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

On the exponent governing the correlation decay of the Airy1 process

Riddhipratim Basu International Centre for Theoretical Sciences, Tata Institute of Fundamental Research, Bangalore, India; E-mail: rbasu@icts.res.in    Ofer Busani Institute for Applied Mathematics, University of Bonn, Endenicher Allee 60,
53115 Bonn, Germany; E-mail: busani@iam.uni-bonn.de
   Patrik L. Ferrari Institute for Applied Mathematics, University of Bonn, Endenicher Allee 60,
53115 Bonn, Germany; E-mail: ferrari@uni-bonn.de
(June 15, 2022)
Abstract

We study the decay of the covariance of the Airy1 process, 𝒜1{\cal A}_{1}, a stationary stochastic process on \mathbb{R} that arises as a universal scaling limit in the Kardar-Parisi-Zhang (KPZ) universality class. We show that the decay is super-exponential and determine the leading order term in the exponent by showing that Cov(𝒜1(0),𝒜1(u))=e(43+o(1))u3\operatorname*{Cov}({\cal A}_{1}(0),{\cal A}_{1}(u))=e^{-(\frac{4}{3}+o(1))u^{3}} as uu\to\infty. The proof employs a combination of probabilistic techniques and integrable probability estimates. The upper bound uses the connection of 𝒜1{\cal A}_{1} to planar exponential last passage percolation and several new results on the geometry of point-to-line geodesics in the latter model which are of independent interest; while the lower bound is primarily analytic, using the Fredholm determinant expressions for the two point function of the Airy1 process together with the FKG inequality.

1 Introduction and the main result

The one-dimensional Kardar-Parisi-Zhang (KPZ) universality class [41] of stochastic growth models has received a lot of attention in recent years, see e.g. the surveys and lecture notes [34, 25, 50, 21, 49, 30, 55, 58]. Two of the most studied models in this class are the exponential/geometric last passage percolation (LPP) and the totally asymmetric simple exclusion process (TASEP). In both cases, one can define a height function h(x,t)h(x,t), where xx stands for space (one-dimensional in our case) and tt for time.

At a large time tt, under the 2/31/32/3-1/3 scaling, one expects to see a non-trivial limit process. To illustrate it, consider the scaling around the origin

htresc(u)=h(ut2/3,t)thma(ut1/3)t1/3h_{t}^{\rm resc}(u)=\frac{h(ut^{2/3},t)-th_{\rm ma}(ut^{-1/3})}{t^{1/3}} (1.1)

with hma(ξ)=limtt1h(ξt)h_{\rm ma}(\xi)=\lim_{t\to\infty}t^{-1}h(\xi t) being the (deterministic) macroscopic limit shape.

The limit process depends on the geometry of the initial condition. One natural initial condition is the stationary one and the limit process in this case, called Airystat, has been determined in [6]. For non-random initial conditions, the two main cases are:

  1. (a)

    curved limit shape hmah_{\rm ma}: one expects the weak limit limthtresc(u)=a1𝒜2(a2u)\lim_{t\to\infty}h^{\rm resc}_{t}(u)=a_{1}{\cal A}_{2}(a_{2}u), with 𝒜2{\cal A}_{2} being the Airy2 process [48] and a1,a2a_{1},a_{2} are model-dependent parameters (see [48, 39, 17] for LPP and TASEP setting and [27] for a non-determinantal case),

  2. (b)

    flat limit shape hmah_{\rm ma}: one expects the weak limit limthtresc(u)=a1𝒜1(a2u)\lim_{t\to\infty}h^{\rm resc}_{t}(u)=a_{1}^{\prime}{\cal A}_{1}(a_{2}^{\prime}u), with 𝒜1{\cal A}_{1} being the Airy1 process [52], with again a1,a2a_{1}^{\prime},a_{2}^{\prime} model-dependent parameters (see [52, 19, 18]).

As universal limit objects in the KPZ universality class, the AirystatAiry_{\rm stat}, as well as 𝒜1{\cal A}_{1} and 𝒜2{\cal A}_{2} (which also are stationary stochastic processes in \mathbb{R}) have attracted much attention. It is known that the one point marginal for 𝒜2{\cal A}_{2} is the GUE Tracy-Widom distribution from random matrix theory [48], whereas the one point marginal for 𝒜1{\cal A}_{1} is a scalar multiple of the GOE Tracy-Widom distribution [52, 19]. The next fundamental question is naturally to understand the two point functions for these processes. Although there are explicit formulae available for the multi-point distributions, extracting asymptotics from these complicated formulae is non-trivial. Widom in [56] (see also [2] for a conditional result) proved that

Cov(𝒜2(0),𝒜2(u))=2u2+𝒪(u4) as u.{\rm Cov}({\cal A}_{2}(0),{\cal A}_{2}(u))=2u^{-2}+\mathcal{O}(u^{-4})\textrm{ as }u\to\infty. (1.2)

Although algebraically there are many similarities between the processes 𝒜1{\cal A}_{1} and 𝒜2{\cal A}_{2} (see the review [29]), the method used in [56] can not be directly applied to the case of the Airy1 process, and the question of understanding the decay of correlations in the Airy1 process had remained open until now.

A numerical study [16] clearly showed that the decay of the covariance for the Airy1 process is very different from that of Airy2, in that it decays super-exponentially fast, i.e., lnCov(𝒜1(0),𝒜1(u))uδ-\ln\operatorname*{Cov}({\cal A}_{1}(0),{\cal A}_{1}(u))\sim u^{\delta} for some δ>1\delta>1. Unfortunately, the numerical data of [16] are coming from a Matlab program developed in [15] and uses the 10-digits machine precision. From the data it was not possible to conjecture the true value of δ\delta.

The reason behind the difference in the decay of the covariances of 𝒜2{\cal A}_{2} and 𝒜1{\cal A}_{1} can be explained as follows. In the curved limit shape situation, the space-time regions which essentially determine the values of h(0,t)h(0,t) and h(ut2/3,t)h(ut^{2/3},t) have an intersection whose size decays polynomially in uu. In contrast, for the flat limit shape, except on a set whose probability goes to zero super-exponentially fast in uu, these regions are disjoint.

The goal of this paper is to prove that the decay of covariance for the Airy1 process is super-exponential with δ=3\delta=3. More precisely, we prove upper and lower bounds of the covariance where exponents have a matching leading order term. The following theorem is the main result of this paper.

Theorem 1.1.

There exist constants c,c>0c,c^{\prime}>0 such that for u>1u>1

eculn(u)e43u3Cov(𝒜1(0),𝒜1(u))ecu2e43u3.e^{-cu\ln(u)}e^{-\frac{4}{3}u^{3}}\leq\operatorname*{Cov}({\cal A}_{1}(0),{\cal A}_{1}(u))\leq e^{c^{\prime}u^{2}}e^{-\frac{4}{3}u^{3}}. (1.3)

Clearly, the threshold u>1u>1 above is arbitrary, and by changing the constants c,cc,c^{\prime} we can get the same bounds for any uu bounded away from 0.

The upper and the lower bounds in Theorem 1.1 are proved separately with very different arguments.

In Section 2 we prove the upper bound; see Corollary 2.2. For this purpose we consider the point-to-line exponential last passage percolation (LPP) which is known to converge to the Airy1 process under an appropriate scaling limit. Corollary 2.2 is an immediate consequence of Theorem 2.1 which proves the corresponding decorrelation statement in the LPP setting. The strategy for the upper bound follows the intuition that the decorrelation comes from the fact that the point-to-line geodesics for two initial points far from each other use mostly disjoint sets of random variables. To make this precise, we prove and use results controlling the transversal fluctuations of the point-to-line geodesics and their coalescence probabilities (see Theorem 2.3 and Theorem 2.7) that are of independent interest. The proof is mainly based on probabilistic arguments, but uses one point moderate deviation estimates for the point-to-point and point-to-line exponential LPP with optimal exponents. Such results have previously been proved in [45], but an estimate with the correct leading order term in the upper tail exponent is required for our purposes. These are obtained in Lemma A.3 and Lemma A.4 by using asymptotic analysis.

In Section 3 we prove the lower bound; see Theorem 3.1. For the lower bound we start with Hoeffding’s covariance formula, which says that the covariance of two random variables is given by the double integral of the difference between their joint distribution and the product of the two marginals; see (3.2). The joint distribution of the Airy1 process is given in terms of a Fredholm determinant (see (3.10)) and the proof uses analytic arguments to obtain precise estimates for these Fredholm determinants. A crucial probabilistic step here, however, is the use of the FKG inequality applied in the LPP setting, which, upon taking an appropriate scaling limit yields that the aforementioned integrand is always non-negative; see Lemma 3.2. This allows one to lower bound the covariance by estimating the integrand only on a suitably chosen compact set, which nonetheless leads to a lower bound with the correct value of the leading order exponent.

We finish this section with a brief discussion of some related works. Studying the decay of correlations in exponential LPP has recently received considerable attention. Following the conjectures in the partly rigorous work [35], the decay of correlations in the time direction has been studied for the stationary and droplet initial conditions in [32], where precise first order asymptotics were obtained (see also [33, 10] for works on the half-space geometry). Similar, but less precise, estimates for the droplet and flat initial conditions were obtained in [11, 12]. All these works also rely on understanding the localization and geometry of geodesics in LPP, some of those results are also useful for us. The lower bounds in [11, 12] also use the FKG inequality in the LPP setting and provide bounds valid in the pre-limit. One might expect that similar arguments can lead to a bound similar, but quantitatively weaker, to Theorem 2.1 valid in the LPP setting.

Acknowledgements.

The work of O. Busani and P.L. Ferrari was partly funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Germany’s Excellence Strategy - GZ 2047/1, projekt-id 390685813. P.L. Ferrari was also supported by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) - Projektnummer 211504053 - SFB 1060. R. Basu is partially supported by a Ramanujan Fellowship (SB/S2/RJN-097/2017) and a MATRICS grant (MTR/2021/000093) from SERB, Govt. of India, DAE project no. RTI4001 via ICTS, and the Infosys Foundation via the Infosys-Chandrasekharan Virtual Centre for Random Geometry of TIFR.

2 Upper Bound

In this section we prove the upper bound of Theorem 1.1.

2.1 Last passage percolation setting

We consider exponential last passage percolation (LPP) on 2\mathbb{Z}^{2}. Let ωi,jexp(1)\omega_{i,j}\sim\exp(1), i,ji,j\in\mathbb{Z}, be independent exponentially distributed random variables with parameter 11. For points u,v2u,v\in\mathbb{Z}^{2} with uvu\prec v, i.e., u1v1u_{1}\leq v_{1} and u2v2u_{2}\leq v_{2}, we denote the passage time between the points uu and vv by

Lu,v=maxπ:uv(i,j)πωi,j,L_{u,v}=\max_{\pi:u\to v}\sum_{(i,j)\in\pi}\omega_{i,j}, (2.1)

where the maximum is taken over all up-right paths from uu to vv in 2\mathbb{Z}^{2}. Denote by Γu,v\Gamma_{u,v} the geodesic from uu to vv, that is, the path π\pi maximizing the above sum. Furthermore, let n={(x,y)2|x+y=n}\mathcal{L}_{n}=\{(x,y)\in\mathbb{Z}^{2}\,|\,x+y=n\} and denote by

Lu,n=maxπ:un(i,j)πωi,jL_{u,\mathcal{L}_{n}}=\max_{\pi:u\to\mathcal{L}_{n}}\sum_{(i,j)\in\pi}\omega_{i,j} (2.2)

the point-to-line last passage time, where the maximum is taken over all up-right paths going from uu to n\mathcal{L}_{n}.

Let us mention some known limiting results of exponential LPP. Let111We do not write explicitly the rounding to integer values, i.e., (x,y)(x,y) stands for (x,y)(\lfloor x\rfloor,\lfloor y\rfloor).

I(u)=u(2N)2/3(1,1),J(u)=(N,N)+u(2N)2/3(1,1)I(u)=u(2N)^{2/3}(1,-1),\quad J(u)=(N,N)+u(2N)^{2/3}(1,-1) (2.3)

and define the rescaled LPP

LN(u)=LI(u),2N4N24/3N1/3,LN(u)=LI(u),(N,N)4N24/3N1/3.L_{N}^{*}(u)=\frac{L_{I(u),\mathcal{L}_{2N}}-4N}{2^{4/3}N^{1/3}},\quad L_{N}(u)=\frac{L_{I(u),(N,N)}-4N}{2^{4/3}N^{1/3}}. (2.4)

Then, by the result on TASEP with density 1/21/2 [52, 19] which can be transferred to LPP using slow decorrelation [26], we know that

limNLN(u)=21/3𝒜1(22/3u),\lim_{N\to\infty}L_{N}^{*}(u)=2^{1/3}{\cal A}_{1}(2^{-2/3}u), (2.5)

where 𝒜1{\cal A}_{1} is the Airy1 process, in the sense of finite-dimensional distributions. Similarly, (see [39] for the geometric case and [22] for a two-parameter extension)

limNLN(u)=𝒜2(u)u2,\lim_{N\to\infty}L_{N}(u)={\cal A}_{2}(u)-u^{2}, (2.6)

with 𝒜2{\cal A}_{2} is the Airy2 process [48], where in [40] the convergence is weak convergence on compact sets.

Refer to caption
Figure 1: (Left): In the typical scenario, the geodesics associated with LN(0)L_{N}^{*}(0) (blue) and LN(u)L_{N}^{*}(u) (green) do not cross the straight line connecting I(u/2)I(u/2) and J(u/2)J(u/2). This suggests that LN(0)L_{N}^{*}(0) and LN(u)L_{N}^{*}(u) are almost independent. (Right) In the untypical case where the blue and green geodesics meet. In that case the geodesics will meet around the point J(u/2)J(u/2). The probability of this event is the main contributor to the covariance of LN(0)L_{N}^{*}(0) and LN(u)L_{N}^{*}(u).

We also denote by ΓN\Gamma_{N} (resp. ΓN\Gamma^{*}_{N}) the (almost surely unique) geodesic attaining L(0,0),(N,N)L_{(0,0),(N,N)} (resp. L(0,0),2NL_{(0,0),\mathcal{L}_{2N}}). For a directed path π\pi, we denote by π(t)=(xy)/2\pi(t)=(x-y)/2 where (x,y)(x,y) is the unique point (if it exits) where π\pi intersect t\mathcal{L}_{t}. The parameter tt will be thought of as time and π(t)\pi(t) will be the position of the path at time tt. For instance, if ΓN\Gamma^{*}_{N} ends at J(u)J(u), then ΓN(2N)=u(2N)2/3\Gamma^{*}_{N}(2N)=u(2N)^{2/3}, see also Figure 1. We shall also denote by L(π)L(\pi) the passage time of the path π\pi, i.e., the sum of the weights on π\pi.

The rescaled last passage times LN(u)L_{N}^{*}(u) and LN(u)L_{N}(u) have super-exponential upper and lower tails (see e.g. Appendices of [32] for a collection of such results and references), which implies that the limit of their covariance is the covariance of their limit, i.e.,

limNCov(LN(u),LN(0))=22/3Cov(𝒜1(22/3u),𝒜1(0)).\lim_{N\to\infty}{\rm Cov}\left(L_{N}^{*}(u),L_{N}^{*}(0)\right)=2^{2/3}{\rm Cov}\left({\cal A}_{1}(2^{-2/3}u),{\cal A}_{1}(0)\right). (2.7)

So if Cov(LN(u),LN(0))eβu3{\rm Cov}\left(L_{N}^{*}(u),L_{N}^{*}(0)\right)\sim e^{-\beta u^{3}}, then Cov(𝒜1(u),𝒜1(0))e4βu3{\rm Cov}\left({\cal A}_{1}(u),{\cal A}_{1}(0)\right)\sim e^{-4\beta u^{3}}.

We first state the upper bound on Cov(LN(u),LN(0)){\rm Cov}\left(L_{N}^{*}(u),L_{N}^{*}(0)\right) which is the main result in this section.

Theorem 2.1.

For N1/14u>1N^{1/14}\gg u>1,

Cov(LN(u),LN(0))ecu2e13u3{\rm Cov}\left(L_{N}^{*}(u),L_{N}^{*}(0)\right)\leq e^{cu^{2}}e^{-\frac{1}{3}u^{3}} (2.8)

for some c>0c>0.

The following corollary, proving the upper bound in Theorem 1.1, is immediate from (2.7) and the above theorem.

Corollary 2.2.

For u>1u>1, we have

Cov(𝒜1(u),𝒜1(0))ecu2e43u3{\rm Cov}\left({\cal A}_{1}(u),{\cal A}_{1}(0)\right)\leq e^{cu^{2}}e^{-\frac{4}{3}u^{3}} (2.9)

for some c>0c>0.

Before proceeding further, let us explain the heuristic idea behind the proof of Theorem 2.1. Let TT denote the straight line joining I(u/2)I(u/2) and J(u/2)J(u/2). Let L~N(0)\tilde{L}_{N}(0) (resp. L~N(u)\tilde{L}_{N}(u)) denote the rescaled last passage time from (0,0)(0,0) to 2N\mathcal{L}_{2N} (resp. from I(u)I(u) to 2N\mathcal{L}_{2N}) in the LPP restricted to use the randomness only to the left (resp. to the right) of TT. Since L~N(0)\tilde{L}_{N}(0) and L~N(u)\tilde{L}_{N}(u) depend on disjoint sets of vertex weights and hence are independent, one expects that the leading order behaviour of the covariance is given by the probability that LN(0)L~N(0)L_{N}^{*}(0)\neq\tilde{L}_{N}(0) and LN(u)L~N(u)L_{N}^{*}(u)\neq\tilde{L}_{N}(u) (parts of the sample space where only one of these two events hold can also contribute, but our arguments will show that these contributions are not of a higher order, see Figure 1). Now,

(LN(0)L~N(0))=(LN(u)L~N(u))=(sup0t2NΓN(t)12u(2N)2/3)\operatorname{\mathbbm{P}}(L_{N}^{*}(0)\neq\tilde{L}_{N}(0))=\operatorname{\mathbbm{P}}(L_{N}^{*}(u)\neq\tilde{L}_{N}(u))=\operatorname{\mathbbm{P}}\Big{(}\sup_{0\leq t\leq 2N}\Gamma^{*}_{N}(t)\geq\tfrac{1}{2}u(2N)^{2/3}\Big{)}

and the probability of the last event is e16u3\lesssim e^{-\frac{1}{6}u^{3}} by Theorem 2.3 below. The proof is completed by showing that the probability of the intersection of the two events has an upper bound which is of the same order (at the level of exponents) as their product. This final step is obtained by considering the two cases, one where the point-to-line geodesics do not intersect, and the second where they do. The first part is bounded using the BK inequality where the probability of the geodesics intersecting is upper bounded separately in Theorem 2.7.

2.2 Localization estimates of geodesics

As explained above, a key step in the proof is to get precise estimates for the probability that the geodesic behaves atypically, i.e., it exits certain given regions. To this end, the main result of this subsection provides the following localization estimate for ΓN\Gamma^{*}_{N} that is of independent interest.

Theorem 2.3.

For N1/14u>1N^{1/14}\gg u>1 we have

(sup0t2NΓN(t)u(2N)2/3)ecu2e43u3\operatorname{\mathbbm{P}}\Big{(}\sup_{0\leq t\leq 2N}\Gamma^{*}_{N}(t)\geq u(2N)^{2/3}\Big{)}\leq e^{cu^{2}}e^{-\frac{4}{3}u^{3}} (2.10)

for some constant c>0c>0.

Although we do not prove a matching lower bound, the constant 43\frac{4}{3} is expected to be optimal; see the discussion following Lemma 2.5. Transversal fluctuation estimates for geodesics in LPP are of substantial interest and have found many applications. This is a first optimal upper bound in this direction for point-to-line geodesic. For point-to-point geodesics, similar estimates (albeit with unspecified constants in front of the cubic exponent) are proved for Poissonian LPP [14] and exponential LPP [12], see also [36] for a lower bound. In fact, we shall need to use the following estimate from [12, 23].

Lemma 2.4 (Proposition C.9 of [12]).

For N1/3u>1N^{1/3}\gg u>1 we have

(sup0t2NΓN(t)u(2N)2/3)ecu3\operatorname{\mathbbm{P}}\Big{(}\sup_{0\leq t\leq 2N}\Gamma_{N}(t)\geq u(2N)^{2/3}\Big{)}\leq e^{-cu^{3}} (2.11)

for some constant c>0c>0.

For the proof of Theorem 2.3 as well as the results in the subsequent subsections of this section, we shall use as input some lower and upper tail estimates of various LPP, which are collected and, if needed, proved in Appendix A.2.

The first step is to prove the special case t=2Nt=2N of Theorem 2.3; we get an estimate of the probability that ΓN\Gamma^{*}_{N} ends in 𝒟u=vuJ(v){\cal D}_{u}=\cup_{v\geq u}J(v), that is, ΓN(2N)u(2N)2/3\Gamma^{*}_{N}(2N)\geq u(2N)^{2/3}.

Lemma 2.5.

For all N1/9u>1N^{1/9}\gg u>1, we have

(ΓN(2N)u(2N)2/3)ecu2e43u3\operatorname{\mathbbm{P}}(\Gamma^{*}_{N}(2N)\geq u(2N)^{2/3})\leq e^{cu^{2}}e^{-\frac{4}{3}u^{3}} (2.12)

for some c>0c>0.

Again, we do not prove a matching lower bound, but the constant 43\frac{4}{3} should be optimal. Indeed, notice that by (2.6), one expects that (2N)2/3ΓN(2N)(2N)^{-2/3}\Gamma^{*}_{N}(2N) weakly converges to the almost surely unique maximizer {\cal M} of 𝒜2(u)u2{\cal A}_{2}(u)-u^{2}. The distribution of {\cal M} has been studied in [53, 7] whence it is known that (||u)e43u3\operatorname{\mathbbm{P}}(|{\cal M}|\geq u)\sim e^{-\frac{4}{3}u^{3}}.

Proof of Lemma 2.5.

Let C0C_{0} be such that (using Lemma A.1)

(L(0,0),(N,N)<4NC0u24/3N1/3)e43u3.\operatorname{\mathbbm{P}}(L_{(0,0),(N,N)}<4N-C_{0}u2^{4/3}N^{1/3})\leq e^{-\frac{4}{3}u^{3}}. (2.13)

We have

(ΓN(2N)u(2N)2/3)\displaystyle\operatorname{\mathbbm{P}}(\Gamma^{*}_{N}(2N)\leq u(2N)^{2/3})\geq  1(L(0,0),𝒟u4NC0u24/3N1/3)\displaystyle\,1-\operatorname{\mathbbm{P}}(L_{(0,0),{\cal D}_{u}}\geq 4N-C_{0}u2^{4/3}N^{1/3}) (2.14)
(L(0,0),(N,N)<4NC0u24/3N1/3).\displaystyle-\operatorname{\mathbbm{P}}(L_{(0,0),(N,N)}<4N-C_{0}u2^{4/3}N^{1/3}).

Using our definition of C0C_{0} and Lemma A.3, we get for N1/3uC0+1N^{1/3}\gg u\geq C_{0}+1

(ΓN(2N)u(2N)2/3)Ce43(u2C0u)3/2+e43u3ecu2e43u3.\operatorname{\mathbbm{P}}(\Gamma^{*}_{N}(2N)\geq u(2N)^{2/3})\leq Ce^{-\frac{4}{3}(u^{2}-C_{0}u)^{3/2}}+e^{-\frac{4}{3}u^{3}}\leq e^{cu^{2}}e^{-\frac{4}{3}u^{3}}. (2.15)

By adjusting the constant cc if necessary, we get the same conclusion for u[1,C0+1]u\in[1,C_{0}+1]. ∎

Our next result is a similar localization estimate for the point-to-line geodesic ΓN\Gamma^{*}_{N} at an intermediate time tt.

Lemma 2.6.

Let t=2τNt=2\tau N. Fix any θ1\theta\geq 1 and take min{τ,1τ}uθ\min\{\tau,1-\tau\}\geq u^{-\theta}. Assume that N1/(9+3θ)u>max{1,43τC0}N^{1/(9+3\theta)}\gg u>\max\{1,\tfrac{4}{3}\sqrt{\tau}C_{0}\} with C0C_{0} as in (2.13). Then there exists a constant c>0c>0 such that

(ΓN(t)u(2N)2/3)e43u3τ3/2e12uτ3/2eu2(c+2C0τ1).\operatorname{\mathbbm{P}}(\Gamma^{*}_{N}(t)\geq u(2N)^{2/3})\leq e^{-\frac{4}{3}u^{3}\tau^{-3/2}}e^{\frac{1}{2}u\tau^{-3/2}}e^{u^{2}(c+2C_{0}\tau^{-1})}. (2.16)

In particular, by choosing another constant c>0c^{\prime}>0,

(ΓN(t)u(2N)2/3)e43u3+cu2\operatorname{\mathbbm{P}}(\Gamma^{*}_{N}(t)\geq u(2N)^{2/3})\leq e^{-\frac{4}{3}u^{3}+c^{\prime}u^{2}} (2.17)

for all N1/(9+3θ)u>1N^{1/(9+3\theta)}\gg u>1.

Notice that (2.16) provides a stronger bound compared to (2.17) for small τ\tau which is expected as the transversal fluctuation should grow with τ\tau. Indeed, for τ1\tau\ll 1, one expects even stronger bounds, see Theorem 3 of [13]. For the proof of Theorem 2.3, (2.17) would suffice but we record the stronger estimate (2.16) as it would be used in the next subsection.

Proof of Lemma 2.6.

