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

A non-oriented first passage percolation model and statistical invariance by time reversal

Alejandro F. Ramírezlabel=e1 [    mark][email protected]    Santiago Sagliettilabel=e2 [    mark][email protected]    Lingyun Shaolabel=e3 [    mark][email protected] NYU-ECNU Institute of Mathematical Sciences, NYU Shanghai, Facultad de Matemáticas, Pontificia Universidad Católica de Chile, Department of Mathematics, University of Toronto
Abstract

We introduce and study a non-oriented first passage percolation model having a property of statistical invariance by time reversal. This model is defined in a graph having directed edges and the passage times associated with each set of outgoing edges from a given vertex are distributed according to a generalized Bernoulli–Exponential law and i.i.d. among vertices. We derive the statistical invariance property by time reversal through a zero-temperature limit of the random walk in Dirichlet environment model.

60K35, 60 K37, 82B43,
First passage percolation,
Statistical invariance by time reversal,
keywords:
[class=MSC]
keywords:
\startlocaldefs\endlocaldefs

, and

1 Introduction

We present a non-oriented first passage percolation (FPP) model satisfying a statistical invariance property by time reversal. This property is analogous to the statistical invariance property satisfied by the random walk in Dirichlet random environment [4].

Random walk in random environment (or RWRE, for short) is a challenging model for which several fundamental questions remain open, including the proof of a law of large numbers [6, 2]. In particular, in the multidimensional case, very few explicit computations can be made, so that for example, there are no general formulas for the velocity, asymptotic direction, the diffusion matrix in the case of Gaussian fluctuations, or the large deviation rate functions. The random walk in Dirichlet environment (RWDE) is a particular case of RWRE in which the marginal law of the environment at each site is given by a Dirichlet distribution [4]. Sabot proved in [3] that RWDE satisfies a property of satistical invariance by time reversal. This was used in several works (see [4]), to prove fundamental properties of this model, including explicit conditions for transient-recurrent behavior in dimension d=3d=3 [3] and an explicit formula for the asymptotic direction in [5]. Also, a particular case of the RWDE model corresponding to parameters which allow only oriented (or directed) movements of the random walk, called the Beta random walk, was studied by Barraquand and Corwin in [1], where they were able to proved Tracy–Widom GUE fluctuations of the quenched large deviation principle, thus establishing the belonging of this model to the KPZ universality class. Furthermore, the zero-temperature limit of the Beta random walk, which is a first passage percolation model called Exponential first passage percolation, was also studied in [1], where they showed that the fluctuations of the first passage times are also of the Tracy–Widom GUE type.

In this article we introduce a zero-temperature version of general RWDE, which we call the Bernoulli–Exponential first passage percolation model, which turns out to be a non-oriented extension of the exponential first passage percolation model from [1]. The main result of this article is the proof of a zero-temperature version of the statistical reversibility under time reversal property of this model. To our knowledge, this is the first model of FPP type satisfying such a property.

In principle, the definition of the Bernoulli–Exponential first passage percolation model can be understood following the principles of tropical geometry: starting from RWDE, addition is replaced with minimization and multiplication with addition. However, due to the presence of infinite sums in the probabilities involved in the RWDE, this correspondence cannot be made rigorous, and so in the end here the Bernoulli–Exponential first passage percolation model is defined and studied directly, without any specific reference to the RWDE. An exception to this is the proof of the statistical invariance property under time reversal for the Bernoulli–Exponential first passage percolation model, which is derived through a limiting procedure from the RWDE.

In what follows, in Section 2, we give a precise statement of the main results of the article. In Section 3, we derive the Bernoulli–Exponential first passage exponential model as a zero-temperature limit of the RWDE. In Section 4, we present a proof of the statistical invariance by time reversal of the model. Finally in Section 5, we apply the statistical invariance by time reversal property to compute some first passage times in specific graphs.

2 Main results

In order to define this model, we need to introduce first a multinomial version of the Bernoulli–Exponential random variable. To this end, let us recall the generalized Bernoulli distribution: given MM\in\mathbb{N} and a probability vector 𝐩=(p1,,pM)M\mathbf{p}=(p_{1},\dots,p_{M})\in\mathbb{R}^{M}, we say that a random vector B=(B1,,BM)B=(B_{1},\dots,B_{M}) has generalized Bernoulli distribution of order MM with parameters 𝐩\mathbf{p}, and denote it by BBeM(𝐩)B\sim\textrm{Be}_{M}(\mathbf{p}), if it satisfies

(B=ei)=pi,i=1,,M,\mathbb{P}(B=\mathrm{e}_{i})=p_{i},\quad i=1,\dots,M,

where ei\mathrm{e}_{i} denotes the ii-th canonical vector in M\mathbb{R}^{M}. We next define the Bernoulli–Exponential distribution of order MM.

Definition 2.1.

Given MM\in\mathbb{N} and a vector 𝐚=(a1,,aM)\mathbf{a}=(a_{1},\dots,a_{M}) with real positive entries, we define the generalized Bernoulli–Exponential distribution of order MM with parameters 𝐚\mathbf{a} as the distribution of the random vector

X=((1B1)E1,,(1BM)EM),X=((1-B_{1})E_{1},\ldots,(1-B_{M})E_{M}), (1)

where:

  1. \bullet

    B=(B1,,BM)B=(B_{1},\dots,B_{M}) has generalized Bernoulli distribution with parameters MM and 𝐩(𝐚)\mathbf{p}^{(\mathbf{a})}, where 𝐩(𝐚)=(p1(𝐚),,pM(𝐚))\mathbf{p}^{(\mathbf{a})}=(p_{1}^{(\mathbf{a})},\dots,p_{M}^{(\mathbf{a})}) is given by

    pi(𝐚):=aij=1Maj;p_{i}^{(\mathbf{a})}:=\frac{a_{i}}{\sum_{j=1}^{M}a_{j}}; (2)
  2. \bullet

    E=(E1,,EM)E=(E_{1},\dots,E_{M}) is a random vector independent of BB which has independent entries, with each entry EiE_{i} having an Exponential distribution of parameter aia_{i}, respectively.

