in (5.55) and write ^t\widehat{\mathcal{F}}_{t} below when adding hats to all processes involved. Finally, for all i{0,,k1}i\in\{0,\dots,k-1\}, let

(5.117) Bi=def.j=0i{(η^jT+(+Y^jT+),η^jT(+Y^jT) is (Mj,L)-balanced}B_{i}\stackrel{{\scriptstyle\text{def.}}}{{=}}\bigcap_{j=0}^{i}\left\{(\widehat{\eta}^{+}_{jT}(\cdot+\widehat{Y}^{+}_{jT}),\widehat{\eta}^{-}_{jT}(\cdot+\widehat{Y}^{-}_{jT})\text{ is }(M_{j},L)\text{-balanced}\right\}

(see items (i)-(iii) above Lemma 5.2 for notation).

By successive extensions of ^\widehat{\mathbb{Q}}, we will construct a coupling such that the following hold for all LL2L\geq L_{2} (as supplied by Lemma 5.2) and i{0,,k}i\in\{0,\dots,k\}:

  • (^\widehat{\mathbb{Q}}-1)

    The processes (η^t+,η^t)0tiT(\widehat{\eta}^{+}_{t},\widehat{\eta}^{-}_{t})_{0\leq t\leq iT}, (U^w)w𝕃,π2(w)iT1(\widehat{U}_{w})_{w\in\mathbb{L},\pi_{2}(w)\leq iT-1} (absent when i=0i=0) and (X^t±)0tiT(\widehat{X}^{\pm}_{t})_{0\leq t\leq iT} are defined under ^\widehat{\mathbb{Q}} with the correct marginal laws. That is, η^0+μρ+ϵ\widehat{\eta}^{+}_{0}\sim\mu_{\rho+\epsilon}, η^0μρ\widehat{\eta}^{-}_{0}\sim\mu_{\rho} and (η^t±)0<tiT(\widehat{\eta}^{\pm}_{t})_{0<t\leq iT} has the same law as (the restriction to [0,iT][0,iT]) of η±\eta^{\pm} under 𝐏η^0±\mathbf{P}^{\widehat{\eta}_{0}^{\pm}}. Moreover, (U^w)w𝕃,π2(w)iT1(\widehat{U}_{w})_{w\in\mathbb{L},\pi_{2}(w)\leq iT-1} are i.i.d. uniform variables on [0,1][0,1], and given η^0±\widehat{\eta}_{0}^{\pm}, (X^t±)0tiT(\widehat{X}^{\pm}_{t})_{0\leq t\leq iT} is ^iT\widehat{\mathcal{F}}_{iT}-measurable and has the law of (the restriction to [0,iT][0,iT] of) X±X^{\pm} under η^0±\mathbb{P}^{\widehat{\eta}_{0}^{\pm}}.

  • (^\widehat{\mathbb{Q}}-2)

    The processes (Y^t±)0tiT(\widehat{Y}^{\pm}_{t})_{0\leq t\leq iT} are ^iT\widehat{\mathcal{F}}_{iT}-measurable and X^iTY^iT\widehat{X}^{-}_{iT}\leq\widehat{Y}^{-}_{iT} and X^iT+Y^iT+\widehat{X}^{+}_{iT}\geq\widehat{Y}^{+}_{iT} hold ^\widehat{\mathbb{Q}}-a.s.

  • (^\widehat{\mathbb{Q}}-3)

    Y^0+=Y^0=0\widehat{Y}^{+}_{0}=\widehat{Y}^{-}_{0}=0, and if i1i\geq 1, with BiB_{i} as in (5.117),

    (5.118) 𝔼^[(Y^iT+Y^(i1)T+)(Y^iTY^(i1)T)|^(i1)T]1Bi1exp((logL)1/20).\mathbb{E}^{\widehat{\mathbb{Q}}}\left[\left.\big{(}\widehat{Y}^{+}_{iT}-\widehat{Y}^{+}_{(i-1)T}\big{)}-\big{(}\widehat{Y}^{-}_{iT}-\widehat{Y}^{-}_{(i-1)T}\big{)}\right|\widehat{\mathcal{F}}_{(i-1)T}\right]1_{B_{i-1}}\geq\exp(-(\log L)^{1/20}).
  • (^\widehat{\mathbb{Q}}-4)

    ^(B0c)=0\widehat{\mathbb{Q}}(B_{0}^{c})=0 and if i1i\geq 1, ^(Bic|Bi1)L100\widehat{\mathbb{Q}}(B_{i}^{c}|B_{i-1})\leq L^{-100}.

For i=0,i=0, we simply couple under ^\widehat{\mathbb{Q}} two configurations η^0+μρ+ϵ\widehat{\eta}^{+}_{0}\sim\mu_{\rho+\epsilon} and η^0μρ\widehat{\eta}^{-}_{0}\sim\mu_{\rho} such that a.s. η^0+(x)η^0(x)\widehat{\eta}^{+}_{0}(x)\geq\widehat{\eta}^{-}_{0}(x) for all xx\in\mathbb{Z}, which we can do with probability one by (P.3). We set X^0+=X^0=Y^0+=Y^0=0\widehat{X}^{+}_{0}=\widehat{X}^{-}_{0}=\widehat{Y}^{+}_{0}=\widehat{Y}^{-}_{0}=0. Thus (^\widehat{\mathbb{Q}}-1) and (^\widehat{\mathbb{Q}}-2) are satisfied, and (^\widehat{\mathbb{Q}}-3) is trivial. Finally ^(B0c)=0\widehat{\mathbb{Q}}(B_{0}^{c})=0 since η^0±\widehat{\eta}^{\pm}_{0} are in particular (M0,L)(M_{0},L)-balanced, whence (^\widehat{\mathbb{Q}}-4) holds.

Assume by induction that for some i{0,,k1}i\in\{0,\dots,k-1\}, we have constructed a coupling ^\widehat{\mathbb{Q}} with the above properties. We now proceed to extend ^\widehat{\mathbb{Q}} so as to have (^\widehat{\mathbb{Q}}-1)-(^\widehat{\mathbb{Q}}-4) with (i+1)(i+1) in place of ii. We first specify matters on the event BicB_{i}^{c}. Conditionally on ^iT\widehat{\mathcal{F}}_{iT}, if BicB_{i}^{c} occurs, we let η^t+\widehat{\eta}^{+}_{t} and η^t\widehat{\eta}^{-}_{t} evolve independently for iT<t(i+1)TiT<t\leq(i+1)T according to 𝐏η^iT+\mathbf{P}^{\widehat{\eta}^{+}_{iT}} and 𝐏η^iT\mathbf{P}^{\widehat{\eta}^{-}_{iT}} respectively, and independently of this, we choose U^w\widehat{U}_{w} as uniform random variables on [0,1][0,1] in an i.i.d. manner, for ww such that iTπ2(w)(i+1)T1iT\leq\pi_{2}(w)\leq(i+1)T-1. On BicB_{i}^{c}, we further let Y^iT+t+=Y^iT+t\widehat{Y}^{+}_{iT+t}=\widehat{Y}^{+}_{iT}-t and Y^iT+t=Y^iT+t\widehat{Y}^{-}_{iT+t}=\widehat{Y}^{-}_{iT}+t, for all 0<tT0<t\leq T and (Xt±)iTt(i+1)T(X^{\pm}_{t})_{iT\leq t\leq(i+1)T} evolve as in (2.5) and (2.6) with (η±,U)(\eta^{\pm},U) instead of (η,U)(\eta,U). With these choices it is clear that the inequalities in (^\widehat{\mathbb{Q}}-2) hold on BicB_{i}^{c}, since for instance Y^(i+1)T+Y^iT+TXiT+TX(i+1)T+\widehat{Y}^{+}_{(i+1)T}\leq\widehat{Y}^{+}_{iT}-T\leq X^{+}_{iT}-T\leq X^{+}_{(i+1)T} ^\widehat{\mathbb{Q}}-a.s., using the induction hypothesis and the fact that increments of X+X^{+} are bounded from below by 1-1. The inequality X^(i+1)TY^(i+1)T\widehat{X}^{-}_{(i+1)T}\leq\widehat{Y}^{-}_{(i+1)T} is derived similarly.

We now turn to the case that BiB_{i} occurs, which brings into play Lemma 5.2. Conditionally on ^iT\widehat{\mathcal{F}}_{iT} and on the event BiB_{i}, we couple (x,t)η^iT+t+(x+Y^iT+)(x,t)\mapsto\widehat{\eta}^{+}_{iT+t}(x+\widehat{Y}^{+}_{iT}) and (x,t)η^iT+t(x+Y^iT)(x,t)\mapsto\widehat{\eta}^{-}_{iT+t}(x+\widehat{Y}^{-}_{iT}) for xx\in\mathbb{Z} and t[0,iT]t\in[0,iT], as well as (U^w+(0,iT):w𝕃,π2(w)T1)(\widehat{U}_{w+(0,iT)}:w\in\mathbb{L},\,\pi_{2}(w)\leq T-1) following the coupling of (η+,η,U)(\eta^{+},\eta^{-},U) provided by Lemma 5.2, with the choice M=MjM=M_{j}. The requirement of (M,L)(M,L)-balancedness of the initial condition needed for Lemma 5.2 to apply is precisely provided by BiB_{i}, cf. (5.117).

Combining (5.51), the Markov property (P.1) applied at time iTiT, and in view of the choices made on BicB_{i}^{c}, it readily follows that the processes (η^t+,η^t)0t(i+1)T(\widehat{\eta}^{+}_{t},\widehat{\eta}^{-}_{t})_{0\leq t\leq(i+1)T}, (U^w)w𝕃,π2(w)(i+1)T1(\widehat{U}_{w})_{w\in\mathbb{L},\pi_{2}(w)\leq(i+1)T-1} thereby defined have the marginal laws prescribed in (^\widehat{\mathbb{Q}}-1). Moreover, by above application of Lemma 5.2, the processes X±X^{\pm} and Y±Y^{\pm} satisfying all of (5.52)-(5.54) are declared. Thus, setting

(5.119) X^iT+t±=X^iT±+Xt±,Y^iT+t±=Y^iT±+Yt±,for all 0tT,\begin{split}&\widehat{X}^{\pm}_{iT+t}=\widehat{X}^{\pm}_{iT}+X^{\pm}_{t},\quad\widehat{Y}^{\pm}_{iT+t}=\widehat{Y}^{\pm}_{iT}+Y^{\pm}_{t},\quad\text{for all $0\leq t\leq T$},\end{split}

it readily follows, combining (5.119) and the induction assumption on the law of (η^t±)0tiT(\widehat{\eta}^{\pm}_{t})_{0\leq t\leq iT}, (X^t±)0tiT(\widehat{X}^{\pm}_{t})_{0\leq t\leq iT}, combined with the Markov property (P.1) and that of the quenched law, that (X^t±)0t(i+1)T(\widehat{X}^{\pm}_{t})_{0\leq t\leq(i+1)T} declared by (5.119) has the desired marginal law, thus completing the verification of (^\widehat{\mathbb{Q}}-1) with (i+1)(i+1) in place of ii. Next, we show (^\widehat{\mathbb{Q}}-3) and (^\widehat{\mathbb{Q}}-4), before returning to (^\widehat{\mathbb{Q}}-2). Since Y^(i+1)T±Y^iT±=YT±\widehat{Y}^{\pm}_{(i+1)T}-\widehat{Y}^{\pm}_{iT}=Y^{\pm}_{T} by (5.119), the inequality (5.118) with (i+1)(i+1) in place of ii is an immediate consequence of (5.53). Hence (^\widehat{\mathbb{Q}}-3) holds. Finally, by construction of the coupling extension on the event BiB_{i}, which uses Lemma 5.2, and in view of (5.119) and (5.117), the failure of Bi+1B_{i+1} on the event BiB_{i} amounts to the failure of ErestartE_{\textnormal{restart}} in (5.54), from which (^\widehat{\mathbb{Q}}-4) follows with (i+1)(i+1) in place of ii.

It remains to show that (^\widehat{\mathbb{Q}}-2) holds with (i+1)(i+1) in place of ii. To this effect, we introduce two auxiliary processes (Z^t±)0t(i+1)T(\widehat{Z}^{\pm}_{t})_{0\leq t\leq(i+1)T}, defined as

(5.120) Z^t±={Y^t±, if 0tiT,Y^iT±+Xt±, if 0<tT.\widehat{Z}^{\pm}_{t}=\begin{cases}\widehat{Y}^{\pm}_{t},&\text{ if }0\leq t\leq iT,\\ \widehat{Y}^{\pm}_{iT}+X_{t}^{\pm},&\text{ if }0<t\leq T.\end{cases}

Combining the induction assumption (^\widehat{\mathbb{Q}}-2), the definition of Y^t±\widehat{Y}^{\pm}_{t} and Z^t±\widehat{Z}^{\pm}_{t} in (5.119) and (5.120) (the latter implying in particular that Z^iT±=Y^iT±\widehat{Z}^{\pm}_{iT}=\widehat{Y}^{\pm}_{iT}), one readily deduces from property (5.52) the ^\widehat{\mathbb{Q}}-almost sure inequalities

(5.121) Z^(i+1)TY^(i+1)T and Y^(i+1)T+Z^(i+1)T+.\widehat{Z}^{-}_{(i+1)T}\leq\widehat{Y}^{-}_{(i+1)T}\text{ and }\widehat{Y}^{+}_{(i+1)T}\leq\widehat{Z}^{+}_{(i+1)T}.

To deduce from this the analogous inequalities with X^\widehat{X} in place of Z^\widehat{Z}, we apply Lemma 2.2, with (X,X~)=(Z^+,X^+)(X,\widetilde{X})=(\widehat{Z}^{+},\widehat{X}^{+}), η=η~=η+\eta=\widetilde{\eta}=\eta^{+} (whence (2.15) plainly holds), π1(w)=Z^iT+=Y^iT+\pi_{1}(w^{\prime})=\widehat{Z}^{+}_{iT}=\widehat{Y}^{+}_{iT}, π1(w)=X^iT+\pi_{1}(w)=\widehat{X}^{+}_{iT}, π2(w)=π2(w)\pi_{2}(w)=\pi_{2}(w^{\prime}) and K=×[iT,(i+1)T]K=\mathbb{Z}\times[iT,(i+1)T] to deduce that Z^(i+1)T+X^(i+1)T+\widehat{Z}^{+}_{(i+1)T}\leq\widehat{X}^{+}_{(i+1)T} holds ^\widehat{\mathbb{Q}}-a.s. Note that the condition π1(w)π1(w)\pi_{1}(w^{\prime})\leq\pi_{1}(w) necessary for Lemma 2.2 to apply is in force by induction hypothesis in (^\widehat{\mathbb{Q}}-2). Together with (5.121), this yields one of the desired inequalities in (^\widehat{\mathbb{Q}}-2) with i+1i+1 in place of ii. The other one is obtained in a similar way using Lemma 2.2. This completes the proof of the induction step.

We now use the coupling ^\widehat{\mathbb{Q}}, which satisfies (^\widehat{\mathbb{Q}}-1)-(^\widehat{\mathbb{Q}}-4) for all 0ik0\leq i\leq k (and LL2L\geq L_{2}), to complete the proof of (3.23). To this end, we first extend the laws of X^t±\widehat{X}_{t}^{\pm} to all t>kTt>kT using the Markov property, by sampling X^kT+±\widehat{X}^{\pm}_{kT+\cdot} independently conditionally on ^kT.\widehat{\mathcal{F}}_{kT}. In particular, recalling that k=L/Tk=\lfloor L/T\rfloor, this implies that X^L±\widehat{X}_{L}^{\pm} is declared under ^\widehat{\mathbb{Q}}. We thus proceed to derive a suitable lower bound on 𝔼^[X^L+X^L]\mathbb{E}^{\widehat{\mathbb{Q}}}[\widehat{X}_{L}^{+}-\widehat{X}_{L}^{-}], which is well-defined, and from which (3.23) will follow.

Using that |X^n+1±X^n±|1|\widehat{X}_{n+1}^{\pm}-\widehat{X}_{n}^{\pm}|\leq 1 for any n0n\geq 0, we obtain (with k=L/Tk=\lfloor L/T\rfloor) that

(5.125) 𝔼^[X^L+X^L]+2T𝔼^[X^L+X^L]+2(LkT)𝔼^[X^kT+X^kT](^-2)𝔼^[Y^kT+Y^kT]1ik𝔼^[(Y^iT+Y^(i1)T+)(Y^iTY^(i1)T)](^-3)ke(logL)1/202T1ik^(Bi1c)L2Te(logL)1/202L^(Bk1c),\mathbb{E}^{\widehat{\mathbb{Q}}}[\widehat{X}_{L}^{+}-\widehat{X}_{L}^{-}]+2T\geq\mathbb{E}^{\widehat{\mathbb{Q}}}[\widehat{X}_{L}^{+}-\widehat{X}_{L}^{-}]+2(L-kT)\geq\mathbb{E}^{\widehat{\mathbb{Q}}}\big{[}\widehat{X}^{+}_{kT}-\widehat{X}^{-}_{kT}\big{]}\\ \stackrel{{\scriptstyle\textnormal{($\widehat{\mathbb{Q}}$-2)}}}{{\geq}}\mathbb{E}^{\widehat{\mathbb{Q}}}\big{[}\widehat{Y}^{+}_{kT}-\widehat{Y}^{-}_{kT}\big{]}{\geq}\sum_{1\leq i\leq k}\mathbb{E}^{\widehat{\mathbb{Q}}}\Big{[}\big{(}\widehat{Y}^{+}_{iT}-\widehat{Y}^{+}_{(i-1)T}\big{)}-\big{(}\widehat{Y}^{-}_{iT}-\widehat{Y}^{-}_{(i-1)T}\big{)}\Big{]}\\ \stackrel{{\scriptstyle\textnormal{($\widehat{\mathbb{Q}}$-3)}}}{{\geq}}ke^{-(\log L)^{1/20}}-2T\sum_{1\leq i\leq k}\widehat{\mathbb{Q}}(B_{i-1}^{c})\geq\frac{L}{2T}e^{-(\log L)^{1/20}}-2L\widehat{\mathbb{Q}}(B_{k-1}^{c}),

for large enough LL, where, in the second inequality of the second line, we have used that Y^0±=0\widehat{Y}^{\pm}_{0}=0, see (^\widehat{\mathbb{Q}}-3), and in the last line, we have first used that the difference of increments is deterministically bounded from below by 2T-2T (see (5.119) and Lemma 5.1), and for the last inequality that the events BicB_{i}^{c} are increasing in ii, cf. (5.117). We also used in various places that LTkTLL-T\leq kT\leq L.

It remains to suitably estimate the probability ^(Bk1c)\widehat{\mathbb{Q}}(B_{k-1}^{c}) appearing in the last line of (5.125), for which we use (^\widehat{\mathbb{Q}}-4) and a straightforward induction argument to bound

^(Bk1c)^(Bk1c|Bk2)+^(Bk2c)(^-4)L100+^(Bk2c)(k1)L100L99,\widehat{\mathbb{Q}}(B_{k-1}^{c})\leq\widehat{\mathbb{Q}}(B_{k-1}^{c}|B_{k-2})+\widehat{\mathbb{Q}}(B_{k-2}^{c})\stackrel{{\scriptstyle\text{($\widehat{\mathbb{Q}}$-4)}}}{{\leq}}L^{-100}+\widehat{\mathbb{Q}}(B_{k-2}^{c})\leq\dots\leq(k-1)L^{-100}\leq L^{-99},

using also in the penultimate step that ^(B0c)=0\widehat{\mathbb{Q}}(B_{0}^{c})=0. Feeding this into (5.125) yields that

(5.126) 𝔼^[X^L+X^L]L2Te(logL)1/202L982T3\CrC:approx(logL)100,\begin{split}\mathbb{E}^{\widehat{\mathbb{Q}}}[\widehat{X}_{L}^{+}-\widehat{X}_{L}^{-}]&\geq\frac{L}{2T}e^{-(\log L)^{1/20}}-2L^{-98}-2T\geq 3\Cr{C:approx}(\log L)^{100},\end{split}

as soon as LL is large enough (recall TT from (5.2)). Dividing by LL and applying (3.19) whilst observing that X^L+\widehat{X}_{L}^{+} has the same law under ^\widehat{\mathbb{Q}} as XLX_{L} under ρ+ϵ,L\mathbb{P}^{\rho+\epsilon,L} and X^L\widehat{X}_{L}^{-} has the same law under ^\widehat{\mathbb{Q}} as XLX_{L} under ρ,L\mathbb{P}^{\rho,L}, (5.126) implies that

(5.127) vL(ρ+δ)vL(ρ)3\CrC:approxL1(logL)100,v_{L}(\rho+\delta)-v_{L}(\rho)\geq{3\Cr{C:approx}L^{-1}(\log L)^{100}},

which concludes the proof of Proposition 3.4.

Remark 5.4.

As mentioned at (1.17), we prove in fact a much stronger statement than Proposition 3.4, owing to (5.126), namely that

(5.128) vL(ρ+δ)vL(ρ)Lo(1)v_{L}(\rho+\delta)-v_{L}(\rho)\geq L^{o(1)}

where the o(1)o(1) denotes a negative quantity that goes to 0 as LL\rightarrow\infty. This is however insufficient to imply directly Theorem 3.1, and the renormalisation in Section 4 is still essential to improve (5.128) to a right-hand side bounded away from 0 as LL\rightarrow\infty.

Remark 5.5 (Necessity for couplings with quenched initial condition).

We explain here the main reason why we need quenched conditions for our couplings. In short, this is due to the lack of an invariant measure (or reasonable proxy thereof) from the point of view of the walk.

More precisely, abbreviating t=s2s1t=s_{2}-s_{1}, the only a-priori lower bound we have for the probability that XρX^{\rho} and Xρ+εX^{\rho+\varepsilon} drift away linearly from each other during [s1,s2][s_{1},s_{2}] (corresponding to (E6)\mathbb{Q}(E_{6}) at (5.90)) is exp(Ct)\exp(-Ct) for some large constant CC by uniform ellipticity – we are in fact precisely trying to derive a better bound in this section.

But this gap is necessary to create a difference between Xρ+ϵXρX^{\rho+\epsilon}-X^{\rho} on a time interval of length TT. Hence, to accrue a significant gain in expectation between Xρ+ϵXρX^{\rho+\epsilon}-X^{\rho}, we need to repeat this at least exp(Ct)\exp(Ct) times. Thus, during that time, XρX^{\rho} and Xρ+ϵX^{\rho+\epsilon} could straddle an interval of width at least exp(Ct)\exp(Ct). The main issue is that we have no a priori information on their local environment (which would not be the case if we had access to an invariant measure and could estimate the speed of convergence to it). Hence if the coupling (C.2.2), that we use between s1s_{1} and s2s_{2} on a interval much narrower than exp(Ct)\exp(Ct), was only valid under the annealed product Bernoulli initial condition (which is a priori not what the walk sees), we could resort to the annealed-to-quenched trick (via Markov’s inequality) and a union bound over exp(Ct)\exp(Ct) intervals to control the probability that the coupling fails. However, the failure probability at (C.2.2) is exp(Ct1/4)exp(Ct)\exp(-C^{\prime}t^{1/4})\gg\exp(-Ct), hence we cannot obtain any non-trivial bound this way. Note that this does not depend on the choice of tt. Furthermore due to the diffusivity of the environment particles and large deviation considerations, it seems unlikely that one could improve the bound of (C.2.2) beyond exp(Ct1/2)\exp(-Ct^{1/2}).

This is why we resort to some quenched control, cf. items (i)-(iii) above Lemma 5.2, and also (5.117). The empirical density was the most accessible and relevant statistic (in particular if the environment is conservative, as is the case of SEP). For similar reasons, we had to establish (C.1) in a quenched setting, to ensure that whatever the distribution of the environment around the walker at time iTiT for some i1i\geq 1 (as long as it is balanced), there is still a uniformly low probability not to have the required empirical density at time s1s_{1} (see (5.88)) to perform the coupling of (C.2.2). Of course the necessity for quenched couplings encapsulated in (C.1) and in particular (C.2), means that we have (more) work to do in order to verify this in specific instances, as we do for SEP in the next section.

6 Exclusion process and couplings

We start by giving in §6.1 a formal definition of the main environment η\eta of interest in this article, the symmetric simple exclusion process (SEP), and first check that it fits the setup of §2.2, in particular, that the basic properties (P.1)–(2.2) listed in §2.2 hold. The main result of this section, proved in §6.2, is to show that the SEP satisfies the conditions (C.1)-(C.3) stated in §3.1; see Proposition 6.3 below. This implies that our main result, Theorem 3.1, applies in this case; cf. also Theorem 1.2 and its proof in §3.2. We refer to Appendix B for another environment η\eta of interest which fits this framework.