We start with a simple rough bound. Notice that since the geodesics are almost surely unique, by planarity they cannot cross each other multiple times. Consequently, if ΓN(2N)A(2N)3\Gamma^{*}_{N}(2N)\leq A(2N)^{3} for some A>0A>0, the geodesic ΓN\Gamma^{*}_{N} lies to the left of the point-to-point geodesic from I(A)I(A) to J(A)J(A). Therefore, the maximal transversal fluctuation of the point-to line geodesic can be upper bounded by the sum of the fluctuation at the endpoint plus the maximal transversal fluctuation of a point-to-point geodesic. Similar arguments will be used multiple times in the sequel and will be referred to as the ordering of geodesics. It follows that

(ΓN(t)C1uτ1/2(2N)2/3)\displaystyle\operatorname{\mathbbm{P}}(\Gamma^{*}_{N}(t)\geq C_{1}u\tau^{-1/2}(2N)^{2/3}) (ΓN(2N)12C1uτ1/2(2N)2/3)\displaystyle\leq\operatorname{\mathbbm{P}}(\Gamma^{*}_{N}(2N)\geq\tfrac{1}{2}C_{1}u\tau^{-1/2}(2N)^{2/3}) (2.18)
+(ΓN(t)12C1uτ1/2(2N)2/3).\displaystyle+\operatorname{\mathbbm{P}}(\Gamma_{N}(t)\geq\tfrac{1}{2}C_{1}u\tau^{-1/2}(2N)^{2/3}).

Applying the bounds of Lemma 2.5 and Lemma 2.4 with the constant C1C_{1} large enough, we get that

(ΓN(t)C1uτ1/2(2N)2/3)e43u3τ3/2.\operatorname{\mathbbm{P}}(\Gamma^{*}_{N}(t)\geq C_{1}u\tau^{-1/2}(2N)^{2/3})\leq e^{-\frac{4}{3}u^{3}\tau^{-3/2}}. (2.19)

Next we need to bound the probability that ΓN(t)\Gamma^{*}_{N}(t) is in [u(2N)2/3,C1uτ1/2(2N)2/3][u(2N)^{2/3},C_{1}u\tau^{-1/2}(2N)^{2/3}]. Define K(v)=(t/2,t/2)+v(2N)2/3(1,1)K(v)=(t/2,t/2)+v(2N)^{2/3}(1,-1). Then, for any SS\in\mathbb{R},

(C1uτ1/2(2N)2/3>ΓN(t)u(2N)2/3)(L(0,0),2NS)+(supuvC1uτ1/2(L(0,0),K(v)+L~K(v),2N)>S),\operatorname{\mathbbm{P}}(C_{1}u\tau^{-1/2}(2N)^{2/3}>\Gamma^{*}_{N}(t)\geq u(2N)^{2/3})\leq\operatorname{\mathbbm{P}}(L_{(0,0),\mathcal{L}_{2N}}\leq S)\\ +\operatorname{\mathbbm{P}}\Big{(}\sup_{u\leq v\leq C_{1}u\tau^{-1/2}}(L_{(0,0),K(v)}+\tilde{L}_{K(v),\mathcal{L}_{2N}})>S\Big{)}, (2.20)

where L~K(v),2N=LK(v),2NωK(v)\tilde{L}_{K(v),\mathcal{L}_{2N}}=L_{K(v),\mathcal{L}_{2N}}-\omega_{K(v)} is the LPP without the first point222The tail bounds in Corollary A.2 clearly continue to hold even after removing the random variable at K(v)K(v) which is Exp(1){\rm Exp}(1)-distributed. The advantage is that in this way L(0,0),K(v)L_{(0,0),K(v)} and L~K(v),2N\tilde{L}_{K(v),\mathcal{L}_{2N}} are independent random variables.. We set

S=4Nau224/3N1/3.S=4N-au^{2}2^{4/3}N^{1/3}. (2.21)

By Corollary A.2, setting a=C0τ1/2u1N2/3a=C_{0}\tau^{-1/2}u^{-1}\ll N^{2/3} with C0C_{0} as in (2.13), we see that

(L(0,0),2NS)e43u3τ3/2.\operatorname{\mathbbm{P}}(L_{(0,0),\mathcal{L}_{2N}}\leq S)\leq e^{-\frac{4}{3}u^{3}\tau^{-3/2}}. (2.22)

It remains to bound the last term in (2.20). We have

(supuvC1uτ1/2(L(0,0),K(v)+L~K(v),2N)>S)\displaystyle\operatorname{\mathbbm{P}}\Big{(}\sup_{u\leq v\leq C_{1}u\tau^{-1/2}}(L_{(0,0),K(v)}+\tilde{L}_{K(v),\mathcal{L}_{2N}})>S\Big{)} (2.23)
k=0C1u/(τδ)u1(supu+kδvu+(k+1)δ(L(0,0),K(v)+L~K(v),2N)>S)\displaystyle\leq\sum_{k=0}^{C_{1}u/(\sqrt{\tau}\delta)-u-1}\operatorname{\mathbbm{P}}\Big{(}\sup_{u+k\delta\leq v\leq u+(k+1)\delta}(L_{(0,0),K(v)}+\tilde{L}_{K(v),\mathcal{L}_{2N}})>S\Big{)}
k=0C1u/(τδ)u1(supvu+kδL(0,0),K(v)+supu+kδvu+(k+1)δL~K(v),2N>S),\displaystyle\leq\sum_{k=0}^{C_{1}u/(\sqrt{\tau}\delta)-u-1}\operatorname{\mathbbm{P}}\Big{(}\sup_{v\geq u+k\delta}L_{(0,0),K(v)}+\sup_{u+k\delta\leq v\leq u+(k+1)\delta}\tilde{L}_{K(v),\mathcal{L}_{2N}}>S\Big{)},

where δ>0\delta>0 will be chosen later. Note that the two supremums above are independent. Denoting Xk=supvu+kδL(0,0),K(v)X_{k}=\sup_{v\geq u+k\delta}L_{(0,0),K(v)} and Y=supδ/2vδ/2L~K(v),2NY=\sup_{-\delta/2\leq v\leq\delta/2}\tilde{L}_{K(v),\mathcal{L}_{2N}}, note also that supu+kδvu+(k+1)δL~K(v),2N\sup_{u+k\delta\leq v\leq u+(k+1)\delta}\tilde{L}_{K(v),\mathcal{L}_{2N}} has the same law as YY for each kk. Hence,

(2.23)=k=0C1u/(τδ)u1(Xk+Y>S)C1uτδ(X0+Y>S)\eqref{eqB6}=\sum_{k=0}^{C_{1}u/(\sqrt{\tau}\delta)-u-1}\operatorname{\mathbbm{P}}(X_{k}+Y>S)\leq\frac{C_{1}u}{\sqrt{\tau}\delta}\operatorname{\mathbbm{P}}(X_{0}+Y>S) (2.24)

since XkX0X_{k}\leq X_{0} for any k0k\geq 0.

As X0X_{0} and YY are independent, it is expected that the leading term should behave as

(X0>S)(Y>SS)\operatorname{\mathbbm{P}}(X_{0}>S^{*})\operatorname{\mathbbm{P}}(Y>S-S^{*}) (2.25)

with SS^{*} chosen such that (2.25) is maximal. Since we want to minimize over a finite number of points (not going to infinity as NN does), we instead look for the maximum of

(X0>S)(Y>SSη24/3N1/3)\operatorname{\mathbbm{P}}(X_{0}>S^{*})\operatorname{\mathbbm{P}}(Y>S-S^{*}-\eta 2^{4/3}N^{1/3}) (2.26)

for some small positive discretization step η\eta; see Figure 2. The natural scale of the fluctuation of X0X_{0} is τ1/324/3N1/3\tau^{1/3}2^{4/3}N^{1/3} and the one for YY is (1τ)1/324/3N1/3(1-\tau)^{1/3}2^{4/3}N^{1/3}, so we choose η\eta at most 14min{(1τ)1/3,τ1/3}\tfrac{1}{4}\min\{(1-\tau)^{1/3},\tau^{1/3}\}.

Refer to caption
Figure 2: The probability (X0+Y>S)\operatorname{\mathbbm{P}}(X_{0}+Y>S) is smaller than the sum of the probabilities (X0A1,YA2)\operatorname{\mathbbm{P}}(X_{0}\geq A_{1},Y\geq A_{2}) with (A1,A2)(A_{1},A_{2}) being the discretized points on A1+A2=Sη24/3N1/3A_{1}+A_{2}=S-\eta 2^{4/3}N^{1/3}.

We want to discretize the interval [4τNu2τ24/3N1/3,4τNau224/3N1/3][4\tau N-\frac{u^{2}}{\tau}2^{4/3}N^{1/3},4\tau N-au^{2}2^{4/3}N^{1/3}] into pieces of size η24/3N1/3\eta 2^{4/3}N^{1/3}. The interval shall be non-empty, which can be ensured if uu is not too small. For that purpose let us assume that u>max{1,43τC0}u>\max\{1,\frac{4}{3}\sqrt{\tau}C_{0}\}. Define the number of discretized points and then η\eta by

M=u2(τ1a)4min{(1τ)1/3,τ1/3},η=u2(τ1a)1M.M=\left\lceil u^{2}(\tau^{-1}-a)\frac{4}{\min\{(1-\tau)^{1/3},\tau^{1/3}\}}\right\rceil,\quad\eta=u^{2}(\tau^{-1}-a)\frac{1}{M}. (2.27)

Our assumption on uu ensures that 1τa14τ\frac{1}{\tau}-a\geq\frac{1}{4\tau}, or equivalently τa34\tau a\leq\frac{3}{4}, as well as η/u21421/3\eta/u^{2}\leq\frac{1}{4}2^{-1/3}. Therefore τa+η/u2<1\tau a+\eta/u^{2}<1.

Let S1=4τNu2τ24/3N1/3S_{1}=4\tau N-\frac{u^{2}}{\tau}2^{4/3}N^{1/3} and S2=SS1=4(1τ)N+1τaτu224/3N1/3S_{2}=S-S_{1}=4(1-\tau)N+\frac{1-\tau a}{\tau}u^{2}2^{4/3}N^{1/3}, S3=4τNau224/3N1/3S_{3}=4\tau N-au^{2}2^{4/3}N^{1/3}. Let SS^{*} be the maximizer of (2.26). Then

(2.24)\displaystyle\eqref{eqB7} C1uτδk=0M1(X0>S1+kη24/3N1/3,Y>S2(k+1)η24/3N1/3)\displaystyle\leq\frac{C_{1}u}{\sqrt{\tau}\delta}\sum_{k=0}^{M-1}\operatorname{\mathbbm{P}}\left(X_{0}>S_{1}+k\eta 2^{4/3}N^{1/3},Y>S_{2}-(k+1)\eta 2^{4/3}N^{1/3}\right) (2.28)
+C1uτδ(X0>S3)+C1uτδ(Y>S2)\displaystyle+\frac{C_{1}u}{\sqrt{\tau}\delta}\operatorname{\mathbbm{P}}\left(X_{0}>S_{3}\right)+\frac{C_{1}u}{\sqrt{\tau}\delta}\operatorname{\mathbbm{P}}\left(Y>S_{2}\right)
C1uτδM(X0>S)(Y>SSη24/3N1/3)\displaystyle\leq\frac{C_{1}u}{\sqrt{\tau}\delta}M\operatorname{\mathbbm{P}}\left(X_{0}>S^{*}\right)\operatorname{\mathbbm{P}}\left(Y>S-S^{*}-\eta 2^{4/3}N^{1/3}\right)
+C1uτδ(X0>S3)+C1uτδ(Y>S2).\displaystyle+\frac{C_{1}u}{\sqrt{\tau}\delta}\operatorname{\mathbbm{P}}\left(X_{0}>S_{3}\right)+\frac{C_{1}u}{\sqrt{\tau}\delta}\operatorname{\mathbbm{P}}\left(Y>S_{2}\right).

Instead of finding SS^{*} that maximizes (2.26), we look to find an upper bound of (2.26) by maximizing the product of the upper bounds of the individual terms. For this, we take δ=(1τ)2/3\delta=(1-\tau)^{2/3}. Then by rescaling Lemma A.3, for s2(τN)2/9s_{2}\ll(\tau N)^{2/9} and u(τN)1/9u\ll(\tau N)^{1/9}

(X0>4τNu2τ24/3N1/3+s1τ1/324/3N1/3)Ce43s13/2\operatorname{\mathbbm{P}}\Big{(}X_{0}>4\tau N-\frac{u^{2}}{\tau}2^{4/3}N^{1/3}+s_{1}\tau^{1/3}2^{4/3}N^{1/3}\Big{)}\leq Ce^{-\frac{4}{3}s_{1}^{3/2}} (2.29)

and by rescaling Proposition A.5, for s2(1τ)2/3N2/3s_{2}\ll(1-\tau)^{2/3}N^{2/3}, we get

(Y>4(1τ)N+s2(1τ)1/324/3N1/3)Cs2e43s23/2.\operatorname{\mathbbm{P}}\Big{(}Y>4(1-\tau)N+s_{2}(1-\tau)^{1/3}2^{4/3}N^{1/3}\Big{)}\leq Cs_{2}e^{-\frac{4}{3}s_{2}^{3/2}}. (2.30)

Thus we need to maximize the product of the terms in (2.29) and (2.30). All the terms in the sum in (2.28) corresponds to s1s_{1} and s2s_{2} being non-negative and also of 𝒪(u2τmin{τ1/3,(1τ)1/3})\mathcal{O}(\frac{u^{2}}{\tau\min\{\tau^{1/3},(1-\tau)^{1/3}\}}). Therefore we can ignore the polynomial pre-factor in (2.30) and look to find s1,s20s_{1},s_{2}\geq 0 such that

u2τ+s1τ1/3+s2(1τ)1/3=(au2+η),s13/2+s23/2 is minimal.-\frac{u^{2}}{\tau}+s_{1}\tau^{1/3}+s_{2}(1-\tau)^{1/3}=-(au^{2}+\eta),\quad s_{1}^{3/2}+s_{2}^{3/2}\textrm{ is minimal}. (2.31)

Denote a~=a+η/u2\tilde{a}=a+\eta/u^{2}, which satisfied τa~<1\tau\tilde{a}<1 by the above assumptions. Plugging in the value of s2s_{2} as function of s1s_{1} into s13/2+s23/2s_{1}^{3/2}+s_{2}^{3/2} and computing its minimum through the derivatives we get

s1=(1τa~)τ2/3u2,s2=(1τa~)(1τ)2/3u2,s13/2+s23/2=u3(1τa~)3/2.s_{1}=\left(\tfrac{1}{\tau}-\tilde{a}\right)\tau^{2/3}u^{2},\quad s_{2}=\left(\tfrac{1}{\tau}-\tilde{a}\right)(1-\tau)^{2/3}u^{2},\quad s_{1}^{3/2}+s_{2}^{3/2}=u^{3}\left(\tfrac{1}{\tau}-\tilde{a}\right)^{3/2}. (2.32)

With this choice of s1s_{1} and s2s_{2}, we write with a minor abuse of notation

S=4τNu2τ24/3N1/3+s1τ1/324/3N1/3,\displaystyle S^{*}=4\tau N-\frac{u^{2}}{\tau}2^{4/3}N^{1/3}+s_{1}\tau^{1/3}2^{4/3}N^{1/3}, (2.33)
SSη24/3N1/3=4(1τ)N+s2(1τ)1/324/3N1/3,\displaystyle S-S^{*}-\eta 2^{4/3}N^{1/3}=4(1-\tau)N+s_{2}(1-\tau)^{1/3}2^{4/3}N^{1/3},

and

(X0>S)(Y>SSη24/3N1/3)Cu2+4θ/3e43u3(1τa~)3/2\operatorname{\mathbbm{P}}\left(X_{0}>S^{*}\right)\operatorname{\mathbbm{P}}\left(Y>S-S^{*}-\eta 2^{4/3}N^{1/3}\right)\leq Cu^{2+4\theta/3}e^{-\frac{4}{3}u^{3}\left(\tfrac{1}{\tau}-\tilde{a}\right)^{3/2}} (2.34)

for some constant C>0C>0, where we have used the a priori bound on s2s_{2} together with the assumption on τ\tau to get the polynomial pre-factor. For τa~<1\tau\tilde{a}<1, (τ1a~)3/2τ3/2(132τa38u2)\left(\tau^{-1}-\tilde{a}\right)^{3/2}\geq\tau^{-3/2}(1-\tfrac{3}{2}\tau a-\frac{3}{8}u^{-2}) so that we get

(2.34)Cu2+4θ/3e43u3τ3/2e12uτ3/2e2u2C0τ1.\eqref{eqB12}\leq Cu^{2+4\theta/3}e^{-\frac{4}{3}u^{3}\tau^{-3/2}}e^{\frac{1}{2}u\tau^{-3/2}}e^{2u^{2}C_{0}\tau^{-1}}. (2.35)

This is the τ\tau-dependent bound which is useful for small enough τ\tau, but the exponent is minimal when τ1\tau\to 1.

Finally, notice that the bound on (X0>S3)\operatorname{\mathbbm{P}}(X_{0}>S_{3}) (resp. (Y>S2)\operatorname{\mathbbm{P}}(Y>S_{2})) corresponds to the bound with the value s2=0s_{2}=0 (resp. s1=0s_{1}=0), thus they are also smaller than (2.34). Thus we have shown that (2.28)(M+2)C1uτδ×(2.35)\eqref{eqB17}\leq(M+2)\frac{C_{1}u}{\sqrt{\tau}\delta}\times\eqref{eq2.36}. The prefactor is only a polynomial in uu (using again the assumptions on τ\tau) and can be absorbed in the u2u^{2} term in the exponent by adjusting the constant. Thus we have proved (2.16), and (2.17) follows by observing that τ1\tau\to 1 is the worst case. The condition uN1/(9+3θ)u\ll N^{1/(9+3\theta)} ensures that all the conditions on s1s_{1}, s2s_{2}, uu mentioned above are satisfied. ∎

We can now prove Theorem 2.3.

Proof of Theorem 2.3.

We shall prove the result for uu sufficiently large, the result for all u>1u>1 shall follow by adjusting the constant cc. Let us set ε=δu3/2\varepsilon=\delta u^{-3/2} for some δ>0\delta>0 to be chosen later (δ\delta will be small but fixed and in particular will not depend on uu or NN). Without loss of generality let us also assume that εN\varepsilon N and 1/ε1/\varepsilon are both integers. Let us define the sequence of points

v0=I(u1),,vj=I(u1)+(jεN,jεN),,vε1=J(u1).v_{0}=I(u-1),\cdots,v_{j}=I(u-1)+(j\varepsilon N,j\varepsilon N),\cdots,v_{\varepsilon^{-1}}=J(u-1). (2.36)

Let AjA_{j} denote the event that ΓN(2jεN)(u1)(2N)2/3\Gamma^{*}_{N}(2j\varepsilon N)\geq(u-1)(2N)^{2/3} for j=1,2,,ε1j=1,2,\ldots,\varepsilon^{-1}, and let BjB_{j} denote the event that suptΓvj1,vju(2N)2/3\sup_{t}\Gamma_{v_{j-1},v_{j}}\geq u(2N)^{2/3} for j=1,2,,ε1j=1,2,\ldots,\varepsilon^{-1} where Γvj1,vj\Gamma_{v_{j-1},v_{j}} denotes the geodesic from vj1v_{j-1} to vjv_{j}. By ordering of geodesics, it follows that on

(jAjc)(jBjc)\Big{(}\bigcap_{j}A_{j}^{c}\Big{)}\cap\Big{(}\bigcap_{j}B_{j}^{c}\Big{)} (2.37)

one has sup0t2NΓN(t)u(2N)2/3\sup_{0\leq t\leq 2N}\Gamma^{*}_{N}(t)\leq u(2N)^{2/3}; see Figure 3.

Refer to caption
Figure 3: The thick blue line is the geodesic ΓN\Gamma^{*}_{N}. It passes to the left of the v1,v2,,vε1v_{1},v_{2},\ldots,v_{\varepsilon^{-1}}. The green thin lines are the point-to-point geodesics from vjv_{j} to vj+1v_{j+1}, j=0,,ε11j=0,\ldots,\varepsilon^{-1}-1, which stay to the left of the dashed line joining I(u)I(u) and J(u)J(u).

Hence,

(sup0t2NΓN(t)u(2N)2/3)j(Aj)+j(Bj).\operatorname{\mathbbm{P}}\Big{(}\sup_{0\leq t\leq 2N}\Gamma^{*}_{N}(t)\geq u(2N)^{2/3}\Big{)}\leq\sum_{j}\operatorname{\mathbbm{P}}(A_{j})+\sum_{j}\operatorname{\mathbbm{P}}(B_{j}). (2.38)

It follows from Lemma 2.6 that for each jj,

(Aj)ec(u1)2e43(u1)3ecu2e43u3\operatorname{\mathbbm{P}}(A_{j})\leq e^{c(u-1)^{2}}e^{-\frac{4}{3}(u-1)^{3}}\leq e^{c^{\prime}u^{2}}e^{-\frac{4}{3}u^{3}} (2.39)

for some new constant c>0c^{\prime}>0. Notice now that

(Bj)=(sup0t2εNΓεN(t)δ2/3u(2εN)2/3).\operatorname{\mathbbm{P}}(B_{j})=\operatorname{\mathbbm{P}}\Big{(}\sup_{0\leq t\leq 2\varepsilon N}\Gamma_{\varepsilon N}(t)\geq\delta^{-2/3}u(2\varepsilon N)^{2/3}\Big{)}. (2.40)

We now use uN2/3u\ll N^{2/3} (which implies u(εN)1/3u\ll(\varepsilon N)^{1/3}) and choose δ\delta sufficiently small such that by Lemma 2.4 we have

(Bj)e43u3.\operatorname{\mathbbm{P}}(B_{j})\leq e^{-\frac{4}{3}u^{3}}. (2.41)

With this choice it follows that

(sup0t2NΓN(t)u(2N)2/3)2ε1ecu2e43u3.\operatorname{\mathbbm{P}}\Big{(}\sup_{0\leq t\leq 2N}\Gamma^{*}_{N}(t)\geq u(2N)^{2/3}\Big{)}\leq 2\varepsilon^{-1}e^{cu^{2}}e^{-\frac{4}{3}u^{3}}. (2.42)

Since ε1=𝒪(u3/2)\varepsilon^{-1}=\mathcal{O}(u^{3/2}), the result follows by adjusting the value of cc. ∎

2.3 Coalescence probability

Consider the point-to-line geodesics ΓN\Gamma^{*}_{N} and Γ~N\tilde{\Gamma}^{*}_{N} to 2N\mathcal{L}_{2N} from (0,0)(0,0) and I(u)I(u) respectively. Owing to the almost sure uniqueness of geodesics, if ΓN\Gamma^{*}_{N} and Γ~N\tilde{\Gamma}^{*}_{N} meet, they coalesce almost surely. Coalescence of geodesics is an important phenomenon in random growth models including first and last passage percolation and has attracted a lot of attention. For exponential LPP, for point-to-point geodesics started at distinct points and ending at a common far away point (or semi-infinite geodesics going in the same direction) tail estimates for distance to coalescence has been obtained; see [13, 46, 57] for more on this. For point-to-line geodesics started at initial points that are far (on-scale), one expects the probability of coalescence to be small. Our next result proves an upper bound to this effect and is of independent interest.

Theorem 2.7.

In the above set-up, for N1/14u>1N^{1/14}\gg u>1

(ΓNΓ~N)e13u3+cu2\operatorname{\mathbbm{P}}(\Gamma^{*}_{N}\cap\tilde{\Gamma}^{*}_{N}\neq\emptyset)\leq e^{-\frac{1}{3}u^{3}+cu^{2}} (2.43)

for some constant c>0c>0.

The rest of this section deals with the proof of Theorem 2.7. We divide it into several smaller results. As always, we shall assume without loss of generality that uu is sufficiently large, extending the results to all u>1u>1 is achieved by adjusting constants.

First of all, due to Theorem 2.3, the probability that the two geodesics in the statement of Theorem 2.7 meet outside the rectangle (u)\mathcal{R}(u) with corners (0,0)(0,0), I(u)I(u), J(u)J(u), (N,N)(N,N) is smaller than the estimate we want to prove. Thus we can restrict to bounding the probability that the two geodesics intersect in (u)\mathcal{R}(u); a stronger result is proved in Lemma 2.8 below. As the number of points in (u)\mathcal{R}(u) where the geodesics Γ~N\tilde{\Gamma}^{*}_{N} and ΓN\Gamma^{*}_{N} could meet is 𝒪(N5/3)\mathcal{O}(N^{5/3}), we need to discretize space. We therefore divide (u)\mathcal{R}(u) into a grid of size εN×(2εN)2/3\varepsilon N\times(2\varepsilon N)^{2/3}, where ε\varepsilon will be taken small enough (but not too small, namely of order u2u^{-2}); see Figure 4.

Refer to caption
Figure 4: Illustration of the grid as discretization of space-time. In the space direction the length is (2εN)2/3(2\varepsilon N)^{2/3} and in the time direction εN\varepsilon N.

For τ\tau an integer multiple of ε\varepsilon, let A(τ,v)A(\tau,v) be the event that the first intersection of ΓN\Gamma^{*}_{N} and Γ~N\tilde{\Gamma}^{*}_{N} occurs at time t(2τN,2τN+2εN]t\in(2\tau N,2\tau N+2\varepsilon N], and they then cross the anti-diagonal grid segment (of length (2εN)2/3(2\varepsilon N)^{2/3}) at time 2τN+2εN2\tau N+2\varepsilon N with mid-point given by

P(τ,v)=(τN+εN+v(2N)2/3,τN+εNv(2N)2/3).P(\tau,v)=(\tau N+\varepsilon N+v(2N)^{2/3},\tau N+\varepsilon N-v(2N)^{2/3}). (2.44)

Notice that, the number of choices of τ\tau and vv is 𝒪(ε5/3)\mathcal{O}(\varepsilon^{-5/3}), which by our choice of ε\varepsilon is at most a polynomial of uu. Thus we need to prove the (A(τ,v))\operatorname{\mathbbm{P}}(A(\tau,v)) is at most e13u3+cu2e^{-\frac{1}{3}u^{3}+cu^{2}} for any τ,v\tau,v. The proof of Theorem 2.7 is completed by taking a union bound.