In the sequel, we will write XBe-ExpM(𝐚)X\sim\textrm{Be-Exp}_{M}(\mathbf{a}) to indicate that XX has this distribution.

With Definition 2.1 at our disposal, we now define the generalized Bernoulli–Exponential first passage percolation model as follows. Let G=(V,)G=(V,\mathcal{E}) be a directed graph endowed with a family of positive weights 𝐚:=(ae)e\mathbf{a}:=(a_{e})_{e\in\mathcal{E}}. Assume that GG is finite and strongly connected, i.e., that for any x,yVx,y\in V there exists a path going from xx to yy, where by path we understand a finite vector π=(x0,,xn)\pi=(x_{0},\ldots,x_{n}) with x0,,xnVx_{0},\ldots,x_{n}\in V and (xi,xi+1)(x_{i},x_{i+1})\in\mathcal{E} for all 0in10\leq i\leq n-1. For each xVx\in V, define x:={e:e=(x,y) for some yV}\mathcal{E}_{x}:=\{e\in\mathcal{E}:e=(x,y)\text{ for some }y\in V\} to be the set of edges in \mathcal{E} leaving xx and write Mx:=#cM_{x}:=\#\mathcal{E}_{c} for their total number. Now consider a family ω=(ω(x))xV\omega=(\omega(x))_{x\in V} of independent random vectors where, for each xVx\in V, the vector ω(x)=(ω(x,e))ex\omega(x)=(\omega(x,e))_{e\in\mathcal{E}_{x}} has generalized Bernolli–Exponential distribution of order MxM_{x} with parameters 𝐚x:=(ae)ex\mathbf{a}_{x}:=(a_{e})_{e\in\mathcal{E}_{x}}. We will call the family ω\omega a generalized Bernoulli–Exponential environment on the graph GG with parameters 𝐚\mathbf{a}, and denote its law by BE(𝐚)\mathbb{P}_{\text{BE}}^{(\mathbf{a})}. Finally, we denote by Π(x,y)\Pi(x,y) the set of paths going from xx to yy. Now, we define the first passage time between two points x,yGx,y\in G by

T(x,y):=infπΠ(x,y)i=0n1w(xi,Δxi),T(x,y):=\inf_{\pi\in\Pi(x,y)}\sum_{i=0}^{n-1}w(x_{i},\Delta x_{i}), (3)

where Δxi:=(xi,xi+1)\Delta x_{i}:=(x_{i},x_{i+1}). We will call (T(x,y))x,yV(T(x,y))_{x,y\in V} the generalized Bernoulli-Exponential first passage percolation model on GG.

In order to state our main result, we first need to introduce the dual graph of G=(V,)G=(V,\mathcal{E}), which is defined as Gˇ:=(V,ˇ)\check{G}:=(V,\check{\mathcal{E}}), where

ˇ={(y,x):(x,y)},\check{\mathcal{E}}=\{(y,x):(x,y)\in\mathcal{E}\},

is the set of reversed edges of \mathcal{E}. In the following, we will write eˇ=(y,x)\check{e}=(y,x) for the reversal of the edge e=(x,y)e=(x,y). We endow Gˇ\check{G} with the family of reversed weights 𝐚ˇ:=(aˇeˇ)eˇˇ\check{\mathbf{a}}:=(\check{a}_{\check{e}})_{\check{e}\in\check{\mathcal{E}}}, where

aˇeˇ:=ae.\check{a}_{\check{e}}:=a_{e}. (4)

Furthermore, define the divergence operator on the graph GG as the function div:V\text{div}:{\mathbb{R}}^{\mathcal{E}}\to{\mathbb{R}}^{V} given by

div(θ)(x):=exθeexθe,\text{div}(\theta)(x):=\sum_{e\in\mathcal{E}^{x}}\theta_{e}-\sum_{e\in\mathcal{E}_{x}}\theta_{e},

where x:={e:e=(y,x) for some yV}\mathcal{E}^{x}:=\{e\in\mathcal{E}:e=(y,x)\text{ for some }y\in V\} is the set of edges in \mathcal{E} arriving at xx. Finally, we shall say that a function f:f:\mathcal{E}\to\mathbb{R} satisfies the closed loop property if for every path π=(x0,,xn)\pi=(x_{0},\ldots,x_{n}) which is closed, i.e., which satisfies xn=x0x_{n}=x_{0}, we have that

i=0n1f(Δxi)=0.\sum_{i=0}^{n-1}f(\Delta x_{i})=0.

Our main result is then the following.

Theorem 2.2.

Let G=(V,)G=(V,\mathcal{E}) be a finite strongly connected directed graph endowed with a family of positive weights 𝐚=(ae)e\mathbf{a}=(a_{e})_{e\in\mathcal{E}} with div(𝐚)=0\text{div}(\mathbf{a})=0. Then, there exists a coupling Υ=(ω,ωˇ,φ)\Upsilon=(\omega,\check{\omega},\varphi) such that:

  1. i.

    ω\omega is a generalized Bernoulli–Exponential environment on GG with parameters 𝐚\mathbf{a};

  2. ii.

    ωˇ\check{\omega} is a generalized Bernoulli–Exponential environment on Gˇ\check{G} with parameters 𝐚ˇ\check{\mathbf{a}};

  3. iii.

    φ=(φ(e))e\varphi=(\varphi(e))_{e\in\mathcal{E}} is a random function satisfying the closed loop property almost surely;

  4. iv.