6.1. Definition of SEP and basic properties

We fix a parameter ν>0\nu>0, which will be constant throughout this section and often implicit in our notation. The (rate ν\nu) symmetric simple exclusion process (SEP) is the Markov process on the state space {0,1}\{0,1\}^{\mathbb{Z}} (tacitly viewed as a subset of Σ\Sigma, cf. §2.2) with (pre-)generator

(6.1) Lf(η)=x,y:|xy|=1𝟙{η(x)=1,η(y)=0}ν2(f(ηxy)f(η)),\displaystyle Lf(\eta)=\sum_{x,y\in\mathbb{Z}:|x-y|=1}\mathds{1}_{\{\eta(x)=1,\eta(y)=0\}}\frac{\nu}{2}\left(f(\eta_{xy})-f(\eta)\right),

for η{0,1}\eta\in\{0,1\}^{\mathbb{Z}} and ff in the domain of LL, where ηxy\eta_{xy} is the configuration obtained from η\eta by exchanging the states of xx and yy, i.e. such that ηxy(x)=η(y)\eta_{xy}(x)=\eta(y), ηxy(y)=η(x)\eta_{xy}(y)=\eta(x) and ηxy(z)=η(z)\eta_{xy}(z)=\eta(z) for all z{x,y}z\in\mathbb{Z}\setminus\{x,y\}; see [51, Chap. VIII]. We denote 𝐏SEPη0\mathbf{P}_{\textrm{SEP}}^{\eta_{0}} its canonical law with initial configuration η0\eta_{0} and drop the subscript SEP whenever there is no risk of confusion. In words, (6.1) entails that the vertices xx such that η(x)=1\eta(x)=1, which can be seen as the locations of particles evolve like continuous-time symmetric simple random walks on \mathbb{Z} with rate ν\nu that obey the exclusion rule; that is, particles are only allowed to jump onto empty locations.

It will often be useful to consider the interchange process on \mathbb{Z}, with generator L^\widehat{L} defined as in (6.1) but omitting the exclusion constraint {η(x)=1,η(y)=0}\{\eta(x)=1,\eta(y)=0\}, which interchanges the state of neighbors xx and yy independently at rate ν/2\nu/2. We will use the following specific construction of this process. Let E={{x,x+1}:x}E=\{\{x,x+1\}:x\in\mathbb{Z}\} denote the set of edges on \mathbb{Z} and 𝐏^\widehat{\mathbf{P}} be a probability governing independent Poisson counting processes 𝒫e\mathcal{P}_{e} of intensity ν/2\nu/2 on +\mathbb{R}_{+} attached to every edge eEe\in E. For any given η0{0,1}{\eta_{0}}\in\{0,1\}^{\mathbb{Z}}, one defines (ηt)t0(\eta_{t})_{t\geq 0} under 𝐏^\widehat{\mathbf{P}} by exchanging the states of η\eta at xx and yy every time the ‘clock rings’ for 𝒫e\mathcal{P}_{e}, where e={x,y}e=\{x,y\}. This is well-defined up to a set of measure zero. Then for every η0{0,1}\eta_{0}\in\{0,1\}^{\mathbb{Z}},

(6.2) (ηt)t0 has the same law under 𝐏^ and 𝐏SEPη0.\text{$(\eta_{t})_{t\geq 0}$ has the same law under $\widehat{\mathbf{P}}$ and $\mathbf{P}_{\textrm{SEP}}^{\eta_{0}}$}.

This follows upon observing that the states of neighboring sites suffering the exclusion constraint can also be exchanged. For our purpose, these two processes are equivalent, but note that they differ when one distinguishes the particles of the system (for instance studying the motion of a tagged particle).

A useful feature of this alternative description is the following. A particle trajectory of the interchange process is obtained by following the trajectory of a state xx\in\mathbb{Z} such that η0(x)=1\eta_{0}(x)=1 (a particle) under 𝐏^\widehat{\mathbf{P}}. We won’t define this formally but roughly speaking, if ee and ee^{\prime} are the two edges incident on xx, one waits until the minimum of the first arrival times of these two processes (which is an exponential variable with parameter ν\nu) and jumps across the corresponding edge. Then one repeats this procedure. In particular, it immediately follows that

(6.3) for each xx such that η0(x)=1\eta_{0}(x)=1, the particle trajectory of xx under 𝐏^\widehat{\mathbf{P}} follows the law of a continuous time simple random walk with jump rate ν\nu.

Recalling the properties (P.1)–(2.2) from §2.2, we first record the following fact.

Lemma 6.1.

With

(6.4) μρ=((1ρ)δ0+ρδ1),ρJ=def.(0,1),\mu_{\rho}=\big{(}(1-\rho)\delta_{0}+\rho\delta_{1}\big{)}^{\otimes\mathbb{Z}},\quad\rho\in J\stackrel{{\scriptstyle\textnormal{def.}}}{{=}}(0,1),

the measures (𝐏η0:η0{0,1})(\mathbf{P}^{\eta_{0}}:\eta_{0}\in\{0,1\}^{\mathbb{Z}}) with 𝐏η0=𝐏SEPη0\mathbf{P}^{\eta_{0}}=\mathbf{P}_{\textnormal{SEP}}^{\eta_{0}} and (μρ:ρJ)(\mu_{\rho}:\rho\in J) satisfy all of (P.1)–(2.2).

Proof.

Property (P.1) is classical, see [51, Chap. I, Thm. 3.9, p.27] along with Example 3.1(d), p.21 of the same reference. So is (P.2), i.e. the stationarity of the measure μρ\mu_{\rho} in (6.4), see [51, Chap. VIII, Thm. 1.12, p.369]. The required coupling (for two given initial configurations η0η0\eta_{0}^{\prime}\preccurlyeq\eta_{0}) needed to verify the quenched monotonicity asserted in (P.3),i) is simply obtained by realizing the process η=(ηt)t0\eta=(\eta_{t})_{t\geq 0} under the auxiliary measure 𝐏^η0\widehat{\mathbf{P}}^{\eta_{0}^{\prime}} and 𝐏^η0\widehat{\mathbf{P}}^{\eta_{0}} using the same Poisson processes (𝒫e){(\mathcal{P}_{e})}. In particular this measure yields a coupling over all possible initial distributions, including η0\eta_{0} and η0\eta_{0}^{\prime}, and this coupling is seen to preserve the partial order η0η0\eta_{0}^{\prime}\preccurlyeq\eta_{0} for all t>0t>0. The monotonicity in (P.3),ii) is classical. Lastly, upon observing that η0[0,1]\eta_{0}[0,\ell-1] is a binomial random variable with parameters \ell and ρ\rho under 𝐏ρ\mathbf{P}_{\rho}, property (2.2) is obtained by combining (A.3) and (A.4), which are well-known large deviation estimates. ∎

In anticipation of §6.2, we now collect a simple lemma to bound the linear deviations of an SEP particle, which we will routinely use in our couplings below.

Lemma 6.2.

Let Z=(Zt)t0Z=(Z_{t})_{t\geq 0} denote a simple random walk on \mathbb{Z} with jump rate ν\nu, starting from 0 at time 0, with law denoted by PP. For all t>0t>0 and k,ak,a\in\mathbb{N}:

(6.7) P(max0st|Zs|2kνt+a)P(Z makes more than 2kνt+a jumps during [0,t])e(2kνt+a)/8.P\Big{(}\max_{0\leq s\leq t}|Z_{s}|\geq 2k\nu t+a\Big{)}\\ \leq P(Z\text{ makes more than $2k\nu t+a$ jumps during }[0,t])\leq e^{-(2k\nu t+a)/8}.
Proof.

The first inequality is immediate since ZZ only performs nearest-neighbor jumps. As for the second one, remark that the number NN of jumps performed by ZZ during [0,t][0,t] is a Poisson random variable with parameter νt\nu t. Hence by (A.1) applied with λ=νt\lambda=\nu t and x=(2k1)νt+a(2kνt+a)/2>0x=(2k-1)\nu t+a\geq(2k\nu t+a)/2>0, we obtain that

(6.8) P(N2kνt+a)exp(((2k1)νt+a)22(2kνt+a))exp((2kνt+a)/8),P(N\geq 2k\nu t+a)\leq\exp\left(-\frac{((2k-1)\nu t+a)^{2}}{2(2k\nu t+a)}\right)\leq\exp\left(-(2k\nu t+a)/8\right),

and the conclusion follows. ∎

6.2. Conditions (C.1)(C.3) for SEP

We now proceed to verify that the conditions introduced in §3.1 all hold for the exclusion process introduced in §6.1, as summarized in the next proposition. Its proof occupies the bulk of this section. These properties (above all, (C.1) and (C.2)) are of independent interest.

Proposition 6.3.

For (𝐏η0:η0{0,1}+)(\mathbf{P}^{\eta_{0}}:\eta_{0}\in\{0,1\}^{\mathbb{Z}_{+}}) with 𝐏η0=𝐏SEPη0\mathbf{P}^{\eta_{0}}=\mathbf{P}^{\eta_{0}}_{\textnormal{SEP}}, J=(0,1)J=(0,1) (cf. (6.4)), and with ν\nu as appearing in (6.1), all of (C.1)(C.2)(C.2.1)(C.2.2) and (C.3) hold.

Proof.

This follows directly by combining Lemmas 6.46.56.76.9 and 6.10 below. ∎

We now proceed to investigate each of the relevant conditions separately. Throughout the remainder of this section, we work implicitly under the assumptions of Proposition 6.3. In particular, in stating that some property PP ‘holds for SEP,’ we mean precisely that PP is verified for the choice (𝐏η0:η0{0,1}+)(\mathbf{P}^{\eta_{0}}:\eta_{0}\in\{0,1\}^{\mathbb{Z}_{+}}) with 𝐏η0=𝐏SEPη0\mathbf{P}^{\eta_{0}}=\mathbf{P}^{\eta_{0}}_{\textnormal{SEP}} for all ρJ=(0,1)\rho\in J=(0,1), and with ν\nu the rate parameter underlying the construction of SEP.

Lemma 6.4.

The condition (C.1) holds for SEP, with no restriction on the choice of 1\ell^{\prime}\geq 1.

We give a brief overview of the proof. A key idea is to exploit the fact that SEP particles, although not independent, are in fact ‘more regularly’ spread out than a bunch of independent random walks (starting from the same initial positions), which is due to the inherent negative association of the SEP. This fact is implicit in the bound (6.10) below, which is borrowed from [43] (itself inspired from [70]), the proof of which uses in a crucial way an inequality due to Liggett, see [51, Chap. VIII, Prop. 1.7, p. 366], encapsulating this property. With this observation, it is enough to argue that after time t2t\gg\ell^{2} (where \ell is the precision mesh of the empirical density of the initial configuration, as in (C.1)), random walks have diffused enough to forget their initial positions and average their density, which follows from classical heat kernel estimates.

Proof.

We focus on proving the upper inequality (i.e. when each η0(I)(ρ+ε)\eta_{0}(I)\leq(\rho+\varepsilon)\ell), and comment where necessary on the minor adjustments needed to derive the other inequality in the course of the proof. Let ρ,ε(0,1)\rho,\varepsilon\in(0,1), H,,t1H,\ell,t\geq 1 and η0\eta_{0} be such that the conditions of (C.1) hold. We will in fact show (C.1) for an arbitrary value of 1\ell^{\prime}\geq 1 although the restriction to \ell^{\prime}\leq\ell is sufficient for later purposes (note that the statement is empty if >2(H2νt)\ell^{\prime}>2(H-2\nu t)). Thus, let 1\ell^{\prime}\geq 1 and \mathcal{I} denote the set of intervals of length \ell^{\prime} included in [H+2νt,H2νt][-H+2\nu t,H-2\nu t], and define

(6.9) η¯t(I)=xI𝐏η0(ηt(x)=1),I,\bar{\eta}_{t}(I)=\sum_{x\in I}\mathbf{P}^{\eta_{0}}(\eta_{t}(x)=1),\quad I\in\mathcal{I},

the average number of occupied sites of II after time t0t\geq 0. Note that η¯0(I)=η0(I)\bar{\eta}_{0}(I)=\eta_{0}(I). By Lemma 2.3 of [43] (see also Lemma 5.4 of [70]) one knows that for all t0t\geq 0 and suitable \Crdensitystableexpo(0,)\Cr{densitystableexpo}\in(0,\infty),

(6.10) 𝐏η0(ηt(I)η¯t(I)+ε)exp(\Crdensitystableexpoε2)\mathbf{P}^{\eta_{0}}\left(\eta_{t}(I)\geq\bar{\eta}_{t}(I)+\varepsilon\ell^{\prime}\right)\leq\exp(-\Cr{densitystableexpo}\varepsilon^{2}\ell^{\prime})

(in fact the bound (6.10) holds for any initial configuration η0\eta_{0}). Thus, if

(6.11) maxIη¯t(I)(ρ+2ε),\max_{I\in\mathcal{I}}\bar{\eta}_{t}(I)\leq(\rho+2\varepsilon)\ell^{\prime},

under our assumptions on η0\eta_{0} and tt, then (6.10), a union bound over \mathcal{I}, and the upper bound on the maximum in (6.11) yield that

(6.12) 𝐏η0(maxIηt(I)>(ρ+3ε))2Hexp(\Crdensitystableexpoε2).\mathbf{P}^{\eta_{0}}\big{(}\max_{I\subseteq\mathcal{I}}\eta_{t}(I)>(\rho+3\varepsilon)\ell^{\prime}\big{)}\leq 2H\exp(-\Cr{densitystableexpo}\varepsilon^{2}\ell^{\prime}).

A companion inequality to (6.10) can be deduced in a similar way using the lower bound on the minimum in (6.11) and exploiting symmetry, i.e. rewriting {ηt(I)<(ρ3ε)}={ξt(I)>((1ρ)+3ε)}\{\eta_{t}(I)<(\rho-3\varepsilon)\ell^{\prime}\}=\{\xi_{t}(I)>((1-\rho)+3\varepsilon)\ell^{\prime}\} where ξt=1ηt\xi_{t}=1-\eta_{t}, while observing that (ξt)t0(\xi_{t})_{t\geq 0} has law 𝐏ξ0\mathbf{P}^{\xi_{0}}, cf. (6.1). The conclusion (C.1) then follows.

Therefore, we are left with showing (6.11). In view of (6.9), it is enough to prove that for every x[H+2νt,H2νt]x\in[-H+2\nu t,H-2\nu t], and under our assumptions on η0\eta_{0} and tt,

(6.13) 𝐏η0(ηt(x)=1)ρ+2ε.\mathbf{P}^{\eta_{0}}(\eta_{t}(x)=1)\leq\rho+2\varepsilon.

Let (Zs)s0(Z_{s})_{s\geq 0} denote a continuous-time symmetric simple random walk on \mathbb{Z} with jump rate ν\nu, defined under an auxiliary probability PP, and let ps(x,y)=P(Zs=y|Z0=x)p_{s}(x,y)={P}(Z_{s}=y\,|Z_{0}=x), for x,yx,y\in\mathbb{Z} and s0s\geq 0 denote its transition probabilities, which are symmetric in xx and yy. Writing {ηt(x)=1}\{\eta_{t}(x)=1\} as the disjoint (due to the exclusion constraint) union over yy\in\mathbb{Z} of the event that η0(y)=1\eta_{0}(y)=1 and the particle starting at yy is located at xx at time tt, it follows using (6.3) that for all xx\in\mathbb{Z},

(6.14) 𝐏η0(ηt(x)=1)=ypt(y,x)η0(y)=ySpt(x,y),\mathbf{P}^{\eta_{0}}(\eta_{t}(x)=1)=\sum_{y\in\mathbb{Z}}p_{t}(y,x)\eta_{0}(y)=\sum_{y\in S}p_{t}(x,y),

where S:={y:η0(y)=1}S:=\{y\in\mathbb{Z}:\eta_{0}(y)=1\} and we used reversibility in the last step. Using the rewrite (6.14), we first show the upper bound in (6.13). Let us define ct=\ClC:ctνtlog(νt)c_{t}=\Cl{C:c-t}\sqrt{\nu t\log(\nu t)}, where the constant \CrC:ct>0\Cr{C:c-t}>0 will be chosen below. Recalling that x[H+2νt,H2νt]x\in[-H+2\nu t,H-2\nu t], we cover the sites of [xct,x+ct][H,H][x-c_{t},x+c_{t}]\subseteq[-H,H] using intervals I1,,IqI_{1},\ldots,I_{q} all contained in this interval and each containing \ell sites, with q=(2ct+1)/q=\lceil(2c_{t}+1)/\ell\rceil. We may assume that all but the last interval IqI_{q} are disjoint. For later reference, we note that |IrS|(ϱ+ε)|I_{r}\cap S|\leq(\varrho+\varepsilon)\ell for all 1rq1\leq r\leq q by assumption on η0\eta_{0}. By (6.14), we have that

(6.15) 𝐏η0(ηt(x)=1)y[xct,x+ct]pt(x,y)+r=1qySIrpt(x,y)\mathbf{P}^{\eta_{0}}(\eta_{t}(x)=1)\leq\sum_{y\in\mathbb{Z}\setminus[x-c_{t},x+c_{t}]}p_{t}(x,y)+\sum_{r=1}^{q}\sum_{y\in S\cap I_{r}}p_{t}(x,y)

We will look at the two terms on the right-hand side separately. Using (A.1)), one knows that with NN denoting the number of jumps of a continuous time random walk with jump rate ν\nu until time t>0t>0, one has (A)1Ceνt/C\mathbb{P}(A)\geq 1-Ce^{-\nu t/C} with A{2νt/3N4νt/3}A\coloneqq\{2\nu t/3\leq N\leq 4\nu t/3\}. Combining this with the deviation estimate (A.3) yields that

(6.19) y[xct,x+ct]pt(x,y)=(Ac)+2k=2νt/34νt/3(N=k)yct(Bin(k,12)=y+k2)Ceνt/C+2k=2νt/34νt/3(N=k)exp(ct216νt)Ceνt/C+2exp(\CrC:ct2log(νt)16)1νtε3,\sum_{y\in\mathbb{Z}\setminus[x-c_{t},x+c_{t}]}p_{t}(x,y)=\mathbb{P}(A^{c})+2\sum_{k=\lceil 2\nu t/3\rceil}^{\lfloor 4\nu t/3\rfloor}\mathbb{P}(N=k)\sum_{y^{\prime}\geq c_{t}}\mathbb{P}\left(\text{Bin}(k,\tfrac{1}{2})=\tfrac{y^{\prime}+k}{2}\right)\\ \leq Ce^{-\nu t/C}+2\sum_{k=\lceil 2\nu t/3\rceil}^{\lfloor 4\nu t/3\rfloor}\mathbb{P}(N=k)\exp\left(-\frac{c_{t}^{2}}{16\nu t}\right)\\ \leq Ce^{-\nu t/C}+2\exp\left(-\frac{\Cr{C:c-t}^{2}\log(\nu t)}{16}\right)\leq\frac{1}{\nu t}\leq\frac{\varepsilon}{3},

where we choose \CrC:ct11\Cr{C:c-t}\geq\sqrt{11}, use that νt\Crdensitystable\nu t\geq\Cr{densitystable}, thus choosing \Crdensitystable\Cr{densitystable} large enough, and use the conditions from (C.1) for the last inequality.

Next, we want to deal with the points in the interval [xct,x+ct][x-c_{t},x+c_{t}]. Recall that a continuous-time simple random walk with rate ν\nu at time tt has the same law as a rate 11 continuous-time simple random walk at time νt\nu t. Hence, by [50, Theorem 2.5.6], for all x,y,zx,y,z such that y,z[xct,x+ct]y,z\in[x-c_{t},x+c_{t}] and |zy||z-y|\leq\ell, using that ct<νt/2c_{t}<\nu t/2 for \Crdensitystable\Cr{densitystable} large enough, with CC changing from line to line,

(6.22) pt(x,y)pt(x,z)exp(|xz|2|xy|22νt+C(1νt+|xy|3+|xz|3(νt)2))exp(clog(νt)νt+C1+log3/2(νt)νt)exp(Cε\Crdensitystable)1+ε3,\frac{p_{t}(x,y)}{p_{t}(x,z)}\leq\exp\left(\frac{|x-z|^{2}-|x-y|^{2}}{2\nu t}+C\left(\frac{1}{\sqrt{\nu t}}+\frac{|x-y|^{3}+|x-z|^{3}}{(\nu t)^{2}}\right)\right)\\ \leq\exp\bigg{(}c\frac{\ell\sqrt{\log(\nu t)}}{\sqrt{\nu t}}+C\frac{1+\log^{3/2}(\nu t)}{\sqrt{\nu t}}\bigg{)}\leq\exp\left(C\frac{\varepsilon}{\sqrt{\Cr{densitystable}}}\right)\leq 1+\frac{\varepsilon}{3},

where we used that 4νt>\Crdensitystable2ε2log3(νt)4\nu t>\Cr{densitystable}\ell^{2}\varepsilon^{-2}\log^{3}(\nu t) and 1\ell\geq 1, from (C.1), and chose \Crdensitystable\Cr{densitystable} large enough. Using [50, Theorem 2.5.6] again, we also have that, as soon as \Crdensitystable\Cr{densitystable} is large enough, using the conditions from (C.1),

(6.23) pt(x,z)1νtε3.p_{t}(x,z)\leq\frac{1}{\sqrt{\nu t}}\leq\frac{\varepsilon}{3\ell}.

Now, recalling that the intervals IrI_{r} are disjoint for 1rq11\leq r\leq q-1 and that they all have cardinality \ell, using (6.22) and (6.23), we have that

(6.27) r=1qySIrpt(x,y)r=1q1ySIrpt(x,y)+ySIqpt(x,y)(1+ε3)r=1q1ySIrminzIrpt(x,z)+ySIqε3(1+ε3)(ρ+ε)r=1q1|Ir|minzIrpt(x,z)+ε|Iq|3(1+ε3)(ρ+ε)r=1q1ySIrpt(x,y)+ε3(1+ε3)(ρ+ε)+ε3.\sum_{r=1}^{q}\sum_{y\in S\cap I_{r}}p_{t}(x,y)\leq\sum_{r=1}^{q-1}\sum_{y\in S\cap I_{r}}p_{t}(x,y)+\sum_{y\in S\cap I_{q}}p_{t}(x,y)\\ \leq\left(1+\frac{\varepsilon}{3}\right)\sum_{r=1}^{q-1}\sum_{y\in S\cap I_{r}}\min_{z\in I_{r}}p_{t}(x,z)+\sum_{y\in S\cap I_{q}}\frac{\varepsilon}{3\ell}\leq\left(1+\frac{\varepsilon}{3}\right)(\rho+\varepsilon)\sum_{r=1}^{q-1}|I_{r}|\min_{z\in I_{r}}p_{t}(x,z)+\frac{\varepsilon|I_{q}|}{3\ell}\\ \leq\left(1+\frac{\varepsilon}{3}\right)(\rho+\varepsilon)\sum_{r=1}^{q-1}\sum_{y\in S\cap I_{r}}p_{t}(x,y)+\frac{\varepsilon}{3}\leq\left(1+\frac{\varepsilon}{3}\right)(\rho+\varepsilon)+\frac{\varepsilon}{3}.

Putting together (6.15), (6.19) and (6.27), and using that ρ+ε1\rho+\varepsilon\leq 1, we obtain

(6.28) 𝐏η0(ηt(x)=1)ε3+(1+ε3)(ρ+ε)+ε3ρ+2ε.\mathbf{P}^{\eta_{0}}(\eta_{t}(x)=1)\leq\frac{\varepsilon}{3}+\left(1+\frac{\varepsilon}{3}\right)(\rho+\varepsilon)+\frac{\varepsilon}{3}\leq\rho+2\varepsilon.

This yields the upper bound in (6.13), and the upper inequality in (C.1) follows.

For the lower bound, which requires to change (6.13) to ρ2ε𝐏η0(ηt(x)=1)\rho-2\varepsilon\leq\mathbf{P}^{\eta_{0}}(\eta_{t}(x)=1), one uses the estimate 𝐏η0(ηt(x)=1)r=1q1ySIrpt(x,y)\mathbf{P}^{\eta_{0}}(\eta_{t}(x)=1)\geq\sum_{r=1}^{q-1}\sum_{y\in S\cap I_{r}}p_{t}(x,y) instead of (6.15) and proceeds similarly as in (6.27). ∎

Lemma 6.5.

The condition (C.2.1) holds for SEP.

Proof.