Our first rough estimate deals with values of vv which are close to 0 or uu and also small values of τ\tau. The basic idea is that in these cases the probability bounds coming from considering the transversal fluctuation of a single geodesic is sufficient.

Lemma 2.8.

Let N1/14u>1N^{1/14}\gg u>1. For any vv satisfying min{v,uv}u(122/3)ε2/3\min\{v,u-v\}\leq u(1-2^{-2/3})-\varepsilon^{2/3} and for any τ22/3ε\tau\leq 2^{-2/3}-\varepsilon,

(A(τ,v))e13u3+cu2\operatorname{\mathbbm{P}}(A(\tau,v))\leq e^{-\frac{1}{3}u^{3}+cu^{2}} (2.45)

for some constant c>0c>0.

Proof.

For uvu(122/3)ε2/3u-v\leq u(1-2^{-2/3})-\varepsilon^{2/3}, we have

(A(τ,v))(suptΓN(t)22/3u(2N)2/3)e13u3+cu2,\operatorname{\mathbbm{P}}(A(\tau,v))\leq\operatorname{\mathbbm{P}}\Big{(}\sup_{t}\Gamma^{*}_{N}(t)\geq 2^{-2/3}u(2N)^{2/3}\Big{)}\leq e^{-\frac{1}{3}u^{3}+cu^{2}}, (2.46)

where the last inequality follows from Theorem 2.3. The same argument gives the desired result for v(122/3)uv\leq(1-2^{-2/3})u by considering the transversal fluctuation of the geodesic Γ~N\tilde{\Gamma}_{N}^{*}.

Next, notice that A(τ,v)A(\tau,v) implies that either ΓN(2τN+2εN)12u(2N)2/3\Gamma^{*}_{N}(2\tau N+2\varepsilon N)\geq\frac{1}{2}u(2N)^{2/3} or Γ~N(2τN+2εN)12u(2N)2/3\tilde{\Gamma}^{*}_{N}(2\tau N+2\varepsilon N)\leq\frac{1}{2}u(2N)^{2/3}, since after meeting they follow the same path. For τ+ε22/3\tau+\varepsilon\leq 2^{-2/3}, by Lemma 2.6 (use the first inequality with ττ+ε22/3\tau\mapsto\tau+\varepsilon\leq 2^{-2/3}) each of these events have probability bounded by e13u3+cu2e^{-\frac{1}{3}u^{3}+cu^{2}} for some constant c>0c>0, completing the proof. ∎

We now proceed towards dealing with the remaining case. Define the segment

𝒮v={(τN+k,τNk)|(v1)(2N)2/3k(v+1)(2N)2/3}.{\cal S}_{v}=\{(\tau N+k,\tau N-k)|(v-1)(2N)^{2/3}\leq k\leq(v+1)(2N)^{2/3}\}. (2.47)

Let C2C_{2} be large enough such that

(L(0,0),2N4NC2u24/3N1/3)e13u3.\operatorname{\mathbbm{P}}(L_{(0,0),\mathcal{L}_{2N}}\leq 4N-C_{2}u2^{4/3}N^{1/3})\leq e^{-\frac{1}{3}u^{3}}. (2.48)

For a path γ\gamma, recall that L(γ)L(\gamma) denotes the passage time of that path. Define the event

B(τ,v)={γ1,γ2|γ1:(0,0)𝒮v,γ2:I(u)𝒮v,γ1γ2= and min{L(γ1)+L~𝒮v,2N,L(γ2)+L~𝒮v,2N}4NC2u24/3N1/3},B(\tau,v)=\{\exists\,\gamma_{1},\gamma_{2}\,|\,\gamma_{1}:(0,0)\to{\cal S}_{v},\gamma_{2}:I(u)\to{\cal S}_{v},\gamma_{1}\cap\gamma_{2}=\emptyset\,\textrm{ and }\\ \min\{L(\gamma_{1})+\tilde{L}_{{\cal S}_{v},\mathcal{L}_{2N}},L(\gamma_{2})+\tilde{L}_{{\cal S}_{v},\mathcal{L}_{2N}}\}\geq 4N-C_{2}u2^{4/3}N^{1/3}\}, (2.49)

where in L~\tilde{L} we remove the first point. Then we have the following estimate.

Refer to caption
Figure 5: Magnification of the local geometry of geodesics used in the sandwitching of Lemma 2.9. The segment 𝒮v{\cal S}_{v} is the dashed one. Notice that the lower line is not τ=0\tau=0.
Lemma 2.9.

Assume (122/3)uε2/3<v<22/3u+ε2/3(1-2^{-2/3})u-\varepsilon^{2/3}<v<2^{-2/3}u+\varepsilon^{2/3} and 22/3ετ1ε2^{-2/3}-\varepsilon\leq\tau\leq 1-\varepsilon (notice that if the geodesics coalesce then A(τ,v)A(\tau,v) must hold for some τ1ε\tau\leq 1-\varepsilon, the case τ=1\tau=1 need not be considered). For N1/11u>1N^{1/11}\gg u>1, there exists a δ>0\delta>0 small enough (not depending on uu and NN) such that with ε=δu2\varepsilon=\delta u^{-2},

(A(τ,v))(B(τ,v))+4Ce13u3.\operatorname{\mathbbm{P}}(A(\tau,v))\leq\operatorname{\mathbbm{P}}(B(\tau,v))+4Ce^{-\frac{1}{3}u^{3}}. (2.50)

for some constant C>0C>0.

Proof.

We prove it for u>4(1+C2)u>4(1+C_{2}) with C2C_{2} as in (2.48). Then by adjusting the constant CC it is true also for u>1u>1. Denote by Γ1\Gamma_{1} and Γ2\Gamma_{2} the geodesics from (0,0)(0,0) to P(τ,v12ε2/3)P(\tau,v-\frac{1}{2}\varepsilon^{2/3}) and from I(u)I(u) to P(τ,v+12ε2/3)P(\tau,v+\frac{1}{2}\varepsilon^{2/3}) respectively. These two points are the end point of the grid interval whose midpoint is P(τ,v)P(\tau,v); see also Figure 5. Define the events

B1\displaystyle B_{1} ={Γ1(2τN)(v1)(2N)2/3},\displaystyle=\{\Gamma_{1}(2\tau N)\leq(v-1)(2N)^{2/3}\}, (2.51)
B2\displaystyle B_{2} ={Γ2(2τN)(v+1)(2N)2/3},\displaystyle=\{\Gamma_{2}(2\tau N)\geq(v+1)(2N)^{2/3}\},
B3\displaystyle B_{3} ={L(0,0),2N4NC2u24/3N1/3},\displaystyle=\{L_{(0,0),\mathcal{L}_{2N}}\leq 4N-C_{2}u2^{4/3}N^{1/3}\},
B4\displaystyle B_{4} ={LI(u),2N4NC2u24/3N1/3}.\displaystyle=\{L_{I(u),\mathcal{L}_{2N}}\leq 4N-C_{2}u2^{4/3}N^{1/3}\}.

Let us show that

A(τ,v)B(τ,v)B1B2B3B4.A(\tau,v)\subseteq B(\tau,v)\cup B_{1}\cup B_{2}\cup B_{3}\cup B_{4}. (2.52)

Observe that

A(τ,v)B1B3B3B4(A(τ,v)B1cB2cB3cB4c)A(\tau,v)\subseteq B_{1}\cup B_{3}\cup B_{3}\cup B_{4}\cup(A(\tau,v)\cap B_{1}^{c}\cap B_{2}^{c}\cap B_{3}^{c}\cap B_{4}^{c}) (2.53)

and the last event is included in B(τ,v)B(\tau,v). Indeed, on A(τ,v)B1cB2cA(\tau,v)\cap B_{1}^{c}\cap B_{2}^{c}, the geodesics ΓN\Gamma^{*}_{N} and Γ~N\tilde{\Gamma}^{*}_{N} must cross 𝒮v{\cal S}_{v}. Now, let γ1\gamma_{1} and γ2\gamma_{2} be the portions of ΓN\Gamma^{*}_{N} and Γ~N\tilde{\Gamma}^{*}_{N} respectively before time τN\tau N. On A(τ,v)A(\tau,v), by definition, γ1\gamma_{1} and γ2\gamma_{2} must be disjoint. On B3cA(τ,v)B_{3}^{c}\cap A(\tau,v) it holds that L(γ1)+L𝒮v,2NL(0,0),2N4NC2u24/3N1/3L(\gamma_{1})+L_{{\cal S}_{v},\mathcal{L}_{2N}}\geq L_{(0,0),\mathcal{L}_{2N}}\geq 4N-C_{2}u2^{4/3}N^{1/3}, and similar inequality holds on B4cA(τ,v)B_{4}^{c}\cap A(\tau,v) replacing γ1\gamma_{1} by γ2\gamma_{2}. Thus the event B(τ,v)B(\tau,v) is satisfied.

To complete the proof, we apply union bound to (2.52) and bound the probabilities (Bi)\operatorname{\mathbbm{P}}(B_{i}). By the choice of C2C_{2}, (B3)\operatorname{\mathbbm{P}}(B_{3}) and (B4)\operatorname{\mathbbm{P}}(B_{4}) are both bounded by e13u3e^{-\frac{1}{3}u^{3}}. Lemma 2.10 below shows that (B1)Ce13u3\operatorname{\mathbbm{P}}(B_{1})\leq Ce^{-\frac{1}{3}u^{3}} for some C>0C>0 and by symmetry (B2)Ce13u3\operatorname{\mathbbm{P}}(B_{2})\leq Ce^{-\frac{1}{3}u^{3}} as well. This completes the proof. ∎

Lemma 2.10.

Assume (122/3)uε2/3<v<22/3u+ε2/3(1-2^{-2/3})u-\varepsilon^{2/3}<v<2^{-2/3}u+\varepsilon^{2/3} and 22/3ετ12^{-2/3}-\varepsilon\leq\tau\leq 1. For N1/14u>1N^{1/14}\gg u>1, there exists a δ>0\delta>0 small enough (not depending on uu and NN) such that with ε=δu2\varepsilon=\delta u^{-2},

(Γ1(2τN)(v1)(2N)2/3)Ce13u3\operatorname{\mathbbm{P}}(\Gamma_{1}(2\tau N)\leq(v-1)(2N)^{2/3})\leq Ce^{-\frac{1}{3}u^{3}} (2.54)

for some C>0C>0.

Proof.

Γ1\Gamma_{1} is the geodesic from (0,0)(0,0) to Q1=P(τ,v12ε2/3)Q_{1}=P(\tau,v-\frac{1}{2}\varepsilon^{2/3}). We want to bound the probability that Γ1(2τN)(v1)(2N)2/3\Gamma_{1}(2\tau N)\leq(v-1)(2N)^{2/3}.

For x1,x2,x3>0x_{1},x_{2},x_{3}>0, define the events

E1\displaystyle E_{1} ={L(0,0),Q14(τ+ε)Nv2τ+ε24/3N1/3x124/3N1/3},\displaystyle=\{L_{(0,0),Q_{1}}\leq 4(\tau+\varepsilon)N-\tfrac{v^{2}}{\tau+\varepsilon}2^{4/3}N^{1/3}-x_{1}2^{4/3}N^{1/3}\}, (2.55)
E2\displaystyle E_{2} ={L(0,0),2τN4τN+x224/3N1/3},\displaystyle=\{L_{(0,0),\mathcal{L}_{2\tau N}}\geq 4\tau N+x_{2}2^{4/3}N^{1/3}\},
E3\displaystyle E_{3} ={supw(v1)(2N)2/3L(τN+w,τNw),Q14εN\displaystyle=\Big{\{}\sup_{w\leq(v-1)(2N)^{2/3}}L_{(\tau N+w,\tau N-w),Q_{1}}\geq 4\varepsilon N
1ε(112ε2/3)224/3N1/3+x324/3N1/3}.\displaystyle\hskip 110.00017pt-\tfrac{1}{\varepsilon}(1-\tfrac{1}{2}\varepsilon^{2/3})^{2}2^{4/3}N^{1/3}+x_{3}2^{4/3}N^{1/3}\Big{\}}.

By a first order approximation, we have L(0,0),Q14(τ+ε)Nv2τ+ε24/3N1/3+𝒪(vε2/3N1/3)L_{(0,0),Q_{1}}\simeq 4(\tau+\varepsilon)N-\frac{v^{2}}{\tau+\varepsilon}2^{4/3}N^{1/3}+\mathcal{O}(v\varepsilon^{2/3}N^{1/3}). So, by Lemma A.1, we have (E1)Cecx13\operatorname{\mathbbm{P}}(E_{1})\leq Ce^{-cx_{1}^{3}} for x1x_{1} of at least the order of uu, and such that x1N2/3x_{1}\ll N^{2/3}. Next, by Lemma A.4, we have (E2)Ce43x23/2τ1/2\operatorname{\mathbbm{P}}(E_{2})\leq Ce^{-\frac{4}{3}x_{2}^{3/2}\tau^{-1/2}} for x2N2/9x_{2}\ll N^{2/9}. Finally, using Lemma A.3 (with the variables (N,u)(N,u) in Lemma A.3 replaced by (εN,ε2/3(112ε2/3))(\varepsilon N,\varepsilon^{-2/3}(1-\tfrac{1}{2}\varepsilon^{2/3})), we get (E3)Ce43x33/2ε1/2\operatorname{\mathbbm{P}}(E_{3})\leq Ce^{-\frac{4}{3}x_{3}^{3/2}\varepsilon^{-1/2}} provided

x3ε1/3(εN)2/9 and N1/7ε.x_{3}\varepsilon^{-1/3}\ll(\varepsilon N)^{2/9}\text{\,\, and \,\, }N^{-1/7}\ll\varepsilon. (2.56)

Under the condition

v2τ+εx1x21ε(112ε2/3)2+x3,-\frac{v^{2}}{\tau+\varepsilon}-x_{1}\geq x_{2}-\frac{1}{\varepsilon}(1-\tfrac{1}{2}\varepsilon^{2/3})^{2}+x_{3}, (2.57)

we have

(Γ1(2τN)(v1)(2N)2/3)(E1)+(E2)+(E3).\operatorname{\mathbbm{P}}(\Gamma_{1}(2\tau N)\leq(v-1)(2N)^{2/3})\leq\operatorname{\mathbbm{P}}(E_{1})+\operatorname{\mathbbm{P}}(E_{2})+\operatorname{\mathbbm{P}}(E_{3}). (2.58)

We assume already that ε\varepsilon is small enough so that τ1/2\tau\geq 1/2. First take x1=u/(3c)x_{1}=u/(3c) so that (E1)Ce13u3\operatorname{\mathbbm{P}}(E_{1})\leq Ce^{-\frac{1}{3}u^{3}}. Then take x2=u2x_{2}=u^{2} which ensures (E2)Ce13u3\operatorname{\mathbbm{P}}(E_{2})\leq Ce^{-\frac{1}{3}u^{3}} as well. Finally we take x3=u2ε1/3x_{3}=u^{2}\varepsilon^{1/3} that gives (E3)Ce13u3\operatorname{\mathbbm{P}}(E_{3})\leq Ce^{-\frac{1}{3}u^{3}}. To satisfy the condition (2.57), it is enough to take ε=δu2\varepsilon=\delta u^{-2} with δ\delta small enough (independent of uu). Finally, note that (2.56) implies that uN114u\ll N^{\tfrac{1}{14}}. ∎

To complete the proof of Theorem 2.7 we need to obtain a bound on the event B(τ,v)B(\tau,v).

Lemma 2.11.

Assume 22/3uε2/3<v<(122/3)u+ε2/32^{-2/3}u-\varepsilon^{2/3}<v<(1-2^{-2/3})u+\varepsilon^{2/3} and 22/3ετ1ε2^{-2/3}-\varepsilon\leq\tau\leq 1-\varepsilon. For N1/9u>4(1+C2)N^{1/9}\gg u>4(1+C_{2}), there exists a constant δ>0\delta>0 small enough such that with ε=δu2\varepsilon=\delta u^{-2},

(B(τ,v))e13u3+cu2\operatorname{\mathbbm{P}}(B(\tau,v))\leq e^{-\frac{1}{3}u^{3}+cu^{2}} (2.59)

for some constant c>0c>0 independent of τ,v\tau,v. The constant C2C_{2} is as in (2.48).

Proof.

For any s1,s2+{}s_{1},s_{2}\in\mathbb{R}_{+}\cup\{-\infty\} we define D(s1,s2)D(s_{1},s_{2}) to be the event that there exist disjoint paths γ1\gamma_{1} and γ2\gamma_{2} as in the definition of B(τ,v)B(\tau,v), such that

L(γ1)\displaystyle L(\gamma_{1}) 4τN(v1)2τ24/3N1/3+s124/3N1/3,\displaystyle\geq 4\tau N-\frac{(v-1)^{2}}{\tau}2^{4/3}N^{1/3}+s_{1}2^{4/3}N^{1/3}, (2.60)
L(γ2)\displaystyle L(\gamma_{2}) 4τN(uv1)2τ24/3N1/3+s224/3N1/3.\displaystyle\geq 4\tau N-\frac{(u-v-1)^{2}}{\tau}2^{4/3}N^{1/3}+s_{2}2^{4/3}N^{1/3}.

For s3+{}s_{3}\in\mathbb{R}_{+}\cup\{-\infty\}, we define the event

C(s3)={L~𝒮v,2N4(1τ)N+s324/3N1/3}.C(s_{3})=\{\tilde{L}_{{\cal S}_{v},\mathcal{L}_{2N}}\geq 4(1-\tau)N+s_{3}2^{4/3}N^{1/3}\}. (2.61)

Recall the constant C2C_{2} from (2.48). Like in the proof of Lemma 2.6 we do a discretization with a fixed width 0<η<10<\eta<1 and thus we will not write all the details. The minor difference is that now we a couple of constraints:

s1+s3=1τ(v1)2ηC2u,s2+s3=1τ(uv1)2ηC2u.s_{1}+s_{3}=\tfrac{1}{\tau}(v-1)^{2}-\eta-C_{2}u,\quad s_{2}+s_{3}=\tfrac{1}{\tau}(u-v-1)^{2}-\eta-C_{2}u. (2.62)

In the discretization of Lemma 2.6, see (2.28), we separated explicitly two terms, which corresponds taking S2=S_{2}=-\infty and S3=S_{3}=-\infty. Here we do the same, but instead of writing those terms separately, we consider subsets allowing positive numbers and -\infty. More precisely, define the set

Θ={\displaystyle\Theta=\{ s1,s2,s3+{}|s3η,s10+s30=1τ(v1)2ηC2u\displaystyle s_{1},s_{2},s_{3}\in\mathbb{R}_{+}\cup\{-\infty\}|\,s_{3}\in\eta\mathbb{Z},s_{1}\vee 0+s_{3}\vee 0=\tfrac{1}{\tau}(v-1)^{2}-\eta-C_{2}u (2.63)
and s20+s30=1τ(uv1)2ηC2u}.\displaystyle\textrm{and }s_{2}\vee 0+s_{3}\vee 0=\tfrac{1}{\tau}(u-v-1)^{2}-\eta-C_{2}u\}.

Then

B(τ,v)s1,s2,s3ΘC(s3)D(s1,s2).B(\tau,v)\subset\bigcup_{s_{1},s_{2},s_{3}\in\Theta}C(s_{3})\cap D(s_{1},s_{2}). (2.64)

The number of elements is, for any vv with min{v,uv}u(122/3)+ε2/3\min\{v,u-v\}\leq u(1-2^{2/3})+\varepsilon^{2/3} of order u2/τu^{2}/\tau. Since τ1/2\tau\geq 1/2 (for ε22/31/2\varepsilon\leq 2^{-2/3}-1/2), the sum contains 𝒪(u2)\mathcal{O}(u^{2}) many terms. Therefore, using the independence of C(s3)C(s_{3}) and D(s1,s2)D(s_{1},s_{2}),

(B(τ,v))Cu2maxs1,s2,s3Θ(C(s3))(D(s1,s2)).\operatorname{\mathbbm{P}}(B(\tau,v))\leq Cu^{2}\max_{s_{1},s_{2},s_{3}\in\Theta}\operatorname{\mathbbm{P}}(C(s_{3}))\operatorname{\mathbbm{P}}(D(s_{1},s_{2})). (2.65)

As γ1\gamma_{1} and γ2\gamma_{2} “occur disjointly”, by the BK (Berg-Kesten) inequality (see e.g. Theorem 7 of [3] for a statement applicable in the above scenario) we get

(D(s1,s2))\displaystyle\operatorname{\mathbbm{P}}(D(s_{1},s_{2})) (L(0,0),𝒮v4τN1τ(v1)224/3N1/3+s124/3N1/3)\displaystyle\leq\operatorname{\mathbbm{P}}(L_{(0,0),{\cal S}_{v}}\geq 4\tau N-\tfrac{1}{\tau}(v-1)^{2}2^{4/3}N^{1/3}+s_{1}2^{4/3}N^{1/3}) (2.66)
×(LI(u),𝒮v4τN1τ(uv1)224/3N1/3+s224/3N1/3).\displaystyle\times\operatorname{\mathbbm{P}}(L_{I(u),{\cal S}_{v}}\geq 4\tau N-\tfrac{1}{\tau}(u-v-1)^{2}2^{4/3}N^{1/3}+s_{2}2^{4/3}N^{1/3}).

Set 𝒟τ,v={(τN+k,τNk)|k(v1)(2N)2/3}{\cal D}_{\tau,v}=\{(\tau N+k,\tau N-k)|k\geq(v-1)(2N)^{2/3}\}. Then L(0,0),𝒮vL(0,0),𝒟τ,vL_{(0,0),{\cal S}_{v}}\leq L_{(0,0),{\cal D}_{\tau,v}} and Lemma A.3 (after rescaling NτNN\to\tau N) leads to

(L(0,0),𝒮v4τN1τ(v1)224/3N1/3+s124/3N1/3)Ce43s13/2τ1/2\operatorname{\mathbbm{P}}(L_{(0,0),{\cal S}_{v}}\geq 4\tau N-\tfrac{1}{\tau}(v-1)^{2}2^{4/3}N^{1/3}+s_{1}2^{4/3}N^{1/3})\leq Ce^{-\frac{4}{3}\frac{s_{1}^{3/2}}{\tau^{1/2}}} (2.67)

and similarly for the bound on the second term in (2.66), so that we have

(D(s1,s2))Ce43s13/2+s23/2τ1/2,\operatorname{\mathbbm{P}}(D(s_{1},s_{2}))\leq Ce^{-\frac{4}{3}\frac{s_{1}^{3/2}+s_{2}^{3/2}}{\tau^{1/2}}}, (2.68)

for 2<uN1/92<u\ll N^{1/9}, 0<s1,s2N2/90<s_{1},s_{2}\ll N^{2/9}.

To estimate (C(s3))\operatorname{\mathbbm{P}}(C(s_{3})) we divide the segment 𝒮v{\cal S}_{v} into pieces of length (1τ)2/3(2N)2/3(1-\tau)^{2/3}(2N)^{2/3} to which we can apply a rescaled version of Proposition A.5. We have 2/(1τ)2/32/ε2/32/(1-\tau)^{2/3}\leq 2/\varepsilon^{2/3} such pieces (we used 1τε1-\tau\geq\varepsilon). Using union bound we then get, for s3(1τ)1/3N2/3s_{3}(1-\tau)^{-1/3}\ll N^{2/3},

(C(s3))Cmax{1,s3(1τ)1/3}ε2/3e43s33/2(1τ)1/2.\operatorname{\mathbbm{P}}(C(s_{3}))\leq\frac{C\max\{1,s_{3}(1-\tau)^{-1/3}\}}{\varepsilon^{2/3}}e^{-\frac{4}{3}\frac{s_{3}^{3/2}}{(1-\tau)^{1/2}}}. (2.69)

Let ε=δu2\varepsilon=\delta u^{-2} for δ>0\delta>0 small enough as in the proof of Lemma 2.10. We have

(B(τ,v))Cu4δ1maxs1,s2,s3Θe43s13/2+s23/2τ1/2e43s33/2(1τ)1/2.\operatorname{\mathbbm{P}}(B(\tau,v))\leq Cu^{4}\delta^{-1}\max_{s_{1},s_{2},s_{3}\in\Theta}e^{-\frac{4}{3}\frac{s_{1}^{3/2}+s_{2}^{3/2}}{\tau^{1/2}}}e^{-\frac{4}{3}\frac{s_{3}^{3/2}}{(1-\tau)^{1/2}}}. (2.70)

Therefore we concentrate now on finding the maximum of

e43s13/2+s23/2τ1/2e43s33/2(1τ)1/2e^{-\frac{4}{3}\frac{s_{1}^{3/2}+s_{2}^{3/2}}{\tau^{1/2}}}e^{-\frac{4}{3}\frac{s_{3}^{3/2}}{(1-\tau)^{1/2}}} (2.71)

for s1,s2,s3Θs_{1},s_{2},s_{3}\in\Theta. Define s~3=s3+η+C2u\tilde{s}_{3}=s_{3}+\eta+C_{2}u. Then for given s3s_{3} on Θ\Theta, we have

s1=(v1)2τs~3,s2=(uv1)2τs~3.s_{1}=\frac{(v-1)^{2}}{\tau}-\tilde{s}_{3},\quad s_{2}=\frac{(u-v-1)^{2}}{\tau}-\tilde{s}_{3}. (2.72)

So we need to maximize

M(v,τ,s3)=43((v1)2/τs~3)3/2+((uv1)2/τs~3)3/2τ43s33/21τ.M(v,\tau,s_{3})=-\frac{4}{3}\frac{((v-1)^{2}/\tau-\tilde{s}_{3})^{3/2}+((u-v-1)^{2}/\tau-\tilde{s}_{3})^{3/2}}{\sqrt{\tau}}-\frac{4}{3}\frac{s_{3}^{3/2}}{\sqrt{1-\tau}}. (2.73)

In principle, to get the bound on B(τ,v)B(\tau,v), we would need to find s3s_{3} maximizing M(v,τ,s3)M(v,\tau,s_{3}). In the statement we want a bound uniform in τ,v\tau,v. This means that we need to maximize the result over τ,v\tau,v as well. In short, we maximize MM for s3,v,τs_{3},v,\tau and thus we do it in another order. First notice that for given τ,s3\tau,s_{3}, M(v,τ,s3)M(v,\tau,s_{3}) is maximized at v=u/2v=u/2, for which

M(u/2,τ,s3)=432((u2)2/(4τ)s~3)3/2τ43s33/21τ.M(u/2,\tau,s_{3})=-\frac{4}{3}\frac{2((u-2)^{2}/(4\tau)-\tilde{s}_{3})^{3/2}}{\sqrt{\tau}}-\frac{4}{3}\frac{s_{3}^{3/2}}{\sqrt{1-\tau}}. (2.74)

Computing the derivative with respect to s3s_{3} we get that, for a given τ\tau, the maximum is at s3=[(u2)24τ(η+C2u)](1τ)τ(43τ)>0s_{3}^{*}=\frac{[(u-2)^{2}-4\tau(\eta+C_{2}u)](1-\tau)}{\tau(4-3\tau)}>0 under the assumption u4(1+C2)u\geq 4(1+C_{2}) and η<1\eta<1. So we get

M(u/2,τ,s3)\displaystyle M(u/2,\tau,s_{3}^{*}) =43[(u2)24τ(η+C2u)]3/24τ3/243τ\displaystyle=-\frac{4}{3}\frac{[(u-2)^{2}-4\tau(\eta+C_{2}u)]^{3/2}}{4\tau^{3/2}\sqrt{4-3\tau}} (2.75)
13[(u2)24τ(η+C2u)]3/213u3+cu2\displaystyle\leq-\frac{1}{3}[(u-2)^{2}-4\tau(\eta+C_{2}u)]^{3/2}\leq-\frac{1}{3}u^{3}+cu^{2}

for some constant c>0c>0. Inserting (2.75) into (2.70) and choosing an appropriate new constant cc leads to the claimed result. ∎

2.4 Proof of Theorem 2.1

As before, we shall prove the bound first for sufficiently large uu, and adjust cc later to deduce the same for all u>1u>1.