    ωˇ(y,eˇ)=φ(e)+ω(x,e)\check{\omega}(y,\check{e})=\varphi(e)+\omega(x,e) for all e=(x,y)e=(x,y)\in\mathcal{E} almost surely.

In particular, if π=(x0,x1,,xn)\pi=(x_{0},x_{1},\dots,x_{n}) is any closed path and πˇ:=(xn,xn1,,x0)\check{\pi}:=(x_{n},x_{n-1},\dots,x_{0}) denotes its time reversal then

i=0n1ω(xi,Δxi)=i=0n1ωˇ(xˇi,Δxˇi),\sum_{i=0}^{n-1}\omega(x_{i},\Delta x_{i})=\sum_{i=0}^{n-1}\check{\omega}(\check{x}_{i},\Delta\check{x}_{i}),

where, for each i=0,,nei=0,\dots,n-e, we set xˇi:=xn1\check{x}_{i}:=x_{n-1} and Δxˇi:=(xˇi,xˇi+1)\Delta\check{x}_{i}:=(\check{x}_{i},\check{x}_{i+1}).

In Section 5, we will see how we can use Theorem 2.2 to compute certain passage times.

3 The Bernoulli–Exponential distribution as the limit of Dirichlet random vectors

To establish Theorem 2.2, an important step will be to show that the Bernoulli–Exponential distribution can be obtained as the weak limit of (suitably rescaled) Dirichlet random vectors. For convenience of the reader, we recall the definition of the Dirichlet distribution next.

Definition 3.1.

Given MM\in\mathbb{N} and a vector 𝐚=(a1,,aM)\mathbf{a}=(a_{1},\dots,a_{M}) with real positive entries, we define the Dirichlet distribution of order MM with parameters 𝐚\mathbf{a}, to be denoted by 𝒟M(𝐚)\mathcal{D}_{M}(\mathbf{a}), as the distribution of the random vector U=(U1,,UM)U=(U_{1},\dots,U_{M}) given by

Ui:=Wij=1MWj,i=1,,M,U_{i}:=\frac{W_{i}}{\sum_{j=1}^{M}W_{j}},\quad i=1,\dots,M, (5)

where W1,,WMW_{1},\dots,W_{M} are independent random variables with WiΓ(ai,1)W_{i}\sim\Gamma(a_{i},1) for i=1,,Mi=1,\dots,M, i.e., WiW_{i} has density function

fWi(w)=1Γ(ai)wai1ew1(0,)(w),f_{W_{i}}(w)=\frac{1}{\Gamma(a_{i})}w^{a_{i}-1}\mathrm{e}^{-w}\mathrm{1}_{(0,\infty)}(w),

where Γ\Gamma denotes the Gamma function.

Remark 3.2.

Whenever M=2M=2, we have that UiBeta(ai,a1+a2)U_{i}\sim\textrm{Beta}(a_{i},a_{1}+a_{2}) and U2=1U1U_{2}=1-U_{1}.

The main result of this section is the following proposition.

Proposition 3.3.

Fix MM\in\mathbb{N} and a vector 𝐚=(a1,,aM)\mathbf{a}=(a_{1},\dots,a_{M}) with real positive entries, and for each ε>0\varepsilon>0 let us consider a random vector U(ε)=(U1(ε),,UM(ε))𝒟M(ε𝐚)U^{(\varepsilon)}=(U_{1}^{(\varepsilon)},\dots,U_{M}^{(\varepsilon)})\sim\mathcal{D}_{M}(\varepsilon\mathbf{a}). Then, as ε0+\varepsilon\to 0^{+},

(εlogU1(ε),,εlogUM(ε))𝑑XBe-ExpM(𝐚).(-\varepsilon\log U_{1}^{(\varepsilon)},\dots,-\varepsilon\log U_{M}^{(\varepsilon)})\overset{d}{\longrightarrow}X\sim\textrm{Be-Exp}_{M}(\mathbf{a}). (6)

To prove Proposition 3.3, we will need two auxiliary results. The first is a standard property of Dirichlet random vectors, contained in Proposition 3.4 below, whose proof is omitted.

Proposition 3.4.

For any MM\in\mathbb{N} and a1,,aM>0a_{1},\dots,a_{M}>0, if W1,,WMW_{1},\dots,W_{M} are independent random variables with WiΓ(ai,1)W_{i}\sim\Gamma(a_{i},1) for i=1,,Mi=1,\dots,M, then the random vector UU defined in (5) is independent of j=1MWj\sum_{j=1}^{M}W_{j}.

The second auxiliary result gives two key properties of the Bernoulli–Exponential law.

Lemma 3.5.

For any M2M\in\mathbb{N}_{\geq 2} and vector 𝐚=(a1,,aM)\mathbf{a}=(a_{1},\dots,a_{M}) with real positive entries, if E1,,EME_{1},\dots,E_{M} are independent random variables with EiExp(ai)E_{i}\sim\textrm{Exp}(a_{i}) for each i=1,,Mi=1,\dots,M and we set E:=min{E1,,EM}E^{*}:=\min\{E_{1},\dots,E_{M}\}, then

(E,E1E,,EME)Exp(j=1Maj)×Be-ExpM(𝐚).(E^{*},E_{1}-E^{*},\dots,E_{M}-E^{*})\sim\textrm{Exp}(\sum_{j=1}^{M}a_{j})\times\textrm{Be-Exp}_{M}(\mathbf{a}). (7)

In particular, if E¯Exp(j=1Maj)\overline{E}\sim\textrm{Exp}(\sum_{j=1}^{M}a_{j}) and Z=(Z1,,ZM)Be-ExpM(𝐚)Z=(Z_{1},\dots,Z_{M})\sim\textrm{Be-Exp}_{M}(\mathbf{a}) are independent then