Recall the construction in (6.2) of the SEP using 𝐏^\widehat{\mathbf{P}}. Setting =𝐏^\mathbb{Q}=\widehat{\mathbf{P}}, this yields a natural coupling of η\eta and η\eta^{\prime} with marginal laws 𝐏η0\mathbf{P}^{\eta_{0}} and 𝐏η0\mathbf{P}^{\eta^{\prime}_{0}}, respectively, for any choice of initial distribution η0\eta_{0} and η0\eta^{\prime}_{0}. In words, the coupling \mathbb{Q} identifies η\eta and η\eta^{\prime} as interchange processes and uses the same Poisson processes (𝒫e)(\mathcal{P}_{e}) on the edges of \mathbb{Z}. Now if η0|[H,H]η0|[H,H]\eta_{0}|_{[-H,H]}\succcurlyeq\eta_{0}^{\prime}|_{[-H,H]}, then under \mathbb{Q}, we claim that

(6.29) {x[H+2kνt,H2kνt],s[0,t],ηs(x)<ηs(x)}y:η0(y)=1,|y|>HE(y),\big{\{}\exists x\in[-H+2k\nu t,H-2k\nu t],s\in[0,t],\,\eta_{s}(x)<\eta^{\prime}_{s}(x)\big{\}}\subseteq\bigcup_{y:\,\eta^{\prime}_{0}(y)=1,|y|>H}{E}(y),

where E(y){E}(y) is the event that the particle at yy enters the interval [H+2kνt,H2kνt][-H+2k\nu t,H-2k\nu t] before time (s)t(s\leq)t. By (6.7) with a=|y|Ha=|y|-H, we have

(6.30) (E(y))exp((2kνt+|y|H)/8).\mathbb{Q}({E}(y))\leq\exp\left(-(2k\nu t+|y|-H)/8\right).

Applying a union bound to (6.29) and feeding (6.30) thus yields that the probability of the complement of the event appearing on the left-hand side of (3.6) is bounded from above by

2y=H+(E(y))2y=0+exp(y/8kνt/4)20exp(kνt/4).2\sum_{y=H}^{+\infty}\mathbb{Q}({E}(y))\leq 2\sum_{y^{\prime}=0}^{+\infty}\exp(-y^{\prime}/8-k\nu t/4)\leq 20\exp(-k\nu t/4).

Remark 6.6 (Locality in (C.2.1)).

For later reference, we record the following locality property of the coupling \mathbb{Q} constructed in the course of proving Lemma 6.5. Let EHEE_{H}\subset E denote the edges having both endpoints in [H,H][-H,H]. Then the above argument continues to work for any specification of clock processes 𝒫e\mathcal{P}_{e}, 𝒫e\mathcal{P}_{e}^{\prime} for eEHe\notin E_{H} for η\eta and η\eta^{\prime}, respectively, so long as (𝒫e)eE(\mathcal{P}_{e})_{e\in E} and (𝒫e)eE(\mathcal{P}_{e}^{\prime})_{e\in E} end up having the correct law. This observation will be important when several couplings are ‘concatenated,’ as in the proof of Lemma 6.9 below (see also Figure 7).

Lemma 6.7.

The condition (C.2.2) holds for SEP.

We first give a brief overview of the argument. Lemma 6.7 corresponds to a quenched version of [15, Lemma 3.2] by Baldasso and Teixeira, in which η0\eta_{0} and η0\eta^{\prime}_{0} are sampled under 𝐏ρ+ε\mathbf{P}^{\rho+\varepsilon} and 𝐏ρ\mathbf{P}^{\rho} respectively. We mostly follow their argument. Since fixing the initial environments induces some changes, and because we will use a variation of this proof to show (C.2) for the SEP (Lemma 6.9), we detail below the coupling of η\eta and η\eta^{\prime}, seen as interchange processes. In doing so we also clarify an essential aspect of this coupling; see in particular (6.33)-(6.34) below.

In a nutshell, the idea is to pair injectively each particle of η\eta^{\prime} with one of η\eta at a relatively small distance (of order t1/4t^{1/4}), during a relatively long time (of order t3/4t^{3/4}), so that they perform independent random walks until they meet (which has a large probability to happen since t3/4(t1/4)2t^{3/4}\gg(t^{1/4})^{2}), after which they coalesce, i.e. follow the same evolution. To make such a matching possible, we ensure that with large probability, η\eta has more particles than η\eta^{\prime} on each interval of length roughly t1/4t^{1/4}. Within the present quenched framework, this property is now obtained by means of (C.1), which has already been proved; see Lemma 6.4. Since the probability that at least one particle of η\eta^{\prime} does not get paired, i.e. does not meet its match, is relatively high (polynomially small in tt), we repeat this coupling; see also Figure 6.

Refer to caption
Figure 6: Coupling performed in Lemma 6.7. The \bullet are particles of η\eta, the \circ are particles of η\eta^{\prime}. The red links between pairs of particles represent matchings. On the (bad) event B2B_{2}, the particle with blue trajectory, in spite of having found a (nearby) match at all stages ii, remains unpaired at time tt.
Proof of Lemma 6.7.

Let ρJ\rho\in J, ε(0,1)\varepsilon\in(0,1), and H,t1H,t\geq 1 and η0,η0\eta_{0},\eta^{\prime}_{0} satisfy the conditions appearing in (C.2.2). Recall that

=t1/4 and let τ=3.\text{$\ell=\lfloor t^{1/4}\rfloor$ and let $\tau=\ell^{3}$}.

We abbreviate Is=[H+2νs,H2νs]I_{s}=[-H+2\nu s,H-2\nu s] for s0s\geq 0 in the sequel.

Let =𝐏^𝐏^\mathbb{Q}=\widehat{\mathbf{P}}\otimes\widehat{\mathbf{P}}^{\prime}, where 𝐏^\widehat{\mathbf{P}}^{\prime} is a copy of 𝐏^\widehat{\mathbf{P}}, with 𝐏^\widehat{\mathbf{P}}, 𝐏^\widehat{\mathbf{P}}^{\prime} governing the independent processes (𝒫e)(\mathcal{P}_{e}), (𝒫e)(\mathcal{P}_{e}^{\prime}), respectively, cf. above (6.2). We will define a coupling of (η,η)(\eta,\eta^{\prime}) under \mathbb{Q} (to be precise, a suitable extension of \mathbb{Q} carrying additional independent randomness), inductively in ii over the time interval (iτ,(i+1)τ](i\tau,(i+1)\tau], for 0i<0\leq i<\ell. Suppose (η[0,iτ],η[0,iτ])(\eta_{[0,i\tau]},\eta_{[0,i\tau]}^{\prime}) have been declared under \mathbb{Q} for some 0i<0\leq i<\ell, with the correct marginal laws for η[0,iτ]\eta_{[0,i\tau]} and η[0,iτ]\eta_{[0,i\tau]}^{\prime} (the case i=0i=0 of this induction assumption holds trivially). We start by controlling the empirical densities of η\eta and η\eta^{\prime} at time iτi\tau, which are already defined under \mathbb{Q}. The original proof of [15] uses the stationarity of 𝐏ρ\mathbf{P}^{\rho} and 𝐏ρ+ε\mathbf{P}^{\rho+\varepsilon} and (2.2), while we resort to (C.1) (cf. (6.39) below). Let

(6.31) G1,i=def.{for all I with IIiτ of length /2|I|,one has that ηiτ(I)(ρ+ε/2)|I|ηiτ(I) }.G_{1,i}\stackrel{{\scriptstyle\text{def.}}}{{=}}\left\{\begin{aligned} &\text{for all $I$ with $I\subset I_{i\tau}$ of length $\lfloor\ell/2\rfloor\leq|I|\leq\ell$,}\\ &\text{one has that $\eta_{i\tau}(I)\geq(\rho+\varepsilon/2)|I|\geq\eta^{\prime}_{i\tau}(I)$ }\end{aligned}\right\}.

Observe that G1,0G_{1,0} is automatically satisfied by assumption on η0\eta_{0} and η0\eta_{0}^{\prime}.

We proceed to define (ηiτ+t)0<tτ(\eta_{i\tau+t})_{0<t\leq\tau} and (ηiτ+t)0<tτ(\eta^{\prime}_{i\tau+t})_{0<t\leq\tau} under \mathbb{Q}. If G1,iG_{1,i} does not occur, we sample (ηt+iτ)0<tτ(\eta^{\prime}_{t+i\tau})_{0<t\leq\tau} (and in fact for all t>0t>0) in a manner as in (6.2) using the processes (𝒫e)(\mathcal{P}_{e}^{\prime}) after time τi\tau_{i}, and similarly for (ηt+iτ)0<tτ(\eta_{t+i\tau})_{0<t\leq\tau} using (𝒫e)(\mathcal{P}_{e}) instead. Thus in this case η\eta^{\prime} evolves independently from η\eta from time τi\tau_{i} on.

If on the other hand G1,iG_{1,i} occurs, we proceed as follows. Let Pais[H,H]\text{Pai}_{s}\subset[-H,H] denote the set

(6.32) Pais={xIs:ηs(x)=ηs(x)=1},\text{Pai}_{s}=\{x\in I_{s}:\,\eta_{s}(x)=\eta_{s}^{\prime}(x)=1\},

so that Paiiτ\text{Pai}_{i\tau} is measurable relative to (ηiτ,ηiτ)(\eta_{i\tau},\eta_{i\tau}^{\prime}). We refer to Pais\text{Pai}_{s} as the set of paired particles (at time ss). Let Πs={xIs:ηs(x)=1}\Pi_{s}=\{x\in I_{s}:\eta_{s}(x)=1\} and Πs={xIs:ηs(x)=1}\Pi_{s}^{\prime}=\{x\in I_{s}:\eta_{s}^{\prime}(x)=1\}. Observe that PaisΠs\text{Pai}_{s}\subset\Pi_{s}^{\prime}. Our goal is to reduce the size of their difference as s=iτs=i\tau for i=1,2,i=1,2,\dots and eventually achieve equality when i=i=\ell.

To this effect, we first define a matching, i.e. an injective map ψi:ΠiτΠiτ\psi_{i}:\Pi_{i\tau}^{\prime}\to\Pi_{i\tau}, still measurable relative to (ηiτ,ηiτ)(\eta_{i\tau},\eta_{i\tau}^{\prime}), as follows. The map ψi\psi_{i} acts as identity map on Paiiτ\text{Pai}_{i\tau}, a subset of both Πiτ\Pi_{i\tau}^{\prime} and Πiτ\Pi_{i\tau}, cf. (6.32). For each xΠiτPaiiτx\in\Pi_{i\tau}^{\prime}\setminus\text{Pai}_{i\tau}, ψi(x)\psi_{i}(x) is a point in ΠiτPaiiτ\Pi_{i\tau}\setminus\text{Pai}_{i\tau} at distance at most \ell from xx. As we now briefly explain, owing to the occurrence of G1,iG_{1,i}, this can be achieved in such a way that ψi\psi_{i} is injective. To see this, first note that one can write IiτI_{i\tau} as disjoint union of intervals of length |I||I| ranging in /2|I|\lfloor\ell/2\rfloor\leq|I|\leq\ell, as follows. One covers IiτI_{i\tau} with contiguous intervals of length \ell starting at one boundary, leaving a remaining interval IrI_{r} at the other boundary of length less than \ell. If IrI_{r} has length at least /2\lfloor\ell/2\rfloor, one simply adds it, else unless IrI_{r} is empty one cuts the penultimate interval into two halves of length at least /2\lfloor\ell/2\rfloor each and merges IrI_{r} with the last of them. By construction any of the disjoint intervals II thereby obtained has length /2|I|\lfloor\ell/2\rfloor\leq|I|\leq\ell as required, and thus on the event G1,iG_{1,i}, see (6.31), one knows that ηiτ(I)ηiτ(I)\eta_{i\tau}(I)\geq\eta^{\prime}_{i\tau}(I). Since the II’s are disjoint and their union is IiτI_{i\tau}, it follows that we can pair injectively each particle of ΠiτPaiiτ\Pi_{i\tau}^{\prime}\setminus\text{Pai}_{i\tau} with a particle of ΠiτPaiiτ\Pi_{i\tau}\setminus\text{Pai}_{i\tau} within the same interval. We now fix any such matching ψi\psi_{i} and call any two particles (x,ψi(x))Πiτ×Πiτ(x,\psi_{i}(x))\in\Pi_{i\tau}^{\prime}\times\Pi_{i\tau} matched. We note that |xψi(x)||x-\psi_{i}(x)|\leq\ell by construction.

The evolution for (ηiτ+s)0<sτ(\eta_{i\tau+s})_{0<s\leq\tau} and (ηiτ+s)0<sτ(\eta^{\prime}_{i\tau+s})_{0<s\leq\tau} under \mathbb{Q} (and on G1,iG_{1,i}) is now prescribed as follows. Both (ηiτ+s)0<sτ(\eta_{i\tau+s})_{0<s\leq\tau} and (ηiτ+s)0<sτ(\eta^{\prime}_{i\tau+s})_{0<s\leq\tau} will be realized as interchange processes as in (6.2), thus it is sufficient to specify the relevant (Poisson) clock processes attached to each edge of \mathbb{Z}. Let EHE_{H} denote the set of edges of \mathbb{Z} having both endpoints in [H,H][-H,H]. For eEHe\notin E_{H}, (ηiτ+s)0<sτ(\eta_{i\tau+s})_{0<s\leq\tau} and (ηiτ+s)0<sτ(\eta^{\prime}_{i\tau+s})_{0<s\leq\tau} simply use the clocks of 𝒫eθiτ\mathcal{P}_{e}\circ\theta_{i\tau} and 𝒫eθiτ\mathcal{P}_{e}^{\prime}\circ\theta_{i\tau}, respectively, where θs\theta_{s} denotes the canonical time-shift of the process by ss. It remains to specify the clock processes for eEHe\in E_{H}. Let 𝒫^e=𝒫e+𝒫e\widehat{\mathcal{P}}_{e}=\mathcal{P}_{e}+\mathcal{P}_{e}^{\prime}. All clock processes attached to edges eEHe\in E_{H} will be defined via suitable thinning of 𝒫^e\widehat{\mathcal{P}}_{e}.

First, one orders chronologically all arrivals for the processes 𝒫^eθiτ=(𝒫^e(s+iτ))s0\widehat{\mathcal{P}}_{e}\circ\theta_{i\tau}=(\widehat{\mathcal{P}}_{e}(s+i\tau))_{s\geq 0} as eEHe\in E_{H} varies (there are countably many such times and they are a.s. different so this is well-defined on a set of full measure). Let σ0=0\sigma_{0}=0 and σ1,σ2\sigma_{1},\sigma_{2} etc. denote the chronologically ordered times thereby obtained. By suitable extension, \mathbb{Q} is assumed to carry a family {Xn:n0}\{X_{n}:\,n\geq 0\} of i.i.d. Bernoulli variables with (Xn=1)=1(Xn=0)=12\mathbb{Q}(X_{n}=1)=1-\mathbb{Q}(X_{n}=0)=\frac{1}{2}. We regard XnX_{n} as the label attached to σn\sigma_{n}. Let e\mathcal{R}_{e} be the thinned process obtained from 𝒫^e\widehat{\mathcal{P}}_{e} by only retaining arrivals with label Xn=1X_{n}=1. Then,

(6.33) η+iτ\eta_{\cdot+i\tau} uses the clock process e\mathcal{R}_{e} for each eEHe\in E_{H} (and 𝒫e\mathcal{P}_{e} for each eEHe\notin E_{H}).

By elementary properties of Poisson processes and applying (6.2), it follows that (ηs+iτ)0<sτ(\eta_{s+i\tau})_{0<s\leq\tau} has the correct conditional law given ηiτ\eta_{i\tau}; indeed by construction to each edge ee of \mathbb{Z} one has associated independent Poisson processes having the correct intensity.

The definition of (ηiτ+s)0<sτ(\eta^{\prime}_{i\tau+s})_{0<s\leq\tau} is analogous to (6.33), and the clock process e\mathcal{R}_{e}^{\prime} for eEHe\in E_{H} underlying the definition of (ηiτ+s)0<sτ(\eta^{\prime}_{i\tau+s})_{0<s\leq\tau} is specified as follows. For a particle xΠiτx\in\Pi_{i\tau}^{\prime} (resp. Πiτ\Pi_{i\tau}), let γi;(x)\gamma_{i;\cdot}^{\prime}(x) (resp. γi;(x)\gamma_{i;\cdot}(x)) denote its evolution under ηiτ+\eta^{\prime}_{i\tau+\cdot} (resp. ηiτ+\eta_{i\tau+\cdot}). Proceeding chronologically starting at n=1n=1, one chooses whether the clock σn\sigma_{n} is retained or not according to the following rule. With e={x,y}EHe=\{x,y\}\in E_{H} denoting the edge of σn\sigma_{n},

(6.34) if for some zΠiτγi;σn1(z)=γi;σn1(ψi(z)){x,y},then σn is retained iff Xn=1, otherwise iff Xn=0;\begin{split}&\text{if for some $z\in\Pi_{i\tau}^{\prime}$, $\gamma^{\prime}_{i;\sigma_{n-1}}(z)=\gamma_{i;\sigma_{n-1}}(\psi_{i}(z))\in\{x,y\}$,}\\ &\text{then $\sigma_{n}$ is retained iff $X_{n}=1$, otherwise iff $X_{n}=0$;}\end{split}

here we think of right-continuous trajectories so γi;σn1(z)\gamma^{\prime}_{i;\sigma_{n-1}}(z) is the position of the particle zz after the (n1)(n-1)-th jump; in fact one could replace each occurrence of σn1\sigma_{n-1} in (6.34) by an arbitrary time ss with σn1s<σn\sigma_{n-1}\leq s<\sigma_{n}, since there is no jump between those times. In words, at time σn1\sigma_{n-1}, one inspects if at least one endpoint of ee contains (the evolution to time ss of) two matched particles, in which case the clock σn\sigma_{n} is retained if it has label 11 only. If no endpoint of ee contains matched particles the clock is retained if it has label 0. The process (ηiτ+s)0<sτ(\eta^{\prime}_{i\tau+s})_{0<s\leq\tau} then simply uses the clocks e\mathcal{R}_{e}^{\prime} on edges eEHe\in E_{H} that are retained according to (6.34) and such that σn<τ\sigma_{n}<\tau. We will now argue that

(6.35) given ηiτ\eta^{\prime}_{i\tau}, the process (ηiτ+s)0<sτ(\eta^{\prime}_{i\tau+s})_{0<s\leq\tau} has law 𝐏ηiτ\mathbf{P}^{\eta^{\prime}_{i\tau}} under \mathbb{Q}.

To see this, one simply notes using a straightforward induction argument that the conditional law of ξn=1{the n-th arrival in (𝒫^)eEH is retained}\xi_{n}=1\{\text{the $n$-th arrival in $(\widehat{\mathcal{P}})_{e\in E_{H}}$ is retained}\} given ξ1,ξn1\xi_{1},\dots\xi_{n-1}, σ1,,σn\sigma_{1},\dots,\sigma_{n}, X1,,Xn1X_{1},\dots,X_{n-1} and (ηiτ+t,ηiτ+s)0<sσn1(\eta_{i\tau+t},\eta^{\prime}_{i\tau+s})_{0<s\leq\sigma_{n-1}} is that of a Bernoulli-12\frac{1}{2} random variable. From this and the thinning property for Poisson processes it readily follows that \mathcal{R}^{\prime} has the right law.

Overall we have now defined a coupling of (η,η)(\eta,\eta^{\prime}) until time τt\ell\tau\leq t. In case τ<t\ell\tau<t we simply use the same process (𝒫e)(\mathcal{P}_{e}) to define the evolution of both η\eta and η\eta^{\prime} in the remaining time interval (τ,t](\ell\tau,t]. The Markov property (P.1), (6.33) and (6.35) ensure that η,η\eta,\eta^{\prime} indeed have the desired marginals during [0,t][0,t]. In view of (6.32), this immediately yields that

(6.36) {ηt|[H+4νt,H4νt]ηt|[H+4νt,H4νt]}c{(ΠtPait)[H+4νt,H4νt]}.\big{\{}\eta_{t}|_{[-H+4{\nu}t,H-4{\nu}t]}\succcurlyeq\eta_{t}^{\prime}|_{[-H+4{\nu}t,H-4{\nu}t]}\big{\}}^{c}\subset\big{\{}(\Pi_{t}^{\prime}\setminus\text{Pai}_{t})\cap[-H+4{\nu}t,H-4{\nu}t]\neq\emptyset\big{\}}.

The key of the above construction is that the latter event forces one of three possible unlikely scenarios. Namely, as we explain below, one has that

(6.37) {(ΠtPait)[H+4νt,H4νt]}B1B2B3,\big{\{}(\Pi_{t}^{\prime}\setminus\text{Pai}_{t})\cap[-H+4{\nu}t,H-4{\nu}t]\neq\emptyset\big{\}}\subseteq B_{1}\cup B_{2}\cup B_{3},

where B1=i=11G1,icB_{1}=\bigcup_{i=1}^{\ell-1}G_{1,i}^{c},

(6.38) B2=s=0t1{one particle of ηs(Is) ends upin [H+4νt,H4νt] at time t },B3=B1c{xΠ0 s.t. γiτ(x)Iiτ and infs[0,τ]Zsi(x)>0 for all 0i<};\begin{split}B_{2}&=\bigcup_{s=0}^{t-1}\left\{\begin{array}[]{c}\text{one particle of $\eta^{\prime}_{s}(\mathbb{Z}\setminus I_{s})$ ends up}\\ \text{in $[-H+4\nu t,H-4\nu t]$ at time $t$ }\end{array}\right\},\\ B_{3}&=B_{1}^{c}\cap\left\{\begin{array}[]{c}\text{$\exists x\in\Pi_{0}^{\prime}$ s.t.~$\gamma_{i\tau}^{\prime}(x)\in I_{i\tau}$ and }\\ \text{$\inf_{s\in[0,\tau]}Z_{s}^{i}(x)>0$ for all $0\leq i<\ell$}\end{array}\right\};\end{split}

here γ(x)\gamma_{\cdot}^{\prime}(x) refers to the evolution of particle xx under η\eta^{\prime} and with xi=γiτ(x)x_{i}=\gamma_{i\tau}^{\prime}(x), one sets Zsi(x)=|γi;s(xi)γi;s(ψi(xi))|Z_{s}^{i}(x)=|\gamma_{i;s}^{\prime}(x_{i})-\gamma_{i;s}(\psi_{i}(x_{i}))|. In words ZsiZ_{s}^{i} follows the evolution of the difference between xix_{i}, which is not paired at time iτi\tau since Z0i(x)0Z_{0}^{i}(x)\neq 0, and its match ψi(xi)\psi_{i}(x_{i}), which is well-defined on the event B1cB_{1}^{c}. Thus B3B_{3} refers to the event that some particle xΠ0x\in\Pi_{0}^{\prime} is found in IiτI_{i\tau} at time iτi\tau for all ii and never meets its match during the time interval (iτ,(i+1)τ](i\tau,(i+1)\tau].

We now explain (6.37). To this effect we first observe that (6.33) and (6.34) ensure that two matched particles at some stage ii follow the same evolution once they meet (and thus belong to Pais\text{Pai}_{s} for all later times ss) as long as they stay in IsI_{s}. Therefore, on the event B1cB2cB_{1}^{c}\cap B_{2}^{c}, on which (due to occurrence of B2cB_{2}^{c}) no unpaired η\eta^{\prime}-particle at time tt can arise by drifting in from the side, meaning that such a particle cannot be seen in ηs(Is)\eta^{\prime}_{s}(\mathbb{Z}\setminus I_{s}) at any time 0s<t0\leq s<t, the set (ΠtPait)[H+4νt,H4νt](\Pi_{t}^{\prime}\setminus\text{Pai}_{t})\cap[-H+4{\nu}t,H-4{\nu}t] being non-empty requires at least one particle from Π0\Pi_{0}^{\prime} to never meet its match at any of the stages 1i1\leq i\leq\ell (matching happens at all stages due to occurrence of B1cB_{1}^{c}). That is, B3B_{3} occurs, and (6.37) follows.

To finish the proof, we now bound the (bad) events appearing in (6.37) separately. In view of (6.31), we apply (C.1) (which now holds on account of Lemma 6.4) to η\eta^{\prime} (resp. η\eta) at time iτi\tau, with (ρ+ε/8,ε/8)(\rho+\varepsilon/8,\varepsilon/8) (resp. (ρ+7ε/8,ε/8)(\rho+7\varepsilon/8,\varepsilon/8)) instead of (ρ,ε)(\rho,\varepsilon), with (H,,iτ)(H,\ell,i\tau) instead of (H,,t)(H,\ell,t) and with \ell^{\prime} ranging from /2\lfloor\ell/2\rfloor to \ell, which fulfils the conditions of (C.1) if \CrSEPcoupling\Cr{SEPcoupling} is large enough so that H>4νtH>4\nu t and

min1i414νiτ4ντ>\Crdensitystable2ε2(1+|log3(νt)|)max1i41\Crdensitystable2ε2(1+|log3(νiτ)|).\min_{1\leq i\leq\ell^{4}-1}4\nu i\tau\geq 4\nu\tau>\Cr{densitystable}\ell^{2}\varepsilon^{-2}(1+|\log^{3}(\nu t)|)\geq\max_{1\leq i\leq\ell^{4}-1}\Cr{densitystable}\ell^{2}\varepsilon^{-2}(1+|\log^{3}(\nu i\tau)|).