Recall that LN(u)L_{N}^{*}(u) is the rescaled LPP from I(u)I(u) to 2N\mathcal{L}_{2N}, see (2.4). Let us use the notations X=LN(0)X=L_{N}^{*}(0) and Y=LN(u)Y=L_{N}^{*}(u). For j,j{1,2,,u1}j,j^{\prime}\in\{1,2,\ldots,u-1\}, let SNjS^{j}_{N} denote the weight of the maximum weight path πj\pi^{j} from (0,0)(0,0) to 2n\mathcal{L}_{2n} such that π(t)<j(2N)2/3\pi(t)<j(2N)^{2/3} for all tt, and similarly, let SNu,jS^{u,j^{\prime}}_{N} denote the weight of the maximum weight path π~j\tilde{\pi}^{j^{\prime}} from I(u)I(u) to 2n\mathcal{L}_{2n} such that π~j(t)>(uj)(2N)2/3\tilde{\pi}^{j^{\prime}}(t)>(u-j^{\prime})(2N)^{2/3} for all tt. Let us also set

Xj=24/3N1/3(SNj4N),Yj=24/3N1/3(SNu,j4N).X_{j}=2^{-4/3}N^{-1/3}(S^{j}_{N}-4N),\quad Y_{j^{\prime}}=2^{-4/3}N^{-1/3}(S_{N}^{u,j^{\prime}}-4N). (2.76)

For notational convenience, we shall also write X0=Y0=0X_{0}=Y_{0}=0 and Xu=XX_{u}=X, Yu=YY_{u}=Y. Now, writing

X=j=0u1Xj+1Xj,Y=j=0u1Yj+1Yj,X=\sum_{j=0}^{u-1}X_{j+1}-X_{j},\quad Y=\sum_{j^{\prime}=0}^{u-1}Y_{j^{\prime}+1}-Y_{j^{\prime}}, (2.77)

and using the bilinearity of covariance it is enough to prove that for some c>0c>0 and for all j,j{0,1,,u1}j,j^{\prime}\in\{0,1,\ldots,u-1\}

Cov(Xj+1Xj,Yj+1Yj)ecu2e13u3.{\rm Cov}(X_{j+1}-X_{j},Y_{j^{\prime}+1}-Y_{j^{\prime}})\leq e^{cu^{2}}e^{-\frac{1}{3}u^{3}}. (2.78)

Notice first that Xj+1XjX_{j+1}-X_{j} and Yj+1YjY_{j^{\prime}+1}-Y_{j^{\prime}} depend on disjoint sets of vertex weights and hence are independent unless j+ju1j+j^{\prime}\geq u-1. Hence we only need to consider (j,j)(j,j^{\prime}) such that j+ju1j+j^{\prime}\geq u-1. For such a pair, noticing XXj+1XjX\geq X_{j+1}\geq X_{j} and YYj+1YjY\geq Y_{j+1}\geq Y_{j} it follows that

Cov(Xj+1Xj,Yj+1Yj)𝔼[(XXj)(YYj)].{\rm Cov}(X_{j+1}-X_{j},Y_{j^{\prime}+1}-Y_{j^{\prime}})\leq\mathbbm{E}[(X-X_{j})(Y-Y_{j^{\prime}})]. (2.79)

For convenience of notation, let Γ1\Gamma_{1} and Γ2\Gamma_{2} locally denote the geodesics from (0,0)(0,0) and I(u)I(u) respectively to 2N\mathcal{L}_{2N}. We define

Aj={sup0t2NΓ1(t)j(2N)2/3},Bj={inf0t2NΓ2(t)(uj)(2N)2/3}.A_{j}=\Big{\{}\sup_{0\leq t\leq 2N}\Gamma_{1}(t)\geq j(2N)^{2/3}\Big{\}},\quad B_{j^{\prime}}=\Big{\{}\inf_{0\leq t\leq 2N}\Gamma_{2}(t)\leq(u-j^{\prime})(2N)^{2/3}\Big{\}}. (2.80)

Clearly, (XXj)(YYj)=0(X-X_{j})(Y-Y_{j^{\prime}})=0 on the complement of AjBjA_{j}\cap B_{j^{\prime}} and XXjX-X_{j} and YYjY-Y_{j^{\prime}} are positive random variable with super-exponential (uniform in j,jj,j^{\prime}) tails (indeed we can just use the upper tail bounds for XX and YY). Using the notation Xp=𝔼(|X|p)1/p\|X\|_{p}=\mathbbm{E}(|X|^{p})^{1/p} and the fact that the pp-th norm of the of random variables with super-exponential tails can grow at most linearly in pp, we know that there exists a constant CC such that for all j,jj,j^{\prime} and all p1p\geq 1 XXjp,YYjpCp.||X-X_{j}||_{p},||Y-Y_{j^{\prime}}||_{p}\leq Cp. Using the Hölder inequality we have

𝔼[(XXj)(YYj)]𝟙AjBjq(XXj)(YYj)p\mathbbm{E}[(X-X_{j})(Y-Y_{j^{\prime}})]\leq||\mathbbm{1}_{A_{j}\cap B_{j^{\prime}}}||_{q}||(X-X_{j})(Y-Y_{j^{\prime}})||_{p} (2.81)

where p1+q1=1p^{-1}+q^{-1}=1. By the Cauchy-Schwarz inequality

(XXj)(YYj)pXXj2pYYj2pCp2||(X-X_{j})(Y-Y_{j^{\prime}})||_{p}\leq||X-X_{j}||_{2p}||Y-Y_{j^{\prime}}||_{2p}\leq Cp^{2} (2.82)

for some new constant C>0C>0. It therefore follows that

𝔼[(XXj)(YYj)]Cp2(AjBj)1/q.\mathbbm{E}[(X-X_{j})(Y-Y_{j^{\prime}})]\leq C^{\prime}p^{2}\operatorname{\mathbbm{P}}(A_{j}\cap B_{j^{\prime}})^{1/q}. (2.83)

for p,q1p,q\geq 1 with p1+q1=1p^{-1}+q^{-1}=1.

By Lemma 2.12 below, we have

(AjBj)e13u3+cu2\operatorname{\mathbbm{P}}(A_{j}\cap B_{j^{\prime}})\leq e^{-\frac{1}{3}u^{3}+cu^{2}} (2.84)

for some c>0c>0. We choose p=up=u so that 1/q=11u1/q=1-\frac{1}{u}. Therefore, plugging (2.84) in (2.83) it follows that

𝔼[(XXj)(YYj)]Cu2e(cu2u3/3)(11u)e13u3+cu2\mathbbm{E}[(X-X_{j})(Y-Y_{j^{\prime}})]\leq Cu^{2}e^{(cu^{2}-u^{3}/3)(1-\frac{1}{u})}\leq e^{-\frac{1}{3}u^{3}+c^{\prime}u^{2}} (2.85)

for some new constant cc^{\prime}. This establishes (2.78) and Theorem 2.1 follows by summing over (j,j)(j,j^{\prime}). ∎

Lemma 2.12.

In the above set-up, for uu large enough and j+ju1j+j^{\prime}\geq u-1 we have

(AjBj)e13u3+cu2\operatorname{\mathbbm{P}}(A_{j}\cap B_{j^{\prime}})\leq e^{-\frac{1}{3}u^{3}+cu^{2}} (2.86)

for some c>0c>0.

Proof.

Notice first that arguing as in the proof of Lemma 2.8, if j22/3uj\geq 2^{-2/3}u then (Aj)e13u3+cu2\operatorname{\mathbbm{P}}(A_{j})\leq e^{-\frac{1}{3}u^{3}+cu^{2}} and similarly if j22/3uj^{\prime}\geq 2^{-2/3}u then (Bj)e13u3+cu2\operatorname{\mathbbm{P}}(B_{j^{\prime}})\leq e^{-\frac{1}{3}u^{3}+cu^{2}}. Therefore it suffices to consider only the cases max{j,j}22/3u\max\{j,j^{\prime}\}\leq 2^{-2/3}u. This, together with j+ju1j+j^{\prime}\geq u-1 also implies that min{j,j}(122/3)u1>0\min\{j,j^{\prime}\}\geq(1-2^{-2/3})u-1>0 for uu sufficiently large.

Observe now that

(AjBj)(Γ1Γ2)+(AjBj{Γ1Γ2=}).\operatorname{\mathbbm{P}}(A_{j}\cap B_{j^{\prime}})\leq\operatorname{\mathbbm{P}}(\Gamma_{1}\cap\Gamma_{2}\neq\emptyset)+\operatorname{\mathbbm{P}}(A_{j}\cap B_{j^{\prime}}\cap\{\Gamma_{1}\cap\Gamma_{2}=\emptyset\}). (2.87)

By Theorem 2.7 it follows that the first term is upper bounded by e13u3+cu2e^{-\frac{1}{3}u^{3}+cu^{2}} and hence it suffices to show that

(AjBj{Γ1Γ2=})e13u3+cu2\operatorname{\mathbbm{P}}(A_{j}\cap B_{j^{\prime}}\cap\{\Gamma_{1}\cap\Gamma_{2}=\emptyset\})\leq e^{-\frac{1}{3}u^{3}+cu^{2}} (2.88)

for some c>0c>0.

Since the two geodesics do not intersect, we would like to use the BK inequality to get an upper bound. However, we can not do it directly, since the property of being a geodesic depends on all the random variables and not on subsets. We therefore would show something similar for any paths which are of typical length. First, we need to approximate the events AjA_{j} and BjB_{j^{\prime}}.

For ε>0\varepsilon>0 to be chosen later, let A~j\tilde{A}_{j} denote the event that there exists k{1,2,,1ε}k\in\{1,2,\ldots,\frac{1}{\varepsilon}\} such that Γ1(2kεN)(j1)(2N)2/3\Gamma_{1}(2k\varepsilon N)\geq(j-1)(2N)^{2/3}. Similarly, let B~j\tilde{B}_{j^{\prime}} denote the event that there exists k{1,2,,1ε}k\in\{1,2,\ldots,\frac{1}{\varepsilon}\} such that Γ2(2kεN)(uj+1)(2N)2/3\Gamma_{2}(2k\varepsilon N)\leq(u-j^{\prime}+1)(2N)^{2/3}. By choosing ε=δu3/2\varepsilon=\delta u^{-3/2} for δ\delta sufficiently small but fixed and arguing as in the proof of Theorem 2.3 it follows that

(AjA~j),(BjB~j)e13u3+cu2.\operatorname{\mathbbm{P}}(A_{j}\setminus\tilde{A}_{j}),\operatorname{\mathbbm{P}}(B_{j^{\prime}}\setminus\tilde{B}_{j^{\prime}})\leq e^{-\frac{1}{3}u^{3}+cu^{2}}. (2.89)

Therefore (2.88) reduces to showing

(A~jB~j{Γ1Γ2=})e13u3+cu2\operatorname{\mathbbm{P}}(\tilde{A}_{j}\cap\tilde{B}_{j^{\prime}}\cap\{\Gamma_{1}\cap\Gamma_{2}=\emptyset\})\leq e^{-\frac{1}{3}u^{3}+cu^{2}} (2.90)

for some c>0c>0.

Let C2C_{2} be such that (using Lemma A.1)

(XC2u)=(YC2u)e13u3.\operatorname{\mathbbm{P}}(X\leq-C_{2}u)=\operatorname{\mathbbm{P}}(Y\leq-C_{2}u)\leq e^{-\frac{1}{3}u^{3}}. (2.91)

Observe that on A~jB~j{Γ1Γ2=}{X>C2u}{Y>C2u}\tilde{A}_{j}\cap\tilde{B}_{j^{\prime}}\cap\{\Gamma_{1}\cap\Gamma_{2}=\emptyset\}\cap\{X>-C_{2}u\}\cap\{Y>-C_{2}u\} there exist disjoint paths γ1\gamma_{1} and γ2\gamma_{2} from (0,0)(0,0) and I(u)I(u) respectively to 2N\mathcal{L}_{2N} with L(γ1),L(γ2)4NC2u24/3N1/3L(\gamma_{1}),L(\gamma_{2})\geq 4N-C_{2}u2^{4/3}N^{1/3} such that there exist k1,k2{1,2,,1ε}k_{1},k_{2}\in\{1,2,\ldots,\frac{1}{\varepsilon}\} with γ1(2k1εN)(j1)(2N)2/3\gamma_{1}(2k_{1}\varepsilon N)\geq(j-1)(2N)^{2/3} and γ2(2k2εN)(uj+1)(2N)2/3\gamma_{2}(2k_{2}\varepsilon N)\leq(u-j^{\prime}+1)(2N)^{2/3}. By using the BK inequality as before we get that

(A~jB~j{Γ1Γ2=})(A^j)(B^j)+2e13u3\operatorname{\mathbbm{P}}(\tilde{A}_{j}\cap\tilde{B}_{j^{\prime}}\cap\{\Gamma_{1}\cap\Gamma_{2}=\emptyset\})\leq\operatorname{\mathbbm{P}}(\hat{A}_{j})\operatorname{\mathbbm{P}}(\hat{B}_{j^{\prime}})+2e^{-\frac{1}{3}u^{3}} (2.92)

where A^j\hat{A}_{j} denotes the event that there exists a path γ1\gamma_{1} from (0,0)(0,0) to 2N\mathcal{L}_{2N} satisfying L(γ1)>4NC2u24/3N1/3L(\gamma_{1})>4N-C_{2}u2^{4/3}N^{1/3} and γ1(2kεN)(j1)(2N)2/3\gamma_{1}(2k\varepsilon N)\geq(j-1)(2N)^{2/3} for some kk and B^j\hat{B}_{j^{\prime}} denotes the event that there exists a path γ2\gamma_{2} from I(u)I(u) to 2N\mathcal{L}_{2N} satisfying L(γ2)>4NC2u24/3N1/3L(\gamma_{2})>4N-C_{2}u2^{4/3}N^{1/3} and γ2(2kεN)(uj1)(2N)2/3\gamma_{2}(2k\varepsilon N)\leq(u-j-1)(2N)^{2/3} for some kk. We claim that

(A^j)ecu2e43(j1)3,(B^j)ecu2e43(j1)3.\operatorname{\mathbbm{P}}(\hat{A}_{j})\leq e^{cu^{2}}e^{-\frac{4}{3}(j-1)^{3}},\quad\operatorname{\mathbbm{P}}(\hat{B}_{j})\leq e^{cu^{2}}e^{-\frac{4}{3}(j^{\prime}-1)^{3}}. (2.93)

Postponing the proof of (2.93) for now, let us first complete the proof of the lemma. By Jensen’s inequality together with j+ju1j+j^{\prime}\geq u-1 it follows that

(j1)3+(j1)314(j+j2)3(u3)34(j-1)^{3}+(j^{\prime}-1)^{3}\geq\frac{1}{4}(j+j^{\prime}-2)^{3}\geq\frac{(u-3)^{3}}{4} (2.94)

and hence

(A^j)(B^j)e13u3+cu2\operatorname{\mathbbm{P}}(\hat{A}_{j})\operatorname{\mathbbm{P}}(\hat{B}_{j^{\prime}})\leq e^{-\frac{1}{3}u^{3}+cu^{2}} (2.95)

for some c>0c>0. This, together with (2.92) establishes (2.90).

To conclude the proof we show (2.93). The idea is to follow the proof of Lemma 2.6. However the first bound (2.19) in that proof applies only to geodesics, while here we have to show it for any paths with a length larger that 4NC2u24/3N1/34N-C_{2}u2^{4/3}N^{1/3}. We will prove that for any path γ1\gamma_{1} satisfying the conditions of A^j\hat{A}_{j}, for any τ{ε,2ε,,1}\tau\in\{\varepsilon,2\varepsilon,\ldots,1\}

(γ1(2τN)M(2N)1/3)Cu2e43u3\operatorname{\mathbbm{P}}(\gamma_{1}(2\tau N)\geq M(2N)^{1/3})\leq Cu^{2}e^{-\frac{4}{3}u^{3}} (2.96)

for M=C2u+u2M=\sqrt{C_{2}u+u^{2}}. Then the rest of the proof of Lemma 2.6 applies, except that the sum in (2.24) goes until Mu1M-u-1.

Denote K(v)=(τN,τN)+v(2N)2/3(1,1)K(v)=(\tau N,\tau N)+v(2N)^{2/3}(1,-1) and divide the possible points where γ1\gamma_{1} crosses the line 2τN\mathcal{L}_{2\tau N} as

{K(v),v[M,τN(2N)2/3]}=0=MN1/101,\{K(v),v\in[M,\tfrac{\tau N}{(2N)^{2/3}}]\}={\cal I}_{0}\bigcup_{\ell=M}^{N^{1/10}-1}{\cal I}_{\ell}, (2.97)

with 0={K(v),v[N1/10,τN(2N)2/3]}{\cal I}_{0}=\{K(v),v\in[N^{1/10},\tfrac{\tau N}{(2N)^{2/3}}]\} and ={K(v),v[,+1)}{\cal I}_{\ell}=\{K(v),v\in[\ell,\ell+1)\}. Then we have, for any choice of AA_{\ell} and B=4NC2u24/3N1/3AB_{\ell}=4N-C_{2}u2^{4/3}N^{1/3}-A_{\ell},

(γ1(2τN)M(2N)1/3)\displaystyle\operatorname{\mathbbm{P}}(\gamma_{1}(2\tau N)\geq M(2N)^{1/3}) (L(0,0),+L,2N4NC2u24/3N1/3)\displaystyle\leq\sum_{\ell}\operatorname{\mathbbm{P}}\left(L_{(0,0),{\cal I}_{\ell}}+L_{{\cal I}_{\ell},\mathcal{L}_{2N}}\geq 4N-C_{2}u2^{4/3}N^{1/3}\right) (2.98)
(L(0,0),A)+(L(0,0),B).\displaystyle\leq\sum_{\ell}\operatorname{\mathbbm{P}}\left(L_{(0,0),{\cal I}_{\ell}}\geq A_{\ell}\right)+\operatorname{\mathbbm{P}}\left(L_{(0,0),{\cal I}_{\ell}}\geq B_{\ell}\right).

By rescaling Lemma A.3, for s1(τN)2/9s_{1}\ll(\tau N)^{2/9} and u(τN)1/9u\ll(\tau N)^{1/9}

(L(0,0),>4τN2τ24/3N1/3+s1τ1/324/3N1/3)\displaystyle\operatorname{\mathbbm{P}}\Big{(}L_{(0,0),{\cal I}_{\ell}}>4\tau N-\frac{\ell^{2}}{\tau}2^{4/3}N^{1/3}+s_{1}\tau^{1/3}2^{4/3}N^{1/3}\Big{)} Ce43s13/2,\displaystyle\leq Ce^{-\frac{4}{3}s_{1}^{3/2}}, (2.99)
(L(0,0),0>4τNN1/5τ24/3N1/3+s1τ1/324/3N1/3)\displaystyle\operatorname{\mathbbm{P}}\Big{(}L_{(0,0),{\cal I}_{0}}>4\tau N-\frac{N^{1/5}}{\tau}2^{4/3}N^{1/3}+s_{1}\tau^{1/3}2^{4/3}N^{1/3}\Big{)} Ce43s13/2,\displaystyle\leq Ce^{-\frac{4}{3}s_{1}^{3/2}},

and by rescaling Proposition A.5 (and union bound on (1τ)2/3(1-\tau)^{-2/3} subsegments per (2N)2/3(2N)^{2/3}-length), for s2(1τ)2/3N2/3s_{2}\ll(1-\tau)^{2/3}N^{2/3}, we get

(L,2N>4(1τ)N+s2(1τ)1/324/3N1/3)\displaystyle\operatorname{\mathbbm{P}}\Big{(}L_{{\cal I}_{\ell},\mathcal{L}_{2N}}>4(1-\tau)N+s_{2}(1-\tau)^{1/3}2^{4/3}N^{1/3}\Big{)} Cs2(1τ)2/3e43s23/2,\displaystyle\leq Cs_{2}(1-\tau)^{-2/3}e^{-\frac{4}{3}s_{2}^{3/2}}, (2.100)
(L0,2N>4(1τ)N+s2(1τ)1/324/3N1/3)\displaystyle\operatorname{\mathbbm{P}}\Big{(}L_{{\cal I}_{0},\mathcal{L}_{2N}}>4(1-\tau)N+s_{2}(1-\tau)^{1/3}2^{4/3}N^{1/3}\Big{)} Cs2N1/3(1τ)2/3e43s23/2.\displaystyle\leq Cs_{2}N^{1/3}(1-\tau)^{-2/3}e^{-\frac{4}{3}s_{2}^{3/2}}.

We take, with α=2(1τ)1/3+C2uτ4/3((1τ)1/3+τ1/3)τ\alpha_{\ell}=-\frac{\ell^{2}(1-\tau)^{1/3}+C_{2}u\tau^{4/3}}{((1-\tau)^{1/3}+\tau^{1/3})\tau}, A=4τN+α24/3n1/3A_{\ell}=4\tau N+\alpha_{\ell}2^{4/3}n^{1/3}. Setting =M+~\ell=M+\tilde{\ell}, we get

s1=s2=2C2uτ((1τ)1/3+τ1/3)τu2+~2.s_{1}=s_{2}=\frac{\ell^{2}-C_{2}u\tau}{((1-\tau)^{1/3}+\tau^{1/3})\tau}\geq u^{2}+\tilde{\ell}^{2}. (2.101)

From this, it follows that

(L(0,0),A)+(L(0,0),B)2Cδ2/3u2e43u3e43(M)3/2\operatorname{\mathbbm{P}}\left(L_{(0,0),{\cal I}_{\ell}}\geq A_{\ell}\right)+\operatorname{\mathbbm{P}}\left(L_{(0,0),{\cal I}_{\ell}}\geq B_{\ell}\right)\leq 2C\delta^{-2/3}u\ell^{2}e^{-\frac{4}{3}u^{3}}e^{-\frac{4}{3}(\ell-M)^{3/2}} (2.102)

and thus =MN1/101(2.102)Cδ2/3u2e43u3\sum_{\ell=M}^{N^{1/10}-1}\eqref{eq2.100}\leq C^{\prime}\delta^{-2/3}u^{2}e^{-\frac{4}{3}u^{3}}, while for =0\ell=0 it is of an order 𝒪(e43N3/20)\mathcal{O}(e^{-\frac{4}{3}N^{3/20}}) smaller. Applying union bound on the ε1=u3/2/δ\varepsilon^{-1}=u^{3/2}/\delta time intervals and the estimate (2.96), we get that any path in A^j\hat{A}_{j} is localized within a distance M(2N)1/3M(2N)^{1/3} with probability at least 1Ceu31-Ce^{-u^{3}}. ∎

3 Lower Bound

In this section we prove the lower bound of Theorem 1.1.

Theorem 3.1.