(E¯+Z1,,E¯+ZM)=𝑑(E1,,EM).(\overline{E}+Z_{1},\dots,\overline{E}+Z_{M})\overset{d}{=}(E_{1},\dots,E_{M}). (8)

On the other hand, if B=(B1,,BM)BeM(𝐩(𝐚))B=(B_{1},\dots,B_{M})\sim\mathrm{Be}_{M}(\mathbf{p}^{(\mathbf{a})}) with 𝐩(𝐚)=(p1(𝐚),,pM(𝐚))\mathbf{p}^{(\mathbf{a})}=(p_{1}^{(\mathbf{a})},\dots,p_{M}^{(\mathbf{a})}) as in (2) is independent of E=(E1,,En)E=(E_{1},\dots,E_{n}) then, conditional on BM=0B_{M}=0, we have

((1B1)E1,,(1BM1)EM1)Be-ExpM1(a1,,aM1).((1-B_{1})E_{1},\dots,(1-B_{M-1})E_{M-1})\sim\textrm{Be-Exp}_{M-1}(a_{1},\dots,a_{M-1}). (9)
Proof.

To check (7), by separating into the different cases {E=Ei}\{E^{*}=E_{i}\} for i=1,,Mi=1,\dots,M, a straightforward computation using the joint density of EE shows that for any s,t1,,tM0s,t_{1},\dots,t_{M}\geq 0 we have

(Es,E1Et1,,EMEtM)=e(j=1Maj)si=1Mpi(𝐚)jieajtj,\mathbb{P}(E^{*}\geq s\,,\,E_{1}-E^{*}\geq t_{1},\dots,E_{M}-E^{*}\geq t_{M})=\mathrm{e}^{-(\sum_{j=1}^{M}a_{j})s}\sum_{i=1}^{M}p_{i}^{(\mathbf{a})}\prod_{j\neq i}\mathrm{e}^{-a_{j}t_{j}},

from where (7) now immediately follows. Equation (8) is an immediate consequence of (7), so it only remains to show (9). To this end we observe that, since the EiE_{i}’s are independent, we only need to check that, conditional on BM=0B_{M}=0, we have

(B1,,BM1)BeM1(a1,,aM1).(B_{1},\dots,B_{M-1})\sim\textrm{Be}_{M-1}(a_{1},\dots,a_{M-1}). (10)

But this is immediate to verify: indeed, for each i=1,,M1i=1,\dots,M-1 we have

((B1,,BM1)=ei|BM=0)=(B=ei)1(B=eM)=aij=1M1aj\mathbb{P}((B_{1},\dots,B_{M-1})=\mathrm{e}_{i}|B_{M}=0)=\frac{\mathbb{P}(B=\mathrm{e}_{i})}{1-\mathbb{P}(B=\mathrm{e}_{M})}=\frac{a_{i}}{{\sum_{j=1}^{M-1}a_{j}}}

which readily implies (10). This concludes the proof. ∎

We are now ready to prove Proposition 3.3.

Proof of Proposition 3.3.

We proceed by induction on MM. The case M=1M=1 is trivial, so let us consider first the case M=2M=2. By Remark 3.2, the statement for M=2M=2 is exactly that of [1, Lemma 5.1]. Thus, given M>2M>2, let us suppose that (6) holds for any vector with Dirichlet distribution of order M1M-1 and show that then (6) must also hold for all vectors with Dirichlet distribution of order MM.

Let 𝐚=(a1,,aM)\mathbf{a}=(a_{1},\dots,a_{M}) be a vector with real positive entries and, for each ε>0\varepsilon>0, consider a random vector U(ε)=(U1(ε),,UM(ε))𝒟M(ε𝐚)U^{(\varepsilon)}=(U_{1}^{(\varepsilon)},\dots,U_{M}^{(\varepsilon)})\sim\mathcal{D}_{M}(\varepsilon\mathbf{a}). Without loss of generality, we can assume that Ui(ε)=Wij=1MWjU_{i}^{(\varepsilon)}=\frac{W_{i}}{\sum_{j=1}^{M}W_{j}} for each ii, where W1,,WMW_{1},\dots,W_{M} are independent random variables with WjΓ(εai,1)W_{j}\sim\Gamma(\varepsilon a_{i},1) for each jj. In this case, for each i=1,,M1i=1,\dots,M-1 we may write Ui(ε)U_{i}^{(\varepsilon)} as

Ui(ε)=(1B(ε))Vi(ε)U^{(\varepsilon)}_{i}=(1-B^{(\varepsilon)})\cdot V_{i}^{(\varepsilon)}

where

B(ε):=WMj=1M1Wj+WM=UM(ε) and Vi(ε):=Wij=1M1Wj.B^{(\varepsilon)}:=\frac{W_{M}}{\sum_{j=1}^{M-1}W_{j}+W_{M}}=U_{M}^{(\varepsilon)}\qquad\text{ and }\qquad V_{i}^{(\varepsilon)}:=\frac{W_{i}}{\sum_{j=1}^{M-1}W_{j}}.