Recalling that =t1/4\ell=\lfloor t^{1/4}\rfloor and summing over the possible values of \ell^{\prime}, this gives that

(6.39) (B1)i=11(G1,ic)4t1/2Hexp(\Crdensitystableexpoε227(1)).\mathbb{Q}\big{(}B_{1}\big{)}\leq\sum_{i=1}^{\ell-1}\mathbb{Q}\left(G_{1,i}^{c}\right)\leq 4t^{1/2}H\exp\left(-\Cr{densitystableexpo}\varepsilon^{2}2^{-7}(\ell-1)\right).

We deal with (B2)\mathbb{Q}(B_{2}) by applying (C.2.1), which is in force on account of Lemma 6.5. Noticing that the event indexed by ss entering the definition of B2B_{2} in (6.38) implies that at least one particle of ηs(Is)\eta^{\prime}_{s}(\mathbb{Z}\setminus I_{s}) ends up in [H+2νs+2νt,H2νs2νt][-H+2\nu s+2\nu t,H-2\nu s-2\nu t] before time s+ts+t, we get using (3.6) with k=1k=1, η0=ηs0\eta_{0}=\eta_{s}\equiv 0, and IsI_{s} playing the role of [H,H][-H,H] that

(6.43) (B2)0s<t(a particle of ηs(Is) ends up in[H+2ν(s+t),H2ν(s+t)]before time s+t)20texp(νt/4).\mathbb{Q}(B_{2})\leq\sum_{0\leq s<t}\mathbb{Q}\left(\begin{array}[]{c}\text{a particle of $\eta^{\prime}_{s}(\mathbb{Z}\setminus I_{s})$ ends up in}\\ \text{$[-H+2\nu(s+t),H-2\nu(s+t)]$}\\ \text{before time $s+t$}\end{array}\right)\leq 20t\exp(-\nu t/4).

Finally, owing to our coupling in (6.33), (6.34), conditionally on (ηu,ηu)(\eta_{u},\eta^{\prime}_{u}), uiτu\leq i\tau, the process Zsi(x)Z_{s}^{i}(x) is a one-dimensional continuous-time random walk with rate 2ν2\nu started at a point in [0,][0,\ell] (owing to the separation of xix_{i} and its match ψi(xi)\psi_{i}(x_{i})) with absorption at 0. If (Zs)s0(Z_{s})_{s\geq 0} is such a random walk and k\mathbb{P}_{k} denotes the probability for this walk starting at k[0,]k\in[0,\ell], we have by invariance by translation and the reflection principle:

(6.44) max0kk(min0sτZs>0)=(min0sτZs>0)=0(min0sτZs>)20(max0sτ|Zs|<)2(Zτ[,]).\begin{split}\max_{0\leq k\leq\ell}\mathbb{P}_{k}(\min_{0\leq s\leq\tau}Z_{s}>0)&=\mathbb{P}_{\ell}(\min_{0\leq s\leq\tau}Z_{s}>0)=\mathbb{P}_{0}(\min_{0\leq s\leq\tau}Z_{s}>-\ell)\\ &\leq 2\mathbb{P}_{0}(\max_{0\leq s\leq\tau}|Z_{s}|<\ell)\leq 2\mathbb{P}(Z_{\tau}\in[-\ell,\ell]).\end{split}

Using again [50, Theorem 2.5.6] and taking \CrSEPcoupling\Cr{SEPcoupling} large enough (so that in particular 2ντ=t2|x|2\nu\tau=t\geq 2|x| for any x[,]x\in[-\ell,\ell]), we have thus for some universal constant C>0C>0 (changing from one expression to the next):

(6.45) max0kk(min0sτZs>0)2+14πντexp(C(ν1τ1+3ν2τ2))CντCν1/2t1/8.\max_{0\leq k\leq\ell}\mathbb{P}_{k}(\min_{0\leq s\leq\tau}Z_{s}>0)\leq\frac{2\ell+1}{\sqrt{4\pi\nu\tau}}\exp(C(\nu^{-1}\tau^{-1}+\ell^{3}\nu^{-2}\tau^{-2}))\leq\frac{C\ell}{\sqrt{\nu\tau}}\leq C\nu^{-1/2}t^{-1/8}.

Recalling that t>ν8t>\nu^{-8} and taking a union bound over xΠ0x\in\Pi_{0}^{\prime}, the previous estimate applied with the Markov property for (η,η)(\eta,\eta^{\prime}) yields that, as long as \CrSEPcoupling\Cr{SEPcoupling} is large enough:

(6.46) (B3)2H(Ct1/16).\mathbb{Q}(B_{3})\leq 2H(Ct^{-1/16})^{\ell}.

Putting together (6.36), (6.37), (6.39), (6.43) and (6.46), we obtain that

(ηt|[H+4νt,H4νt]ηt|[H+4νt,H4νt])18t1/2Hexp(\Crdensitystableexpoε227(1))20texp(νt/4)2H(Ct1/16),\mathbb{Q}(\eta^{\prime}_{t}|_{[-H+4{\nu}t,H-4{\nu}t]}\preccurlyeq\eta_{t}|_{[-H+4{\nu}t,H-4{\nu}t]})\\ \geq 1-8t^{1/2}H\exp\left(-\Cr{densitystableexpo}\varepsilon^{2}2^{-7}(\ell-1)\right)-20t\exp(-\nu t/4)-2H(Ct^{-1/16})^{\ell},

which is larger than 1\CrSEPcoupling2tHexp(\CrSEPcoupling21νν+1ε2t1/4)1-\Cr{SEPcoupling2}tH\exp(-\Cr{SEPcoupling2}^{-1}\frac{\nu}{\nu+1}\varepsilon^{2}t^{1/4}) as required by (3.7) provided \CrSEPcoupling\Cr{SEPcoupling} and \CrSEPcoupling2\Cr{SEPcoupling2} are chosen large enough. ∎

Remark 6.8 (Locality in (C.2.2)).

Similarly as in Remark 6.6, which exhibits an analogous property for the coupling inherent to (C.2.1), the coupling \mathbb{Q} yielding property (C.2.2) constructed in the proof of Lemma 6.7 can be performed for any specification of clock processes (𝒫e,𝒫e)eEH(\mathcal{P}_{e},\mathcal{P}_{e}^{\prime})_{e\notin E_{H}} used to define η,η\eta,\eta^{\prime}, so long as the marginal laws of (𝒫e)eE(\mathcal{P}_{e})_{e\in E} and (𝒫e)eE(\mathcal{P}_{e}^{\prime})_{e\in E}, are that of independent Poisson processes of intensity ν/2\nu/2. This can be seen by inspection of the proof: the only ‘non-trivial’ joint distribution concerns (𝒫e,𝒫e)eEH(\mathcal{P}_{e},\mathcal{P}_{e}^{\prime})_{e\in E_{H}}, which are obtained by suitable thinning from (𝒫^e)eEH(\widehat{\mathcal{P}}_{e})_{e\in E_{H}}, see in particular the discussion around (6.33) and (6.34).

Combining the couplings supplied by Lemmas 6.5 and 6.7 multiple times, which will be permitted owing to Remarks 6.6 and 6.8, yields the following result.

Lemma 6.9.

The condition (C.2) holds for SEP.

Refer to caption
Figure 7: Couplings used in the proof of Lemma 6.9. If the couplings of Step 1 (operating with time horizon t1t_{1}) are successful, then the only unpaired particles of η\eta^{\prime} at time t1t_{1} are in the two red dashed segments, and there are enough spare particles of η\eta in the blue segments to cover these particles of η\eta^{\prime} before time t2t_{2} (Step 2, using coupling ‘Variation of (C.2.2)’).
Proof.

Let ρ,ε(0,1)\rho,\varepsilon\in(0,1), H1,H2,t,1H_{1},H_{2},t,\ell\geq 1 and η0,η0{0,1}\eta_{0},\eta^{\prime}_{0}\in\{0,1\}^{\mathbb{Z}} be such that the assumptions of (C.2) hold. In particular, note that these entail that H2>H1H_{2}>H_{1}. Define

(6.49) t1=4,t2=t1+19.t_{1}=\ell^{4},\quad t_{2}=t_{1}+\ell^{19}.

We proceed in two steps and refer to Figure 7 for visual aid. In the first step, we apply simultaneously the couplings of Lemma 6.5 on [H1,H1][-H_{1},H_{1}] and of Lemma 6.7 on [H2,H2][H1,H1][-H_{2},H_{2}]\setminus[-H_{1},H_{1}], in the time-interval [0,t1][0,t_{1}] (in doing so we shall explain how this preserves the marginals of η\eta and η\eta^{\prime}). As a result, we get that ηt1(x)ηt1(x)\eta_{t_{1}}(x)\geq\eta^{\prime}_{t_{1}}(x) for all x[H2+2νt1,H22νt1]x\in[-H_{2}+2\nu t_{1},H_{2}-2\nu t_{1}], except possibly around two intervals around H1-H_{1} and H1H_{1}, of width O(νt1)O(\nu t_{1}).

In the second step, we couple the particles of η\eta^{\prime} on these intervals with "additional" particles of ηt1ηt1\eta_{t_{1}}\setminus\eta^{\prime}_{t_{1}} on [H2,H2][-H_{2},H_{2}] (using that the empirical density of η\eta^{\prime} is slightly larger than that of η\eta on [H2,H2][-H_{2},H_{2}] by (C.1)), in a manner similar to that used in the proof of Lemma 6.7. This ensures that with large enough probability, all these particles of η\eta^{\prime} get covered by particles of η\eta within time t2t1t_{2}-t_{1}, without affecting the coupling of the previous step on account of the Markov property. Finally, during the time interval [t2,t][t_{2},t] we use again the natural coupling of Lemma 6.5, and conclude by showing (3.4) and (3.5). We now proceed to make this precise. Step 1: we construct a coupling 1\mathbb{Q}_{1} of the evolutions of η\eta and η\eta^{\prime} during the time-interval [0,t1][0,t_{1}] such that if we define the (good) events

G1={x[H1+2νt1,H12νt1],s[0,t1]:ηs(x)ηs(x)},G2={x[H2+2νt1,H12νt11][H1+2νt1+1,H22νt1]:ηt1(x)ηt1(x)},\begin{split}&G_{1}=\{\forall x\in[-H_{1}+2\nu t_{1},H_{1}-2\nu t_{1}],\,\forall s\in[0,t_{1}]:\eta_{s}(x)\geq\eta^{\prime}_{s}(x)\},\\ &G_{2}=\{\forall x\in[-H_{2}+2\nu t_{1},-H_{1}-2\nu t_{1}-1]\cup[H_{1}+2\nu t_{1}+1,H_{2}-2\nu t_{1}]:\eta_{t_{1}}(x)\geq\eta^{\prime}_{t_{1}}(x)\},\end{split}

then

(6.50) 1(G1G2)13\CrSEPcoupling2t1H2exp(\CrSEPcoupling21νν+1ε2).\mathbb{Q}_{1}(G_{1}\cap G_{2})\geq 1-3\Cr{SEPcoupling2}t_{1}H_{2}\exp\left(-\Cr{SEPcoupling2}^{-1}\textstyle\frac{\nu}{\nu+1}\varepsilon^{2}\ell\right).

The coupling 1\mathbb{Q}_{1} is defined as follows. Let (𝒫e)eE(\mathcal{P}_{e})_{e\in E} be a family of i.i.d. Poisson processes on +\mathbb{R}_{+} of intensity ν/2\nu/2. These processes are used to describe the exchange times for η\eta (seen as an interchange process as in (6.2)), during the time-interval [0,t1][0,t_{1}]. We now define the exchange times (𝒫e)eE(\mathcal{P}_{e}^{\prime})_{e\in E} to be used for η\eta^{\prime} during [0,t1][0,t_{1}] as follows. For an edge ee having at least one endpoint outside [H2,H2][-H_{2},H_{2}] or at least one endpoint inside [H1,H1][-H_{1},H_{1}], set 𝒫e=𝒫e\mathcal{P}_{e}^{\prime}=\mathcal{P}_{e}. It remains to specify 𝒫e\mathcal{P}_{e}^{\prime} for ee with both endpoints in [H2,H2][-H_{2},H_{2}] but outside [H1,H1][-H_{1},H_{1}]. This set splits into two disjoint intervals E±E_{\pm}, which are both dealt with separately and in exactly the same manner. Thus restricting our attention to E+E_{+}, one couples (𝒫e)eE+(\mathcal{P}_{e}^{\prime})_{e\in E_{+}} and (𝒫e)eE+(\mathcal{P}_{e})_{e\in E_{+}} in exactly the same manner as in the proof of Lemma 6.7, with the interval I+I_{+}, defined as the set of all endpoints of edges in E+E_{+}, playing the role of [H,H][-H,H]. The fact that Lemma 6.7 applies even though the processes 𝒫e\mathcal{P}_{e} and 𝒫e\mathcal{P}_{e}^{\prime} have been specified for certain edges eE+e\notin E_{+} is owed to Remark 6.8. Define II_{-} similarly and perform the same coupling of (𝒫e)eE(\mathcal{P}_{e}^{\prime})_{e\in E_{-}} and (𝒫e)eE(\mathcal{P}_{e})_{e\in E_{-}}.

Since the sets of edges EE_{-}, E+E_{+} and E(EE+)E\setminus(E_{-}\cup E_{+}) are disjoint, it readily follows that (𝒫e)eE(\mathcal{P}_{e})_{e\in E} is an i.i.d. family of Poisson processes on +\mathbb{R}_{+} of intensity ν/2\nu/2. This ensures that under 1\mathbb{Q}_{1}, η𝐏η0\eta\sim\mathbf{P}^{\eta_{0}} and η𝐏η0\eta^{\prime}\sim\mathbf{P}^{\eta^{\prime}_{0}}.

Now, by our assumptions on ρ,ε,H,,t1,η0\rho,\varepsilon,H,\ell,t_{1},\eta_{0} and η0\eta^{\prime}_{0}, the above construction of 1\mathbb{Q}_{1} together with Remark 6.6 ensure that Lemma 6.5 applies on the interval [H1,H1][-H_{1},H_{1}] with t=t1t=t_{1} and k=1k=1 (recall to this effect that the relevant coupling for which (C.2.1) is shown to hold is simply =𝐏^\mathbb{Q}=\widehat{\mathbf{P}}, see above (6.29), and that this coupling is also local in the sense of Remark 6.6). This yields that

(6.51) 1(G1)120exp(νt1/4).\mathbb{Q}_{1}(G_{1})\geq 1-20\exp(-{\nu}t_{1}/4).

Second, our assumptions (taking \Crcompatible>\CrSEPcoupling\Cr{compatible}>\Cr{SEPcoupling}) and the above construction of \mathbb{Q} also allow us to apply Lemma 6.7 on EE_{-} and on E+E_{+} instead of [H,H][-H,H], with the same values of ρ,ε\rho,\varepsilon and \ell, and with t=t1t=t_{1} (in particular, ν8t1=(ν2)4>1\nu^{8}t_{1}=(\nu^{2}\ell)^{4}>1). We obtain that

(6.52) 1(G2)12\CrSEPcoupling2t1H2exp(\CrSEPcoupling21νν+1ε2).\mathbb{Q}_{1}(G_{2})\geq 1-2\Cr{SEPcoupling2}t_{1}H_{2}\exp\left(-\Cr{SEPcoupling2}^{-1}\textstyle\frac{\nu}{\nu+1}\varepsilon^{2}\ell\right).

Combining (6.51) and (6.52) yields (6.50), since 20exp(νt1/4)\CrSEPcoupling2t1H2exp(\CrSEPcoupling21νν+1ε2)20\exp(-{\nu}t_{1}/4)\leq\Cr{SEPcoupling2}t_{1}H_{2}\exp(-\Cr{SEPcoupling2}^{-1}\frac{\nu}{\nu+1}\varepsilon^{2}\ell) if \Crcompatible\Cr{compatible} is chosen large enough. This concludes Step 1.

Step 2: We now extend 1\mathbb{Q}_{1} to a coupling 2\mathbb{Q}_{2} up to time tt, using a slight variation of the coupling in the proof of Lemma 6.7 in the time-interval [t1,t2][t_{1},t_{2}]. For definiteness of 2\mathbb{Q}_{2}, on the complement of G1G2G_{1}\cap G_{2}, use (𝒫e)eE(\mathcal{P}_{e})_{e\in E} for the exchange times of both η\eta and η\eta^{\prime} during the time [t1,t][t_{1},t]. Focusing now on the case where G1G2G_{1}\cap G_{2} occurs, the aim of the step is to show that we can have

(6.53) 2(ηt2|[H+4νt2,H4νt2]ηt2|[H+4νt2,H4νt2]|G1G2)1H2exp(νν+1ε24).\mathbb{Q}_{2}(\eta^{\prime}_{t_{2}}|_{[-H+4\nu t_{2},H-4\nu t_{2}]}\preccurlyeq\eta_{t_{2}}|_{[-H+4\nu t_{2},H-4\nu t_{2}]}\,|\,G_{1}\cap G_{2})\geq 1-H_{2}\exp\left(-\textstyle\frac{\nu}{\nu+1}\varepsilon^{2}\ell^{4}\right).

Fix any realization of ηt1,ηt1\eta_{t_{1}},\eta^{\prime}_{t_{1}} such that G1G2G_{1}\cap G_{2} holds. Define Pai0\text{Pai}_{0} (cf. (6.32)) as the set of vertices in I0=def.[H2+2νt1,H22νt1]I_{0}\stackrel{{\scriptstyle\text{def.}}}{{=}}[-H_{2}+2\nu t_{1},H_{2}-2\nu t_{1}] containing both a particle of ηt1\eta_{t_{1}} and of ηt1\eta^{\prime}_{t_{1}}, and think of these particles as being paired. Let U0U_{0} (resp. U0U^{\prime}_{0}) be the set of unpaired particles of ηt1\eta_{t_{1}} (resp. ηt1\eta^{\prime}_{t_{1}}) in I0I_{0}. By definition of G1G_{1} and G2G_{2}, the particles of U0U^{\prime}_{0} must be in [H12νt1,H1+2νt1][H12νt1,H1+2νt1][-H_{1}-2\nu t_{1},-H_{1}+2\nu t_{1}]\cup[H_{1}-2\nu t_{1},H_{1}+2\nu t_{1}] when the event G1G2G_{1}\cap G_{2} occurs; cf. Figure 7. Hence, on G1G2G_{1}\cap G_{2} we have that

(6.54) |U0|8νt1.|U^{\prime}_{0}|\leq 8\nu t_{1}.

Denote B1,0cB_{1,0}^{c} the event that on every interval of length between 52\frac{\ell^{5}}{2} and 5\ell^{5} included in I0I_{0}, ηt1\eta_{t_{1}} has at least ε5/108νt1\varepsilon\ell^{5}/10\geq 8\nu t_{1} unmatched particles (recall that >80ν/ε\ell>80\nu/\varepsilon by assumption). On B1,0B_{1,0}, use (𝒫e)eE(\mathcal{P}_{e})_{e\in E} as exchange time process to define both η\eta and η\eta^{\prime} during [t1,t][t_{1},t] (cf. the discussion leading to (6.2)).

Henceforth, assume that B1,0cG1G2B_{1,0}^{c}\cap G_{1}\cap G_{2} occurs. Following the line of argument in the paragraph after (6.32), one matches injectively each particle of U0U_{0}^{\prime} with a particle of U0U_{0} at distance at most 5\ell^{5}. In the present context this is possible owing to (6.54) and occurrence of B1,0cB_{1,0}^{c}. Then one couples the evolutions of η\eta and η\eta^{\prime}, first during [t1,t1+15][t_{1},t_{1}+\ell^{15}] as done in the proof of Lemma 6.7 during the time interval [0,τ][0,\tau]. Then iteratively at times t1+i15t_{1}+i\ell^{15} for 1i411\leq i\leq\ell^{4}-1, one performs on Ii=[H2+2ν(t1+i15),H22ν(t1+i15)]I_{i}=[-H_{2}+2\nu(t_{1}+i\ell^{15}),H_{2}-2\nu(t_{1}+i\ell^{15})] the same coupling at times it3/4i\lfloor t^{3/4}\rfloor on the event B1,icB_{1,i}^{c} that for any interval IIiI\subseteq I_{i} of length ranging between 52\frac{\ell^{5}}{2} and 5\ell^{5}, one has ηt1+i15(I)ηt1+i15(I)\eta_{t_{1}+i\ell^{15}}(I)\geq\eta^{\prime}_{t_{1}+i\ell^{15}}(I). On B1,iB_{1,i}, use (𝒫e)eE(\mathcal{P}_{e})_{e\in E} for the exchange times of both η\eta and η\eta^{\prime} during the time interval [t1+i15,t][t_{1}+i\ell^{15},t].

Overall, this yields a coupling of (ηt2,ηt2)(\eta_{\cdot\wedge t_{2}},\eta_{\cdot\wedge t_{2}}^{\prime}) with the correct marginal law (as in the proof of Lemma 6.7). Finally one extends this coupling during [t2,t][t_{2},t] on the event G1G2i=041B1,icG_{1}\cap G_{2}\cap\bigcap_{i=0}^{\ell^{4}-1}B_{1,i}^{c} by using the same exchange times for η\eta and η\eta^{\prime}. It follows that η𝐏η0\eta\sim\mathbf{P}^{\eta_{0}} and η𝐏η0\eta^{\prime}\sim\mathbf{P}^{\eta^{\prime}_{0}} under 2\mathbb{Q}_{2}.

In much the same way as in (6.36), it follows from the above construction that the complement of the event on the left-hand side of (6.53) implies that at least one particle of η\eta^{\prime} in the interval [H2+2νt2,H22νt2][-H_{2}+2\nu t_{2},H_{2}-2\nu t_{2}] is unpaired at time t2t_{2}, which in turn (cf. (6.37)-(6.38)) implies the occurrence of

(i=041B1,i)B2B3\left(\bigcup_{i=0}^{\ell^{4}-1}B_{1,i}\right)\cup B_{2}\cup B_{3}

where

B2=s=0191{a particle of ηt1+s([H2+2ν(t1+s),H22ν(t1+s)]) ends up in ηt2([H2+4νt2,H24νt2])}\displaystyle B_{2}=\bigcup_{s=0}^{\ell^{19}-1}\left\{\begin{array}[]{c}\text{a particle of $\eta^{\prime}_{t_{1}+s}(\mathbb{Z}\setminus[-H_{2}+2\nu(t_{1}+s),H_{2}-2\nu(t_{1}+s)])$}\\ \text{ ends up in $\eta^{\prime}_{t_{2}}([-H_{2}+4\nu t_{2},H_{2}-4\nu t_{2}])$}\end{array}\right\}
B3={at time t2, one particle from ηt1([H2+2νt1,H22νt1]) has been in Ii for all 0i<4 and remains unpaired}.\displaystyle B_{3}=\left\{\begin{aligned} &\text{at time $t_{2}$, one particle from $\eta^{\prime}_{t_{1}}([-H_{2}+2\nu t_{1},H_{2}-2\nu t_{1}])$}\\ &\text{ has been in $I_{i}$ for all $0\leq i<\ell^{4}$ and remains unpaired}\end{aligned}\right\}.

We now mimic (6.39), (6.43) and (6.46) to handle i=0412(B1,i)\sum_{i=0}^{\ell^{4}-1}\mathbb{Q}_{2}(B_{1,i}), 2(B2)\mathbb{Q}_{2}(B_{2}) and 2(B3)\mathbb{Q}_{2}(B_{3}) respectively. In detail, for the first term, we apply (C.1) with (H,t)=(H2,i15)(H,t)=(H_{2},i\ell^{15}) for 1i411\leq i\leq\ell^{4}-1, and \ell^{\prime} ranging from 52\frac{\ell^{5}}{2} to 5\ell^{5}, noting that

(6.55) H2>4νi154ν15\Crdensitystable19/2ε2(1+|log3(ν19)|)\Crdensitystable(i15)1/2ε2(1+|log3(νi15)|).H_{2}>4\nu i\ell^{15}\geq 4\nu\ell^{15}\geq\Cr{densitystable}\ell^{19/2}\varepsilon^{-2}(1+|\log^{3}(\nu\ell^{19})|)\geq\Cr{densitystable}(i\ell^{15})^{1/2}\varepsilon^{-2}(1+|\log^{3}(\nu i\ell^{15})|).