There exist a constant c>0c>0 such that

Cov(𝒜1(0),𝒜1(u))eculn(u)e43u3.\operatorname*{Cov}({\cal A}_{1}(0),{\cal A}_{1}(u))\geq e^{-cu\ln(u)}e^{-\frac{4}{3}u^{3}}. (3.1)

We begin explaining the strategy of the proof. Hoeffding’s covariance identity [37], which comes from integration by parts on +\mathbb{R}_{+} and \mathbb{R}_{-} separately, states that

Cov(X,Y)=𝑑s1𝑑s2[(Xs1,Ys2)(Xs1)(Ys2)].\operatorname*{Cov}(X,Y)=\int_{\mathbb{R}}ds_{1}\int_{\mathbb{R}}ds_{2}\left[\operatorname{\mathbbm{P}}(X\leq s_{1},Y\leq s_{2})-\operatorname{\mathbbm{P}}(X\leq s_{1})\operatorname{\mathbbm{P}}(Y\leq s_{2})\right]. (3.2)

We therefore define the following functions

F(u;s1,s2)\displaystyle F(u;s_{1},s_{2}) =(𝒜1(0)s1,𝒜1(u)s2),\displaystyle=\operatorname{\mathbbm{P}}\left({\cal A}_{1}(0)\leq s_{1},{\cal A}_{1}(u)\leq s_{2}\right), (3.3)
f(s)\displaystyle f(s) =(𝒜1(0)s)=(𝒜1(u)s),\displaystyle=\operatorname{\mathbbm{P}}\left({\cal A}_{1}(0)\leq s\right)=\operatorname{\mathbbm{P}}\left({\cal A}_{1}(u)\leq s\right),

where in the last equality we used the stationarity of 𝒜1{\cal A}_{1}. As we would like to use (3.2) with X=𝒜1(0)X={\cal A}_{1}(0) and Y=𝒜1(u)Y={\cal A}_{1}(u), we are interested in finding the asymptotic behavior of

F(u;s1,s2)f(s1)f(s2) as u.F(u;s_{1},s_{2})-f(s_{1})f(s_{2})\quad\text{ as }u\rightarrow\infty. (3.4)

As uu\to\infty, the random variables 𝒜1(0){\cal A}_{1}(0) and 𝒜1(u){\cal A}_{1}(u) become independent of each other; thus we define \cal E through

F(u;s1,s2)=f(s1)f(s2)(1+(u;s1,s2)),F(u;s_{1},s_{2})=f(s_{1})f(s_{2})(1+{\cal E}(u;s_{1},s_{2})), (3.5)

where 0{\cal E}\rightarrow 0 when uu\rightarrow\infty (at least for s1s_{1} and s2s_{2} independent of uu). Using (3.5) in (3.4) and (3.2) we obtain

Cov(𝒜1(0),𝒜1(u))=𝑑s1𝑑s2f(s1)f(s2)(u;s1,s2)\text{Cov}({\cal A}_{1}(0),{\cal A}_{1}(u))=\int_{\mathbb{R}}ds_{1}\int_{\mathbb{R}}ds_{2}f(s_{1})f(s_{2}){\cal E}(u;s_{1},s_{2}) (3.6)

Next, by the FKG inequality, see Lemma 3.2 below, the integrand in (3.6) is positive for all u0u\geq 0. We can therefore restrict the integration in (3.6) to a compact subset of 2\mathbb{R}^{2} to obtain the following lower bound

Cov(𝒜1(0),𝒜1(u))αβ𝑑s1αβ𝑑s2f(s1)f(s2)(u;s1,s2)\text{Cov}({\cal A}_{1}(0),{\cal A}_{1}(u))\geq\int_{\alpha}^{\beta}ds_{1}\int_{\alpha}^{\beta}ds_{2}f(s_{1})f(s_{2}){\cal E}(u;s_{1},s_{2}) (3.7)

for any choice of α<β\alpha<\beta.

Thus the goal of the computations below is to show that \cal E is of order e43u3e^{-\frac{4}{3}u^{3}} times a subleading term for s1,s2s_{1},s_{2} in some chosen intervals of size 𝒪(1)\mathcal{O}(1) where f(s1)f(s_{1}) and f(s2)f(s_{2}) are bounded away from 0.

Lemma 3.2.

For any s1,s2s_{1},s_{2}\in\mathbb{R},

(𝒜1(0)s1,𝒜1(u)s2)(𝒜1(0)s1)(𝒜1(u)s2)0.\operatorname{\mathbbm{P}}({\cal A}_{1}(0)\leq s_{1},{\cal A}_{1}(u)\leq s_{2})-\operatorname{\mathbbm{P}}({\cal A}_{1}(0)\leq s_{1})\operatorname{\mathbbm{P}}({\cal A}_{1}(u)\leq s_{2})\geq 0. (3.8)
Proof.

Recalling the notation from (2.4), notice that both LN(0)L^{*}_{N}(0) and LN(u)L_{N}^{*}(u) are increasing function of the weights ωi,jexp(1)\omega_{i,j}\sim\exp(1) (and they depend only on finitely many vertex weights). For any t1,t2t_{1},t_{2} it therefore follows that the events {LN(0)t1}\{L^{*}_{N}(0)\leq t_{1}\} and {LN(u)t2}\{L^{*}_{N}(u)\leq t_{2}\} are both decreasing and hence by the FKG inequality they are positively correlated (note that the FKG inequality is often stated for measures on finite distributive lattices satisfying the FKG lattice condition, but more general versions for product measures on finite products of totally ordered measure spaces applicable in the above scenario are available; see e.g. Lemma 2.1 of [43] or Corollary 2 of [42]), and therefore

(LN(0)t1,LN(u)t2)(LN(0)t1)(LN(u)t2)0.\operatorname{\mathbbm{P}}(L^{*}_{N}(0)\leq t_{1},L^{*}_{N}(u)\leq t_{2})-\operatorname{\mathbbm{P}}(L^{*}_{N}(0)\leq t_{1})\operatorname{\mathbbm{P}}(L^{*}_{N}(u)\leq t_{2})\geq 0. (3.9)

Using that the Airy1 process is a scaling limit of LL^{*} (see (2.5)), the proof is complete. ∎

We first derive an expression for (u;s1,s2){\cal E}(u;s_{1},s_{2}). Let us begin with the Fredholm representation of the function FF. We have from [52, 19, 29]

F(u;s1,s2)=det(𝟙K)F(u;s_{1},s_{2})=\det(\mathbbm{1}-K) (3.10)

where KK is a 2×22\times 2 matrix kernel

K=(K1,1K1,2K2,1K2,2)K=\left(\begin{array}[]{cc}K_{1,1}&K_{1,2}\\ K_{2,1}&K_{2,2}\\ \end{array}\right) (3.11)

with entries given by the extended kernel of the Airy1 process [52, 19, 29]

K1,1(x,y)\displaystyle K_{1,1}(x,y) =𝟙[x>s1]𝟙[y>s1]Ai(x+y),\displaystyle=\mathbbm{1}_{[x>s_{1}]}\mathbbm{1}_{[y>s_{1}]}\operatorname{Ai}(x+y), (3.12)
K1,2(x,y)\displaystyle K_{1,2}(x,y) =𝟙[x>s1]𝟙[y>s2](Ai(x+y+u2)e(x+y)u+23u3e(xy)2/4u4πu),\displaystyle=\mathbbm{1}_{[x>s_{1}]}\mathbbm{1}_{[y>s_{2}]}\Big{(}\operatorname{Ai}(x+y+u^{2})e^{(x+y)u+\tfrac{2}{3}u^{3}}-\frac{e^{-(x-y)^{2}/4u}}{\sqrt{4\pi u}}\Big{)},
K2,1(x,y)\displaystyle K_{2,1}(x,y) =𝟙[x>s2]𝟙[y>s1]Ai(x+y+u2)e(x+y)u23u3,\displaystyle=\mathbbm{1}_{[x>s_{2}]}\mathbbm{1}_{[y>s_{1}]}\operatorname{Ai}(x+y+u^{2})e^{-(x+y)u-\tfrac{2}{3}u^{3}},
K2,2(x,y)\displaystyle K_{2,2}(x,y) =𝟙[x>s2]𝟙[y>s2]Ai(x+y),\displaystyle=\mathbbm{1}_{[x>s_{2}]}\mathbbm{1}_{[y>s_{2}]}\operatorname{Ai}(x+y),

where Ai\operatorname{Ai} denotes the Airy function. Also, for the one-point distributions, we have f(si)=det(𝟙Ki,i)f(s_{i})=\det(\mathbbm{1}-K_{i,i}) for i=1,2i=1,2.

The first step is the following result.

Lemma 3.3.

With the above notations

1+(u;s1,s2)=det(𝟙K~)1+{\cal E}(u;s_{1},s_{2})=\det(\mathbbm{1}-\widetilde{K}) (3.13)

where

K~=(𝟙K1,1)1K1,2(𝟙K2,2)1K2,1.\widetilde{K}=(\mathbbm{1}-K_{1,1})^{-1}K_{1,2}(\mathbbm{1}-K_{2,2})^{-1}K_{2,1}. (3.14)
Proof.

Similar to [56], we compute

det(𝟙(K1,1K1,2K2,1K2,2))=det(𝟙(K1,100K2,2)(0K1,2K2,10))\displaystyle\det\left(\mathbbm{1}-\left(\begin{array}[]{cc}K_{1,1}&K_{1,2}\\ K_{2,1}&K_{2,2}\\ \end{array}\right)\right)=\det\left(\mathbbm{1}-\left(\begin{array}[]{cc}K_{1,1}&0\\ 0&K_{2,2}\\ \end{array}\right)-\left(\begin{array}[]{cc}0&K_{1,2}\\ K_{2,1}&0\\ \end{array}\right)\right) (3.15)
=det(𝟙(K1,100K2,2))\displaystyle=\det\left(\mathbbm{1}-\left(\begin{array}[]{cc}K_{1,1}&0\\ 0&K_{2,2}\\ \end{array}\right)\right)
×det(𝟙((𝟙K1,1)100(𝟙K2,2)1)(0K1,2K2,10))\displaystyle\qquad\times\det\left(\mathbbm{1}-\left(\begin{array}[]{cc}(\mathbbm{1}-K_{1,1})^{-1}&0\\ 0&(\mathbbm{1}-K_{2,2})^{-1}\\ \end{array}\right)\left(\begin{array}[]{cc}0&K_{1,2}\\ K_{2,1}&0\\ \end{array}\right)\right)
=det(𝟙K1,1)det(𝟙K2,2)det(𝟙(0GH0)),\displaystyle=\det(\mathbbm{1}-K_{1,1})\det(\mathbbm{1}-K_{2,2})\det\left(\mathbbm{1}-\left(\begin{array}[]{cc}0&-G\\ -H&0\\ \end{array}\right)\right),

where we set G=(𝟙K1,1)1K1,2G=-(\mathbbm{1}-K_{1,1})^{-1}K_{1,2} and H=(𝟙K2,2)1K2,1H=-(\mathbbm{1}-K_{2,2})^{-1}K_{2,1}. Moreover,

det(𝟙GH𝟙)=det((𝟙GH𝟙)(𝟙0H𝟙))=det(𝟙GH),\displaystyle\det\left(\begin{array}[]{cc}\mathbbm{1}&G\\ H&\mathbbm{1}\\ \end{array}\right)=\det\left(\left(\begin{array}[]{cc}\mathbbm{1}&G\\ H&\mathbbm{1}\\ \end{array}\right)\left(\begin{array}[]{cc}\mathbbm{1}&0\\ -H&\mathbbm{1}\\ \end{array}\right)\right)=\det(\mathbbm{1}-G\,H), (3.16)

where

GH=(𝟙K1,1)1K1,2(𝟙K2,2)1K2,1.G\,H=(\mathbbm{1}-K_{1,1})^{-1}K_{1,2}(\mathbbm{1}-K_{2,2})^{-1}K_{2,1}. (3.17)

Since det(𝟙K,)=f(s)\det(\mathbbm{1}-K_{\ell,\ell})=f(s_{\ell}), for =1,2\ell=1,2 as we mentioned above, (3.13) follows from (3.5). ∎

Next we would like to approximate the Fredholm determinant in (3.13) by that of a simpler kernel.

Proposition 3.4.

Let us define

R1(u;s1,s2):=det(𝟙K~)det(𝟙K1,2K2,1).R_{1}(u;s_{1},s_{2}):=\det(\mathbbm{1}-\tilde{K})-\det(\mathbbm{1}-K_{1,2}K_{2,1}). (3.18)

Then, for any s1,s20s_{1},s_{2}\geq 0, there exists a constant C2>0C_{2}>0 such that

|R1(u;s1,s2)|C2emin{s1,s2}u2e43u32(s1+s2)u.|R_{1}(u;s_{1},s_{2})|\leq\frac{C_{2}e^{-\min\{s_{1},s_{2}\}}}{u^{2}}e^{-\tfrac{4}{3}u^{3}-2(s_{1}+s_{2})u}. (3.19)

for any umax{12,s1+s2}u\geq\max\{\frac{1}{2},\sqrt{s_{1}+s_{2}}\}.

The proof of this proposition is in Section 3.3.

3.1 The leading term

Lemma 3.3 and Proposition 3.4 suggest

(u;s1,s1)det(𝟙K1,2K2,1)1 as u.{\cal E}(u;s_{1},s_{1})\sim\det(\mathbbm{1}-K_{1,2}K_{2,1})-1\quad\text{ as }u\rightarrow\infty. (3.20)

For a trace class operator 𝒦\cal K, the Fredholm determinant is given by

det(𝟙𝒦)=n=0(1)nn!n𝑑x1𝑑xndet[𝒦(xi,xj)]i,j=1n.\det(\mathbbm{1}-{\cal K})=\sum_{n=0}^{\infty}\frac{(-1)^{n}}{n!}\int_{\mathbb{R}^{n}}dx_{1}\cdots dx_{n}\det[{\cal K}(x_{i},x_{j})]_{i,j=1}^{n}. (3.21)

When 𝒦=K1,2K2,1{\cal K}=K_{1,2}K_{2,1}, this translates to

det(𝟙K1,2K2,1)=1tr(K1,2K2,1)+R2(u;s1,s2),\det(\mathbbm{1}-K_{1,2}K_{2,1})=1-\operatorname*{tr}\big{(}K_{1,2}K_{2,1}\big{)}+R_{2}(u;s_{1},s_{2}), (3.22)

where tr(K1,2K2,1)=𝑑x𝑑yK1,2(x,y)K2,1(y,x)\operatorname*{tr}\big{(}K_{1,2}K_{2,1}\big{)}=\int_{\mathbb{R}}dx\int_{\mathbb{R}}dyK_{1,2}(x,y)K_{2,1}(y,x) and

R2(u;s1,s2)=n=2(1)nn!s1s1𝑑x1𝑑xndet[K1,2K2,1(xi,xj)]i,j=1n.R_{2}(u;s_{1},s_{2})=\sum_{n=2}^{\infty}\frac{(-1)^{n}}{n!}\int_{s_{1}}^{\infty}\cdots\int_{s_{1}}^{\infty}dx_{1}\cdots dx_{n}\det[K_{1,2}K_{2,1}(x_{i},x_{j})]_{i,j=1}^{n}. (3.23)

From (3.12), it is clear that the upper tail of K1,2K2,1K_{1,2}K_{2,1} in either variables, is determined by that of the function Ai\operatorname{Ai}, which is known to decay super-exponentially, see (A.2). As each of the determinants in (3.23) is a sum of products of elements of the order of tr(K1,2K2,1)\operatorname*{tr}\big{(}K_{1,2}K_{2,1}\big{)}, one expects the latter to dominate R2R_{2} and therefore that

det(𝟙K1,2K2,1)1tr(K1,2K2,1)\det(\mathbbm{1}-K_{1,2}K_{2,1})\sim 1-\operatorname*{tr}(K_{1,2}K_{2,1}) (3.24)

if tr(K1,2K2,1)\operatorname*{tr}(K_{1,2}K_{2,1}) is small.

Let us move on to the computation of tr(K1,2K2,1)\operatorname*{tr}(K_{1,2}K_{2,1}). We write the kernel entries (K1,2K2,1)(x,y)(K_{1,2}K_{2,1})(x,y) as well as its trace using complex integral representations, which will then be analyzed. We start with the following identities (see e.g. Appendix A of [6] for the first and last, while the second is a standard Gaussian integral)

12πiγa𝑑ξeξ3/3+uξ2+xξ=Ai(x+u2)e23u3+ux,\displaystyle\frac{1}{2\pi{\rm i}}\int_{\gamma_{a}}d\xi e^{-\xi^{3}/3+u\xi^{2}+x\xi}=\operatorname{Ai}(x+u^{2})e^{\frac{2}{3}u^{3}+ux}, (3.25)
12πiγa𝑑ξeuξ2+xξ=ex24u4πu,\displaystyle\frac{1}{2\pi{\rm i}}\int_{\gamma_{a}}d\xi e^{u\xi^{2}+x\xi}=\frac{e^{-\frac{x^{2}}{4u}}}{\sqrt{4\pi u}},
12πiγb𝑑ηeη33uη2xη=Ai(x+u2)e23u3ux,\displaystyle\frac{1}{2\pi{\rm i}}\int_{\gamma_{b}}d\eta e^{\frac{\eta^{3}}{3}-u\eta^{2}-x\eta}=\operatorname{Ai}(x+u^{2})e^{-\frac{2}{3}u^{3}-ux},

where

γa\displaystyle\gamma_{a} =a+ifor a<u\displaystyle=a+{\rm i}\mathbb{R}\quad\text{for $a<u$} (3.26)
γb\displaystyle\gamma_{b} =b+ifor b>u.\displaystyle=b+{\rm i}\mathbb{R}\quad\text{for $b>u$}. (3.27)

Plugging these identities into (3.12) we can write

K1,2(x,y)\displaystyle K_{1,2}(x,y) =𝟙x>s1𝟙y>s212πiγa𝑑ξ[eξ3/3+uξ2+(x+y)ξeuξ2+(xy)ξ],\displaystyle=\mathbbm{1}_{x>s_{1}}\mathbbm{1}_{y>s_{2}}\frac{1}{2\pi{\rm i}}\int_{\gamma_{a}}d\xi\Big{[}e^{-\xi^{3}/3+u\xi^{2}+(x+y)\xi}-e^{u\xi^{2}+(x-y)\xi}\Big{]}, (3.28)
K2,1(x,y)\displaystyle K_{2,1}(x,y) =𝟙x>s2𝟙y>s112πiγb𝑑ηeη33uη2(x+y)η.\displaystyle=\mathbbm{1}_{x>s_{2}}\mathbbm{1}_{y>s_{1}}\frac{1}{2\pi{\rm i}}\int_{\gamma_{b}}d\eta e^{\frac{\eta^{3}}{3}-u\eta^{2}-(x+y)\eta}.

This leads to the following representation of the kernel K1,2K2,1K_{1,2}K_{2,1}:

(K1,2K2,1)(x,y)=𝑑zK1,2(x,z)K2,1(z,y)\displaystyle(K_{1,2}K_{2,1})(x,y)=\int_{\mathbb{R}}dzK_{1,2}(x,z)K_{2,1}(z,y) (3.29)
=𝟙x>s1𝟙y>s1(2πi)2γa𝑑ξγb𝑑ηeuξ2+xξeη33uη2yη[eξ33s2(ηξ)ηξes2(η+ξ)η+ξ].\displaystyle=\frac{\mathbbm{1}_{x>s_{1}}\mathbbm{1}_{y>s_{1}}}{(2\pi{\rm i})^{2}}\int_{\gamma_{a}}d\xi\int_{\gamma_{b}}d\eta e^{u\xi^{2}+x\xi}e^{\frac{\eta^{3}}{3}-u\eta^{2}-y\eta}\Bigg{[}\frac{e^{-\frac{\xi^{3}}{3}-s_{2}(\eta-\xi)}}{\eta-\xi}-\frac{e^{-s_{2}(\eta+\xi)}}{\eta+\xi}\Bigg{]}.

Setting x=yx=y and integrating over xx (on [s1,)[s_{1},\infty) due to the indicator functions) we get

tr(K1,2K2,1)=\displaystyle\operatorname*{tr}(K_{1,2}K_{2,1})= 1(2πi)2γa𝑑ξγb𝑑ηeη33uη2(s1+s2)η1ηξ\displaystyle\frac{1}{(2\pi{\rm i})^{2}}\int_{\gamma_{a}}d\xi\int_{\gamma_{b}}d\eta\,e^{\frac{\eta^{3}}{3}-u\eta^{2}-(s_{1}+s_{2})\eta}\frac{1}{\eta-\xi} (3.30)
×[eξ3/3+uξ2+(s1+s2)ξ1ηξeuξ2+(s1s2)ξ1ξ+η].\displaystyle\times\Big{[}e^{-\xi^{3}/3+u\xi^{2}+(s_{1}+s_{2})\xi}\frac{1}{\eta-\xi}-e^{u\xi^{2}+(s_{1}-s_{2})\xi}\frac{1}{\xi+\eta}\Big{]}.

We begin with determining the asymptotic behavior of tr(K1,2K2,1)\operatorname*{tr}(K_{1,2}K_{2,1}).

Proposition 3.5.

For all 0s1,s2u0\leq s_{1},s_{2}\leq\sqrt{u},

tr(K1,2K2,1)=116πu4e2(s1+s2)u43u3[s1s2+𝒪(u1/4)]-\operatorname*{tr}(K_{1,2}K_{2,1})=\frac{1}{16\pi u^{4}}e^{-2(s_{1}+s_{2})u-\frac{4}{3}u^{3}}\left[s_{1}s_{2}+\mathcal{O}(u^{-1/4})\right] (3.31)

as uu\to\infty.

Proof.

To get the asymptotic behavior of the trace, we need to choose the parameters a,ba,b in the integration contours. We do it in a way that the dominant parts in the exponents of the contour integrals in (3.30) are minimized.

(3.30)\displaystyle\eqref{K12K21} =1(2πi)2γa𝑑ξ1γb𝑑ηeη33uη2(s1+s2)ηeξ13/3+uξ12+(s1+s2)ξ11(ηξ1)2\displaystyle=\frac{1}{(2\pi{\rm i})^{2}}\int_{\gamma_{a}}d\xi_{1}\int_{\gamma_{b}}d\eta\,e^{\frac{\eta^{3}}{3}-u\eta^{2}-(s_{1}+s_{2})\eta}e^{-\xi_{1}^{3}/3+u\xi_{1}^{2}+(s_{1}+s_{2})\xi_{1}}\frac{1}{(\eta-\xi_{1})^{2}} (3.32)
1(2πi)2γa𝑑ξ2γb𝑑ηeη33uη2(s1+s2)ηeuξ22+(s1s2)ξ21(ξ2+η)(ηξ2).\displaystyle-\frac{1}{(2\pi{\rm i})^{2}}\int_{\gamma_{a}}d\xi_{2}\int_{\gamma_{b}}d\eta\,e^{\frac{\eta^{3}}{3}-u\eta^{2}-(s_{1}+s_{2})\eta}e^{u\xi_{2}^{2}+(s_{1}-s_{2})\xi_{2}}\frac{1}{(\xi_{2}+\eta)(\eta-\xi_{2})}.

Now we need to choose the parameters a,ba,b. For that reason we search for the minimizers of the different exponents in (3.32):

  1. (a)

    for

    ddξ1(ξ133+uξ12+(s1+s2)ξ1)=2uξ1ξ12=0,\frac{d}{d\xi_{1}}\Big{(}-\frac{\xi_{1}^{3}}{3}+u\xi_{1}^{2}+(s_{1}+s_{2})\xi_{1}\Big{)}=2u\xi_{1}-\xi_{1}^{2}=0, (3.33)

    which is solved for ξ1=uu2+s1+s2=:a1\xi_{1}=u-\sqrt{u^{2}+s_{1}+s_{2}}=:a_{1} or ξ1=u+u2+s1+s2\xi_{1}=u+\sqrt{u^{2}+s_{1}+s_{2}}. The solution which satisfy the constraint Re(ξ1)<u\operatorname{Re}(\xi_{1})<u in (3.26) is also the minimum.

  2. (b)

    For

    ddξ2(uξ22+(s1s2)ξ2)=2uξ2+(s1s2)\frac{d}{d\xi_{2}}\Big{(}u\xi_{2}^{2}+(s_{1}-s_{2})\xi_{2}\Big{)}=2u\xi_{2}+(s_{1}-s_{2}) (3.34)

    we see that ξ2=(s2s1)/(2u)=:a2\xi_{2}=(s_{2}-s_{1})/(2u)=:a_{2} is the minimum.

  3. (c)

    Similarly,

    ddη(η33uη2(s1+s2)η)=η22uη(s1+s2)=0,\frac{d}{d\eta}\Big{(}\frac{\eta^{3}}{3}-u\eta^{2}-(s_{1}+s_{2})\eta\Big{)}=\eta^{2}-2u\eta-(s_{1}+s_{2})=0, (3.35)

    has the minimum is at η=u+u2+s1+s2=:b\eta=u+\sqrt{u^{2}+s_{1}+s_{2}}=:b satisfies Re(η)>u{\rm{Re}(\eta)}>u.

So let us use the following change of variables

ξ1=a1+zu,ξ2=a2+zu,η=b+wu\xi_{1}=a_{1}+\frac{z}{\sqrt{u}},\quad\xi_{2}=a_{2}+\frac{z}{\sqrt{u}},\quad\eta=b+\frac{w}{\sqrt{u}} (3.36)

with z,wiz,w\in{\rm i}\mathbb{R} into (3.32). The two terms are analyzed in the same way, thus we write the details only for the first one.

Denote σ=(s1+s2)/u22u3/2\sigma=(s_{1}+s_{2})/u^{2}\leq 2u^{-3/2} and consider the first term in (3.32). We have

eη33uη2(s1+s2)ηe13ξ13+uξ12+(s1+s2)ξ1=e43u3(1+σ)3/2e1+σ(w2+z2)ew3z33u3/2\displaystyle e^{\frac{\eta^{3}}{3}-u\eta^{2}-(s_{1}+s_{2})\eta}e^{-\frac{1}{3}\xi_{1}^{3}+u\xi_{1}^{2}+(s_{1}+s_{2})\xi_{1}}=e^{-\frac{4}{3}u^{3}(1+\sigma)^{3/2}}e^{\sqrt{1+\sigma}(w^{2}+z^{2})}e^{\frac{w^{3}-z^{3}}{3u^{3/2}}} (3.37)
=e43u32(s1+s2)u(s1+s2)22u(1+𝒪(σ))e(z2+w2)(1+𝒪(σ;zu3/2;wu3/2))\displaystyle=e^{-\frac{4}{3}u^{3}-2(s_{1}+s_{2})u-\frac{(s_{1}+s_{2})^{2}}{2u}(1+\mathcal{O}(\sigma))}e^{(z^{2}+w^{2})(1+\mathcal{O}(\sigma;zu^{-3/2};wu^{-3/2}))}

and for the prefactor333The notation 𝒪(a1;;ak)\mathcal{O}(a_{1};...\,;a_{k}) stands for 𝒪(a1)++𝒪(ak)\mathcal{O}(a_{1})+...+\mathcal{O}(a_{k}).