Observe that, by definition of Dirichlet distribution together with Remark 3.2, Proposition 3.4 and the independence of the variables WjW_{j}, we have that B(ε)B^{(\varepsilon)} and V(ε)V^{(\varepsilon)} are independent, with distributions B(ε)Beta(εaM,εj=1M1aj)B^{(\varepsilon)}\sim\textrm{Beta}(\varepsilon a_{M},\varepsilon\sum_{j=1}^{M-1}a_{j}) and V(ε)𝒟M(εa1,,εaM1)V^{(\varepsilon)}\sim\mathcal{D}_{M}(\varepsilon a_{1},\dots,\varepsilon a_{M-1}). Therefore, by inductive hypothesis and the case M=2M=2, we conclude that

(εlogU1(ε),,εlogUM(ε))𝑑(Y1+Z1,Y1+Z2,,Y1+ZM1,Y2)=:X~(-\varepsilon\log U_{1}^{(\varepsilon)},\dots,-\varepsilon\log U_{M}^{(\varepsilon)})\overset{d}{\longrightarrow}(Y_{1}+Z_{1},Y_{1}+Z_{2},\dots,Y_{1}+Z_{M-1},Y_{2})=:\widetilde{X}

where Y=(Y1,Y2)Y=(Y_{1},Y_{2}) and Z=(Z1,,ZM1)Z=(Z_{1},\dots,Z_{M-1}) are independent and respectively given by

(εlog(1B(ε)),εlogB(ε))𝑑YBe-Exp2(j=1M1aj,aM)(-\varepsilon\log(1-B^{(\varepsilon)}),-\varepsilon\log B^{(\varepsilon)})\overset{d}{\longrightarrow}Y\sim\textrm{Be-Exp}_{2}\Big{(}\sum_{j=1}^{M-1}a_{j},a_{M}\Big{)}

and

(εlogV1(ε),,εlogVM(ε))𝑑ZBe-ExpM1(a1,,aM1).(-\varepsilon\log V_{1}^{(\varepsilon)},\dots,-\varepsilon\log V_{M}^{(\varepsilon)})\overset{d}{\longrightarrow}Z\sim\textrm{Be-Exp}_{M-1}(a_{1},\dots,a_{M-1}). (11)

Thus, to conclude the proof we only need to show that X~Be-ExpM(𝐚)\widetilde{X}\sim\textrm{Be-Exp}_{M}(\mathbf{a}).

To this end, we observe that, by conditioning on whether X~M=0\widetilde{X}_{M}=0 (that is, Y2=0Y_{2}=0) or not, if pM(𝐚)p^{(\mathbf{a})}_{M} is as in (2), then for any 𝐱=(x1,,xM)\mathbf{x}=(x_{1},\dots,x_{M}) with nonnegative entries we can write

FX~(𝐱)=pM(𝐚)F(E¯+Z1,,E¯+ZM1)(x1,,xM1)+(1pM(𝐚))FZ(x1,,xM1)FEM(xM),F_{\widetilde{X}}(\mathbf{x})=p^{(\mathbf{a})}_{M}F_{(\overline{E}+Z_{1},\dots,\overline{E}+Z_{M-1})}(x_{1},\dots,x_{M-1})+(1-p^{(\mathbf{a})}_{M})F_{Z}(x_{1},\dots,x_{M-1})F_{E_{M}}(x_{M}),

where ZZ is as in (11), E¯Exp(j=1M1aj)\overline{E}\sim\textrm{Exp}(\sum_{j=1}^{M-1}a_{j}) and EMExp(aM)E_{M}\sim\textrm{Exp}(a_{M}) are independent of ZZ and, given any random vector HH, we write FHF_{H} to denote its cumulative distribution function.

On the other hand, if we take XBe-ExpM(𝐚)X\sim\textrm{Be-Exp}_{M}(\mathbf{a}) to be the random vector from (1), then, by conditioning on whether XM=0X_{M}=0 (that is, BM=1B_{M}=1) or not, we have

FX(𝐱)=pM(𝐚)FE|M1(x1,,xM1)+(1pM(𝐚))FX(M1)(x1,,xM1)FEM(xM),F_{X}(\mathbf{x})=p^{(\mathbf{a})}_{M}F_{E|_{M-1}}(x_{1},\dots,x_{M-1})+(1-p^{(\mathbf{a})}_{M})F_{X^{(M-1)}}(x_{1},\dots,x_{M-1})F_{E_{M}}(x_{M}),

where E|M1:=(E1,,EM1)E|_{M-1}:=(E_{1},\dots,E_{M-1}) are the first M1M-1 coordinates of the vector EE from (1), and X(M1)X^{(M-1)} is distributed as (X1,,XM1)(X_{1},\dots,X_{M-1}) conditioned on BM=0B_{M}=0.

The result now follows immediately from Lemma 3.5, which in particular tells us that E|M1=𝑑(E¯+Z1,,E¯+ZM)E|_{M-1}\overset{d}{=}(\overline{E}+Z_{1},\dots,\overline{E}+Z_{M}) and X(M1)=𝑑ZX^{(M-1)}\overset{d}{=}Z. ∎

4 Proof of Theorem 2.2

We will now use Proposition 3.3 together with the property of statistical reversibility of random walks in Dirichlet environments to prove our Theorem 2.2. We begin by recalling the model of random walks in Dirichlet environments.

Given a strongly connected finite directed graph G=(V,)G=(V,\mathcal{E}), for each xVx\in V let

𝒫x={𝐩x=(p(x,e))ex:p(x,e)0 for all ex,exp(x,e)=1}\mathcal{P}_{x}=\{\mathbf{p}_{x}=(p(x,e))_{e\in\mathcal{E}_{x}}:p(x,e)\geq 0\text{ for all }e\in\mathcal{E}_{x}\,,\,\sum_{e\in\mathcal{E}_{x}}p(x,e)=1\}

be the space of all probability vectors on x\mathcal{E}_{x} and consider the product space Ω:=xV𝒫x\Omega:=\prod_{x\in V}\mathcal{P}_{x} endowed with the usual product topology. Any element ωΩ\omega\in\Omega will be called an environment, i.e., each ω=(ω(x))xV\omega=(\omega(x))_{x\in V} is a finite family of probability vectors ω(x)=(ω(x,e))ex𝒫x\omega(x)=(\omega(x,e))_{e\in\mathcal{E}_{x}}\in\mathcal{P}_{x}. Given any zVz\in V and ωΩ\omega\in\Omega, we define the random walk in the environment ω\omega starting at zz as the Markov chain (Xn)n0(X_{n})_{n\in\mathbb{N}_{0}} on GG whose law Pz,ωP_{z,\omega} is given by