For the third inequality, remark that ν>\Crdensitystableε2(1+|log3(ν4)|)\nu\ell>\Cr{densitystable}\varepsilon^{-2}(1+|\log^{3}(\nu\ell^{4})|) (taking \Crcompatible>\Crdensitystable\Cr{compatible}>\Cr{densitystable}), hence it is enough to show that 9/2(1+|log3(ν4)|)1+|log3(ν19)|\ell^{9/2}(1+|\log^{3}(\nu\ell^{4})|)\geq 1+|\log^{3}(\nu\ell^{19})|. But 1+|log3(ν19)|1+4|log3(ν4)|+60log1+|\log^{3}(\nu\ell^{19})|\leq 1+4|\log^{3}(\nu\ell^{4})|+60\log\ell (since (a+b)34a3+4b3(a+b)^{3}\leq 4a^{3}+4b^{3} for a,b0a,b\geq 0). Taking \Crcompatible\Cr{compatible} large enough so that ν100\nu\ell^{100} and thus \ell is large enough (recall that >80ν/ε>ν\ell>80\nu/\varepsilon>\nu), we have 1+4|log3(ν4)|+60log1+4|log3(ν4)|+9/2(1+|log3(ν4)|)1+4|\log^{3}(\nu\ell^{4})|+60\log\ell\leq 1+4|\log^{3}(\nu\ell^{4})|+\ell\leq\ell^{9/2}(1+|\log^{3}(\nu\ell^{4})|) as desired. For the second term (B2)\mathbb{Q}(B_{2}), we use (C.2.1), and for the third term (B3)\mathbb{Q}(B_{3}), we note that a continuous-time random walk with rate 2ν2\nu started in [0,5][0,\ell^{5}] will hit 0 before time 15\ell^{15} with probability at least 1C5/(ν15)1/21C21-C\ell^{5}/(\nu\ell^{15})^{1/2}\geq 1-C\ell^{-2} (note that ν\Crcompatible>1\nu\ell\geq\Cr{compatible}>1 if we take \Crcompatible>1\Cr{compatible}>1). We obtain that

2(x[H2+4νt2,H24νt2],ηt2(x)ηt2(x))\displaystyle\mathbb{Q}_{2}(\forall x\in[-H_{2}+4\nu t_{2},H_{2}-4\nu t_{2}],\,\eta_{t_{2}}(x)\geq\eta^{\prime}_{t_{2}}(x))
1i=0412(B1,i)2(B2)2(B3)\displaystyle\qquad\qquad\geq 1-\sum_{i=0}^{\ell^{4}-1}\mathbb{Q}_{2}(B_{1,i})-\mathbb{Q}_{2}(B_{2})-\mathbb{Q}_{2}(B_{3})
149H2exp(\Crdensitystableexpoε227(51))2019exp(ν19/4)2H2(C2)4\displaystyle\qquad\qquad\geq 1-4\ell^{9}H_{2}\exp\left(-\Cr{densitystableexpo}\varepsilon^{2}2^{-7}(\ell^{5}-1)\right)-20\ell^{19}\exp(-\nu\ell^{19}/4)-2H_{2}(C\ell^{-2})^{\ell^{4}}
1H2exp(νν+1ε24),\displaystyle\qquad\qquad\geq 1-H_{2}\exp\left(\textstyle-\frac{\nu}{\nu+1}\varepsilon^{2}\ell^{4}\right),

if \Crcompatible\Cr{compatible} in (C.2) is chosen large enough. This yields (6.53).

Let =def.2\mathbb{Q}\stackrel{{\scriptstyle\text{def.}}}{{=}}\mathbb{Q}_{2}. It remains to establish (3.4) and (3.5). Owing to the way the coupling \mathbb{Q} is defined during the intervals [0,t1],[t1,t2][0,t_{1}],[t_{1},t_{2}] and [t2,t][t_{2},t], \mathbb{Q} has the following property: in [H1,H1][-H_{1},H_{1}], any particle of η\eta^{\prime} that is covered at some time s<ts<t by a particle of η\eta will be covered by this particle until time tt, or until it leaves [H1,H1][-H_{1},H_{1}]. Moreover, by assumption every particle of η0([H1,H1])\eta^{\prime}_{0}([-H_{1},H_{1}]) is covered by a particle of η0\eta_{0}. Therefore,

(6.56) {s[0,t],ηs|[H1+4νt,H14νt]ηs|[H1+4νt,H14νt]}cs=1t{a particle of ηs([H1,H1]) enters [H1+4νt,H14νt] before or at time t}.\{\forall s\in[0,t],\,\eta_{s}|_{[-H_{1}+4{\nu}t,H_{1}-4{\nu}t]}\succcurlyeq\eta^{\prime}_{s}|_{[-H_{1}+4{\nu}t,H_{1}-4{\nu}t]}\}^{c}\\ \subseteq\bigcup_{s=1}^{t}\{\text{a particle of $\eta^{\prime}_{s}(\mathbb{Z}\setminus[-H_{1},H_{1}])$ enters $[-H_{1}+4{\nu}t,H_{1}-4{\nu}t]$ before or at time $t$}\}.

By Lemma 6.2 applied with (t,k,a)=(ts,2,4νt+x)(t,k,a)=(t-s,2,4\nu t+x) for every sts\leq t and x0x\geq 0 if ηs(±(H1+x))=1\eta^{\prime}_{s}(\pm(H_{1}+x))=1, a union bound over all particles appearing in the event on the right-hand side of (6.56) leads to the bound on the \mathbb{Q}-probability of the left-hand side by 2tx0exp((4νt+x)/8)20texp(νt/4),2t\sum_{x\geq 0}\exp(-(4\nu t+x)/8)\leq 20t\exp(-\nu t/4), and (3.4) follows.

As for (3.5), first note that by (6.50) and (6.53), and for \Crcompatible\Cr{compatible} large enough, we have that

(6.57) (ηt2|[H2+4νt2,H24νt2]ηt2|[H2+4νt2,H24νt2])14\CrSEPcoupling2t1H2exp(\CrSEPcoupling21νν+1ε2).\mathbb{Q}(\eta^{\prime}_{t_{2}}|_{[-H_{2}+4\nu t_{2},H_{2}-4\nu t_{2}]}\preccurlyeq\eta_{t_{2}}|_{[-H_{2}+4\nu t_{2},H_{2}-4\nu t_{2}]})\geq 1-4\Cr{SEPcoupling2}t_{1}H_{2}\exp\left(-\textstyle\Cr{SEPcoupling2}^{-1}\frac{\nu}{\nu+1}\varepsilon^{2}\ell\right).

Since we use, during [t2,t][t_{2},t], the same natural coupling as in the proof of Lemma 6.5, letting B4B_{4} denote the event that a particle of ηt2([H2+4νt2,H24νt2])\eta^{\prime}_{t_{2}}(\mathbb{Z}\setminus[-H_{2}+4\nu t_{2},H_{2}-4\nu t_{2}]) ends up in [H2+6νt,H26νt][H2+4νt2+2νt,H24νt22νt][-H_{2}+6\nu t,H_{2}-6\nu t]\subseteq[-H_{2}+4\nu t_{2}+2\nu t,H_{2}-4\nu t_{2}-2\nu t] in time at most tt, we obtain, provided \Crcompatible\Cr{compatible} is large enough and abbreviating ξ=\CrSEPcoupling21νν+1ε2\xi=\Cr{SEPcoupling2}^{-1}\frac{\nu}{\nu+1}\varepsilon^{2}\ell, that

(ηt|[H2+6νt,H26νt]ηt|[H2+6νt,H26νt])14\CrSEPcoupling2t1H2eξ(B4)14\CrSEPcoupling2t1H2eξ20eνt/415\CrSEPcoupling2t1H2eξ.\mathbb{Q}(\eta^{\prime}_{t}|_{[-H_{2}+6\nu t,H_{2}-6\nu t]}\preccurlyeq\eta_{t}|_{[-H_{2}+6\nu t,H_{2}-6\nu t]})\\ \geq 1-4\Cr{SEPcoupling2}t_{1}H_{2}e^{-\xi}-\mathbb{Q}(B_{4})\geq 1-4\Cr{SEPcoupling2}t_{1}H_{2}e^{-\xi}-20e^{-\nu t/4}\geq 1-5\Cr{SEPcoupling2}t_{1}H_{2}e^{-\xi}.

This shows (3.5) (recalling that t1=4t_{1}=\ell^{4} by (6.49)), and concludes the proof.

The final result which feeds into the proof of Proposition 6.3 is the following.

Lemma 6.10.

The condition (C.3) holds for SEP.

Proof.

Let 1\ell\geq 1, ρ(0,1)\rho\in(0,1) and η0,η0\eta_{0},\eta^{\prime}_{0} be such that the conditions of (C.3) hold. In particular, these imply the existence of x0[0,]x_{0}\in[0,\ell] such that η0(x0)=1\eta_{0}(x_{0})=1 and η0(x0)=0\eta^{\prime}_{0}(x_{0})=0. Let =𝐏^\mathbb{Q}=\widehat{\mathbf{P}} as above (6.2), by which (η,η)(\eta,\eta^{\prime}) are coupled as interchange processes using the same exchange times (𝒫e)eE(\mathcal{P}_{e})_{e\in E}. As observed in (6.3), under \mathbb{Q} the particle of η0\eta_{0} starting at x0x_{0} moves like a simple random walk Z=(Zt)t0Z=(Z_{t})_{t\geq 0} with jump rate ν\nu, and we have ηt(Zt)=0\eta^{\prime}_{t}(Z_{t})=0 for all t0t\geq 0 by construction of 𝐏^\widehat{\mathbf{P}}. Hence, (η(x)>0,η(x)=0)(Z=x)\mathbb{Q}(\eta_{\ell}(x)>0,\,\eta^{\prime}_{\ell}(x)=0)\geq\mathbb{Q}(Z_{\ell}=x) and therefore, in order to deduce (3.8) it is enough to argue that

(6.60) (Z=x)(ν2eν)6(ρ+1),x=0,1.\mathbb{Q}(Z_{\ell}=x)\geq\big{(}\tfrac{\nu}{2e^{\nu}}\big{)}^{6(\rho+1)\ell},\quad x=0,1.

Indeed, let ZLZ^{L}_{\ell} (resp. ZRZ^{R}_{\ell}) denote the number of jumps of ZZ to the left (resp. right) during the time interval [0,][0,\ell]. These two variables are independent and distributed as Poi(ν/2)\text{Poi}(\nu\ell/2), which entails that, for x=0,1x=0,1, and x0xx_{0}\geq x,

(Z=x)(ZL=x0x)(ZR=0)(ν2)x0xeν(x0x)!(ν2)x0xeν(ν2eν)x0xe((x0x))ν,\mathbb{Q}(Z_{\ell}=x)\geq\mathbb{Q}(Z^{L}_{\ell}=x_{0}-x)\mathbb{Q}(Z^{R}_{\ell}=0)\geq\left(\frac{\ell\nu}{2}\right)^{x_{0}-x}\frac{e^{-\ell\nu}}{(x_{0}-x)!}\geq\left(\frac{\nu}{2}\right)^{x_{0}-x}e^{-\ell\nu}\\ \geq\left(\frac{\nu}{2e^{\nu}}\right)^{x_{0}-x}e^{-(\ell-(x_{0}-x))\nu},

using that (x0x)!x0x0xx0x(x_{0}-x)!\leq x_{0}^{x_{0}-x}\leq\ell^{x_{0}-x} in the third step. Since 0x0x0\leq x_{0}-x\leq\ell and ν2eν\nu\leq 2e^{\nu}, (6.60) follows for all x0xx_{0}\geq x (and x=0,1x=0,1), and it is easy to see that the bound remains true in the remaining case, i.e. when x0=0=1xx_{0}=0=1-x (now forcing ZL=0Z^{L}_{\ell}=0 and ZR=1Z^{R}_{\ell}=1 instead, which by symmetry of ZLZ^{L}_{\ell} and ZRZ^{R}_{\ell} yields the same bound as when x0=x+1x_{0}=x+1). Overall this yields (3.8). Finally, since the marginal of the coupling \mathbb{Q} for (η,η)(\eta,\eta^{\prime}) is that of Lemma 6.5, we get the first inequality of (3.9) from (C.2.1) with t=t=\ell. The second inequality follows from our condition on kk (using that ρ1\rho\leq 1). ∎

Appendix A Concentration estimates

We collect here a few classical facts on concentration of Poisson and binomial distributions, that are repeatedly use to control the probability that the environment could have an abnormal empirical density. The PCRW case corresponds to Poisson distributions, and the SSEP to binomial distributions).

Lemma A.1.

Let λ,x>0\lambda,x>0, and XPoisson(λ)X\sim\text{Poisson}(\lambda). Then

(A.1) (Xλ+x)exp(x22(λ+x)).\mathbb{P}(X\geq\lambda+x)\leq\exp\left(-\frac{x^{2}}{2(\lambda+x)}\right).

If x[0,λ]x\in[0,\lambda], then

(A.2) (Xλx)exp(x22(λ+x)).\mathbb{P}(X\leq\lambda-x)\leq\exp\left(-\frac{x^{2}}{2(\lambda+x)}\right).

Let m,p(0,1)m\in\mathbb{N},p\in(0,1) and XBin(m,p)X\sim\text{Bin}(m,p). Then for all q(0,1p)q\in(0,1-p), we have

(A.3) (X(p+q)m)exp(mq2/3)\mathbb{P}(X\geq(p+q)m)\leq\exp(-mq^{2}/3)

and for all q(0,p)q\in(0,p):

(A.4) (X(pq)m)exp(mq2/2).\mathbb{P}(X\leq(p-q)m)\leq\exp(-mq^{2}/2).
Proof.

We get (A.3) (resp. (A.4)) from Theorem 4.4 (resp. Theorem 4.5) of [54] with μ=mp\mu=mp and δ=q/pq\delta=q/p\geq q in both cases. We now turn to (A.1). By a classical application (and optimization) of Chernoff’s bound (see Theorem 5.4 of [54]), one shows that

(Xλ+x)eλ(eλ)λ+x(λ+x)λ+x=exp(λ(λ+x)(log(λ+x)logλ1)).\mathbb{P}(X\geq\lambda+x)\leq\frac{e^{-\lambda}(e\lambda)^{\lambda+x}}{(\lambda+x)^{\lambda+x}}=\exp\left(-\lambda-(\lambda+x)(\log(\lambda+x)-\log\lambda\,-1)\right).

Hence for (A.1), it remains to show that

λ+(λ+x)(log(λ+x)logλ1)x22(λ+x).\lambda+(\lambda+x)(\log(\lambda+x)-\log\lambda\,-1)\geq\frac{x^{2}}{2(\lambda+x)}.

Setting u=1+x/λu=1+x/\lambda and multiplying both sides by u/λu/\lambda this amounts to show that the map f:[1,)f:[1,\infty)\to\mathbb{R} defined by

f(u)=u+u2(logu1)(u1)22f(u)=u+u^{2}(\log u\,-1)-\frac{(u-1)^{2}}{2}

remains non-negative. Indeed, f(1)=0f(1)=0 and f(u)=1+2u(logu1)+u(u1)=2(1+u(logu1))f^{\prime}(u)=1+2u(\log u\,-1)+u-(u-1)=2(1+u(\log u\,-1)). One checks that logu11/u\log u\,-1\geq-1/u for all u1u\geq 1 (with equality iff u=1u=1) so that ff^{\prime} is nonnegative on [1,)[1,\infty), and this concludes the proof of (A.1). One proves (A.2) via the same method. ∎

Appendix B The PCRW environment

In this Appendix, we consider another environment, the Poisson Cloud of Random Walks (PCRW) previously considered in [41, 40] among others; we refer to the introduction for a more complete list of references. For this environment, the parameter ρ(0,)\rho\in(0,\infty) governs the intensity of walks entering the picture. We show below, see Lemma B.1 and Proposition B.2, that the PCRW also fits the setup of §2.2 and satisfies (C.1)-(C.3). Hence, this environment yields another example to which the conclusions of our main result, Theorem 3.1, apply. The proof is virtually the same as that of Theorem 1.2 given in §3.2, upon noting that the relevant law of large numbers (3.10) in this context was also shown in [40], yielding the existence of the speed v(ρ)v(\rho) at all but at most two values ρ=ρ±\rho=\rho_{\pm} (as for SEP). The PCRW is defined below using random walks evolving in discrete time, following the practice of [40] and other previous works, but the following results could easily be adapted to random walks in continuous time (with exponential holding times of mean one).

B.1. Definition of the PCRW

The PCRW is a stochastic process η=(ηt(x);x,t+)\eta=(\eta_{t}(x);\,x\in\mathbb{Z},t\in\mathbb{R_{+}}) with state space Σ=+\Sigma=\mathbb{Z}^{\mathbb{Z}_{+}} defined as follows: for any given initial configuration η0Σ\eta_{0}\in\Sigma and every xx\in\mathbb{Z}, place η0(x)\eta_{0}(x) particles at xx. Then, let all the particles follow independent discrete-time lazy simple random walks, i.e. at each integer time, any given particle stays put with probability 1/21/2, or jumps to its left or right neighbour with the same probability 1/41/4. For t0t\geq 0 and xx\in\mathbb{Z}, let ηt(x)\eta_{t}(x) be the number of particles located at xx. We denote by 𝐏PCRWη0\mathbf{P}^{\eta_{0}}_{\text{PCRW}} the canonical law of this environment with initial state η0\eta_{0}, and frequently abbreviate 𝐏η0=𝐏PCRWη0\mathbf{P}^{\eta_{0}}=\mathbf{P}^{\eta_{0}}_{\text{PCRW}} below.

B.2. Properties of the PCRW

We proceed to show that the PCRW environment has the desired features, i.e. that the properties (P.1)-(2.2) listed in §2.2 as well as the conditions (C.1)-(C.2) appearing in §3.1 all hold. The parameter ρ\rho indexing the stationary measures will naturally vary in (0,)(0,\infty). We will in practice always consider a bounded open interval J=(K1,K)J={\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}(K^{-1}},K) for arbitrary K>1K>1 below. The constants \Crdensitydev,,\CrSEPcoupling\Cr{densitydev},\ldots,\Cr{SEPcoupling} appearing as part of the conditions we aim to verify will henceforth be allowed to tacitly depend on KK. Note that this is inconsequential for the purposes of deriving monotonicity of v()v(\cdot) on (0,)(0,\infty) (cf. (3.11)) since this is a local property: to check monotonicity at ρ\rho one simply picks KK large enough such that ρ(K1,K)=J\rho\in{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}(K^{-1}},K)=J.

For the remainder of this appendix, let K>1K>1 be arbitrary and J=def.(K1,K)J\stackrel{{\scriptstyle\text{def.}}}{{=}}(K^{-1},K). We start by verifying properties (P.1)-(2.2).

Lemma B.1.

With

(B.1) μρ=Poi(ρ),ρ(0,+),\mu_{\rho}=\emph{Poi}(\rho)^{\otimes\mathbb{Z}},\quad\rho\in(0,+\infty),

the measures (𝐏η0:η0{0,1})(\mathbf{P}^{\eta_{0}}:\eta_{0}\in\{0,1\}^{\mathbb{Z}}) with 𝐏η0=𝐏PCRWη0\mathbf{P}^{\eta_{0}}=\mathbf{P}_{\textnormal{PCRW}}^{\eta_{0}} and (μρ:ρJ)(\mu_{\rho}:\rho\in J) satisfy all of (P.1)-(2.2).

Proof.

Fix ρ>0\rho>0. Property (P.1) is classical, and follows readily from the time-homogeneity, translation invariance and axial symmetry of the lazy simple random walk. Property (P.2) is also standard: if η0μρ\eta_{0}\sim\mu_{\rho}, by suitably thinning the Poisson process one can realize η0\eta_{0} by decomposing η0(x)=l(x)+c(x)+r(x)\eta_{0}(x)=l(x)+c(x)+r(x) for all xx\in\mathbb{Z}, where l(x)l(x) (resp. c(x)c(x), r(x)r(x)) is the number of particles starting from xx that make their first move to the left (resp. stay put, and make their first move to the right), with l(x),r(x)Poi(ρ/4)l(x),r(x)\sim\text{Poi}(\rho/4), c(x)Poi(ρ/2)c(x)\sim\text{Poi}(\rho/2) and the family of variables (l(x),c(x),r(x))x(l(x),c(x),r(x))_{x\in\mathbb{Z}} is independent. From this one infers that η1μρ\eta_{1}\sim\mu_{\rho}, as η1(x)=l(x+1)+c(x)+r(x1)Poi(ρ)\eta_{1}(x)=l(x+1)+c(x)+r(x-1)\sim\text{Poi}(\rho) for all xx\in\mathbb{Z}, and the variables η1(x)\eta_{1}(x) are independent as xx varies.

As for property (P.3), it holds with the following natural coupling (which straightforwardly yields the correct marginal laws for η\eta and η\eta^{\prime}): if η0(x)η0(x)\eta^{\prime}_{0}(x)\leq\eta_{0}(x) for all xx\in\mathbb{Z}, then one matches injectively each particle of η0\eta^{\prime}_{0} to a particle of η0\eta_{0} located at the same position. The coupling imposes that matched particles follow the same trajectory, and the remaining particles of η0\eta_{0} (if any) follow independent lazy simple random walks, independently of the matched particles. Finally, Property (2.2) is a consequence of the fact that under 𝐏ρ\mathbf{P}^{\rho}, for every finite interval II and every time t0t\geq 0, ηt(I)Poi(ρ|I|)\eta_{t}(I)\sim\text{Poi}(\rho|I|), and combining with the tail estimates (A.1)-(A.2). ∎

We now establish the conditions (C.1)-(C.3).

Proposition B.2.

For (𝐏η0:η0{0,1}+)(\mathbf{P}^{\eta_{0}}:\eta_{0}\in\{0,1\}^{\mathbb{Z}_{+}}) with 𝐏η0=𝐏PCRWη0\mathbf{P}^{\eta_{0}}=\mathbf{P}^{\eta_{0}}_{\textnormal{PCRW}}, ρJ\rho\in J and with ν=1\nu=1, all of (C.1)(C.2)(C.2.1)(C.2.2) and (C.3) hold.

The proof of Proposition B.2 is given in §B.3 below. We start with a coupling result (which for instance readily implies (C.1) as shall be seen), similar in spirit to Lemma B.3 of [41], stating that the evolution of the PCRW with a sufficiently regular deterministic initial condition can be approximated by a product of independent Poisson variables. This is of independent interest (and lurks in the background of various more elaborate coupling constructions employed in §B.3).

Proposition B.3.

There exist positive and finite constants \Clc:diffusive\Cl{c:diffusive}, \Cl[c]EGrestesmallerthanepsilon\Cl[c]{EGrestesmallerthanepsilon} and \Cl[c]coupling1\Cl[c]{coupling-1} such that the following holds. Let ρ(0,K)\rho\in(0,K), ε(0,(Kρ)ρ1)\varepsilon\in(0,(K-\rho)\wedge\rho\wedge 1) and H,,tH,\ell,t\in\mathbb{N} be such that \Crc:diffusive2<t<H/2\Cr{c:diffusive}\ell^{2}<t<H/2 and (ρ+ε)(t1logt)1/2<\CrEGrestesmallerthanepsilonε(\rho+\varepsilon){(t^{-1}{\log t})^{1/2}}\ell<\Cr{EGrestesmallerthanepsilon}\varepsilon. There exists a coupling \mathbb{Q} of (ηρε,ηt,ηρ+ε)(\eta^{\rho-\varepsilon},\eta_{t},\eta^{\rho+\varepsilon}) with ηρ±εμρ±ε\eta^{\rho\pm\varepsilon}\sim\mu_{\rho\pm\varepsilon} (and ηt\eta_{t} sampled under 𝐏η0\mathbf{P}^{\eta_{0}}) such that, if η0Σ\eta_{0}\in\Sigma is such that for any interval I[0,H]I\subseteq[0,H] with |I|=|I|=\ell,

(B.2) (ρε2)η0(I)(resp. η0(I)(ρ+ε2)),\textstyle(\rho-\frac{\varepsilon}{2})\ell\leq\eta_{0}(I)\quad(\text{resp.~}\eta_{0}(I)\leq(\rho+\frac{\varepsilon}{2})\ell),

then with G={ηρε|[t,Ht]ηt|[t,Ht]}G=\{\eta^{\rho-\varepsilon}|_{[t,H-t]}\preccurlyeq\eta_{t}|_{[t,H-t]}\} (resp. G={ηt|[t,Ht]ηρ+ε|[t,Ht]}G=\{\eta_{t}|_{[t,H-t]}\preccurlyeq\eta^{\rho+\varepsilon}|_{[t,H-t]}\}),

(B.3) (G)1Hexp(\Crcoupling1(ρ+ε)1ε2t).\begin{split}&\mathbb{Q}(G)\geq 1-H\exp\big{(}-\Cr{coupling-1}(\rho+\varepsilon)^{-1}\varepsilon^{2}\sqrt{t}\big{)}.\end{split}

Moreover, the coupling \mathbb{Q} is local in the sense that (ηρε,ηt,ηρ+ε)|[t,Ht](\eta^{\rho-\varepsilon},\eta_{t},\eta^{\rho+\varepsilon})|_{[t,H-t]} depends on the initial condition η0\eta_{0} through η0(x)\eta_{0}(x), x[0,H]x\in[0,H], alone.