1(ηξ1)2=14u2(1+𝒪(zu3/2;wu3/2;σ)).\frac{1}{(\eta-\xi_{1})^{2}}=\frac{1}{4u^{2}}(1+\mathcal{O}(zu^{-3/2};wu^{-3/2};\sigma)). (3.38)

Each term in the exponential which is cubic in z,wiz,w\in{\rm i}\mathbb{R} is purely imaginary, thus its exponential is bounded by 11. Furthermore, the quadratic terms in zz and ww have a positive prefactor 1+σ1\sqrt{1+\sigma}\geq 1. Thus integrating over |z|>u1/4|z|>u^{1/4} and/or |w|>u1/4|w|>u^{1/4} we get a correction term of order eue^{-\sqrt{u}} times the value of the integrand at z=w=0z=w=0. For the rest of the integral, for which |z|,|w|u1/4|z|,|w|\leq u^{1/4}, the error terms 𝒪(zu3/2;wu3/2;σ)=𝒪(u5/4)\mathcal{O}(zu^{-3/2};wu^{-3/2};\sigma)=\mathcal{O}(u^{-5/4}).

So the first term in (3.32) is given by

e43u32(s1+s2)u(s1+s2)22u(1+𝒪(σ))4u3[𝒪(eu)+(1+𝒪(u5/4))(2πi)2iu1/4iu1/4dziu1/4iu1/4dwe(z2+w2)(1+𝒪(u5/4))].\frac{e^{-\frac{4}{3}u^{3}-2(s_{1}+s_{2})u-\frac{(s_{1}+s_{2})^{2}}{2u}(1+\mathcal{O}(\sigma))}}{4u^{3}}\bigg{[}\mathcal{O}(e^{-\sqrt{u}})\\ +\frac{(1+\mathcal{O}(u^{-5/4}))}{(2\pi{\rm i})^{2}}\int_{-{\rm i}u^{1/4}}^{{\rm i}u^{1/4}}dz\int_{-{\rm i}u^{1/4}}^{{\rm i}u^{1/4}}dwe^{(z^{2}+w^{2})(1+\mathcal{O}(u^{-5/4}))}\bigg{]}. (3.39)

Finally, extending the Gaussian integrals to i{\rm i}\mathbb{R}, we get an error term 𝒪(eu)\mathcal{O}(e^{-\sqrt{u}}) only and using

1(2πi)2i𝑑zi𝑑wew2+z2=14π\frac{1}{(2\pi{\rm i})^{2}}\int_{{\rm i}\mathbb{R}}dz\int_{{\rm i}\mathbb{R}}dw\,e^{w^{2}+z^{2}}=\frac{1}{4\pi} (3.40)

we get

(3.39)\displaystyle\eqref{eq3.39} =e43u32(s1+s2)u(s1+s2)22u(1+𝒪(σ))16πu3(1+𝒪(u5/4))\displaystyle=\frac{e^{-\frac{4}{3}u^{3}-2(s_{1}+s_{2})u-\frac{(s_{1}+s_{2})^{2}}{2u}(1+\mathcal{O}(\sigma))}}{16\pi u^{3}}(1+\mathcal{O}(u^{-5/4})) (3.41)
=e43u32(s1+s2)u16πu3[1(s1+s2)22u+𝒪(u5/4)]\displaystyle=\frac{e^{-\frac{4}{3}u^{3}-2(s_{1}+s_{2})u}}{16\pi u^{3}}\left[1-\frac{(s_{1}+s_{2})^{2}}{2u}+\mathcal{O}(u^{-5/4})\right]

A similar computation for the second term in (3.32) leads to

e43u32(s1+s2)u16πu3[1s12+s222u+𝒪(u5/4)].-\frac{e^{-\frac{4}{3}u^{3}-2(s_{1}+s_{2})u}}{16\pi u^{3}}\left[1-\frac{s_{1}^{2}+s_{2}^{2}}{2u}+\mathcal{O}(u^{-5/4})\right]. (3.42)

Summing up (3.41) and (3.42) we obtain

tr(K1,2K2,1)=e43u32(s1+s2)u16πu4[s1s2+𝒪(u1/4)].-\operatorname*{tr}(K_{1,2}K_{2,1})=\frac{e^{-\frac{4}{3}u^{3}-2(s_{1}+s_{2})u}}{16\pi u^{4}}\left[s_{1}s_{2}+\mathcal{O}(u^{-1/4})\right]. (3.43)

3.2 Bounding lower order terms

To show (3.24) we need to get a bound on R2(u;s1,s2)R_{2}(u;s_{1},s_{2}) from (3.23), which is o(tr(K1,2K2,1))o(\operatorname*{tr}\big{(}K_{1,2}K_{2,1}\big{)}).

Proposition 3.6.

There exists a constant C>0C>0 such that

|R2(u;s1,s2)|Cu6e4(s1+s2)ue83u3|R_{2}(u;s_{1},s_{2})|\leq\frac{C}{u^{6}}e^{-4(s_{1}+s_{2})u}e^{-\frac{8}{3}u^{3}} (3.44)

uniformly in s1,s2s_{1},s_{2}\in\mathbb{R}.

Proof.

In (LABEL:eq7) we use the change of variables

ξ=zu,η=2u+wu,\xi=\frac{z}{\sqrt{u}},\quad\eta=2u+\frac{w}{\sqrt{u}}, (3.45)

where z,wiz,w\in{\rm i}\mathbb{R}. This leads to

(LABEL:eq7)=𝟙x>s1𝟙y>s1e43u32u(s2+y)(2πi)2ui𝑑zi𝑑wez2+w2P(z,w,x,y)\eqref{eq7}=\mathbbm{1}_{x>s_{1}}\mathbbm{1}_{y>s_{1}}\frac{e^{-\frac{4}{3}u^{3}-2u(s_{2}+y)}}{(2\pi{\rm i})^{2}u}\int_{{\rm i}\mathbb{R}}dz\int_{{\rm i}\mathbb{R}}dwe^{z^{2}+w^{2}}P(z,w,x,y) (3.46)

with

P(z,w,x,y)=exzu+w33u3/2wyu[ez33u3/2s2wzu2u+wzues2z+wu2u+z+wu].P(z,w,x,y)=e^{\frac{xz}{\sqrt{u}}+\frac{w^{3}}{3u^{3/2}}-\frac{wy}{\sqrt{u}}}\bigg{[}\frac{e^{-\frac{z^{3}}{3u^{3/2}}-s_{2}\frac{w-z}{\sqrt{u}}}}{2u+\frac{w-z}{\sqrt{u}}}-\frac{e^{-s_{2}\frac{z+w}{\sqrt{u}}}}{2u+\frac{z+w}{\sqrt{u}}}\bigg{]}. (3.47)

Since x,yx,y\in\mathbb{R} and z,wiz,w\in{\rm i}\mathbb{R}, we get the simple bound |P(z,w,x,y)|1u|P(z,w,x,y)|\leq\frac{1}{u}, while ez2+w2e^{z^{2}+w^{2}} is real. Performing the Gaussian integral we then get

|(3.46)|𝟙x>s1𝟙y>s1e43u32(s2+y)u4πu2.|\eqref{eq1.40}|\leq\mathbbm{1}_{x>s_{1}}\mathbbm{1}_{y>s_{1}}\frac{e^{-\frac{4}{3}u^{3}-2(s_{2}+y)u}}{4\pi u^{2}}. (3.48)

Let K=K1,2K2,1K=K_{1,2}K_{2,1}. Hadamard’s inequality444Let AA be a n×nn\times n matrix with |Ai,j|1|A_{i,j}|\leq 1. Then |det(A)|nn/2|\det(A)|\leq n^{n/2}. gives

|det[K(xi,xj)]i,j=1n|nn/2j=1ne4u332(xj+s2)u𝟙xj>s14πu2|\det[K(x_{i},x_{j})]_{i,j=1}^{n}|\leq n^{n/2}\prod_{j=1}^{n}\frac{e^{-\frac{4u^{3}}{3}-2(x_{j}+s_{2})u}\mathbbm{1}_{x_{j}>s_{1}}}{4\pi u^{2}} (3.49)

so that

|x1s1𝑑x1xns1𝑑xndet[K(xi,xj)]i,j=1n|nn/2(e4u332(s1+s2)u8πu3)n.\bigg{|}\int_{x_{1}\geq s_{1}}dx_{1}\cdots\int_{x_{n}\geq s_{1}}dx_{n}\det[K(x_{i},x_{j})]_{i,j=1}^{n}\bigg{|}\leq n^{n/2}\bigg{(}\frac{e^{-\frac{{4u^{3}}}{3}-2(s_{1}+s_{2})u}}{8\pi u^{3}}\bigg{)}^{n}. (3.50)

Denote M=e4u332(s1+s2)u8πu3M=\frac{e^{-\frac{{4u^{3}}}{3}-2(s_{1}+s_{2})u}}{8\pi u^{3}}. Then there exists C>0C>0 such that

|R2(u;s1,s2)|n=2Mnnn/2n!CM2Cu6e8u334(s1+s2)u.|R_{2}(u;s_{1},s_{2})|\leq\sum_{n=2}^{\infty}\frac{M^{n}n^{n/2}}{n!}\leq CM^{2}\leq Cu^{-6}e^{-\frac{{8u^{3}}}{3}-4(s_{1}+s_{2})u}. (3.51)

This completes the proof. ∎

We have now all the ingredients to complete the proof of Theorem 3.1.

Proof of Theorem 3.1.

We have

F(u;s1,s2)=(3.13)\displaystyle F(u;s_{1},s_{2})\stackrel{{\scriptstyle\eqref{FdK}}}{{=}} f(s1)f(s2)det(𝟙K~)\displaystyle f(s_{1})f(s_{2})\det(\mathbbm{1}-\tilde{K}) (3.52)
=(3.19)\displaystyle\stackrel{{\scriptstyle\eqref{ubd}}}{{=}} f(s1)f(s2)[det(𝟙K1,2K2,1)+R1(u;s1,s2)]\displaystyle f(s_{1})f(s_{2})[\det(\mathbbm{1}-K_{1,2}K_{2,1})+R_{1}(u;s_{1},s_{2})]
=(3.22)\displaystyle\stackrel{{\scriptstyle\eqref{asy2}}}{{=}} f(s1)f(s2)[1tr(K1,2K2,1)+R(u;s1,s2)],\displaystyle f(s_{1})f(s_{2})[1-\operatorname*{tr}(K_{1,2}K_{2,1})+R(u;s_{1},s_{2})],

where

R(u;s1,s2)=R1(u;s1,s2)+R2(u;s1,s2).R(u;s_{1},s_{2})=R_{1}(u;s_{1},s_{2})+R_{2}(u;s_{1},s_{2}). (3.53)

It follows that

F(u;s1,s2)f(s1)f(s2)=f(s1)f(s2)tr(K1,2K2,1)+f(s1)f(s2)R(u;s1,s2).F(u;s_{1},s_{2})-f(s_{1})f(s_{2})=-f(s_{1})f(s_{2})\operatorname*{tr}(K_{1,2}K_{2,1})+f(s_{1})f(s_{2})R(u;s_{1},s_{2}). (3.54)

We shall apply the lower bound (3.7) for the covariance for an appropriate choice of α=α(u)\alpha=\alpha(u) and β=β(u)\beta=\beta(u) such the contribution of error term f(s1)f(s2)R(u;s1,s2)f(s_{1})f(s_{2})R(u;s_{1},s_{2}) is of a subleading order.

We shall choose α>0\alpha>0 depending on uu and for concreteness set β=α+1\beta=\alpha+1. Thus we integrate over a region of area 11. By Proposition 3.6, the error term R2(u;s1,s2)R_{2}(u;s_{1},s_{2}) is much smaller than the leading term coming from the trace, see Proposition 3.5 (of order e43u3e^{-\frac{4}{3}u^{3}} smaller) for s1,s2u2s_{1},s_{2}\ll u^{2}. However, the exponential term in uu in the error bound of R1(u;s1,s2)R_{1}(u;s_{1},s_{2}) coming from (3.19) is of the same order, namely e2(s1+s2)u43u3e^{-2(s_{1}+s_{2})u-\frac{4}{3}u^{3}}. The difference is that in the trace we have a prefactor s1s2u4\sim\frac{s_{1}s_{2}}{u^{4}}, while in the bound for R1R_{1} we have a prefactor emin{s1,s2}u2\sim\frac{e^{-\min\{s_{1},s_{2}\}}}{u^{2}}. Thus, in order to ensure that the contribution from R1R_{1} is subleading, we can take s1,s2cln(u)s_{1},s_{2}\sim c\ln(u) for any c3c\geq 3. Therefore choosing α=3ln(u)\alpha=3\ln(u) and using the fact that for all x0x\geq 0, (x)=FGOE(22/3x)(x)=F_{\rm GOE}(2^{2/3}x) is bounded away from 0 (in fact, one can numerically check f(x)=FGOE(22/3x)[FGOE(0),1]=[0.83,1]f(x)=F_{\rm GOE}(2^{2/3}x)\in[F_{\rm GOE}(0),1]=[0.83...,1]), we finally get the claimed result. ∎

3.3 Proof of Proposition 3.4

To prove Proposition 3.4, we begin with the following well known bound (see e.g. Lemma 4(d), Chapter XIII.17 of [51])

|det(𝟙K~)det(𝟙K1,2K2,1)|Q1eQ1+2K1,2K2,11+1.|\det(\mathbbm{1}-\tilde{K})-\det(\mathbbm{1}-K_{1,2}K_{2,1})|\leq\|Q\|_{1}e^{\|Q\|_{1}+2\|K_{1,2}K_{2,1}\|_{1}+1}. (3.55)

where Q=K~K1,2K2,1Q=\tilde{K}-K_{1,2}K_{2,1}, where 1\|\cdot\|_{1} is the trace-norm (see e.g. [54]). From Lemmas 3.10 and 3.11 below, the exponent in the display above is bounded and the result follows. ∎

In the remainder of this section we prove the two results Lemma 3.10 and Lemma 3.11 used above. We first need to prove some auxiliary bounds. From the identity

(𝟙K1,1)1=𝟙+(𝟙K1,1)1K1,1(\mathbbm{1}-K_{1,1})^{-1}=\mathbbm{1}+(\mathbbm{1}-K_{1,1})^{-1}K_{1,1} (3.56)

we see that

Q=\displaystyle Q= K~K1,2K2,1=(𝟙K1,1)1K1,1K1,2(𝟙K2,2)1K2,2K2,1\displaystyle\tilde{K}-K_{1,2}K_{2,1}=(\mathbbm{1}-K_{1,1})^{-1}K_{1,1}K_{1,2}(\mathbbm{1}-K_{2,2})^{-1}K_{2,2}K_{2,1} (3.57)
+K1,2(𝟙K2,2)1K2,2K2,1+(𝟙K1,1)1K1,1K1,2K2,1.\displaystyle+K_{1,2}(\mathbbm{1}-K_{2,2})^{-1}K_{2,2}K_{2,1}+(\mathbbm{1}-K_{1,1})^{-1}K_{1,1}K_{1,2}K_{2,1}.

Recall that from (3.55) we need to bound Q1\|Q\|_{1}. Thus it is enough to bound the 1\|\cdot\|_{1}-norm of each of the terms on the right hand side of (3.57). Since K1,2(x,y)K_{1,2}(x,y) is not decaying as x=yx=y\to\infty, we could either work in weighted L2L^{2} spaces, or as we do here, introduce some weighting in the kernel elements. Namely define

K¯1,2L(x,y)\displaystyle\bar{K}_{1,2}^{L}(x,y) =ex2K1,2(x,y),K¯1,2R(x,y)=K1,2(x,y)ey2,\displaystyle=e^{-\frac{x}{2}}K_{1,2}(x,y),\quad\bar{K}_{1,2}^{R}(x,y)=K_{1,2}(x,y)e^{-\frac{y}{2}}, (3.58)
K¯1,1(x,y)\displaystyle\bar{K}_{1,1}(x,y) =K1,1(x,y)ey2,K¯2,2(x,y)=ex2K2,2(x,y).\displaystyle=K_{1,1}(x,y)e^{\frac{y}{2}},\quad\bar{K}_{2,2}(x,y)=e^{\frac{x}{2}}K_{2,2}(x,y).

Also, we use the norm inequalities AB1A2B2\|AB\|_{1}\leq\|A\|_{2}\|B\|_{2} and A2A1\|A\|_{2}\leq\|A\|_{1}, with 2\|\cdot\|_{2} the Hilbert-Schmidt norm given by K2=(𝑑x𝑑y(K(x,y))2)1/2\|K\|_{2}=\Big{(}\int dxdy(K(x,y))^{2}\Big{)}^{1/2}.

These norm inequalities, the fact that Ki,iK_{i,i} and (𝟙Ki,i)1(\mathbbm{1}-K_{i,i})^{-1} commute and the identity K1,1K1,2=K¯1,1K¯1,2LK_{1,1}K_{1,2}=\bar{K}_{1,1}\bar{K}_{1,2}^{L} lead to

(𝟙K1,1)1K1,1K1,2K2,11(𝟙K1,1)12K¯1,12K¯1,2L2K2,12.\|(\mathbbm{1}-K_{1,1})^{-1}K_{1,1}K_{1,2}K_{2,1}\|_{1}\leq\|(\mathbbm{1}-K_{1,1})^{-1}\|_{2}\|\bar{K}_{1,1}\|_{2}\|\bar{K}^{L}_{1,2}\|_{2}\|K_{2,1}\|_{2}. (3.59)

Moreover, using K1,2K2,2=K¯1,2RK¯2,2K_{1,2}K_{2,2}=\bar{K}^{R}_{1,2}\bar{K}_{2,2}, we get

K1,2(𝟙K2,2)1K2,2K2,11\displaystyle\|K_{1,2}(\mathbbm{1}-K_{2,2})^{-1}K_{2,2}K_{2,1}\|_{1} =K1,2K2,2(𝟙K2,2)1K2,11\displaystyle=\|K_{1,2}K_{2,2}(\mathbbm{1}-K_{2,2})^{-1}K_{2,1}\|_{1} (3.60)
K¯1,2R2K¯2,22(𝟙K2,2)12K2,12,\displaystyle\leq\|\bar{K}_{1,2}^{R}\|_{2}\|\bar{K}_{2,2}\|_{2}\|(\mathbbm{1}-K_{2,2})^{-1}\|_{2}\|K_{2,1}\|_{2},

and finally

(𝟙K1,1)1K1,1K1,2(𝟙K2,2)1K2,2K2,11\displaystyle\|(\mathbbm{1}-K_{1,1})^{-1}K_{1,1}K_{1,2}(\mathbbm{1}-K_{2,2})^{-1}K_{2,2}K_{2,1}\|_{1} (3.61)
(𝟙K1,1)12K¯1,12K¯1,2L2(𝟙K2,2)12K2,22K2,12.\displaystyle\leq\|(\mathbbm{1}-K_{1,1})^{-1}\|_{2}\|\bar{K}_{1,1}\|_{2}\|\bar{K}^{L}_{1,2}\|_{2}\|(\mathbbm{1}-K_{2,2})^{-1}\|_{2}\|K_{2,2}\|_{2}\|K_{2,1}\|_{2}.

We turn to bound each of the terms on the right hand side of each of the inequalities in (3.59)-(3.61).

We also use the following change of variables: for a,ba,b\in\mathbb{R}, define

ω=(xa)+(yb),ζ=(xa)(yb),\displaystyle\omega=(x-a)+(y-b),\quad\zeta=(x-a)-(y-b), (3.62)

which give x=a+12(ω+ζ)x=a+\frac{1}{2}(\omega+\zeta) and y=b+12(ωζ)y=b+\frac{1}{2}(\omega-\zeta). So the integral over (x,y)[a,)×[b,)(x,y)\in[a,\infty)\times[b,\infty) becomes an integral over (ω,ζ)(\omega,\zeta) with ω0\omega\geq 0 and ζ[ω,ω]\zeta\in[-\omega,\omega].

Lemma 3.7.

Uniformly for all s1,s20s_{1},s_{2}\geq 0,

Ki,i212e2si,K¯i,i2esi,(𝟙Ki,i)122,\|K_{i,i}\|_{2}\leq\tfrac{1}{2}e^{-2s_{i}},\quad\|\bar{K}_{i,i}\|_{2}\leq e^{-s_{i}},\quad\|(\mathbbm{1}-K_{i,i})^{-1}\|_{2}\leq 2, (3.63)

for i{1,2}i\in\{1,2\}.

Proof.

The bounds for i=1i=1 and i=2i=2 are fully analogous, thus we restrict to the case i=1i=1 here. To bound K1,122\|K_{1,1}\|_{2}^{2} we use (A.3) which gives

K1,122\displaystyle\|K_{1,1}\|^{2}_{2} =s1𝑑xs1𝑑y[Ai(x+y)]2s1𝑑xs1𝑑ye2(x+y)14e4s1.\displaystyle=\int_{s_{1}}^{\infty}dx\int_{s_{1}}^{\infty}dy\big{[}\operatorname{Ai}(x+y)\big{]}^{2}\leq\int_{s_{1}}^{\infty}dx\int_{s_{1}}^{\infty}dye^{-2(x+y)}\leq\tfrac{1}{4}e^{-4s_{1}}. (3.64)

Thus K1,1212e2s1\|K_{1,1}\|_{2}\leq\frac{1}{2}e^{-2s_{1}}. Similarly

K¯1,122\displaystyle\|\bar{K}_{1,1}\|^{2}_{2} =s1𝑑xs1𝑑y[ey2Ai(x+y)]2s1𝑑xs1𝑑ye(x+2y)=12e3s1,\displaystyle=\int_{s_{1}}^{\infty}dx\int_{s_{1}}^{\infty}dy\big{[}e^{\frac{y}{2}}\operatorname{Ai}(x+y)\big{]}^{2}\leq\int_{s_{1}}^{\infty}dx\int_{s_{1}}^{\infty}dye^{-(x+2y)}=\tfrac{1}{2}e^{-3s_{1}}, (3.65)

which implies the second bound in (3.63). Finally using

𝟙Ki,i21(1Ki,i2)1,\|\mathbbm{1}-K_{i,i}\|_{2}^{-1}\leq(1-\|K_{i,i}\|_{2})^{-1}, (3.66)

which holds whenever Ki,i2<1\|K_{i,i}\|_{2}<1, we get the last inequality. ∎

Lemma 3.8.

For s1,s20s_{1},s_{2}\geq 0 and u>0u>0

K2,1212u3/2exp[2(s1+s2)u43u3].\displaystyle\|K_{2,1}\|_{2}\leq\frac{1}{2u^{3/2}}\exp\Big{[}-2(s_{1}+s_{2})u-\tfrac{4}{3}u^{3}\Big{]}. (3.67)
Proof.

Using the change of variables (3.62) with a=s2,b=s1a=s_{2},b=s_{1}, it follows that

K2,122\displaystyle\|K_{2,1}\|^{2}_{2} =s2𝑑xs1𝑑y[Ai(x+y+u2)]2e2(x+y)u43u3\displaystyle=\int_{s_{2}}^{\infty}dx\int_{s_{1}}^{\infty}dy\big{[}\operatorname{Ai}(x+y+u^{2})\big{]}^{2}e^{-2(x+y)u-\tfrac{4}{3}u^{3}} (3.68)
=20𝑑ωωω𝑑ζ[Ai(ω+s1+s2+u2)]2e2(ω+s1+s2)u43u3,\displaystyle=2\int_{0}^{\infty}d\omega\int_{-\omega}^{\omega}d\zeta\big{[}\operatorname{Ai}(\omega+s_{1}+s_{2}+u^{2})\big{]}^{2}e^{-2(\omega+s_{1}+s_{2})u-\tfrac{4}{3}u^{3}},

where we used that the Jacobian of the transformation in (3.62) is 2.

Next using (A.5) (with x=ω+s1+s2x=\omega+s_{1}+s_{2}) we get

(3.68)\displaystyle\eqref{eq4} =4e2(s1+s2)u43u30𝑑ω[Ai(ω+s1+s2+u2)]2ωe2ωu\displaystyle=4e^{-2(s_{1}+s_{2})u-\tfrac{4}{3}u^{3}}\int_{0}^{\infty}d\omega\big{[}\operatorname{Ai}(\omega+s_{1}+s_{2}+u^{2})\big{]}^{2}\omega e^{-2\omega u} (3.69)
4e4(s1+s2)u83u3u0𝑑ωωe4ωu=14u3e4(s1+s2)u83u3.\displaystyle\leq\frac{4e^{-4(s_{1}+s_{2})u-\tfrac{8}{3}u^{3}}}{u}\int_{0}^{\infty}d\omega\,\omega e^{-4\omega u}=\frac{1}{4u^{3}}e^{-4(s_{1}+s_{2})u-\tfrac{8}{3}u^{3}}.

Lemma 3.9.

For s1,s20s_{1},s_{2}\geq 0, there exists a constant C1>0C_{1}>0 such that

K¯1,2R2,K¯1,2L2C1ues1/2\displaystyle\|\bar{K}^{R}_{1,2}\|_{2},\|\bar{K}^{L}_{1,2}\|_{2}\leq\frac{C_{1}}{\sqrt{u}}e^{-s_{1}/2} (3.70)

for all us1+s2u\geq\sqrt{s_{1}+s_{2}}.