Pz,ω(X0=z)=1 and Pz,ω(Xn+1=y|Xn=x)=ω(x,(x,y))P_{z,\omega}(X_{0}=z)=1\hskip 28.45274pt\text{ and }\hskip 28.45274ptP_{z,\omega}(X_{n+1}=y|X_{n}=x)=\omega(x,(x,y))

for all x,yVx,y\in V such that (x,y)(x,y)\in\mathcal{E}. If we now let the environment ω\omega be chosen at random according to some Borel probability measure \mathbb{P} on Ω\Omega, we obtain the measure PxP_{x} on Ω×V0\Omega\times V^{\mathbb{N}_{0}} given by

Px(A×B):=APx,ω(B)d,P_{x}(A\times B):=\int_{A}P_{x,\omega}(B)\mathrm{d}\mathbb{P},

for all Borel sets AΩA\subseteq\Omega and BV0B\subseteq V^{\mathbb{N}_{0}}. The process (Xn)n0(X_{n})_{n\in\mathbb{N}_{0}} under the measure PzP_{z} is called the random walk in random environment (RWRE) with environmental law \mathbb{P} (starting at zz). The measure PzP_{z} is known as the annealed law of the RWRE, whereas Pz,ωP_{z,\omega} for a fixed ωΩ\omega\in\Omega is referred to as the quenched law. Finally, if given a family of positive weights 𝐚=(ae)e\mathbf{a}=(a_{e})_{e\in\mathcal{E}} we consider the environmental law

𝒟(𝐚):=xV𝒟Mx(𝐚x),\mathbb{P}_{\mathcal{D}}^{(\mathbf{a})}:=\prod_{x\in V}\mathcal{D}_{M_{x}}(\mathbf{a}_{x}),

where, as before, we write 𝐚x:=(ae)ex\mathbf{a}_{x}:=(a_{e})_{e\in\mathcal{E}_{x}}, then the resulting RWRE is called a random walk in a Dirichlet environment (RWDE) with parameters 𝐚\mathbf{a}.

Now let us fix a family of positive weights 𝐚=(ae)e\mathbf{a}=(a_{e})_{e\in\mathcal{E}} on \mathcal{E} and, for each ε>0\varepsilon>0, consider a Dirichlet random environment ωε\omega_{\varepsilon} on GG with parameter ε𝐚\varepsilon\mathbf{a}, i.e., a family of independent random vectors ωε=(ωε(x))xV\omega_{\varepsilon}=(\omega_{\varepsilon}(x))_{x\in V} where, for each xVx\in V, the vector ωε(x)=(ωε(x,e))ex\omega_{\varepsilon}(x)=(\omega_{\varepsilon}(x,e))_{e\in\mathcal{E}_{x}} has Dirichlet distribution of order MxM_{x} with parameters ε𝐚x:=(εae)ex\varepsilon\mathbf{a}_{x}:=(\varepsilon a_{e})_{e\in\mathcal{E}_{x}}. Then, since εae>0\varepsilon a_{e}>0 for every ee\in\mathcal{E} and GG is strongly connected, for (ε𝐚)\mathbb{P}^{(\varepsilon\mathbf{a})}-almost every ωε\omega_{\varepsilon} the associated walk (Xn)n0(X_{n})_{n\in\mathbb{N}_{0}} is an irreducible finite-state Markov chain under Pz,ωP_{z,\omega} and, as such, admits a unique stationary distribution πωε=(πωε(x))xV\pi^{\omega_{\varepsilon}}=(\pi^{\omega_{\varepsilon}}(x))_{x\in V}. Following [4], we can now define the time-reversed environment ωˇε\check{\omega}_{\varepsilon} in the dual graph Gˇ\check{G} by

ωˇε(y,eˇ)=πωε(x)πωε(y)ωε(x,e),e=(x,y).\check{\omega}_{\varepsilon}(y,\check{e})=\frac{\pi^{\omega_{\varepsilon}}(x)}{\pi^{\omega_{\varepsilon}}(y)}\omega_{\varepsilon}(x,e),\qquad e=(x,y)\in\mathcal{E}. (12)

Since div(𝐚)=0\textrm{div}(\mathbf{a})=0, it follows from [4, Lemma 3] that ωˇε\check{\omega}_{\varepsilon} is a Dirichlet random environment on the dual graph Gˇ\check{G} with parameters ε𝐚ˇ\varepsilon\check{\mathbf{a}}, for 𝐚ˇ\check{\mathbf{a}} as defined in (4).

If for each e=(x,y)e=(x,y)\in\mathcal{E} we define the quantities

ϕε(x,e):=εlogωε(x,e) and ϕˇε(y,eˇ):=εlogωˇε(y,eˇ),\phi_{\varepsilon}(x,e):=-\varepsilon\log\omega_{\varepsilon}(x,e)\qquad\text{ and }\qquad\check{\phi}_{\varepsilon}(y,\check{e}):=-\varepsilon\log\check{\omega}_{\varepsilon}(y,\check{e}),

then (12) gives

ϕˇε(y,eˇ)=υε(e)+ϕε(x,e),\check{\phi}_{\varepsilon}(y,\check{e})=\upsilon_{\varepsilon}(e)+\phi_{\varepsilon}(x,e), (13)

where

υε(e):=εlogπωε(x)+εlogπωε(y).\upsilon_{\varepsilon}(e):=-\varepsilon\log\pi^{\omega_{\varepsilon}}(x)+\varepsilon\log\pi^{\omega_{\varepsilon}}(y).