We now prepare the ground for the proof of Proposition B.3. Let us abbreviate It=[t,Ht]I_{t}=[t,H-t]. We use the framework of soft local times from Appendix A of [41] (the latter following Section 4 of [64]), which we extend to fit our needs. We define a coupling \mathbb{Q} as follows. Let Λ\Lambda (defined under \mathbb{Q}) be a Poisson point process on +\mathbb{Z}\otimes\mathbb{R}_{+} with intensity 1λ1\otimes\lambda where 11 stands for the counting measure and λ\lambda is the Lebesgue measure on \mathbb{R}. For each zItz\in I_{t}, set

(B.4) ηρ±ε=def.Λ({z}×(0,ρ±ε]).\eta^{\rho\pm\varepsilon}\stackrel{{\scriptstyle\text{def.}}}{{=}}\Lambda(\{z\}\times(0,\rho\pm\varepsilon]).

For zItz\in\mathbb{Z}\setminus I_{t}, let independently ηρ±ε(z)Poi(ρ±ε)\eta^{\rho\pm\varepsilon}(z)\sim\text{Poi}({\rho\pm\varepsilon}). This indeed yields the correct marginal distributions μρ±ε\mu_{\rho\pm\varepsilon} in view of (B.1).

As for ηt\eta_{t}, given any initial configuration η0Σ\eta_{0}\in\Sigma let (xi)i1(x_{i})_{i\geq 1} denote an arbitrary ordering of the positions of the (finitely many) particles of η0([0,H])\eta_{0}([0,H]) (counted with multiplicity, hence the sequence (xi)i1(x_{i})_{i\geq 1} is not necessarily injective). Define for i1i\geq 1 and z[t,H+t]z\in[-t,H+t] gi(z):=qt(xi,z)g_{i}(z):=q_{t}(x_{i},z) with (qn)n(q_{n})_{n\in\mathbb{N}} denoting the discrete-time heat kernel for the lazy simple random walk. Let ξ1:=sup{t0:zΛ({z}×(0,tg1(z)])=0}\xi_{1}:=\sup\{t\geq 0:\bigcup_{z\in\mathbb{Z}}\Lambda(\{z\}\times(0,tg_{1}(z)])=0\} and for i2i\geq 2, define recursively

(B.5) ξi=def.sup{t0:zΛ({z}×(0,ξ1g1(z)++ξi1gi1(z)+tgi(z)])=i1},\xi_{i}\stackrel{{\scriptstyle\text{def.}}}{{=}}\sup\big{\{}t\geq 0:\textstyle\bigcup_{z\in\mathbb{Z}}\Lambda\big{(}\{z\}\times(0,\xi_{1}g_{1}(z)+\ldots+\xi_{i-1}g_{i-1}(z)+tg_{i}(z)]\big{)}=i-1\big{\}},

see Figure 5 of [41]. Note that since each gig_{i} has a finite support (included in [yit,yi+t][y_{i}-t,y_{i}+t]), the ξi\xi_{i}’s are well-defined. In fact by Propositions A.1-A.2 of [41], the variables are i.i.d. Exp(1). For all z[t,H+t]z\in[-t,H+t], we define the soft local time

(B.6) Gη0(z)=def.i1ξiqt(xi,z),G_{\eta_{0}}(z)\stackrel{{\scriptstyle\text{def.}}}{{=}}\sum_{i\geq 1}\xi_{i}q_{t}(x_{i},z),

with ξi\xi_{i} as in (B.5). With these definitions, it follows, denoting

(B.7) h(z)=def.Λ({z}×(0,Gη0(z))),h(z)\stackrel{{\scriptstyle\text{def.}}}{{=}}\Lambda(\{z\}\times(0,G_{\eta_{0}}(z))),

that the family (h(z))z(h(z))_{z\in\mathbb{Z}} is distributed as η~t\widetilde{\eta}_{t}, defined as the restriction of ηt\eta_{t} under 𝐏η0\mathbf{P}^{\eta_{0}} restricted to the particles of η0([0,H])\eta_{0}([0,H]). To see this, note that \mathbb{Q} is such that, if uu is the \mathbb{Z}-coordinate of the particle of Λ\Lambda seen when determining ξi\xi_{i} at (B.5), then the particle xix_{i} of η0\eta_{0} moves to uu by time tt. Remark indeed that by (B.5) and the spatial Markov property for Poisson point processes, the choice of uu is proportional to gi()=qt(xi,)g_{i}(\cdot)=q_{t}(x_{i},\cdot) and independent of what happened in the first i1i-1 steps. In particular, (h(z))zIt(h(z))_{z\in I_{t}} is distributed as ηt|It\eta_{t}|_{I_{t}}, since all the particles of ηt|It\eta_{t}|_{I_{t}} perform at most one step per unit of time, and must have been in [0,H][0,H] at time 0. Finally, independently of all this, let all particles of η0([0,H])\eta_{0}(\mathbb{Z}\setminus[0,H]) follow independent lazy random walks.

Overall \mathbb{Q} indeed defines a coupling of (ηρε,ηt,ηρ+ε)(\eta^{\rho-\varepsilon},\eta_{t},\eta^{\rho+\varepsilon}) with the required marginal law, and the desired locality (see below (B.3)) follows immediately from the previous construction. Moreover, (B.4) and (B.7) imply that under \mathbb{Q}, for all tt\in\mathbb{N} and η0Σ\eta_{0}\in\Sigma,

(B.8) {ηρε|Itηt|It}{zIt:ρεGη0(z)},{ηt|Itηρ+ε|It}{zIt:Gη0(z)ρ+ε}.\begin{split}&\left\{\eta^{\rho-\varepsilon}|_{I_{t}}\preccurlyeq\eta_{t}|_{I_{t}}\right\}\subseteq\{\forall z\in I_{t}:\,\rho-\varepsilon\leq G_{\eta_{0}}(z)\},\\[3.00003pt] &\left\{\eta_{t}|_{I_{t}}\preccurlyeq\eta^{\rho+\varepsilon}|_{I_{t}}\right\}\subseteq\{\forall z\in I_{t}:\,G_{\eta_{0}}(z)\leq\rho+\varepsilon\}.\end{split}

It is now clear from (B.8) that the desired high-probability domination in (B.3) hinges on a suitable control of the soft local time Gη0G_{\eta_{0}} defined by (B.6). To this effect we first isolate the following first moment estimate.

Lemma B.4.

Under the assumptions of Proposition B.3, there exists \ClEGheatkernel\Cl{EGheatkernel} such that for all zItz\in I_{t},

(B.9) (ρε2)(1\CrEGheatkernellogtt)𝔼[Gη0(z)],(resp. 𝔼[Gη0(z)](ρ+ε2)(1+\CrEGheatkernellogtt)).\textstyle(\rho-\frac{\varepsilon}{2})\Big{(}1-\Cr{EGheatkernel}{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\ell}\sqrt{\frac{\log t}{t}}}\Big{)}\leq\mathbb{E}^{\mathbb{Q}}[G_{\eta_{0}}(z)],\quad\left(\text{resp.~}\mathbb{E}^{\mathbb{Q}}[G_{\eta_{0}}(z)]\leq(\rho+\frac{\varepsilon}{2})\Big{(}1+\Cr{EGheatkernel}{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\ell}\sqrt{\frac{\log t}{t}}}\Big{)}\right).
Proof.

Let ct=tlogtc_{t}=\sqrt{t\log t} and fix z[H+t,Ht]z\in[-H+t,H-t]. Cover [zct,z+ct][z-\lfloor c_{t}\rfloor,z+\lfloor c_{t}\rfloor] by a family (Ii)1i(2ct+1)/(I_{i})_{1\leq i\leq\lceil(2\lfloor c_{t}\rfloor+1)/\ell\rceil} of intervals of length \ell, all disjoint except possibly I1I_{1} and I2I_{2}. We focus on the upper bound on 𝔼[Gη0(z)]\mathbb{E}^{\mathbb{Q}}[G_{\eta_{0}}(z)] in (B.9) (the lower bound is derived in a similar fashion). By assumption in (B.2), we have that η0(Ii)(ρ+ε2)\eta_{0}(I_{i})\leq(\rho+\frac{\varepsilon}{2})\ell for all ii, hence, recalling that ξi\xi_{i} in (B.5) has unit mean,

(B.11) E[Gη0(z)]=(B.6)i1qt(xi,z)i=1(2ct+1)/(ρ+ε2)|Ii|maxxIiqt(z,x)+2xz+ct1qt(z,x).E^{\mathbb{Q}}[G_{\eta_{0}}(z)]\stackrel{{\scriptstyle\eqref{eq:couple-pf1}}}{{=}}\sum_{i\geq 1}q_{t}(x_{i},z)\leq\sum_{i=1}^{\lceil(2\lfloor c_{t}\rfloor+1)/\ell\rceil}(\rho+\frac{\varepsilon}{2})|I_{i}|\max_{x\in I_{i}}q_{t}(z,x)+2\sum_{x\geq z+c_{t}-1}q_{t}(z,x).

We start by dealing with the last term above. By Azuma’s inequality, we have that

(B.12) xz+ct1qt(z,x)exp((ct1)22t)exp(logt2+logtt)C(ρ+ε2)t,\sum_{x\geq z+c_{t}-1}q_{t}(z,x)\leq\exp\left(-\frac{(c_{t}-1)^{2}}{2t}\right)\leq\exp\bigg{(}-\frac{\log t}{2}+\sqrt{\frac{\log t}{t}}\bigg{)}\leq\frac{C(\rho+\tfrac{\varepsilon}{2})}{\sqrt{t}},

for some constant C>0C>0, depending on KK. Let us now handle the first term in the right-hand side of (B.11). We start by noting that

(B.15) i=1(2ct+1)/(ρ+ε2)|Ii|maxxIiqt(z,x)(ρ+ε2)i=1(2ct+1)/yIiqt(z,y)+(ρ+ε2)i=1(2ct+1)/yIimaxxIi(qt(z,x)qt(z,y))\sum_{i=1}^{\lceil(2\lfloor c_{t}\rfloor+1)/\ell\rceil}(\rho+\tfrac{\varepsilon}{2})|I_{i}|\max_{x\in I_{i}}q_{t}(z,x)\\ \leq(\rho+\tfrac{\varepsilon}{2})\sum_{i=1}^{\lceil(2\lfloor c_{t}\rfloor+1)/\ell\rceil}\sum_{y\in I_{i}}q_{t}(z,y)+(\rho+\tfrac{\varepsilon}{2})\sum_{i=1}^{\lceil(2\lfloor c_{t}\rfloor+1)/\ell\rceil}\sum_{y\in I_{i}}\max_{x\in I_{i}}(q_{t}(z,x)-q_{t}(z,y))

By [50, Proposition 2.4.4], we have that

(B.16) maxyqt(z,y)Ct1/2.\max_{y\in\mathbb{Z}}q_{t}(z,y)\leq{C}t^{-1/2}.

Recalling that at most I1I_{1} and I2I_{2} may overlap, the above implies that

(B.17) i=1(2ct+1)/yIiqt(z,y)1+yI1qt(z,y)1+Ct.\sum_{i=1}^{\lceil(2\lfloor c_{t}\rfloor+1)/\ell\rceil}\sum_{y\in I_{i}}q_{t}(z,y)\leq 1+\sum_{y\in I_{1}}q_{t}(z,y)\leq 1+C\frac{\ell}{\sqrt{t}}.

It remains to deal with the last term in (B.15). By standard heat kernel estimates (see for instance [50, Proposition 2.5.3] and [50, Corollary 2.5.4]) and a computation similar to (6.22), combined with a large deviation estimate on the number NtN_{t} of non-zero steps performed by the lazy random walk up to time tt (using Azuma’s inequality for instance), for all x,y[zct,z+ct]x,y\in[z-c_{t},z+c_{t}] with |xy||x-y|\leq\ell and first assuming that both |xz||x-z| and |yz||y-z| are even, leaving CC be a universal constant changing from line to line, we have that

(B.18) qt(z,x)n=t/4ctt/4+ct(Nt=2n)q2n(z,y)×exp(Cctt)+(|Ntt2|>2ct)qt(z,y)×(1+Clogtt)+2exp(2logt),\begin{split}q_{t}(z,x)&\leq\sum_{n=\lfloor t/4-c_{t}\rfloor}^{\lceil t/4+c_{t}\rceil}\mathbb{P}(N_{t}=2n)q_{2n}(z,y)\times\exp\left(C\frac{\ell c_{t}}{t}\right)+\mathbb{P}(|N_{t}-\tfrac{t}{2}|>2c_{t})\\ &\leq q_{t}(z,y)\times\bigg{(}1+C\ell\sqrt{\frac{\log t}{t}}\bigg{)}+2\exp(-2\log t),\end{split}

where we have to choose \Crc:diffusive\Cr{c:diffusive} (and hence tt) large enough in the assumptions of Proposition B.3 The case where both |xz||x-z| and |yz||y-z| are odd is treated similarly, considering 2n+12n+1 instead of 2n2n. If |xz||x-z| is even and |yz||y-z| is odd, note that for all nn such that t/4ctnt/4+ct\lfloor t/4-c_{t}\rfloor\leq n\leq\lceil t/4+c_{t}\rceil,

(Nt=2n)=2n+1t2n(Nt=2n+1)(1+Clogtt)(Nt=2n+1),\mathbb{P}(N_{t}=2n)=\frac{2n+1}{t-2n}\cdot\mathbb{P}(N_{t}=2n+1)\leq\bigg{(}1+C\sqrt{\frac{\log t}{t}}\bigg{)}\cdot\mathbb{P}(N_{t}=2n+1),

where a corresponding lower bound holds in order to treat the case where |xz||x-z| is odd and |yz||y-z| is even. Hence the result in (B.18) holds regardless of the parity of |xz||x-z| and |yz||y-z|. Using (B.18), we obtain that

(B.21) i=1(2ct+1)/yIimaxxIi(qt(z,x)qt(z,y))i=1(2ct+1)/yIiClogttqt(z,y)+1tClogtt+yI1qt(z,y)Clogtt,\sum_{i=1}^{\lceil(2\lfloor c_{t}\rfloor+1)/\ell\rceil}\sum_{y\in I_{i}}\max_{x\in I_{i}}(q_{t}(z,x)-q_{t}(z,y))\leq\sum_{i=1}^{\lceil(2\lfloor c_{t}\rfloor+1)/\ell\rceil}\sum_{y\in I_{i}}C\ell\sqrt{\frac{\log t}{t}}q_{t}(z,y)+\frac{1}{t}\\ \leq C\ell\sqrt{\frac{\log t}{t}}+\sum_{y\in I_{1}}q_{t}(z,y)\leq C\ell\sqrt{\frac{\log t}{t}},

where we used (B.16) and the fact that 12<t/\Crc:diffusive1\leq\ell^{2}<t/\Cr{c:diffusive}, chose \Crc:diffusive\Cr{c:diffusive} (and hence tt) large enough, and let the value of CC change from one line to the next. Substituting (B.17) and (B.21) into (B.15) and feeding the resulting estimate together with (B.12), into (B.11) yields that

E[Gη0(z)](ρ+ε2)(1+Ct+Clogtt+Ct)(ρ+ε2)(1+Clogtt),E^{\mathbb{Q}}[G_{\eta_{0}}(z)]\leq\Big{(}\rho+\frac{\varepsilon}{2}\Big{)}\bigg{(}1+C\frac{\ell}{\sqrt{t}}+C\ell\sqrt{\frac{\log t}{t}}+\frac{C}{\sqrt{t}}\bigg{)}\leq\Big{(}\rho+\frac{\varepsilon}{2}\Big{)}\bigg{(}1+C\ell\sqrt{\frac{\log t}{t}}\bigg{)},

and the conclusion follows. ∎

We are now ready to give the short proof of Proposition B.3, which combines the above ingredients.

Proof of Proposition B.3.

We use the coupling \mathbb{Q} defined atop Lemma B.4 and show that Gη0G_{\eta_{0}} concentrates in order to exploit (B.8). To this end, first note that for all θ<12minxqt(x,z)1\theta<\frac{1}{2}\min_{x}q_{t}(x,z)^{-1},

(B.22) 𝔼[eθGη0(z)]=i111θqt(xi,z).\mathbb{E}^{\mathbb{Q}}[e^{\theta G_{\eta_{0}}(z)}]=\prod_{i\geq 1}\frac{1}{1-\theta q_{t}(x_{i},z)}.

Observe that only a finite number of factors may differ from 11 (which requires qt(xi,z)>0q_{t}(x_{i},z)>0, hence |xiz|t|x_{i}-z|\leq t). By (B.8), the probability (Gc)\mathbb{Q}(G^{c}) with G={ηt|[t,Ht]ηρ+ε|[t,Ht]}G=\{\eta_{t}|_{[t,H-t]}\preccurlyeq\eta^{\rho+\varepsilon}|_{[t,H-t]}\} is thus bounded from above by

(B.25) [zIt:Gη0(z)>ρ+ε]HsupzIt[Gη0(z)>ρ+ε]Hsupzexp{θ(ρ+εi1qt(xi,z))+i1θ2qt(xi,z)2}Hexp{(ρ+εq)24q2},\mathbb{Q}[\exists z\in I_{t}:\,G_{\eta_{0}}(z)>\rho+\varepsilon]\leq H\sup_{z\in I_{t}}\mathbb{Q}[G_{\eta_{0}}(z)>\rho+\varepsilon]\\ \leq H\sup_{z}\exp\big{\{}-\textstyle\theta\big{(}\rho+\varepsilon-\sum_{i\geq 1}q_{t}(x_{i},z)\big{)}+\sum_{i\geq 1}\theta^{2}q_{t}(x_{i},z)^{2}\big{\}}\leq H\exp\big{\{}\textstyle-\frac{(\rho+\varepsilon-\sum q)^{2}}{4\sum q^{2}}\big{\}},

using (B.22), the exponential Markov inequality and the inequality log(1x)xx2\log(1-x)\geq-x-x^{2} for |x|<12|x|<\frac{1}{2} in the second step and optimizing over θ\theta in the third, and abbreviating qα=i1qt(xi,z)α\sum q^{\alpha}=\sum_{i\geq 1}q_{t}(x_{i},z)^{\alpha}. Using (B.16) and then (B.9), we have

(B.28) i1qt(xi,z)2Cti1qt(xi,z)=CtE[Gη0(z)]Ct(K+1+\CrEGheatkernel(ρ+ε)logtt)Ct((K+1)+\CrEGheatkernel\CrEGrestesmallerthanepsilonε)CKt,\sum_{i\geq 1}q_{t}(x_{i},z)^{2}\leq\frac{C}{\sqrt{t}}\sum_{i\geq 1}q_{t}(x_{i},z)=\frac{C}{\sqrt{t}}E^{\mathbb{Q}}[G_{\eta_{0}}(z)]\\ \leq\frac{C}{\sqrt{t}}\bigg{(}K+1+\Cr{EGheatkernel}(\rho+\varepsilon){{\ell}\sqrt{\frac{\log t}{t}}}\bigg{)}\leq\frac{C}{\sqrt{t}}\left((K+1)+{\Cr{EGheatkernel}}\Cr{EGrestesmallerthanepsilon}\varepsilon\right)\leq\frac{C}{K\sqrt{t}},

where we used the assumptions of Proposition B.3, chose \CrEGrestesmallerthanepsilon\Cr{EGrestesmallerthanepsilon} small enough (depending on \CrEGheatkernel\Cr{EGheatkernel}), and let the value of CC change in the last inequality (depending on K1K\geq 1). Moreover,

(B.29) K+1ρ+εi1qt(xi,z)ε2\CrEGheatkernel(ρ+ε)logttε2\CrEGheatkernel\CrEGrestesmallerthanepsilonεε4,K+1\geq\rho+\varepsilon-\sum_{i\geq 1}q_{t}(x_{i},z)\geq\frac{\varepsilon}{2}-{\Cr{EGheatkernel}}(\rho+\varepsilon){{\ell}\sqrt{\frac{\log t}{t}}}\geq\frac{\varepsilon}{2}-{\Cr{EGheatkernel}}\Cr{EGrestesmallerthanepsilon}\varepsilon\geq\frac{\varepsilon}{4},

provided that \CrEGrestesmallerthanepsilon\Cr{EGrestesmallerthanepsilon} is small enough (depending on \CrEGheatkernel\Cr{EGheatkernel}). Substituting the two displays above into (B.25) provides the asserted upper bound on (Gc)\mathbb{Q}(G^{c}) in (B.3). For the other choice of GG in (B.3), we bound, using that log(1+x)xx2/2\log(1+x)\geq x-x^{2}/2 for all x0x\geq 0,

(B.32) [zIt:Gη0(z)<ρε]HsupzIt[Gη0(z)<ρε]Hsupzexp{θ(ερ+i1qt(xi,z))+12i1θ2qt(xi,z)}Hexp{(ερ+q)22q2},\mathbb{Q}[\exists z\in I_{t}:\,G_{\eta_{0}}(z)<\rho-\varepsilon]\leq H\sup_{z\in I_{t}}\mathbb{Q}[G_{\eta_{0}}(z)<\rho-\varepsilon]\\ \leq H\sup_{z}\exp\big{\{}-\textstyle\theta\big{(}\varepsilon-\rho+\sum_{i\geq 1}q_{t}(x_{i},z)\big{)}+\frac{1}{2}\sum_{i\geq 1}\theta^{2}q_{t}(x_{i},z)\big{\}}\leq H\exp\big{\{}\textstyle-\frac{(\varepsilon-\rho+\sum q)^{2}}{2\sum q^{2}}\big{\}},

optimizing again over θ\theta in the last step. We conclude in the same way as for the upper bound. ∎

B.3. Proof of Proposition B.2

The proof of Proposition B.2 follows immediately by combining Lemmas B.5-B.9 below, each of which focuses on one specific property among (C.1)(C.2), (C.2.1), (C.2.2) and (C.3), which are proved in this order. Recall that J=(K1,K)J=(K^{-1},K) for some K>1K>1 and that constants may implicitly depend on KK.

Lemma B.5.

Condition (C.1) (with ν=1\nu=1) holds for PCRW.

Proof.

Let ρ,ε,H,,t\rho,\varepsilon,H,\ell,t and η0\eta_{0} be such that the conditions of (C.1) hold. It is straightforward to check that these imply the conditions for applying Proposition B.3 with (2H,2ε)(2H,2\varepsilon) instead of (H,ε)(H,\varepsilon), provided that \Crdensitystable\Cr{densitystable} is large enough w.r.t. \Crc:diffusive\Cr{c:diffusive}\CrEGrestesmallerthanepsilon\Cr{EGrestesmallerthanepsilon} and KK. Hence, we can now use Proposition B.3 with (2H,2ε)(2H,2\varepsilon) to show (C.1).

Let 11\leq\ell^{\prime}\leq\ell. By (B.3) (with an appropriate coupling \mathbb{Q} under which η𝐏η0\eta\sim\mathbf{P}^{\eta_{0}} and ηρ±εμρ±ε\eta^{\rho\pm\varepsilon}\sim\mu_{\rho\pm\varepsilon}, and translating [0,2H][0,2H] to [H,H][-H,H] by means of (P.1)), we have that

(B.35) 𝐏η0(for all I[H+2t,H2t]of length ±(ηt(I)ρ)3ε)12exp(\Crcoupling1(ρ+ε)1ε2t)p±\mathbf{P}^{\eta_{0}}\left(\begin{array}[]{c}\text{for all $I^{\prime}\subset[-H+2t,H-2t]$}\\ \text{of length $\ell^{\prime}$: $\pm(\eta_{t}(I^{\prime})-\rho\ell^{\prime})\leq 3\varepsilon\ell^{\prime}$}\end{array}\right)\geq 1-2\exp\left(-\Cr{coupling-1}(\rho+\varepsilon)^{-1}\varepsilon^{2}\sqrt{t}\right)-p_{\pm}

where

p±=def.μρ±ε(there exists an interval I of length  included in[H+2t,H2t] so that: |ηρ±ε(I)(ρ±ε)|2ε).p_{\pm}\stackrel{{\scriptstyle\text{def.}}}{{=}}\mu_{\rho\pm\varepsilon}\left(\begin{array}[]{c}\text{there exists an interval $I^{\prime}$ of length $\ell^{\prime}$ included in}\\ \text{$[-H+2t,H-2t]$ so that: $|\eta^{\rho\pm\varepsilon}(I^{\prime})-(\rho\pm\varepsilon)\ell^{\prime}|\geq 2\varepsilon\ell^{\prime}$}\end{array}\right).