Proof of Lemma 3.9.

By symmetry in x,yx,y the bounds for K¯1,2L22\|\bar{K}^{L}_{1,2}\|_{2}^{2} and for K¯1,2R22\|\bar{K}^{R}_{1,2}\|_{2}^{2} are the same. So we consider K¯1,2L22\|\bar{K}^{L}_{1,2}\|_{2}^{2} only. Using (AB)22(A2+B2)(A-B)^{2}\leq 2(A^{2}+B^{2}) we get

K¯1,2L22\displaystyle\|\bar{K}^{L}_{1,2}\|_{2}^{2} =s1𝑑xs2𝑑yex[Ai(x+y+u2)e(x+y)u+23u3e(xy)2/4u4πu]2\displaystyle=\int_{s_{1}}^{\infty}dx\int_{s_{2}}^{\infty}dye^{-x}\Big{[}\operatorname{Ai}(x+y+u^{2})e^{(x+y)u+\tfrac{2}{3}u^{3}}-\frac{e^{-(x-y)^{2}/4u}}{\sqrt{4\pi u}}\Big{]}^{2} (3.71)
2s1𝑑xs2𝑑yex[(Ai(x+y+u2))2e2(x+y)u+43u3+e(xy)2/2u4πu].\displaystyle\leq 2\int_{s_{1}}^{\infty}dx\int_{s_{2}}^{\infty}dye^{-x}\Big{[}(\operatorname{Ai}(x+y+u^{2}))^{2}e^{2(x+y)u+\tfrac{4}{3}u^{3}}+\frac{e^{-(x-y)^{2}/2u}}{4\pi u}\Big{]}.

The second term in (3.71) can be bounded by integrating in yy over \mathbb{R} and then integrating over xs2x\geq s_{2}. This gives

2s1𝑑xs2𝑑yexe(xy)2/2u4πu12πues1.2\int_{s_{1}}^{\infty}dx\int_{s_{2}}^{\infty}dye^{-x}\frac{e^{-(x-y)^{2}/2u}}{4\pi u}\leq\frac{1}{\sqrt{2\pi u}}e^{-s_{1}}. (3.72)

For the first term in (3.71), we use the change of variable (3.62) with a=s1a=s_{1}, b=s2b=s_{2} and obtain

2s1𝑑xs2𝑑yex(Ai(x+y+u2))2e2(x+y)u+43u3\displaystyle 2\int_{s_{1}}^{\infty}dx\int_{s_{2}}^{\infty}dye^{-x}(\operatorname{Ai}(x+y+u^{2}))^{2}e^{2(x+y)u+\tfrac{4}{3}u^{3}} (3.73)
=4es10𝑑ω(Ai(ω+s1+s2+u2))2e2(ω+s1+s2)u+43u3ωω𝑑ζe12(ω+ζ)\displaystyle=4e^{-s_{1}}\int_{0}^{\infty}d\omega(\operatorname{Ai}(\omega+s_{1}+s_{2}+u^{2}))^{2}e^{2(\omega+s_{1}+s_{2})u+\tfrac{4}{3}u^{3}}\!\int_{-\omega}^{\omega}\!d\zeta e^{-\frac{1}{2}(\omega+\zeta)}
8es10𝑑ω(Ai(ω+s1+s2+u2))2e2(ω+s1+s2)u+43u3\displaystyle\leq 8e^{-s_{1}}\int_{0}^{\infty}d\omega(\operatorname{Ai}(\omega+s_{1}+s_{2}+u^{2}))^{2}e^{2(\omega+s_{1}+s_{2})u+\tfrac{4}{3}u^{3}}
(A.4)8es10𝑑ωe43(ω+s1+s2+u2)3/2+2(ω+s1+s2)u+43u3.\displaystyle\stackrel{{\scriptstyle\eqref{eqAiExpBound2}}}{{\leq}}8e^{-s_{1}}\int_{0}^{\infty}d\omega e^{-\frac{4}{3}(\omega+s_{1}+s_{2}+u^{2})^{3/2}+2(\omega+s_{1}+s_{2})u+\tfrac{4}{3}u^{3}}.

Let A(ω)=43(ω+s1+s2+u2)3/2+2(ω+s1+s2)u+43u3A(\omega)=-\frac{4}{3}(\omega+s_{1}+s_{2}+u^{2})^{3/2}+2(\omega+s_{1}+s_{2})u+\tfrac{4}{3}u^{3}. Then for all s1,s2,u0s_{1},s_{2},u\geq 0 it is a concave function in ω\omega and thus

A(ω)\displaystyle A(\omega) A(0)+A(0)ω\displaystyle\leq A(0)+A^{\prime}(0)\omega (3.74)
=2(s1+s2)u+43u343(s1+s2+u2)3/22(s1+s2+u2u)ω.\displaystyle=2(s_{1}+s_{2})u+\frac{4}{3}u^{3}-\frac{4}{3}(s_{1}+s_{2}+u^{2})^{3/2}-2(\sqrt{s_{1}+s_{2}+u^{2}}-u)\omega.

The term independent of ω\omega is always negative. Indeed, x43(x+u2)3/2x\mapsto\frac{4}{3}(x+u^{2})^{3/2} is convex and thus greater than its linear approximation at x=0x=0, which is 43u3+2ux\frac{4}{3}u^{3}+2ux. So we have

(3.73)4es1s1+s2+u2u4es1(21)u,\displaystyle\eqref{ub9}\leq\frac{4e^{-s_{1}}}{\sqrt{s_{1}+s_{2}+u^{2}}-u}\leq\frac{4e^{-s_{1}}}{(\sqrt{2}-1)u}, (3.75)

where in the last step we used the assumption that us1+s2u\geq\sqrt{s_{1}+s_{2}}. ∎

Lemma 3.10.

There exists a constant C1>0C_{1}>0 such that for s1,s20s_{1},s_{2}\geq 0,

K1,2K2,11C12u2e43u32(s1+s2)ues2C12u2e43u3\|K_{1,2}K_{2,1}\|_{1}\leq\frac{C_{1}}{2u^{2}}e^{-\frac{4}{3}u^{3}-2(s_{1}+s_{2})u}e^{s_{2}}\leq\frac{C_{1}}{2u^{2}}e^{-\frac{4}{3}u^{3}} (3.76)

for all umax{12,s1+s2}u\geq\max\{\frac{1}{2},\sqrt{s_{1}+s_{2}}\}.

Proof.

Denote

K¯2,1(x,y)=ex2K2,1(x,y).\bar{K}_{2,1}(x,y)=e^{\frac{x}{2}}K_{2,1}(x,y). (3.77)

Then we have the bound

K1,2K2,11K¯1,2R2K¯2,12.\displaystyle\|K_{1,2}K_{2,1}\|_{1}\leq\|\bar{K}_{1,2}^{R}\|_{2}\|\bar{K}_{2,1}\|_{2}. (3.78)

From Lemma 3.9, we get K¯1,2R2C1u1/2\|\bar{K}_{1,2}^{R}\|_{2}\leq\frac{C_{1}}{u^{1/2}}. Furthermore, using (A.5), we have

K¯2,122\displaystyle\|\bar{K}_{2,1}\|_{2}^{2} =s2𝑑xs1𝑑yex[Ai(x+y+u2)]2e2(x+y)u43u3\displaystyle=\int_{s_{2}}^{\infty}dx\int_{s_{1}}^{\infty}dy\,e^{x}\big{[}\operatorname{Ai}(x+y+u^{2})\big{]}^{2}e^{-2(x+y)u-\tfrac{4}{3}u^{3}} (3.79)
1us2𝑑xs1𝑑yexe4(x+y)u83u3e83u3e4(s1+s2)u+s214u2(4u1)\displaystyle\leq\frac{1}{u}\int_{s_{2}}^{\infty}dx\int_{s_{1}}^{\infty}dy\,e^{x}e^{-4(x+y)u-\tfrac{8}{3}u^{3}}\leq e^{-\frac{8}{3}u^{3}}e^{-4(s_{1}+s_{2})u+s_{2}}\frac{1}{4u^{2}(4u-1)}
e83u3e4(s1+s2)u+s218u3\displaystyle\leq e^{-\frac{8}{3}u^{3}}e^{-4(s_{1}+s_{2})u+s_{2}}\frac{1}{8u^{3}}

for all u1/2u\geq 1/2. ∎

We are now ready to bound Q1\|Q\|_{1}.

Lemma 3.11.

Uniformly for s1,s20s_{1},s_{2}\geq 0, there exists a constant C1C_{1} such that

Q12C1emin{s1,s2}u2e2(s1+s2)u43u3\|Q\|_{1}\leq\frac{2C_{1}e^{-\min\{s_{1},s_{2}\}}}{u^{2}}e^{-2(s_{1}+s_{2})u-\tfrac{4}{3}u^{3}} (3.80)

for all umax{12,s1+s2}u\geq\max\{\frac{1}{2},\sqrt{s_{1}+s_{2}}\}.

Proof.

Applying the bounds of Lemmas 3.7 and 3.9 to (3.59)-(3.61), we obtain

Q18C1u1/2emin{s1,s2}K2,12\|Q\|_{1}\leq\frac{8C_{1}}{u^{1/2}}e^{-\min\{s_{1},s_{2}\}}\|K_{2,1}\|_{2} (3.81)

for all s1,s20s_{1},s_{2}\geq 0 and uu as mentioned. Then, applying the bound of Lemma 3.8 leads to

Q14C1emin{s1,s2}u2e2(s1+s2)u43u3.\|Q\|_{1}\leq\frac{4C_{1}e^{-\min\{s_{1},s_{2}\}}}{u^{2}}e^{-2(s_{1}+s_{2})u-\tfrac{4}{3}u^{3}}. (3.82)

Appendix A Some tail estimates

Here we collect and, in some cases, prove estimates on tails of the Airy function and tails of some LPPs (point-to-point and point-to-half line) that we have used earlier.

A.1 Bounds on Airy functions

The Airy function Ai\operatorname{Ai} satisfied the following identity (see e.g. Appendix A of [6])

12πiγ𝑑zexp(z33+bz2zx)=Ai(b2+x)e23b3+bx\frac{1}{2\pi{\rm i}}\int_{\gamma}dz\exp\Big{(}\frac{z^{3}}{3}+bz^{2}-zx\Big{)}=\operatorname{Ai}(b^{2}+x)e^{\frac{2}{3}b^{3}+bx} (A.1)

for any γ=α+i\gamma=\alpha+{\rm i}\mathbb{R} with α>b\alpha>-b. The asymptotics of Ai\operatorname{Ai} is given by [1]

Ai(x)=x1/4πexp(23x3/2)[1+O(x1)]x\operatorname{Ai}(x)=\frac{x^{-1/4}}{\sqrt{\pi}}\exp\Big{(}-\frac{2}{3}x^{3/2}\Big{)}[1+O(x^{-1})]\quad x\rightarrow\infty (A.2)

Furthermore, one has the following simple bounds555supx|Ai(x)|c=0.7857\sup_{x\in\mathbb{R}}|\operatorname{Ai}(x)|\leq c=0.7857\ldots follows by limnn1/3J[2n+un1/3u](2n)=Ai(u)\lim_{n\to\infty}n^{1/3}J_{[2n+un^{1/3}u]}(2n)=\operatorname{Ai}(u) (see also (3.2.23) of [1]) and the bound of Landau [44]. For any x0.01x\geq 0.01, the bound (A.4), is better that the bound in (A.3) and e0.01>ce^{-0.01}>c.

|Ai(x)|ex,x|\operatorname{Ai}(x)|\leq e^{-x},\quad x\in\mathbb{R} (A.3)

and, see Equation 9.7.15 of [28],

|Ai(x)|12πx1/4e23x3/2,x0.|\operatorname{Ai}(x)|\leq\frac{1}{2\sqrt{\pi}x^{1/4}}e^{-\frac{2}{3}x^{3/2}},\quad x\geq 0. (A.4)

from which it follows that for all x,u0x,u\geq 0,

|Ai(x+u2)|1ue23(x+u2)3/21ue23u3xu,|\operatorname{Ai}(x+u^{2})|\leq\frac{1}{\sqrt{u}}e^{-\frac{2}{3}(x+u^{2})^{3/2}}\leq\frac{1}{\sqrt{u}}e^{-\frac{2}{3}u^{3}-xu}, (A.5)

using that 23(x+u2)3/223u3+ux\frac{2}{3}(x+u^{2})^{3/2}\geq\frac{2}{3}u^{3}+ux since yy3/2y\mapsto y^{3/2} is convex on the positive real line.

A.2 Bounds on LPP

Lemma A.1 (Theorem 2 of [45]).

There exist constants C,c>0C,c>0 such that

(L(0,0),(N+v(2N)2/3,Nv(2N)2/3)4N24/3v2N1/3x24/3N1/3)Cecx3,\operatorname{\mathbbm{P}}(L_{(0,0),(N+v(2N)^{2/3},N-v(2N)^{2/3})}\leq 4N-2^{4/3}v^{2}N^{1/3}-x2^{4/3}N^{1/3})\leq Ce^{-cx^{3}}, (A.6)

for all N2/3x>0N^{2/3}\gg x>0 and N1/3|v|N^{1/3}\gg|v|.

Using Riemann-Hilbert methods like as in the case of geometric and Poissonian LPP, see [9, 8, 4, 5] for the Riemann-Hilbert problem for exponential LPP, it should be possible to get the optimal constant cc as well (expected to be 112\frac{1}{12} like for the lower tail of the GUE Tracy-Widom distribution function), but we do not require the optimal constant in the lower tail estimate for our purposes. A simple corollary is the following.

The following corollary follows from the inequality L(0,0),2NL(0,0),(N,N)L_{(0,0),\mathcal{L}_{2N}}\geq L_{(0,0),(N,N)}.

Corollary A.2.

There exist constants C,c>0C,c>0 such that

(L(0,0),2N4Nx24/3N1/3)Cecx3,\operatorname{\mathbbm{P}}(L_{(0,0),\mathcal{L}_{2N}}\leq 4N-x2^{4/3}N^{1/3})\leq Ce^{-cx^{3}}, (A.7)

for all N2/3x>0N^{2/3}\gg x>0.

Next we would like to have a sharp upper tail moderate deviation estimate from the last passage time from (0,0)(0,0) to the half line {(N,N)+v(2N)2/3(1,1):vu}\{(N,N)+v(2N)^{2/3}(1,-1):v\geq u\}.

Lemma A.3.

Let 𝒟u=vuJ(v){\cal D}_{u}=\cup_{v\geq u}J(v) with J(v)=(N,N)+v(2N)2/3(1,1)J(v)=(N,N)+v(2N)^{2/3}(1,-1). Then, for N1/9u>0N^{1/9}\gg u>0 and N2/9s>0N^{2/9}\gg s>0,

(L(0,0),𝒟u4Nu224/3N1/3+s24/3N1/3)Ce43s3/2smin{s,u}\operatorname{\mathbbm{P}}(L_{(0,0),{\cal D}_{u}}\geq 4N-u^{2}2^{4/3}N^{1/3}+s2^{4/3}N^{1/3})\leq\frac{Ce^{-\frac{4}{3}s^{3/2}}}{s\min\{\sqrt{s},u\}} (A.8)

for some constant CC.

Proof.

Let ~={(x,y)2|x+y=0,y0}\tilde{\mathcal{L}}=\{(x,y)\in\mathbb{Z}^{2}\,|\,x+y=0,y\geq 0\}. Then for any aa,

(L(0,0),𝒟ua)=(L~,J(u)a).\operatorname{\mathbbm{P}}(L_{(0,0),{\cal D}_{u}}\geq a)=\operatorname{\mathbbm{P}}(L_{\tilde{\mathcal{L}},J(u)}\geq a). (A.9)

Next we use the well-known connection between TASEP and LPP (see e.g. Equation (1.15) of [6], which generalizes [38]). In our case, the connection with TASEP with half-flat initial condition, i.e., xk(0)=2kx_{k}(0)=-2k for k0k\geq 0 are the positions of TASEP particles at time 0. This gives

(L~,J(u)t)\displaystyle\operatorname{\mathbbm{P}}(L_{\tilde{\mathcal{L}},J(u)}\geq t) =(xm(t)2u(2N)2/3)\displaystyle=\operatorname{\mathbbm{P}}(x_{m}(t)\leq 2u(2N)^{2/3}) (A.10)
=n1(1)nn!x1,,xn<2u(2N)2/3det[Km,t(xi,xj)]1i,jn,\displaystyle=-\sum_{n\geq 1}\frac{(-1)^{n}}{n!}\sum_{x_{1},\ldots,x_{n}<2u(2N)^{2/3}}\det[K_{m,t}(x_{i},x_{j})]_{1\leq i,j\leq n},

with m=Nu(2N)2/3m=N-u(2N)^{2/3} and the sum is on x1,,xnx_{1},\ldots,x_{n} being integers below 2u(2N)2/32u(2N)^{2/3}; the kernel is given by [20]

Km,t(x,y)=1(2πi)2Σ0dvv+1Σ1,1v𝑑weΦ(v;m,y)eΦ(w;m,x)(1wv11+w+v),K_{m,t}(x,y)=\frac{1}{(2\pi{\rm i})^{2}}\oint_{\Sigma_{0}}\frac{dv}{v+1}\oint_{\Sigma_{-1,-1-v}}dw\frac{e^{\Phi(v;m,y)}}{e^{\Phi(w;m,x)}}\left(\frac{1}{w-v}-\frac{1}{1+w+v}\right), (A.11)

with Φ(v;m,y)=vt+(y+m)ln(1+v)mln(v)\Phi(v;m,y)=-vt+(y+m)\ln(1+v)-m\ln(v). The contours are simple paths anticlockwise oriented, with Σ0\Sigma_{0} enclosing only the pole at 0, while Σ1,1v\Sigma_{-1,-1-v} enclosing the poles at 1-1 and 1v-1-v. Consider the scaling

xi=2u(2N)2/3ξi24/3N1/3,m=Nu(2N)2/3,\displaystyle x_{i}=2u(2N)^{2/3}-\xi_{i}2^{4/3}N^{1/3},\quad m=N-u(2N)^{2/3}, (A.12)
t=4Nu224/3N1/3+s24/3N1/3,KN(ξi,ξj)=24/3N1/3Km,t(xi,xj).\displaystyle t=4N-u^{2}2^{4/3}N^{1/3}+s2^{4/3}N^{1/3},\quad K_{N}(\xi_{i},\xi_{j})=2^{4/3}N^{1/3}K_{m,t}(x_{i},x_{j}).

Then, with δ=24/3N1/3\delta=2^{-4/3}N^{-1/3}, we have

|(A.10)|n11n!ξ1,,ξnδδn|det[KN(ξi,ξj)]1i,jn|.|\eqref{eq1.10}|\leq\sum_{n\geq 1}\frac{1}{n!}\sum_{\xi_{1},\ldots,\xi_{n}\in\delta\mathbb{N}}\delta^{n}\left|\det[K_{N}(\xi_{i},\xi_{j})]_{1\leq i,j\leq n}\right|. (A.13)

The reason why we singled out the δn\delta^{n} is because one can think that when δ\delta is small, the Riemann sum ξ1,,ξnδδnf(ξ1,,ξn)\sum_{\xi_{1},\ldots,\xi_{n}\in\delta\mathbb{N}}\delta^{n}f(\xi_{1},\ldots,\xi_{n}) is close to the integral +n𝑑ξ1𝑑ξnf(ξ1,,ξn)\int_{\mathbb{R}_{+}^{n}}d\xi_{1}\ldots d\xi_{n}f(\xi_{1},\ldots,\xi_{n}). Actually, from the exponential bound we get for f(ξ1,,ξn)f(\xi_{1},\ldots,\xi_{n}) we can bound the estimate with the integral (starting from δ-\delta instead of 0 to be precise).

Now we give the choice of the paths for vv and ww (which are such that they are steep descent). Since it is quite standard, we just indicate some main steps (see e.g. [20] or Lemma 2.7 of [24] for a case very similar to the one in this paper). Choose the paths as

v\displaystyle v =(12ε1)eiϕ,ϕ[π,π),ε1=24/3N1/3u(1+su1),\displaystyle=-(\tfrac{1}{2}-\varepsilon_{1})e^{{\rm i}\phi},\,\phi\in[-\pi,\pi),\quad\varepsilon_{1}=2^{-4/3}N^{-1/3}u(1+\sqrt{s}u^{-1}), (A.14)
w\displaystyle w =1+(12+ε2)eiθ,θ[π,π),ε2=24/3N1/3u(1su1).\displaystyle=-1+(\tfrac{1}{2}+\varepsilon_{2})e^{{\rm i}\theta},\,\theta\in[-\pi,\pi),\quad\varepsilon_{2}=2^{-4/3}N^{-1/3}u(1-\sqrt{s}u^{-1}).

On the chosen paths, we have, for uN1/3u\ll N^{1/3} and sN2/3s\ll N^{2/3}

Re(Φ(v;m,xj))\displaystyle\operatorname{Re}(\Phi(v;m,x_{j})) Φ(1/2+ε1;m;xj)(1cos(ϕ))2sN2/3,\displaystyle\leq\Phi(-1/2+\varepsilon_{1};m;x_{j})-(1-\cos(\phi))2\sqrt{s}N^{2/3}, (A.15)
Re(Φ(w;m,xi))\displaystyle\operatorname{Re}(-\Phi(w;m,x_{i})) Φ(1/2+ε2;m;xi)(1cos(θ))2sN2/3.\displaystyle\leq-\Phi(-1/2+\varepsilon_{2};m;x_{i})-(1-\cos(\theta))2\sqrt{s}N^{2/3}.

Furthermore

|1v+1|2,|1wv|21/3N1/3s,|11+w+v|22/3N1/3u,\left|\frac{1}{v+1}\right|\leq 2,\quad\left|\frac{1}{w-v}\right|\leq\frac{2^{1/3}N^{1/3}}{\sqrt{s}},\quad\left|\frac{1}{1+w+v}\right|\leq\frac{2^{2/3}N^{1/3}}{u}, (A.16)

as the maximal value of the first expression is obtained for the values θ=ϕ=0\theta=\phi=0, i.e., when vv and ww are closest to each other, while 1+w1+w and vv are two circles around 0 with radius difference ε1ε2\varepsilon_{1}-\varepsilon_{2}. Putting all together and noticing that ππ𝑑θec1(1cos(θ))C/c1\int_{-\pi}^{\pi}d\theta e^{-c_{1}(1-\cos(\theta))}\leq C/\sqrt{c_{1}} we finally obtain

|KN(ξi,ξj)|Csmin{s,u}eΦ(1/2+ε1;m;xj)Φ(1/2+ε2;m;xi).|K_{N}(\xi_{i},\xi_{j})|\leq\frac{C}{\sqrt{s}\min\{\sqrt{s},u\}}e^{\Phi(-1/2+\varepsilon_{1};m;x_{j})-\Phi(-1/2+\varepsilon_{2};m;x_{i})}. (A.17)

Taylor expansion gives

Φ(12+ε1;m;xj)\displaystyle\Phi(-\tfrac{1}{2}+\varepsilon_{1};m;x_{j}) Φ(12+ε2;m;xi)\displaystyle-\Phi(-\tfrac{1}{2}+\varepsilon_{2};m;x_{i}) (A.18)
=g(ξj)g(ξi)43s3/2(1+E2)2s(ξi+ξj)(1+E1)\displaystyle=g(\xi_{j})-g(\xi_{i})-\frac{4}{3}s^{3/2}(1+E_{2})-2\sqrt{s}(\xi_{i}+\xi_{j})(1+E_{1})
g(ξj)g(ξi)43s3/2s(ξi+ξj)+1\displaystyle\leq g(\xi_{j})-g(\xi_{i})-\frac{4}{3}s^{3/2}-\sqrt{s}(\xi_{i}+\xi_{j})+1

for all NN large enough and for some explicit conjugation terms g(ξ)g(\xi) (which cancel out exactly when computing the determinant), where E1=𝒪(sN2/3;uN1/3)=o(1)E_{1}=\mathcal{O}(sN^{-2/3};uN^{-1/3})=o(1) and s3/2E2=𝒪(s5/2N2/3;s1/2u4N2/3)=o(1)s^{3/2}E_{2}=\mathcal{O}(s^{5/2}N^{-2/3};s^{1/2}u^{4}N^{-2/3})=o(1) do not depend on ξi,ξj\xi_{i},\xi_{j}.

Inserting (A.18) and (A.17) in (A.13),using the Hadamard bound and interchanging the order of summation, we get

(A.13)\displaystyle\eqref{eq1.16} n1nn/2n!Cnen(smin{s,u})ne4n3s3/2j=1nξjδδe2sξj\displaystyle\leq\sum_{n\geq 1}\frac{n^{n/2}}{n!}\frac{C^{n}e^{n}}{(\sqrt{s}\min\{\sqrt{s},u\})^{n}}e^{-\frac{4n}{3}s^{3/2}}\prod_{j=1}^{n}\sum_{\xi_{j}\in\delta\mathbb{N}}\delta e^{-2\sqrt{s}\xi_{j}} (A.19)
n1nn/2n!(Csmin{s,u}e43s3/212s)nC′′smin{s,u}e43s3/2,\displaystyle\leq\sum_{n\geq 1}\frac{n^{n/2}}{n!}\left(\frac{C^{\prime}}{\sqrt{s}\min\{\sqrt{s},u\}}e^{-\frac{4}{3}s^{3/2}}\frac{1}{2\sqrt{s}}\right)^{n}\leq\frac{C^{\prime\prime}}{s\min\{\sqrt{s},u\}}e^{-\frac{4}{3}s^{3/2}},

for some constants C,C′′C^{\prime},C^{\prime\prime}. To bound the sum over ξi\xi_{i}, we simply used ξδδeαξ0eαξ𝑑ξ=α1\sum_{\xi\in\delta\mathbb{N}}\delta e^{-\alpha\xi}\leq\int_{0}^{\infty}e^{-\alpha\xi}d\xi=\alpha^{-1}, but one could also compute the geometric sum explicitly. ∎

The next result gives upper tail bounds for point-to-line LPP.