Finally, consider the random vector

Υε=(ϕε(x,e),ϕˇε(y,eˇ),υε(e))e=(x,y).\Upsilon_{\varepsilon}=(\phi_{\varepsilon}(x,e),\check{\phi}_{\varepsilon}(y,\check{e}),\upsilon_{\varepsilon}(e))_{e=(x,y)\in\mathcal{E}}.

It follows from Proposition 3.3 and (13) that the family (Υε)ε>0(\Upsilon_{\varepsilon})_{\varepsilon>0} is tight as ε0\varepsilon\to 0, as shown in the following lemma.

Lemma 4.1.

For any sequence (εn)n(0,)(\varepsilon_{n})_{n\in\mathbb{N}}\subseteq(0,\infty) such that limnεn=0\lim_{n\to\infty}\varepsilon_{n}=0 we have that the sequence (Υεn)n(\Upsilon_{\varepsilon_{n}})_{n\in\mathbb{N}} is tight.

Proof.

Since the graph GG is finite, that the family (ϕεn(x,e))e=(x,y)(\phi_{\varepsilon_{n}}(x,e))_{e=(x,y)\in\mathcal{E}} is tight (in nn\in\mathbb{N}) follows from Proposition 3.3. Since ωˇεn𝒟(εn𝐚ˇ)\check{\omega}_{\varepsilon_{n}}\sim\mathbb{P}^{(\varepsilon_{n}\check{\mathbf{a}})}_{\mathcal{D}}, similarly we have that (ϕˇεn(y,eˇ))e=(x,y)(\check{\phi}_{\varepsilon_{n}}(y,\check{e}))_{e=(x,y)\in\mathcal{E}} is tight. In view of the relation in (13), the tightness of (ϕεn(x,e),ϕˇεn(y,eˇ))e=(x,y)(\phi_{\varepsilon_{n}}(x,e),\check{\phi}_{\varepsilon_{n}}(y,\check{e}))_{e=(x,y)\in\mathcal{E}} implies that of (υεn(e))e(\upsilon_{\varepsilon_{n}}(e))_{e\in\mathcal{E}}. From here the result now follows. ∎

As a consequence of Lemma 4.1 we obtain that there exists some subsequence (εn)n(\varepsilon_{n})_{n\in\mathbb{N}} such that limnεn=0\lim_{n\to\infty}\varepsilon_{n}=0 and a random vector Υ=(Υ(1)(x,e),Υ(2)(y,eˇ),Υ(3)(e))e=(x,y)\Upsilon=(\Upsilon^{(1)}(x,e),\Upsilon^{(2)}(y,\check{e}),\Upsilon^{(3)}(e))_{e=(x,y)\in\mathcal{E}} satisfying that as nn\to\infty,

Υεn𝑑Υ.\Upsilon_{\varepsilon_{n}}\overset{d}{\longrightarrow}\Upsilon. (14)

We claim that Υ\Upsilon is the desired coupling. Indeed, by Proposition 3.3 we have that Υ(1)BE(𝐚)\Upsilon^{(1)}\sim\mathbb{P}^{(\mathbf{a})}_{\text{BE}} and Υ(2)BE(𝐚ˇ)\Upsilon^{(2)}\sim\mathbb{P}^{(\check{\mathbf{a}})}_{\text{BE}}. On the other hand, (13) and (14) together imply that

Υ(2)(y,eˇ)=Υ(3)(e)+Υ(1)(x,e)\Upsilon^{(2)}(y,\check{e})=\Upsilon^{(3)}(e)+\Upsilon^{(1)}(x,e)

for all e=(x,y)e=(x,y)\in\mathcal{E}. Finally, if π=(x0,x1,,xn1,xn)\pi=(x_{0},x_{1},\dots,x_{n-1},x_{n}) is a closed path, then given ε>0\varepsilon>0 since xn=x0x_{n}=x_{0} we have that

i=0n1ϕˇε(xi+1,(xi+1,xi))=εlog(i=0n1πωε(xi)πωε(xi+1)ω(x,(xi,xi+1)))=i=0n1ϕε(xi,(xi+1,xi)),\sum_{i=0}^{n-1}\check{\phi}_{\varepsilon}(x_{i+1},(x_{i+1},x_{i}))=-\varepsilon\log\left(\prod_{i=0}^{n-1}\frac{\pi^{\omega_{\varepsilon}}(x_{i})}{\pi^{\omega_{\varepsilon}}(x_{i+1})}\omega(x,(x_{i},x_{i+1}))\right)=\sum_{i=0}^{n-1}\phi_{\varepsilon}(x_{i},(x_{i+1},x_{i})),

so that by (13) the vector (υε(e))e(\upsilon_{\varepsilon}(e))_{e\in\mathcal{E}} satisfies the closed loop property. It then follows from (14) that (Υ(3)(e))e(\Upsilon^{(3)}(e))_{e\in\mathcal{E}} must satisfy the closed loop property as well. Therefore, we see that Υ\Upsilon satisfies all the required properties and thus constitutes the desired coupling. This concludes the proof of Theorem 2.2.

5 First passage return time

We conclude by giving an example of how Theorem 2.2 can be used to do explicit computations of non-trivial first passage times.