By (2.2) (which holds on account of Lemma B.1) and a union bound over all intervals I[H+2t,H2t]I^{\prime}\subseteq[-H+2t,H-2t] of length \ell^{\prime}, we have that p±2Hexp(\Crdensitydevε2).p_{\pm}\leq 2H\exp(-\Cr{densitydev}\varepsilon^{2}\ell^{\prime}). Combining this and (B.35), noting that \Crcoupling1(ρ+ε)1t\Crdensitystableexpo\Crdensitystableexpo\Cr{coupling-1}(\rho+\varepsilon)^{-1}\sqrt{t}\geq\Cr{densitystableexpo}\ell\geq\Cr{densitystableexpo}\ell^{\prime} if we choose \Crdensitystableexpo\Cr{densitystableexpo} small enough w.r.t. KK and \Crcoupling1\Cr{coupling-1}, yields (C.1). ∎

Lemma B.6.

For every H,t0H,t\geq 0, and all η0,η0Σ\eta_{0},\eta^{\prime}_{0}\in\Sigma such that η0|[0,H]η0|[0,H]\eta_{0}|_{[0,H]}\succcurlyeq\eta_{0}^{\prime}|_{[0,H]}, there exists a coupling \mathbb{Q} of η,η\eta,\eta^{\prime} with respective marginals 𝐏η0\mathbf{P}^{\eta_{0}} and 𝐏η0\mathbf{P}^{\eta^{\prime}_{0}} such that

(B.36) (s[0,t],ηs|[t,Ht]ηs|[t,Ht])=1.\mathbb{Q}\big{(}\forall s\in[0,t],\,\eta_{s}|_{[t,H-t]}\succcurlyeq\eta^{\prime}_{s}|_{[t,H-t]}\big{)}=1.

Therefore, condition (C.2.1) holds for PCRW with ν=1\nu=1.

Proof.

Clearly, (B.36) implies (C.2.1), up to changing HH to 2H2H and translating [0,2H][0,2H] to [H,H][-H,H] (using (P.1), as established in Lemma B.1). We now show (B.36).

Let H,t0H,t\geq 0 and η0,η0Σ\eta_{0},\eta^{\prime}_{0}\in\Sigma be as above. Couple η\eta and η\eta^{\prime} by matching injectively each particle of η0(x)\eta^{\prime}_{0}(x) to a particle of η0(x)\eta_{0}(x), for all x[0,H]x\in[0,H], and by imposing that matched particles follow the same trajectory (and by letting all other particles follow independent lazy random walks). Since particles can make at most one move (to a neighbouring position) per unit of time due to the discrete-time nature of the walks, no particle of η0\eta^{\prime}_{0} outside of [0,H][0,H] can land in [t,Ht][t,H-t] before or at time tt. Thus the event in (B.36) holds with probability 1. ∎

Lemma B.7.

Let ρ(K1,K)\rho\in(K^{-1},K), ε(0,(Kρ)ρ1)\varepsilon\in(0,(K-\rho)\wedge\rho\wedge 1), and H,,tH,\ell,t\in\mathbb{N} be such that \Crc:diffusive2<t<H/2\Cr{c:diffusive}\ell^{2}<t<H/2 and (ρ+32ε)(t1logt)1/2<\CrEGrestesmallerthanepsilonε/4(\rho+\frac{3}{2}\varepsilon){(t^{-1}{\log t})^{1/2}}\ell<\Cr{EGrestesmallerthanepsilon}\varepsilon/4. Let η0,η0Σ\eta_{0},\eta^{\prime}_{0}\in\Sigma be such that for every interval I[0,H]I\subseteq[0,H] of length \ell, we have η0(I)(ρ+3ε/4)\eta_{0}(I)\geq(\rho+3\varepsilon/4)\ell and η0(I)(ρ+ε/4)\eta^{\prime}_{0}(I)\leq(\rho+\varepsilon/4)\ell. Then there exists a coupling \mathbb{Q} of η\eta and η\eta^{\prime} such that

(B.37) (ηt|[t,Ht]ηt|[t,Ht])14Hexp(\Crcoupling1(ρ+ε)1ε2t/4).\mathbb{Q}(\eta^{\prime}_{t}|_{[t,H-t]}\preccurlyeq\eta_{t}|_{[t,H-t]})\geq 1-4H\exp\big{(}-\Cr{coupling-1}(\rho+\varepsilon)^{-1}\varepsilon^{2}\sqrt{t}/4\big{)}.

Consequently, (C.2.2) with ν=1\nu=1 holds for PCRW. Moreover, \mathbb{Q} is local in that (ηt,ηt)|[t,Ht](\eta^{\prime}_{t},\eta_{t})|_{[t,H-t]} depends on the initial conditions (η0,η0)(\eta_{0},\eta_{0}^{\prime}) through η0(x),η0(x)\eta_{0}(x),\eta_{0}^{\prime}(x), x[0,H]x\in[0,H], alone.

Proof.

We first show how (B.37) implies (C.2.2). Let ρ,ε,H,,t,η0\rho,\varepsilon,H,\ell,t,\eta_{0} and η0\eta^{\prime}_{0} satisfy the assumptions of (C.2.2) (in particular, =t1/4\ell=\lfloor t^{1/4}\rfloor). Then they also satisfy the assumptions of Lemma B.7 (with 2H2H instead of HH), upon taking \CrSEPcoupling\Cr{SEPcoupling} large enough in (C.2.2). By (B.37) applied to [H,H][-H,H] instead of [0,2H][0,2H] (again using translation invariance, see (P.1), established in Lemma B.1), (3.7) holds with \CrSEPcoupling2=max(4,\Crcoupling11K/2)\Cr{SEPcoupling2}=\max(4,\Cr{coupling-1}^{-1}K/2) since ρ+εK\rho+\varepsilon\leq K.

We now proceed to the proof of (B.37). Let ρ,ε,H,,t,η0\rho,\varepsilon,H,\ell,t,\eta_{0} and η0\eta^{\prime}_{0} satisfy the assumptions of Lemma B.7. Then we can apply Proposition B.3 to η0\eta_{0} with (ρ+ε,ε/2)(\rho+\varepsilon,\varepsilon/2) instead of (ρ,ε)(\rho,\varepsilon), and the same values of H,,tH,\ell,t. Similarly, we can apply it to η0\eta^{\prime}_{0} with (ρε,ε/2)(\rho-\varepsilon,\varepsilon/2) instead of (ρ,ε)(\rho,\varepsilon). This entails the existence of two couplings 1\mathbb{Q}^{1} of (ηt,ηρ+ε/2)(\eta_{t},\eta^{\rho+\varepsilon/2}) and 2\mathbb{Q}^{2} of (ηρ+ε/2,ηt)(\eta^{\rho+\varepsilon/2},\eta^{\prime}_{t}), where ηρ+ε/2μρ+ε/2\eta^{\rho+\varepsilon/2}\sim\mu_{{\rho+\varepsilon/2}}, such that, abbreviating It=[t,Ht]I_{t}=[t,H-t],

(B.38) 1(ηt|Itηρ+ε/2|It)2(ηρ+ε/2|Itηt|It)1Hexp(\Crcoupling14(ρ+ε/2)1ε2t).\begin{split}&\mathbb{Q}^{1}\big{(}\eta^{\prime}_{t}|_{I_{t}}\preccurlyeq\eta^{\rho+\varepsilon/2}|_{I_{t}}\big{)}\wedge\mathbb{Q}^{2}\big{(}\eta^{\rho+\varepsilon/2}|_{I_{t}}\preccurlyeq\eta_{t}|_{I_{t}}\big{)}\geq 1-H\exp\big{(}-\textstyle\frac{\Cr{coupling-1}}{4}(\rho+\varepsilon/2)^{-1}\varepsilon^{2}\sqrt{t}\big{)}.\end{split}

Applying [32, Lemma 2.4] with (X,Y)=(ηt,ηρ+ε/2)(X,Y)=(\eta_{t},\eta^{\rho+\varepsilon/2}) and (Y,Z)=(ηρ+ε/2,ηt)(Y^{\prime},Z)=(\eta^{\rho+\varepsilon/2},\eta^{\prime}_{t}), one can ‘chain’ 1\mathbb{Q}^{1} and 2\mathbb{Q}^{2}, i.e. one obtains a coupling \mathbb{Q} of (ηt|It,ηρ+ε/2|It,ηt|It)(\eta_{t}|_{I_{t}},\eta^{\rho+\varepsilon/2}|_{I_{t}},\eta^{\prime}_{t}|_{I_{t}}) such that the pair (ηt|It,ηρ+ε/2|It)(\eta_{t}|_{I_{t}},\eta^{\rho+\varepsilon/2}|_{I_{t}}) has the same (marginal) law as under 1\mathbb{Q}^{1} and (ηρ+ε/2|It,ηt|It)(\eta^{\rho+\varepsilon/2}|_{I_{t}},\eta^{\prime}_{t}|_{I_{t}}) has the same law as under 2\mathbb{Q}^{2}; explicitly, a possible choice is

(ηt|It=μA,ηt|It=μB,ηρ+ε/2|It=μC)=1(ηt|It=μA|ηρ+ε/2|It=μC)2(ηt|It=μB|ηρ+ε/2|It=μC)1(ηρ+ε/2|It=μC),\mathbb{Q}\big{(}\eta_{t}|_{I_{t}}=\mu_{A},\eta^{\prime}_{t}|_{I_{t}}=\mu_{B},\eta^{\rho+\varepsilon/2}|_{I_{t}}=\mu_{C}\big{)}\\[5.0pt] =\mathbb{Q}^{1}\big{(}\eta_{t}|_{I_{t}}=\mu_{A}\,\big{|}\,\eta^{\rho+\varepsilon/2}|_{I_{t}}=\mu_{C}\big{)}\cdot\mathbb{Q}^{2}\big{(}\eta^{\prime}_{t}|_{I_{t}}=\mu_{B}\,\big{|}\,\eta^{\rho+\varepsilon/2}|_{I_{t}}=\mu_{C}\big{)}\cdot\mathbb{Q}^{1}\big{(}\eta^{\rho+\varepsilon/2}|_{I_{t}}=\mu_{C}\big{)},

with μA,μB,μC\mu_{A},\mu_{B},\mu_{C} ranging over point measures on ItI_{t}. On account of [32, Remark 2.5,2)] applied with the choices ε1=11(ηt|Itηρ+ε/2|It)\varepsilon_{1}=1-\mathbb{Q}^{1}(\eta^{\prime}_{t}|_{I_{t}}\preccurlyeq\eta^{\rho+\varepsilon/2}|_{I_{t}}) and ε2=12(ηρ+ε/2|Itηt|It)\varepsilon_{2}=1-\mathbb{Q}^{2}(\eta^{\rho+\varepsilon/2}|_{I_{t}}\preccurlyeq\eta_{t}|_{I_{t}}), \mathbb{Q} has the property that (ηt|Itηt|It)1ε1ε2\mathbb{Q}(\eta^{\prime}_{t}|_{I_{t}}\preccurlyeq\eta_{t}|_{I_{t}})\geq 1-\varepsilon_{1}-\varepsilon_{2}. In view of (B.38), (B.37) follows. The asserted locality of \mathbb{Q} is inherited from i\mathbb{Q}^{i}, i=1,2i=1,2, due to Proposition B.3 (used to define i\mathbb{Q}^{i}). ∎

Lemma B.8.

Condition (C.2) with ν=1\nu=1 holds for PCRW.

Proof.

Let ρ(K1,K),ε(0,1)\rho\in(K^{-1},K),\varepsilon\in(0,1), H1,H2,t,H_{1},H_{2},t,\ell\in\mathbb{N}, and η0,η0Σ\eta_{0},\eta^{\prime}_{0}\in\Sigma be such that the conditions of (C.2) hold. We proceed by a two-step coupling similar to the one in Lemma 6.9 and first give a short overview of both steps; cf. also Fig. 7. In the first step, we couple η\eta and η\eta^{\prime} during the time interval [0,t1][0,t_{1}], with t1:=4t_{1}:=\ell^{4}, using the coupling of Lemma B.7 on [H2,H2][H1,H1][-H_{2},H_{2}]\setminus[-H_{1},H_{1}] and the coupling given in Lemma B.6, making sure that these couplings can be simultaneously performed on disjoint intervals. As a result, we get that ηt1(x)ηt1(x)\eta_{t_{1}}(x)\geq\eta^{\prime}_{t_{1}}(x) for all x[H2+t1,H2t1]x\in[-H_{2}+t_{1},H_{2}-t_{1}], except possibly within two intervals around H1-H_{1} and H1H_{1}, of width O(t1)O(t_{1}).

In the second step, during the time interval [t1,t][t_{1},t], we couple the particles of η\eta^{\prime} on these intervals with "additional" particles of ηt1ηt1\eta_{t_{1}}\setminus\eta^{\prime}_{t_{1}} on [H2,H2][-H_{2},H_{2}] (using that the empirical density of η\eta^{\prime} is slightly larger than that of η\eta on [H2,H2][-H_{2},H_{2}] by (C.1)), using Lemma B.7. This ensures that with large enough probability, all these particles of η\eta^{\prime} get covered by particles of η\eta within time tt1t-t_{1}, without affecting the coupling of the previous step by the Markov property. Step 1: choosing \Crcompatible\Cr{compatible} large enough (in a manner depending on K,\Crc:diffusiveK,\Cr{c:diffusive} and \CrEGrestesmallerthanepsilon\Cr{EGrestesmallerthanepsilon}), as we now briefly explain, the conditions of Lemma B.7 hold for η0,η0\eta_{0},\eta^{\prime}_{0}, with (H,,t)=(H2H11,,t1)(H,\ell,t)=(H_{2}-H_{1}-1,\ell,t_{1}) up to translating [0,H][0,H] in either of the intervals [H2,H11][-H_{2},-H_{1}-1] or [H1+1,H2][H_{1}+1,H_{2}]. Indeed, the choice t1=4t_{1}=\ell^{4} and the assumptions in (C.2) yield that

(ρ+3ε/2)(t11logt1)1/2(K+2)1/2(K+2)ε/\Crcompatible\CrEGrestesmallerthanepsilonε/4,(\rho+3\varepsilon/2)(t_{1}^{-1}\log t_{1})^{1/2}\ell\leq(K+2)\ell^{-1/2}\leq(K+2){\varepsilon}/{\sqrt{\Cr{compatible}}}\leq\Cr{EGrestesmallerthanepsilon}\varepsilon/4,

where we choose \Crcompatible\Cr{compatible} large enough depending on \CrEGrestesmallerthanepsilon\Cr{EGrestesmallerthanepsilon} and KK.

During the time interval [0,t1][0,t_{1}], we apply Lemma B.7, which we now know is in force, simultaneously on [H2,H11][-H_{2},-H_{1}-1] and [H1+1,H2][H_{1}+1,H_{2}]. This is possible owing to the locality property stated as part of Lemma B.7 (see below (B.37)), since the couplings involved rely independently on the particles of η0([H2,H11])\eta_{0}([-H_{2},-H_{1}-1]) and η0([H2,H11])\eta^{\prime}_{0}([-H_{2},-H_{1}-1]), and those of η0([H1+1,H2])\eta_{0}([H_{1}+1,H_{2}]) and η0([H1+1,H2])\eta^{\prime}_{0}([H_{1}+1,H_{2}]) respectively. Moreover, by suitable extension of this coupling we can also couple, during the interval [0,t1][0,t_{1}], the particles of η0([H1,H1])\eta_{0}([-H_{1},H_{1}]) and η0([H1,H1])\eta^{\prime}_{0}([-H_{1},H_{1}]) in the following way: match injectively each particle of η0(x)\eta^{\prime}_{0}(x) to a particle of η0(x)\eta_{0}(x), for all x[H1,H1]x\in[-H_{1},H_{1}], and impose that matched particles follow the same trajectory (note that this is precisely the coupling underlying the statement of Lemma B.6).

From the construction in the previous paragraph, the four groups of particles η0([H2,H2])\eta_{0}(\mathbb{Z}\setminus[-H_{2},H_{2}]), η0([H2,H11])\eta_{0}([-H_{2},-H_{1}-1]), η0([H1,H1])\eta_{0}([-H_{1},H_{1}]) and η0([H1+1,H2])\eta_{0}([H_{1}+1,H_{2}]) evolve independently under \mathbb{Q}, and within each group the particles themselves follow independent lazy simple random walks. Hence under \mathbb{Q}, during [0,t1][0,t_{1}], we have indeed η𝐏η0\eta\sim\mathbf{P}^{\eta_{0}}, and by a similar argument that η𝐏η0\eta^{\prime}\sim\mathbf{P}^{\eta^{\prime}_{0}}.

Now consider the events 1={x[H1+t1,H1t1],s[0,t1],ηs(x)ηs(x)}\mathcal{E}_{1}=\{\forall x\in[-H_{1}+t_{1},H_{1}-t_{1}],\,\forall s\in[0,t_{1}],\,\eta_{s}(x)\geq\eta^{\prime}_{s}(x)\} and 2={x[H2+t1,H11t1][H1+1+t1,H2t1],ηt1(x)ηt1(x)}\mathcal{E}_{2}=\{\forall x\in[-H_{2}+t_{1},-H_{1}-1-t_{1}]\cup[H_{1}+1+t_{1},H_{2}-t_{1}],\,\eta_{t_{1}}(x)\geq\eta^{\prime}_{t_{1}}(x)\} declared under \mathbb{Q}. By Lemmas B.6 and B.7, respectively, we have that

(B.41) (1)=1 and (2)18(H2H1)exp(\Crcoupling1(ρ+ε)1ε22/4).\mathbb{Q}(\mathcal{E}_{1})=1\text{ and }\mathbb{Q}(\mathcal{E}_{2})\geq 1-8(H_{2}-H_{1})\exp\big{(}-\Cr{coupling-1}(\rho+\varepsilon)^{-1}\varepsilon^{2}{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\ell^{2}}/4\big{)}.

Then, define (still under \mathbb{Q}) two further events

3={for all intervals I[H2+2t1,H22t1] of length 2:ηt1(I)(ρ+2ε/5)2}\displaystyle\mathcal{E}_{3}=\left\{\text{for all intervals }I\subseteq[-H_{2}+2t_{1},H_{2}-2t_{1}]\text{ of length $\ell^{2}$}:\,\eta^{\prime}_{t_{1}}(I)\leq(\rho+2\varepsilon/5)\ell^{2}\right\}
4={for all intervals I[H2+2t1,H22t1] of length 2:ηt1(I)(ρ+3ε/5)2}.\displaystyle\mathcal{E}_{4}=\big{\{}\text{for all intervals }I\subseteq[-H_{2}+2t_{1},H_{2}-2t_{1}]\text{ of length $\ell^{2}$}:\,\eta_{t_{1}}(I)\geq(\rho+3\varepsilon/5)\ell^{2}\big{\}}.

We apply condition (C.1), which holds by Lemma B.5 to η\eta^{\prime} with (ρ,ε,,,H,t)=(ρ+ε/5,ε/20,,2,2H2,t1)(\rho,\varepsilon,\ell,\ell^{\prime},H,t)=(\rho+\varepsilon/5,\varepsilon/20,\ell,\ell^{2},2H_{2},t_{1}), and to η\eta with (ρ,ε,,,H,t)=(ρ+3ε/4,ε/20,,2,2H2,t1)(\rho,\varepsilon,\ell,\ell^{\prime},H,t)=(\rho+3\varepsilon/4,\varepsilon/20,\ell,\ell^{2},2H_{2},t_{1}). A straightforward computation proves that the necessary conditions are implied by the assumptions in (C.2). This yields

(B.42) (34)116H2exp(\Crdensitystableexpoε22/400).\mathbb{Q}(\mathcal{E}_{3}\cap\mathcal{E}_{4})\geq 1-16H_{2}\exp(-\Cr{densitystableexpo}\varepsilon^{2}\ell^{2}/400).

Step 2: since 1\mathcal{E}_{1} holds with full \mathbb{Q}-measure, we can, as in Lemma B.6, pair injectively each particle of ηt1([H1+t1,H1t1])\eta^{\prime}_{t_{1}}([-H_{1}+t_{1},H_{1}-t_{1}]) to one of ηt1\eta_{t_{1}} on the same site, and impose by suitable extension of \mathbb{Q} that paired particles follow the same trajectory during [t1,t][t_{1},t], all pairs being together independent. We obtain in this way that

(B.43) (s[0,t],ηs|[H1+t,H1t]ηs|[H1+t,H1t])=1.\mathbb{Q}\big{(}\forall s\in[0,t],\,\eta_{s}|_{[-H_{1}+t,H_{1}-t]}\succcurlyeq\eta^{\prime}_{s}|_{[-H_{1}+t,H_{1}-t]}\big{)}=1.

To complete the construction of \mathbb{Q} it remains to describe the trajectory of the particles of ηt1([H1+t1,H1t1])\eta^{\prime}_{t_{1}}(\mathbb{Z}\setminus[-H_{1}+t_{1},H_{1}-t_{1}]) and ηt1([H1+t1,H1t1])\eta_{t_{1}}(\mathbb{Z}\setminus[-H_{1}+t_{1},H_{1}-t_{1}]), and those of ηt1([H1+t1,H1t1])\eta_{t_{1}}([-H_{1}+t_{1},H_{1}-t_{1}]) that were not paired. On (1234)c(\mathcal{E}_{1}\cap\mathcal{E}_{2}\cap\mathcal{E}_{3}\cap\mathcal{E}_{4})^{c}, let all these particles follow independent lazy simple random walks during [t1,t][t_{1},t], independently from the particles paired at (B.43). On (1234)(12)(\mathcal{E}_{1}\cap\mathcal{E}_{2}\cap\mathcal{E}_{3}\cap\mathcal{E}_{4})\subseteq(\mathcal{E}_{1}\cap\mathcal{E}_{2}), note that ηt(x)ηt(x)\eta_{t}(x)\geq\eta^{\prime}_{t}(x) for all x[H2+t1,H2t1]I1x\in[-H_{2}+t_{1},H_{2}-t_{1}]\setminus I_{1}, where

I1=[H1t1,H1+t11][H1t1+1,H1+t1].I_{1}=[-H_{1}-t_{1},-H_{1}+t_{1}-1]\cup[H_{1}-t_{1}+1,H_{1}+t_{1}].

We pair injectively each particle of ηt1([H2+t1,H2t1]I1)\eta^{\prime}_{t_{1}}([-H_{2}+t_{1},H_{2}-t_{1}]\setminus I_{1}) to one of ηt1\eta_{t_{1}} on the same site, and impose by suitable extension of \mathbb{Q} that paired particles follow the same trajectory during [t1,t][t_{1},t], all pairs being together independent.