Lemma A.4.

For N2/9s0N^{2/9}\gg s\geq 0,

(L(0,0),2N4N+s24/3N1/3)Ce43s3/2max{1,s3/4}\operatorname{\mathbbm{P}}(L_{(0,0),\mathcal{L}_{2N}}\geq 4N+s2^{4/3}N^{1/3})\leq\frac{Ce^{-\frac{4}{3}s^{3/2}}}{\max\{1,s^{3/4}\}} (A.20)

for some constant CC.

Proof.

We need only to prove the result for s1s\geq 1 (or any other constant). Then by appropriate choice of the constant CC, the result holds for all s0s\geq 0 (just take CC such that the upper bound is larger than the trivial bound 11).

We have

(L(0,0),2Nt)=(L0,(N,N)t)=(xN(t)0),\operatorname{\mathbbm{P}}(L_{(0,0),\mathcal{L}_{2N}}\geq t)=\operatorname{\mathbbm{P}}(L_{\mathcal{L}_{0},(N,N)}\geq t)=\operatorname{\mathbbm{P}}(x_{N}(t)\leq 0), (A.21)

where xn(t)x_{n}(t) it the position of TASEP particle nn at time tt with initial condition xn(0)=2nx_{n}(0)=-2n, nn\in\mathbb{Z}. The distribution of TASEP particles are given in terms of a Fredholm determinant

(xN(t)0)=n1(1)nn!x1,,xn0det[KN,t(xi,xj)]1i,jn,\operatorname{\mathbbm{P}}(x_{N}(t)\leq 0)=-\sum_{n\geq 1}\frac{(-1)^{n}}{n!}\sum_{x_{1},\ldots,x_{n}\leq 0}\det[K_{N,t}(x_{i},x_{j})]_{1\leq i,j\leq n}, (A.22)

with the kernel given by [19]

KN,t(x,y)=12πiΣ0dvv(1+v)y+2N(v)2N+xet(1+2v)K_{N,t}(x,y)=\frac{1}{2\pi{\rm i}}\oint_{\Sigma_{0}}\frac{dv}{v}\frac{(1+v)^{y+2N}}{(-v)^{2N+x}}e^{-t(1+2v)} (A.23)

with Σ0\Sigma_{0} a simple path anticlockwise oriented enclosing only the pole at 0. Setting xi=ξi24/3N1/3x_{i}=-\xi_{i}2^{4/3}N^{1/3} and t=4N+s24/3N1/3t=4N+s2^{4/3}N^{1/3} we obtain

KN,t(xi,xj)=12πiΣ0dvveNf0(v)+24/3N1/3f1(v)K_{N,t}(x_{i},x_{j})=\frac{1}{2\pi{\rm i}}\oint_{\Sigma_{0}}\frac{dv}{v}e^{Nf_{0}(v)+2^{4/3}N^{1/3}f_{1}(v)} (A.24)

with

f0(v)\displaystyle f_{0}(v) =4(1+2v)+2ln(1+v)2ln(v),\displaystyle=-4(1+2v)+2\ln(1+v)-2\ln(-v), (A.25)
f1(v)\displaystyle f_{1}(v) =s(1+2v)ξiln(1+v)+ξjln(v).\displaystyle=-s(1+2v)-\xi_{i}\ln(1+v)+\xi_{j}\ln(-v).

Consider the path parameterized by v=ρeiθv=-\rho e^{{\rm i}\theta}, θ[π,π)\theta\in[-\pi,\pi). Then due to

dRe(f0(v))dθ=2ρsin(θ)[41|1+v|2],\frac{d\operatorname{Re}(f_{0}(v))}{d\theta}=-2\rho\sin(\theta)\left[4-\frac{1}{|1+v|^{2}}\right], (A.26)

the path is steep descent for any ρ(0,1/2]\rho\in(0,1/2]. Let us choose the radius by

ρ=1/2s24/3N1/3.\rho=1/2-\sqrt{s}2^{-4/3}N^{-1/3}. (A.27)

For any s,ξi,ξj0s,\xi_{i},\xi_{j}\geq 0,

Re(24/3N1/3f1(v))24/3N1/3f1(ρ)\operatorname{Re}(2^{4/3}N^{1/3}f_{1}(v))\leq 2^{4/3}N^{1/3}f_{1}(-\rho) (A.28)

and for 0<sN2/30<s\ll N^{2/3},

Re(Nf0(v))Nf0(ρ)(1cos(θ))4sN2/3.\operatorname{Re}(Nf_{0}(v))\leq Nf_{0}(-\rho)-(1-\cos(\theta))4\sqrt{s}N^{2/3}. (A.29)

Finally, integrating over θ\theta leads to the following estimate on the kernel

24/3N1/3|KN,t(xi,xj)|CeNf0(ρ)+24/3N1/3f1(ρ).2^{4/3}N^{1/3}|K_{N,t}(x_{i},x_{j})|\leq Ce^{Nf_{0}(-\rho)+2^{4/3}N^{1/3}f_{1}(-\rho)}. (A.30)

A computation gives

Nf0(ρ)+24/3N1/3f1(ρ)\displaystyle Nf_{0}(-\rho)+2^{4/3}N^{1/3}f_{1}(-\rho) (A.31)
=g(ξi)g(ξj)43s3/2(1+𝒪(sN2/3))2s(ξi+ξj)(1+𝒪(sN2/3))\displaystyle=g(\xi_{i})-g(\xi_{j})-\frac{4}{3}s^{3/2}(1+\mathcal{O}(sN^{-2/3}))-2\sqrt{s}(\xi_{i}+\xi_{j})(1+\mathcal{O}(sN^{-2/3}))
g(ξi)g(ξj)43s3/2s(ξi+ξj)+1\displaystyle\leq g(\xi_{i})-g(\xi_{j})-\frac{4}{3}s^{3/2}-\sqrt{s}(\xi_{i}+\xi_{j})+1

for some conjugation function gg, where the error terms do not depend on ξi,ξj\xi_{i},\xi_{j}. Thus we have the following estimate on the kernel

24/3N1/3|KN,t(xi,xj)|Cs1/4eg(ξi)g(ξj)43s3/2(ξi+ξj)s2^{4/3}N^{1/3}|K_{N,t}(x_{i},x_{j})|\leq\frac{C}{s^{1/4}}e^{g(\xi_{i})-g(\xi_{j})-\frac{4}{3}s^{3/2}-(\xi_{i}+\xi_{j})\sqrt{s}} (A.32)

for some constant CC.

Plugging this estimate into (A.22) and using Hadamard inequality like in the proof of Lemma A.3 we finally get the claimed bound (A.20). ∎

The final result we need is an upper tail bound for interval-to-line LPP. The proof is similar to that of Lemma 10.6 in [14], but with the optimal exponent.

Proposition A.5.

Let ={(k,k)|12(2N)2/3k12(2N)2/3}{\cal M}=\{(k,-k)|-\tfrac{1}{2}(2N)^{2/3}\leq k\leq\tfrac{1}{2}(2N)^{2/3}\} and 2N={(N+k,Nk),k}\mathcal{L}_{2N}=\{(N+k,N-k),k\in\mathbb{Z}\}. Then, for N2/3s0N^{2/3}\gg s\geq 0,

(suppLp,2N4N+s24/3N1/3)Cmax{1,s}e43s3/2,\operatorname{\mathbbm{P}}\bigg{(}\sup_{p\in\cal M}L_{p,\mathcal{L}_{2N}}\geq 4N+s2^{4/3}N^{1/3}\bigg{)}\leq C\max\{1,s\}e^{-\frac{4}{3}s^{3/2}}, (A.33)

for some constant C>0C>0.

Proof.

Notice that it is enough to prove the bound only for s1s\geq 1 (or any other positive constant).

Recall the notation I(v)=v(2N)2/3(1,1)I(v)=v(2N)^{2/3}(1,-1). So {\cal M} is the union of points I(v)I(v) with |v|12|v|\leq\tfrac{1}{2}. Define the point ww

w=(ε~N,ε~N).w=(-\tilde{\varepsilon}N,-\tilde{\varepsilon}N). (A.34)

Let us divide \cal M into union of ε~2/3\tilde{\varepsilon}^{-2/3} segments of size ε~2/3(2N)2/3\tilde{\varepsilon}^{2/3}(2N)^{2/3}, say =k=1ε~2/3k{\cal M}=\bigcup_{k=1}^{\tilde{\varepsilon}^{-2/3}}{\cal M}_{k}. Then

suppLp,2N=sup1kε~2/3suppkLp,2N.\sup_{p\in{\cal M}}L_{p,\mathcal{L}_{2N}}=\sup_{1\leq k\leq\tilde{\varepsilon}^{-2/3}}\sup_{p\in{\cal M}_{k}}L_{p,\mathcal{L}_{2N}}. (A.35)

Using union bound the the fact that each of the suppkLp,2N\sup_{p\in{\cal M}_{k}}L_{p,\mathcal{L}_{2N}} has the same distribution, we get

(suppLp,2NS)ε~2/3(sup|v|12ε~2/3LI(v),2NS).\operatorname{\mathbbm{P}}\bigg{(}\sup_{p\in\cal M}L_{p,\mathcal{L}_{2N}}\geq S\bigg{)}\leq\tilde{\varepsilon}^{-2/3}\operatorname{\mathbbm{P}}\bigg{(}\sup_{|v|\leq\frac{1}{2}\tilde{\varepsilon}^{2/3}}L_{I(v),\mathcal{L}_{2N}}\geq S\bigg{)}. (A.36)

Next we want to bound the last probability in (A.36).

The idea is the following. For fixed vv, we know that the same tail estimates we want to prove hold for LI(v),2NL_{I(v),\mathcal{L}_{2N}}. We also have an estimate of the upper tail for Lw,2NL_{w,\mathcal{L}_{2N}} and that with positive probability, Lw,I(v)L_{w,I(v)} can not be too small in the N1/3N^{1/3} scale. So, the fluctuations coming from maximizing LI(v),2NL_{I(v),\mathcal{L}_{2N}} over |v|12(1ε~)2/3|v|\leq\tfrac{1}{2}(1-\tilde{\varepsilon})^{2/3} can not be compensated fully from the fluctuations of Lw,I(v)L_{w,I(v)}. This will imply our claim.

Let L^w,I(u)=Lw,I(u)ωI(u)\hat{L}_{w,I(u)}=L_{w,I(u)}-\omega_{I(u)} be the LPP from ww to I(u)I(u) without the random variable at the end-point I(u)I(u) 666Removing the end-point does not influence the asymptotics and bounds, but it has the property that L^w,I(u)\hat{L}_{w,I(u)} and LI(u),2NL_{I(u),\mathcal{L}_{2N}} are independent random variables and the concatenation property holds true.. We know from [31, 47] that

vL^w,I(vε~2/3)4ε~N24/3(ε~N)1/3v\mapsto\frac{\hat{L}_{w,I(v\tilde{\varepsilon}^{2/3})}-4\tilde{\varepsilon}N}{2^{4/3}(\tilde{\varepsilon}N)^{1/3}} (A.37)

is tight in the space of continuous functions on compact sets. As a consequence there exists a constant C>0C^{\prime}>0 such that the event

={inf|v|12ε~2/3L^w,I(v)4ε~NC24/3(ε~N)1/3}{\cal H}=\Big{\{}\inf_{|v|\leq\frac{1}{2}\tilde{\varepsilon}^{2/3}}\hat{L}_{w,I(v)}\geq 4\tilde{\varepsilon}N-C^{\prime}2^{4/3}(\tilde{\varepsilon}N)^{1/3}\Big{\}} (A.38)

satisfies ()1/2\operatorname{\mathbbm{P}}({\cal H})\geq 1/2.

We also have

Lw,2Nsup|v|12ε~2/3(L^w,I(v)+LI(v),2N)inf|v|12ε~2/3L^w,I(v)+sup|v|12ε~2/3LI(v),2N.L_{w,\mathcal{L}_{2N}}\geq\sup_{|v|\leq\frac{1}{2}\tilde{\varepsilon}^{2/3}}(\hat{L}_{w,I(v)}+L_{I(v),\mathcal{L}_{2N}})\geq\inf_{|v|\leq\frac{1}{2}\tilde{\varepsilon}^{2/3}}\hat{L}_{w,I(v)}+\sup_{|v|\leq\frac{1}{2}\tilde{\varepsilon}^{2/3}}L_{I(v),\mathcal{L}_{2N}}. (A.39)

Define the event

𝒢={sup|v|12ε~2/3LI(v),2N4N+s24/3N1/3}.{\cal G}=\Big{\{}\sup_{|v|\leq\frac{1}{2}\tilde{\varepsilon}^{2/3}}L_{I(v),\mathcal{L}_{2N}}\geq 4N+s2^{4/3}N^{1/3}\Big{\}}. (A.40)

Then,

(Lw,2N4(1+ε~)N+(sCε~1/3)N1/3)(𝒢)=()(𝒢),\operatorname{\mathbbm{P}}(L_{w,\mathcal{L}_{2N}}\geq 4(1+\tilde{\varepsilon})N+(s-C^{\prime}\tilde{\varepsilon}^{1/3})N^{1/3})\geq\operatorname{\mathbbm{P}}({\cal H}\cap{\cal G})=\operatorname{\mathbbm{P}}({\cal H})\operatorname{\mathbbm{P}}({\cal G}), (A.41)

where we used independence of \cal H and 𝒢\cal G. Thus we have shown that

(𝒢)2(Lw,2N4(1+ε~)N+(sCε~1/3)N1/3)2Ce43(sCε~1/3)3/2(1+ε~)1/2\operatorname{\mathbbm{P}}({\cal G})\leq 2\operatorname{\mathbbm{P}}(L_{w,\mathcal{L}_{2N}}\geq 4(1+\tilde{\varepsilon})N+(s-C^{\prime}\tilde{\varepsilon}^{1/3})N^{1/3})\leq 2Ce^{-\frac{4}{3}\frac{(s-C^{\prime}\tilde{\varepsilon}^{1/3})^{3/2}}{(1+\tilde{\varepsilon})^{1/2}}} (A.42)

where the last inequality holds for N2/3sCε~1/3>1N^{2/3}\gg s-C^{\prime}\tilde{\varepsilon}^{1/3}>1 by Lemma A.4. Replacing this into (A.36) we get

(suppLp,2N4N+s24/3N1/3)2Cε~2/3e43(sCε~1/3)3/2(1+ε~)1/2.\operatorname{\mathbbm{P}}\bigg{(}\sup_{p\in\cal M}L_{p,\mathcal{L}_{2N}}\geq 4N+s2^{4/3}N^{1/3}\bigg{)}\leq 2C\tilde{\varepsilon}^{-2/3}e^{-\frac{4}{3}\frac{(s-C^{\prime}\tilde{\varepsilon}^{1/3})^{3/2}}{(1+\tilde{\varepsilon})^{1/2}}}. (A.43)

Consider first s1s\geq 1. Then taking ε~=1/s3/2\tilde{\varepsilon}=1/s^{3/2} (notice that ε~N1\tilde{\varepsilon}N\gg 1 since we assumed sN2/3s\ll N^{2/3}), we get

(suppLp,2N4N+s24/3N1/3)Cse43s3/2\operatorname{\mathbbm{P}}\bigg{(}\sup_{p\in\cal M}L_{p,\mathcal{L}_{2N}}\geq 4N+s2^{4/3}N^{1/3}\bigg{)}\leq Cse^{-\frac{4}{3}s^{3/2}} (A.44)

for some new constant C>0C>0. ∎

References

  • [1] M. Abramowitz and I.A. Stegun. Pocketbook of Mathematical Functions. Verlag Harri Deutsch, Thun-Frankfurt am Main, 1984.
  • [2] M. Adler and P. van Moerbeke. PDE’s for the joint distribution of the Dyson, Airy and Sine processes. Ann. Probab., 33:1326–1361, 2005.
  • [3] R. Arratia, S. Garibaldi, and A.W. Hales. The van den Berg–Kesten–Reimer operator and inequality for infinite spaces. Bernoulli, 24(1):433–448, 2018.
  • [4] J. Baik. Painlevé expressions for LOE, LSE and interpolating ensembles. Int. Math. Res. Notices, 2002:1739–1789, 2002.
  • [5] J. Baik, R. Buckingham, and J. DiFranco. Asymptotics of Tracy-Widom distributions and the total integral of a Painleve II function. Comm. Math. Phys., 280:463–497, 2008.
  • [6] J. Baik, P.L. Ferrari, and S. Péché. Limit process of stationary TASEP near the characteristic line. Comm. Pure Appl. Math., 63:1017–1070, 2010.
  • [7] J. Baik, K. Liechty, and G. Schehr. On the joint distribution of the maximum and its position of the Airy2 process minus a parabola. J. Math. Phys., 53:083303, 2012.
  • [8] J. Baik and E.M. Rains. Algebraic aspects of increasing subsequences. Duke Math. J., 109:1–65, 2001.
  • [9] J. Baik and E.M. Rains. The asymptotics of monotone subsequences of involutions. Duke Math. J., 109:205–281, 2001.
  • [10] G. Barraquand, A. Krajenbrink, and P. Le Doussal. Half-space stationary Kardar-Parisi-Zhang equation beyond the Brownian case. arXiv:2202.10487, 2022.
  • [11] R. Basu and S. Ganguly. Time correlation exponents in last passage percolation. In M.E. Vares, R. Fernández, L.R. Fontes, and C.M. Newman, editors, In and Out of Equilibrium 3: Celebrating Vladas Sidoravicius, volume 77 of Progress in Probability. Birkhäuser, 2021.
  • [12] R. Basu, S. Ganguly, and L. Zhang. Temporal correlation in last passage percolation with flat initial condition via brownian comparison. Communications in Mathematical Physics, 383(3):1805–1888, 2021.
  • [13] R. Basu, S. Sarkar, and A. Sly. Coalescence of geodesics in exactly solvable models of last passage percolation. Journal of Mathematical Physics, 60(9):093301, 2019.
  • [14] R. Basu, V. Sidoravicius, and A. Sly. Last passage percolation with a defect line and the solution of the slow bond problem. arXiv:1408.3464, 2014.
  • [15] F. Bornemann. On the numerical evaluation of Fredholm determinants. Math. Comput., 79:871–915, 2009.
  • [16] F. Bornemann, P.L. Ferrari, and M. Prähofer. The Airy1 process is not the limit of the largest eigenvalue in GOE matrix diffusion. J. Stat. Phys., 133:405–415, 2008.
  • [17] A. Borodin and P.L. Ferrari. Large time asymptotics of growth models on space-like paths I: PushASEP. Electron. J. Probab., 13:1380–1418, 2008.
  • [18] A. Borodin, P.L. Ferrari, and M. Prähofer. Fluctuations in the discrete TASEP with periodic initial configurations and the Airy1 process. Int. Math. Res. Papers, 2007:rpm002, 2007.
  • [19] A. Borodin, P.L. Ferrari, M. Prähofer, and T. Sasamoto. Fluctuation properties of the TASEP with periodic initial configuration. J. Stat. Phys., 129:1055–1080, 2007.
  • [20] A. Borodin, P.L. Ferrari, and T. Sasamoto. Transition between Airy1 and Airy2 processes and TASEP fluctuations. Comm. Pure Appl. Math., 61:1603–1629, 2008.
  • [21] A. Borodin and V. Gorin. Lectures on integrable probability. In Probability and Statistical Physics in St. Petersburg, volume 91, pages 155–214. Proceedings of Symposia in Pure Mathematics, AMS, 2016.
  • [22] A. Borodin and S. Péché. Airy Kernel with Two Sets of Parameters in Directed Percolation and Random Matrix Theory. J. Stat. Phys., 132:275–290, 2008.
  • [23] O. Busani and P.L. Ferrari. Universality of the geodesic tree in last passage percolation. Ann. Probab., 50:90–130, 2022.
  • [24] S. Chhita, P.L. Ferrari, and H. Spohn. Limit distributions for KPZ growth models with spatially homogeneous random initial conditions. Ann. Appl. Probab., 28:1573–1603, 2018.
  • [25] I. Corwin. The Kardar-Parisi-Zhang equation and universality class. Random Matrices: Theory Appl., 01:1130001, 2012.
  • [26] I. Corwin, P.L. Ferrari, and S. Péché. Universality of slow decorrelation in KPZ models. Ann. Inst. H. Poincaré Probab. Statist., 48:134–150, 2012.
  • [27] E. Dimitrov. Two-point convergence of the stochastic six-vertex model to the Airy process. arXiv:2006.15934, 2020.
  • [28] NIST Digital Library of Mathematical Functions. http://dlmf.nist.gov/, Release 1.1.5 of 2022-03-15. F. W. J. Olver, A. B. Olde Daalhuis, D. W. Lozier, B. I. Schneider, R. F. Boisvert, C. W. Clark, B. R. Miller, B. V. Saunders, H. S. Cohl, and M. A. McClain, eds.
  • [29] P.L. Ferrari. The universal Airy1 and Airy2 processes in the Totally Asymmetric Simple Exclusion Process. In J. Baik, T. Kriecherbauer, L-C. Li, K. McLaughlin, and C. Tomei, editors, Integrable Systems and Random Matrices: In Honor of Percy Deift, Contemporary Math., pages 321–332. Amer. Math. Soc., 2008.
  • [30] P.L. Ferrari. From interacting particle systems to random matrices. J. Stat. Mech., page P10016, 2010.
  • [31] P.L. Ferrari and A. Occelli. Universality of the GOE Tracy-Widom distribution for TASEP with arbitrary particle density. Eletron. J. Probab., 23(51):1–24, 2018.
  • [32] P.L. Ferrari and A. Occelli. Time-time covariance for last passage percolation with generic initial profile. Math. Phys. Anal. Geom., 22:1, 2019.
  • [33] P.L. Ferrari and A. Occelli. Time-time covariance for last passage percolation in half-space. arXiv:2204.06782, 2022.
  • [34] P.L. Ferrari and H. Spohn. Random Growth Models. In G. Akemann, J. Baik, and P. Di Francesco, editors, The Oxford handbook of random matrix theory, pages 782–801. Oxford Univ. Press, Oxford, 2011.
  • [35] P.L. Ferrari and H. Spohn. On time correlations for KPZ growth in one dimension. SIGMA, 12:074, 2016.
  • [36] A. Hammond and S. Sarkar. Modulus of continuity for polymer fluctuations and weight profiles in Poissonian last passage percolation. Electron. J. Probab., 25:38 pp., 2020.
  • [37] W. Hoeffding. Masstabinvariante Korrelationstheorie. Schriften Math. Inst. Univ. Berlin, 5:181–233, 1940.
  • [38] K. Johansson. Shape fluctuations and random matrices. Comm. Math. Phys., 209:437–476, 2000.
  • [39] K. Johansson. Discrete polynuclear growth and determinantal processes. Comm. Math. Phys., 242:277–329, 2003.
  • [40] K. Johansson. The arctic circle boundary and the Airy process. Ann. Probab., 33:1–30, 2005.
  • [41] M. Kardar, G. Parisi, and Y.Z. Zhang. Dynamic scaling of growing interfaces. Phys. Rev. Lett., 56:889–892, 1986.
  • [42] J.H.B. Kemperman. On the FKG-inequality for measures on a partially ordered space. Indagationes Mathematicae (Proceedings), 80(4):313–331, 1977. North-Holland.
  • [43] H. Kesten. First-passage percolation. In P. Picco and J. San Martin, editors, From Classical to Modern Probability: CIMPA Summer School 2001, pages 93–143. Birkhäuser Basel, Basel, 2003.
  • [44] L.J. Landau. Bessel functions: monotonicity and bounds. J. London Math. Soc., 61:197–215, 2000.
  • [45] M. Ledoux and B. Rider. Small deviations for beta ensembles. Electron. J. Probab., 15:1319–1343, 2010.
  • [46] L.P.R. Pimentel. Duality between coalescence times and exit points in last-passage percolation models. Ann. Probab., 44(5):3187–3206, 2016.
  • [47] L.P.R. Pimentel. Local Behavior of Airy Processes. J. Stat. Phys., 173:1614–1638, 2018.
  • [48] M. Prähofer and H. Spohn. Scale invariance of the PNG droplet and the Airy process. J. Stat. Phys., 108:1071–1106, 2002.
  • [49] J. Quastel. Introduction to KPZ. Current Developments in Mathematics, pages 125–194, 2011.
  • [50] J. Quastel and H. Spohn. The one-dimensional kpz equation and its universality class. J. Stat. Phys., 160:965–984, 2015.
  • [51] M. Reed and B. Simon. Methods of Modern Mathematical Physics III: Scattering theory. Academic Press, New York, 1978.
  • [52] T. Sasamoto. Spatial correlations of the 1D KPZ surface on a flat substrate. J. Phys. A, 38:L549–L556, 2005.
  • [53] G. Schehr. Extremes of NN vicious walkers for large NN: application to the directed polymer and KPZ interfaces. J. Stat. Phys., 149:385–410, 2012.
  • [54] B. Simon. Trace Ideals and Their Applications. American Mathematical Society, second edition edition, 2000.
  • [55] K.A. Takeuchi. An appetizer to modern developments on the Kardar–Parisi–Zhang universality class. Physica A, 504:77–105, 2016.
  • [56] H. Widom. On asymptotic for the Airy process. J. Stat. Phys., 115:1129–1134, 2004.
  • [57] L. Zhang. Optimal exponent for coalescence of finite geodesics in exponential last passage percolation. Electron. Commun. Probab., 25:14 pp., 2020.
  • [58] N. Zygouras. Some algebraic structures in the KPZ universality. arXiv:1812.07204, 2018.