Consider the finite graph GG of Figure 1. The vertices are classified into seven layers, plus an extra vertex which we call yy. The super-index ii in a vertex xkix_{k}^{i} indicates to which layer it belongs, e.g. xkix^{i}_{k} is the kk-th vertex of the ii-th layer. Let us fix three parameters a1,a2,a3>0a_{1},a_{2},a_{3}>0. We give the weight a1+a2+a3a_{1}+a_{2}+a_{3} to the edge from x11x^{1}_{1} to any vertex in layer 22. From layer 22 to 33 we give, for each 1k,j31\leq k,j\leq 3, the weight aja_{j} to the edge connecting xk2x^{2}_{k} to xj3x^{3}_{j}. In general, given j{1,2,3}j\in\{1,2,3\}, for 2i62\leq i\leq 6 we give the weight aja_{j} to any outgoing edge from a vertex xkix^{i}_{k} in layer ii to the jj-th vertex xji+1x^{i+1}_{j} in layer i+1i+1. Every outgoing edge from layer 77 to x11x^{1}_{1} or yy is assigned the weight (a1+a2+a3)/2(a_{1}+a_{2}+a_{3})/2. The edge from yy to x11x^{1}_{1} has weight a1+a2+a3a_{1}+a_{2}+a_{3}, while the edge from x11x_{1}^{1} to yy has weight (a1+a2+a3)/2(a_{1}+a_{2}+a_{3})/2. The weights indicated in the edges of graph GG define a generalized Bernoulli–Exponential first passage percolation model with those weights. Notice that the divergence condition div(𝐚)=0\text{div}({\bf a})=0 is satisfied, where 𝐚\mathbf{a} denotes the vector of all such edge weights.

x11x^{1}_{1}x22x^{2}_{2}x12x^{2}_{1}x32x^{2}_{3}x23x^{3}_{2}x13x^{3}_{1}x33x^{3}_{3}x24x^{4}_{2}x14x^{4}_{1}x34x^{4}_{3}x25x^{5}_{2}x15x^{5}_{1}x35x^{5}_{3}x26x^{6}_{2}x16x^{6}_{1}x36x^{6}_{3}x27x^{7}_{2}x17x^{7}_{1}x37x^{7}_{3}x11x^{1}_{1}yy
Figure 1: Edges from layer ii to jj with 1i,j61\leq i,j\leq 6 are assigned the weights a1,a2a_{1},a_{2} or a3a_{3}.

Define now the passage time,

Tc(x11,x11):=infπΠc(x11,x11){i=0n1w(xi,Δxi)},T_{c}(x_{1}^{1},x_{1}^{1}):=\inf_{\pi\in\Pi_{c}(x^{1}_{1},x^{1}_{1})}\left\{\sum_{i=0}^{n-1}w(x_{i},\Delta x_{i})\right\},

where Πc\Pi_{c} is the set of paths which exit x11x^{1}_{1} through the right in Figure 1 (first jump is through any of the edges from x11x^{1}_{1} to layer 22) and finish entering again x11x^{1}_{1} through layer 77 or vertex yy. By the statistical reversibility of Theorem 2.2, we have that

Tc(x11,x11)=infπΠˇc(x11,x11){i=0n1wˇ(xi,Δxi)},T_{c}(x_{1}^{1},x_{1}^{1})=\inf_{\pi\in\check{\Pi}_{c}(x^{1}_{1},x^{1}_{1})}\left\{\sum_{i=0}^{n-1}\check{w}(x_{i},\Delta x_{i})\right\}, (15)

where Πˇc\check{\Pi}_{c} is the set of paths in the dual graph Gˇ\check{G} which exit x11x^{1}_{1} with a first jump through any of the edges from x11x^{1}_{1} to layer 77 or vertex yy, and finish entering again x11x^{1}_{1} through layer 22. But now note that for the reversed process there is always a path of 0 weight from any vertex in layer 77 to the vertex x11x^{1}_{1}. Hence the minimum in the right-hand side of (15) is just the minimum of the passage time in the dual graph Gˇ\check{G} from x11x^{1}_{1} to layer 77 (without revisiting x11x^{1}_{1}), which can be immediately computed and gives

Tc(x11,x11)=min{wˇx11,x17,wˇx11,x27,wˇx11,x37,wˇx11,y+wˇy,x17,wˇx11,y+wˇy,x27,wˇx11,y+wˇy,x37}.T_{c}(x^{1}_{1},x^{1}_{1})=\min\{\check{w}_{x^{1}_{1},x^{7}_{1}},\check{w}_{x^{1}_{1},x^{7}_{2}},\check{w}_{x^{1}_{1},x^{7}_{3}},\check{w}_{x^{1}_{1},y}+\check{w}_{y,x^{7}_{1}},\check{w}_{x^{1}_{1},y}+\check{w}_{y,x^{7}_{2}},\check{w}_{x^{1}_{1},y}+\check{w}_{y,x^{7}_{3}}\}.

Acknowledgements

Alejandro Ramίrez was supported by Fondecyt 1220396. Santiago Saglietti was supported by Fondecyt 11200690.

References

  • BC [17] G. Barraquand and I. Corwin. Random walk in Beta-distributed random environment. Probab. Theory Related Fields 167, 1057-1116 (2017).
  • DR [14] A. Drewitz and A.F. Ramίrez. Selected topics in random walks in random environment. Topics in percolative and disordered systems, 23-83, Springer Proc. Math. Stat., 69, Springer, New York, (2014).
  • S [11] C. Sabot. Random Dirichlet environment viewed from the particle are transient in dimension d3d\geq 3. Probab. Theory Relat. Fields, 151 297-317 (2011).
  • ST [17] C. Sabot and L. Tournier. Random walks in Dirichlet environment: an overview. Annales de la Faculté des sciences de Toulouse: Mathématiques, Serie 6, Volume 26 (2017) no. 2, pp. 463-509.
  • T [15] L. Tournier. Asymptotic direction of random walks in Dirichlet environment. Ann. Inst. H. Poincaré Probab. Statist. 51 716–726 (2015).
  • Z [06] O. Zeitouni. Random walks in random environments. J. Phys. A, 39, R433-R464 (2006).