For convenience, we introduce

(B.44) η~t1(x)=ηt1(x)ηt1(x)𝟙{x[H2+t1,H2t1]I1}η~t1(x)=ηt1(x)𝟙{xI1},\begin{split}&\widetilde{\eta}_{t_{1}}(x)=\eta_{t_{1}}(x)-\eta^{\prime}_{t_{1}}(x){\mathds{1}}_{\{x\in[-H_{2}+t_{1},H_{2}-t_{1}]\setminus I_{1}\}}\\ &\widetilde{\eta}_{t_{1}}^{\prime}(x)=\eta^{\prime}_{t_{1}}(x){\mathds{1}}_{\{x\in I_{1}\}},\end{split}

which denote the number of particles of ηt1\eta_{t_{1}} and ηt1\eta^{\prime}_{t_{1}} at position xx\in\mathbb{Z} whose trajectory has not yet been described. It thus remains to cover the particles of η~\widetilde{\eta}^{\prime} by those of η~\widetilde{\eta} by time tt. Let us first explain the reasoning to establish this covering. Note that the two intervals making up I1I_{1} escape to our couplings during [0,t1][0,t_{1}], so that we can only guarantee that the empirical density of η~t1\widetilde{\eta}_{t_{1}}^{\prime} is lower than ρ+2ε/5\rho+2\varepsilon/5, see 3\mathcal{E}_{3}. Fortunately, 3\mathcal{E}_{3} and 4\mathcal{E}_{4} ensure that the empirical density of η~t1\widetilde{\eta}_{t_{1}} is at least ε/5\varepsilon/5 on [H2+t1,H2t1]I1[-H_{2}+t_{1},H_{2}-t_{1}]\setminus I_{1}, which is much wider than I1I_{1}. We thus apply Lemma B.7 with a mesh much larger than |I1|=4t1|I_{1}|=4t_{1} in order to ’dilute’ the particles of η~t1\widetilde{\eta}^{\prime}_{t_{1}}. By making this coupling independent of the other particles of η,η\eta,\eta^{\prime} previously paired, we will ensure that both η\eta and η\eta^{\prime} have the correct PCRW marginals. We now formalise this coupling. By choosing \Crcompatible\Cr{compatible} large enough, one can check, via a straightforward computation, that the assumptions in (C.2) imply the necessary conditions to apply Lemma B.7 for η\eta and η\eta^{\prime} of on [H2+2t1,H22t1][-H_{2}+2t_{1},H_{2}-2t_{1}] with (H,,t,ρ,ε)=(2H24t1,5,tt1,ε/40,ε/50)(H,\ell,t,\rho,\varepsilon)=(2H_{2}-4t_{1},\ell^{5},t-t_{1},\varepsilon/40,\varepsilon/50). Hence by Lemma B.7, we can extend \mathbb{Q} such that

(B.45) the trajectories of the particles of η~t1+ andη~t1+ are independent of those paired at (B.43),\begin{split}&\text{the trajectories of the particles of $\widetilde{\eta}_{t_{1}+\cdot}$ and}\\ &\text{$\widetilde{\eta}^{\prime}_{t_{1}+\cdot}$ are independent of those paired at~\eqref{eq:PCRWcentraltrapezoid},}\end{split}

and such that (using that ε<1\varepsilon<1), on the event 1234\mathcal{E}_{1}\cap\mathcal{E}_{2}\cap\mathcal{E}_{3}\cap\mathcal{E}_{4},

(B.48) (η~t|[H2+2t,H22t]η~t|[H2+2t,H22t]|(ηs,ηs)s[0,t1])18H2exp(\Crcoupling1ε2tt1/104).\mathbb{Q}\big{(}\widetilde{\eta}_{t}|_{[-H_{2}+2t,H_{2}-2t]}\succcurlyeq\widetilde{\eta}^{\prime}_{t}\big{|}_{[-H_{2}+2t,H_{2}-2t]}\,|\,(\eta_{s},\eta_{s}^{\prime})_{s\in[0,t_{1}]}\big{)}\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\\ \geq 1-8H_{2}\exp\left(-\Cr{coupling-1}\varepsilon^{2}\sqrt{t-t_{1}}/10^{4}\right).\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}

Let us check that Step 2 yields the marginals η𝐏η1\eta\sim\mathbf{P}^{\eta_{1}} and η𝐏η1\eta\sim\mathbf{P}^{\eta^{\prime}_{1}} during [t1,t][t_{1},t] (we have already seen in Step 1 that η𝐏η0\eta\sim\mathbf{P}^{\eta_{0}} and η𝐏η0\eta\sim\mathbf{P}^{\eta^{\prime}_{0}} during [0,t1][0,t_{1}] so that the Markov property (P.1) will ensure that η𝐏η0\eta\sim\mathbf{P}^{\eta_{0}} and η𝐏η0\eta\sim\mathbf{P}^{\eta^{\prime}_{0}} during [0,t][0,t]). Remark that the events i\mathcal{E}_{i}, 1i41\leq i\leq 4 are measurable w.r.t. the evolution of η\eta and η\eta^{\prime} until time t1t_{1}. On (1234)c(\mathcal{E}_{1}\cap\mathcal{E}_{2}\cap\mathcal{E}_{3}\cap\mathcal{E}_{4})^{c}, by (B.43) and the paragraph below (B.44), it is clear that all particles of ηt1\eta_{t_{1}} follow independent lazy simple random walks, and that the same is true for ηt1\eta^{\prime}_{t_{1}}. On 1234\mathcal{E}_{1}\cap\mathcal{E}_{2}\cap\mathcal{E}_{3}\cap\mathcal{E}_{4}, (B.45) and Lemma B.7 ensure that this is also the case. Therefore, we have indeed that η𝐏η1\eta\sim\mathbf{P}^{\eta_{1}} and η𝐏η1\eta\sim\mathbf{P}^{\eta^{\prime}_{1}}.

Finally, we explain how to derive (C.2) from our construction. Note that (3.4) immediately follows from (B.43) (with full \mathbb{Q}-probability). As for (3.5), we have

(B.51) Q=def.(ηt|[H2+6t,H26t]ηt|[H2+6t,H26t])(η~t|[H2+2t,H22t]η~t|[H2+2t,H22t])Q\stackrel{{\scriptstyle\text{def.}}}{{=}}\mathbb{Q}({\eta}_{t}|_{[-H_{2}+6t,H_{2}-6t]}\succcurlyeq{\eta}^{\prime}_{t}|_{[-H_{2}+6t,H_{2}-6t]})\\ \geq\mathbb{Q}(\widetilde{\eta}_{t}|_{[-H_{2}+2t,H_{2}-2t]}\succcurlyeq\widetilde{\eta}^{\prime}_{t}|_{[-H_{2}+2t,H_{2}-2t]})

by (B.43) and (B.44). Thus, by combining (B.41), (B.42), (B.48) and (B.51), we get

Q1(2c3c4c)8H2exp(\Crcoupling1ε2tt1/10000)132H2exp(\Crdensitystableexpoε2/64)15\CrSEPcoupling24H2exp(ε2/(2\CrSEPcoupling2)),Q\geq 1-\mathbb{Q}(\mathcal{E}_{2}^{c}\cup\mathcal{E}_{3}^{c}\cup\mathcal{E}_{4}^{c})-8H_{2}\exp\left(-\Cr{coupling-1}\varepsilon^{2}\sqrt{t-t_{1}}/10000\right)\\ \geq 1-32H_{2}\exp\left(-\Cr{densitystableexpo}\varepsilon^{2}\ell/64\right)\geq 1-5\Cr{SEPcoupling2}\ell^{4}H_{2}\exp\big{(}-{\varepsilon^{2}\ell}/{(2\Cr{SEPcoupling2})}\big{)},

choosing \Crcompatible\Cr{compatible} and \CrSEPcoupling2\Cr{SEPcoupling2} large enough (w.r.t. KK, \Crdensitystableexpo\Cr{densitystableexpo} and \Crcoupling1\Cr{coupling-1}). This yields (3.5) and concludes the proof, since (3.4) is implied by Lemma B.6. ∎

Lemma B.9.

Condition (C.3) (with ν=1\nu=1) holds for PCRW, the constraint on kk with the pre-factor 4848 now replaced by 24(K+1)24(K+1).

Remark B.10.

The modification of the pre-factor appearing in the constraint on kk is inconsequential for our arguments (and consistent with SEP where one can afford to choose K=1K=1 since J=(0,1)J=(0,1)). Indeed (C.3) is only used at (5.60), where this modified condition on kk clearly holds for LL large enough (KK being fixed).

Proof.

We adapt the proof of Lemma 6.10. Let H,,k1H,\ell,k\geq 1, ρ(0,K),ε(0,Kρ)\rho\in(0,K),\varepsilon\in(0,K-\rho) and η0,η0\eta_{0},\eta^{\prime}_{0} be such that the conditions of (C.3) hold. We can thus pair injectively each particle of η0([H,H])\eta^{\prime}_{0}([-H,H]) to one particle of η0([H,H])\eta_{0}([-H,H]) located at the same position. Moreover, y assumption there is at least one particle of η0([0,])\eta_{0}([0,\ell]) that is not paired. For s0s\geq 0, denote ZsZ_{s} the position of this particle at time ss.

Let \mathbb{Q} be a coupling of η\eta and η\eta^{\prime} during [0,][0,\ell] such that paired particles perform the same lazy simple random walk (independently from all other pairs), and all other particles of η0\eta_{0} and η0\eta^{\prime}_{0} follow independent lazy simple random walks (which yields the marginals η𝐏η0\eta\sim\mathbf{P}^{\eta_{0}} and η𝐏η0\eta^{\prime}\sim\mathbf{P}^{\eta^{\prime}_{0}}). As in Lemma B.6, we get that

(B.54) (s[0,],ηs|[H+,H]ηs|[H+,H])=1\mathbb{Q}\big{(}\forall s\in[0,\ell],\,\eta_{s}|_{[-H+\ell,H-\ell]}\succcurlyeq\eta^{\prime}_{s}|_{[-H+\ell,H-\ell]}\big{)}=1

and (3.9) follows.

It remains to show (3.8). Assume for convenience that \ell is even (the case \ell odd being treated in essentially the same way). Similarly as in the proof of Lemma 6.10, we get that

(B.55) (η(x)>0,η(x)=0)(Z=x,η(x)=0),\mathbb{Q}(\eta_{\ell}(x)>0,\,\eta^{\prime}_{\ell}(x)=0)\geq\mathbb{Q}(Z_{\ell}=x,\,\eta^{\prime}_{\ell}(x)=0),

for x=0,1x=0,1. Note that by our construction, the two events {Z=x}\{Z_{\ell}=x\} and {η(x)=0}\{\eta^{\prime}_{\ell}(x)=0\} are independent. Clearly, we have that for both x=0,1x=0,1,

(B.56) (Z=x)4.\mathbb{Q}(Z_{\ell}=x)\geq 4^{-\ell}.

Note that there is no parity issue in (B.56) because ZZ performs a lazy random walk. Moreover, since η0([3+1,3])6(ρ+1)\eta^{\prime}_{0}([-3\ell+1,3\ell])\leq 6(\rho+1)\ell by assumption and since no particle outside [3+1,3][-3\ell+1,3\ell] can reach 0 by time \ell, it follows that at time 1\ell-1, there are at most 6(ρ+1)6(\rho+1)\ell particles of η0\eta^{\prime}_{0} in [1,1][-1,1], and each of them has probability at least 1/21/2 not to be at x{0,1}x\in\{0,1\} at time \ell. Hence

(B.57) (η(x)=0)26(ρ+1).\mathbb{Q}(\eta^{\prime}_{\ell}(x)=0)\geq 2^{-6(\rho+1)\ell}.

Putting together (B.55), (B.56) and (B.57), we get that the probability on the left of (B.55) is bonded from below by (2e)6(ρ+1)(2e)^{-6(\rho+1)\ell}, whence (3.8). ∎

References

  • [1] M. Aizenman and D. J. Barsky. Sharpness of the phase transition in percolation models. Comm. Math. Phys., 108(3):489–526, 1987.
  • [2] M. Aizenman and G. Grimmett. Strict monotonicity for critical points in percolation and ferromagnetic models. Journal of Statistical Physics, 63(5):817–835, 1991.
  • [3] J. Allasia. Law of large numbers for a finite-range random walk in a dynamic random environment with nonuniform mixing, 2024.
  • [4] C. Alves and S. Popov. Conditional decoupling of random interlacements. ALEA, Lat. Am. J. Probab. Math. Stat., 15(2):1027–1063, 2018.
  • [5] S. Andres, A. Chiarini, J.-D. Deuschel, and M. Slowik. Quenched invariance principle for random walks with time-dependent ergodic degenerate weights. Ann. Probab., 46(1):302–336, 2018.
  • [6] S. Andres, A. Chiarini, and M. Slowik. Quenched local limit theorem for random walks among time-dependent ergodic degenerate weights. Probab. Theory Relat. Fields, 179(3-4):1145–1181, 2021.
  • [7] L. Avena, O. Blondel, and A. Faggionato. A class of random walks in reversible dynamic environments: Antisymmetry and applications to the east model. Journal of Statistical Physics, 165(1):1–23, Oct 2016.
  • [8] L. Avena, F. den Hollander, and F. Redig. Large Deviation Principle for One-Dimensional Random Walk in Dynamic Random Environment: Attractive Spin-Flips and Simple Symmetric Exclusion. Markov Processes and Related Fields, 16(Issue 1):139 – 168, 2010.
  • [9] L. Avena, F. den Hollander, and F. Redig. Law of Large Numbers for a Class of Random Walks in Dynamic Random Environments. Electronic Journal of Probability, 16(none):587 – 617, 2011.
  • [10] L. Avena, R. S. dos Santos, and F. Völlering. Transient random walk in symmetric exclusion: limit theorems and an Einstein relation. ALEA, X(3):693 – 709, 2013.
  • [11] L. Avena, T. Franco, M. Jara, and F. Völlering. Symmetric exclusion as a random environment: Hydrodynamic limits. Annales de l’Institut Henri Poincaré, Probabilités et Statistiques, 51(3):901 – 916, 2015.
  • [12] L. Avena, M. Jara, and F. Völlering. Explicit ldp for a slowed rw driven by a symmetric exclusion process. Probability Theory and Related Fields, 171(3):865–915, Aug 2018.
  • [13] L. Avena, R. Santos, and F. Völlering. Transient random walk in symmetric exclusion: limit theorems and an einstein relation. Latin American journal of probability and mathematical statistics, 10, 02 2011.
  • [14] L. Avena and P. Thomann. Continuity and anomalous fluctuations in random walks in dynamic random environments: Numerics, phase diagrams and conjectures. Journal of Statistical Physics, 147(6):1041–1067, Jul 2012.
  • [15] R. Baldasso and A. Teixeira. How can a clairvoyant particle escape the exclusion process? Annales de l’Institut Henri Poincaré, Probabilités et Statistiques, 54(4):2177 – 2202, 2018.
  • [16] U. Basu and C. Maes. Mobility transition in a dynamic environment. Journal of Physics A: Mathematical and Theoretical, 47(25):255003, jun 2014.
  • [17] V. Beffara and H. Duminil-Copin. The self-dual point of the two-dimensional random-cluster model is critical for q1q\geq 1. Probab. Theory Related Fields, 153(3-4):511–542, 2012.
  • [18] N. Berger, A. Drewitz, and A. F. Ramírez. Effective polynomial ballisticity conditions for random walk in random environment. Commun. Pure Appl. Math., 67(12):1947–1973, 2014.
  • [19] S. A. Bethuelsen and F. Völlering. Absolute continuity and weak uniform mixing of random walk in dynamic random environment. Electronic Journal of Probability, 21(none):1 – 32, 2016.
  • [20] M. Biskup and P.-F. Rodriguez. Limit theory for random walks in degenerate time-dependent random environments. J. Funct. Anal., 274(4):985–1046, 2018.
  • [21] O. Blondel, M. R. Hilario, R. Soares dos Santos, V. Sidoravicius, and A. Teixeira. Random walk on random walks: higher dimensions. arXiv e-prints, page arXiv:1709.01253, Sept. 2017.
  • [22] O. Blondel, M. R. Hilário, and A. Teixeira. Random walks on dynamical random environments with nonuniform mixing. The Annals of Probability, 48(4):2014 – 2051, 2020.
  • [23] B. Bollobás and O. Riordan. The critical probability for random Voronoi percolation in the plane is 1/2. Probab. Theory Relat. Fields, 136(3):417–468, 2006.
  • [24] J. Bricmont and A. Kupiainen. Random walks in asymmetric random environments. Commun. Math. Phys., 142(2):345–420, 1991.
  • [25] F. den Hollander and R. S. dos Santos. Scaling of a random walk on a supercritical contact process. Annales de l’Institut Henri Poincaré, Probabilités et Statistiques, 50(4):1276 – 1300, 2014.
  • [26] F. den Hollander, H. Kesten, and V. Sidoravicius. Random walk in a high density dynamic random environment. Indagationes Mathematicae, 25(4):785–799, 2014. Probability and Numbers.
  • [27] J.-D. Deuschel, T. Kumagai, and M. Slowik. Gradient estimates of the heat kernel for random walks among time-dependent random conductances. Preprint, arXiv:2309.09675, 2023.
  • [28] R. S. dos Santos. Non-trivial linear bounds for a random walk driven by a simple symmetric exclusion process. Electronic Journal of Probability, 19(none):1 – 18, 2014.
  • [29] B. Drossel and M. Kardar. Passive sliders on growing surfaces and advection in burger’s flows. Phys. Rev. B, 66:195414, Nov 2002.
  • [30] H. Duminil-Copin, S. Goswami, A. Raoufi, F. Severo, and A. Yadin. Existence of phase transition for percolation using the Gaussian free field. Duke Math. J., 169(18):3539 – 3563, 2020.
  • [31] H. Duminil-Copin, S. Goswami, P.-F. Rodriguez, and F. Severo. Equality of critical parameters for percolation of Gaussian free field level sets. Duke Math. J., 172(5):839 – 913, 2023.
  • [32] H. Duminil-Copin, S. Goswami, P.-F. Rodriguez, F. Severo, and A. Teixeira. Finite range interlacements and couplings. Ann. Probab. (to appear), available at arXiv:2308.07303, 2023.
  • [33] H. Duminil-Copin, S. Goswami, P.-F. Rodriguez, F. Severo, and A. Teixeira. Phase transition for the vacant set of random walk and random interlacements. Preprint, arXiv:2308.07919, 2023.
  • [34] H. Duminil-Copin, S. Goswami, P.-F. Rodriguez, F. Severo, and A. Teixeira. A characterization of strong percolation via disconnection. Proc. Lond. Math. Soc., 129(2):Paper No. e12622, 1–49, 2024.
  • [35] H. Duminil-Copin, A. Raoufi, and V. Tassion. Sharp phase transition for the random-cluster and potts models via decision trees. Ann. of Math. (2), 189(1):75–99, 2019.
  • [36] H. Duminil-Copin and V. Tassion. A new proof of the sharpness of the phase transition for Bernoulli percolation and the Ising model. Comm. Math. Phys, 343(2):725–745, 2016.
  • [37] N. Gantert, X. Guo, and J. Nagel. Einstein relation and steady states for the random conductance model. Ann. Probab., 45(4):2533–2567, 2017.
  • [38] C. Garban and J. E. Steif. Noise sensitivity of Boolean functions and percolation, volume 5 of IMS Textb. Cambridge: Cambridge University Press, 2015.
  • [39] M. Gopalakrishnan. Dynamics of a passive sliding particle on a randomly fluctuating surface. Phys. Rev. E, 69(1):011105, Jan. 2004.
  • [40] M. R. Hilário, D. Kious, and A. Teixeira. Random walk on the simple symmetric exclusion process. Commun. Math. Phys., 379(1):61–101, 2020.
  • [41] M. Hilário, F. den Hollander, V. Sidoravicius, R. S. dos Santos, and A. Teixeira. Random walk on random walks. Electronic Journal of Probability, 20(none):1 – 35, 2015.
  • [42] F. Huveneers. Response to a small external force and fluctuations of a passive particle in a one-dimensional diffusive environment. Phys. Rev. E, 97:042116, Apr 2018.
  • [43] F. Huveneers and F. Simenhaus. Random walk driven by simple exclusion process. Electron. J. Probab., 20(none):1–42, 2015.
  • [44] M. Jara and O. Menezes. Symmetric exclusion as a random environment: Invariance principle. The Annals of Probability, 48(6):3124 – 3149, 2020.
  • [45] H. Kesten. The critical probability of bond percolation on the square lattice equals 1/2. Commun. Math. Phys., 74:41–59, 1980.
  • [46] H. Kesten, M. V. Kozlov, and F. Spitzer. A limit law for random walk in a random environment. International Journal of Computer Mathematics, 1975.
  • [47] C. Kipnis and S. R. S. Varadhan. Central limit theorem for additive functionals of reversible Markov processes and applications to simple exclusions. Commun. Math. Phys., 104:1–19, 1986.
  • [48] T. Komorowski, C. Landim, and S. Olla. Fluctuations in Markov processes. Time symmetry and martingale approximation., volume 345 of Grundlehren Math. Wiss. Berlin: Springer, 2012.
  • [49] S. M. Kozlov. Averaging method and random walks in inhomogeneous environment. Usp. Mat. Nauk, 40(2(242)):61–120, 1985.
  • [50] G. F. Lawler and V. Limic. Random walk: A modern introduction, volume 123 of Camb. Stud. Adv. Math. Cambridge: Cambridge University Press, 2010.
  • [51] T. M. Liggett. Interacting particle systems, volume 276 of Grundlehren der mathematischen Wissenschaften. Springer-Verlag, New York, 1985.
  • [52] O. Menezes, J. Peterson, and Y. Xie. Variable speed symmetric random walk driven by the simple symmetric exclusion process. Electron. J. Probab., 27:Paper No. 6, 14, 2022.
  • [53] M. V. Menshikov. Coincidence of critical points in percolation problems. Dokl. Akad. Nauk SSSR, 288(6):1308–1311, 1986.
  • [54] M. Mitzenmacher and E. Upfal. Probability and Computing : Randomized Algorithms and Probabilistic Analysis, volume 75. 01 2005.
  • [55] B. Morris. The mixing time for simple exclusion. The Annals of Applied Probability, 16(2):615 – 635, 2006.
  • [56] T. Mountford and M. E. Vares. Random walks generated by equilibrium contact processes. Electronic Journal of Probability, 20(none):1 – 17, 2015.
  • [57] J.-C. Mourrat and F. Otto. Anchored Nash inequalities and heat kernel bounds for static and dynamic degenerate environments. J. Funct. Anal., 270(1):201–228, 2016.
  • [58] S. Muirhead and H. Vanneuville. The sharp phase transition for level set percolation of smooth planar Gaussian fields. Annales de l’Institut Henri Poincaré, Probabilités et Statistiques, 56(2):1358 – 1390, 2020.
  • [59] A. Nagar, S. N. Majumdar, and M. Barma. Strong clustering of noninteracting, sliding passive scalars driven by fluctuating surfaces. Physical review. E, Statistical, nonlinear, and soft matter physics, 74(2 Pt 1):021124, August 2006.
  • [60] A. Nagar, S. N. Majumdar, and M. Barma. Strong clustering of noninteracting, sliding passive scalars driven by fluctuating surfaces. Phys. Rev. E, 74:021124, Aug 2006.
  • [61] T. Orenshtein and R. S. dos Santos. Zero-one law for directional transience of one-dimensional random walks in dynamic random environments. Electronic Communications in Probability, 21(none):1 – 11, 2016.
  • [62] G. C. Papanicolaou and S. R. S. Varadhan. Diffusions with random coefficients. In Collected papers. Volume II: PDE, SDE, diffusions, random media. Edited by Rajendra Bhatia, Abhay Bhatt and K. R. Parthasarathy, pages 533–538. Berlin: Springer; New Dehli: Hindustan Book Agency, 2012.
  • [63] J. Peterson. Lecture notes on random walks in random environments. Notes from a minicourse at IMPA, https://api.semanticscholar.org/CorpusID:124439870, 2013.
  • [64] S. Popov and A. Teixeira. Soft local times and decoupling of random interlacements. J. Eur. Math. Soc., 17:2545–2593, 2015. No 10.
  • [65] A. Prévost, P.-F. Rodriguez, and P. Sousi. Phase transition for the late points of random walk. Preprint, arXiv:2309.03192, 2023.
  • [66] F. Redig and F. Völlering. Random walks in dynamic random environments: A transference principle. The Annals of Probability, 41(5):3157 – 3180, 2013.
  • [67] P.-F. Rodriguez and A.-S. Sznitman. Phase transition and level-set percolation for the Gaussian free field. Comm. Math. Phys., 320(2):571–601, 2013.
  • [68] M. Salvi and F. Simenhaus. Random Walk on a Perturbation of the Infinitely-Fast Mixing Interchange Process. Journal of Statistical Physics, 171:656 – 678, 2018.
  • [69] Y. G. Sinai. The limiting behavior of a one-dimensional random walk in a random medium. Theory of Probability & Its Applications, 27(2):256–268, 1983.
  • [70] R. Soares dos Santos. Non-trivial linear bounds for a random walk driven by a simple symmetric exclusion process. Electron. J. Probab., 19:1–18, 2014. Id/No 49.
  • [71] F. Solomon. Random Walks in a Random Environment. The Annals of Probability, 3(1):1 – 31, 1975.
  • [72] A.-S. Sznitman. Topics in random walks in random environment. School and Conference on Probability Theory, ICTP Lecture Notes, 17:203–266, 2004.
  • [73] A.-S. Sznitman. Vacant set of random interlacements and percolation. Annals of Mathematics, 171(3):2039–2087, 2010.
  • [74] A.-S. Sznitman. Decoupling inequalities and interlacement percolation on G×G\times\mathbb{Z}. Invent. Math., 187(3):645–706, 2012.
  • [75] A.-S. Sznitman and O. Zeitouni. An invariance principle for isotropic diffusions in random environment. Invent. Math., 164(3):455–567, 2006.
  • [76] H. Vanneuville. Sharpness of Bernoulli percolation via couplings. Preprint, arXiv:2201.08223, 2023.
  • [77] O. Zeitouni. Random walks in random environment. In Lectures on probability theory and statistics. Ecole d’Eté de probabilités de Saint-Flour XXXI, pages 191–312. Berlin: Springer, 2004.