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

Extremal Invariant Distributions of Infinite Brownian Particle Systems with Rank Dependent Drifts

Sayan Banerjee, Amarjit Budhiraja
Abstract

Consider an infinite collection of particles on the real line moving according to independent Brownian motions and such that the ii-th particle from the left gets the drift gi1g_{i-1}. The case where g0=1g_{0}=1 and gi=0g_{i}=0 for all ii\in{\mathbb{N}} corresponds to the well studied infinite Atlas model. Under conditions on the drift vector 𝒈=(g0,g1,){\boldsymbol{g}}=(g_{0},g_{1},\ldots)^{\prime} it is known that the Markov process corresponding to the gap sequence of the associated ranked particles has a continuum of product form stationary distributions {πa𝒈,aS𝒈}\{\pi_{a}^{{\boldsymbol{g}}},a\in S^{{\boldsymbol{g}}}\} where S𝒈S^{{\boldsymbol{g}}} is a semi-infinite interval of the real line. In this work we show that all of these stationary distributions are extremal and ergodic. We also prove that any product form stationary distribution of this Markov process that satisfies a mild integrability condition must be πa𝒈\pi_{a}^{{\boldsymbol{g}}} for some aS𝒈a\in S^{{\boldsymbol{g}}}. These results are new even for the infinite Atlas model. The work makes progress on the open problem of characterizing all the invariant distributions of general competing Brownian particle systems interacting through their relative ranks. Proofs rely on synchronous and mirror coupling of Brownian particles and properties of the intersection local times of the various particles in the infinite system.


AMS 2010 subject classifications: 60J60, 60K35, 60J25, 60H10.

Keywords: Reflecting Brownian motions, long time behavior, extremal invariant distributions, infinite Atlas model, collision local time, ergodicity, product-form stationary distributions, synchronous coupling, mirror coupling.

1 Introduction

1.1 Background

Consider a collection (finite or infinite) of particles on the real line moving according to mutually independent Brownian motions and such that the ii-th particle from the left gets a constant drift gi1g_{i-1}. The special case when g0=1g_{0}=1 and gi=0g_{i}=0 for ii\in{\mathbb{N}} is the well studied Atlas model. We refer to the general setting as the 𝒈{\boldsymbol{g}}-Atlas model, where 𝒈=(g0,g1,){\boldsymbol{g}}=(g_{0},g_{1},\ldots)^{\prime}. Such particle systems were originally introduced in stochastic portfolio theory [13, 7, 14] as models for stock growth evolution in equity markets and have been investigated extensively in recent years in several different directions. In particular, characterizations of such particle systems as uniform scaling limits of jump processes with local interactions on integer lattices, such as the totally asymmetric simple exclusion process, have been studied in [24]. Various types of results for the asymptotic behavior of the empirical measure of the particle states have been studied, such as propagation of chaos, characterization of the associated McKean-Vlasov equation and nonlinear Markov processes [39, 21], large deviation principles [12], characterizing the asymptotic density profile and the trajectory of the leftmost particle via Stefan free-boundary problems [10]. These particle systems also have close connections with Aldous’ “Up the river” stochastic control problem [1], recently solved in [40]. Results on wellposedness of the associated stochastic differential equations (in the weak and strong sense) and on absence of triple collisions (three particles at the same place at the same time) have been studied in [8, 38, 18, 17, 31, 20].

One important direction of investigation has been in the study of the long-time behavior of such particle systems. For finite particle systems, under conditions on the drift vector 𝒈{\boldsymbol{g}}, it follows from results of Harrison and Williams [16, 15] that the multidimensional reflected Brownian motion describing the evolution of the gaps between the ranked particles has a unique stationary (invariant) distribution (see [27]). It is known that convergence of the law at time tt to this stationary distribution, as tt\to\infty, occurs at a geometric rate [9]. Rates of convergence to stationarity, depending explicitly on the drift vector 𝒈{\boldsymbol{g}} and dimension have been obtained in [19, 5, 4].

In the current work we are interested in infinite particle systems. One basic result on long-time behavior of such particle systems was obtained in [27] which showed that for the infinite Atlas model, i.e. when 𝒈=𝒈1=(1,0,0,){\boldsymbol{g}}={\boldsymbol{g}}^{1}=(1,0,0,\ldots)^{\prime}, the process describing the gaps between the ranked particles in the system has a simple product form stationary distribution given as π0=n=1Exp(2)\pi_{0}=\otimes_{n=1}^{\infty}\mbox{Exp}(2) (here and later, for a>0a>0, Exp(a)\mbox{Exp}(a) denotes the Exponential distribution with mean 1/a1/a). The paper [27] also conjectured that this is the unique stationary distribution of the (gap sequence in the) infinite Atlas model. However, this was shown to be false in [35] who gave an uncountable collection of product form stationary distributions for the gaps in the infinite Atlas model defined as

πai=1Exp(2+ia),a0.\pi_{a}\doteq\bigotimes_{i=1}^{\infty}\mbox{Exp}(2+ia),\ a\geq 0. (1.1)

As in the finite dimensional settings it is of interest to investigate convergence of the laws at time tt to the stationary distributions as tt\to\infty. Due to the multiplicity of stationary distributions, a meaningful goal is to understand the local stability structure of this infinite dimensional stochastic dynamical system and identify the basins of attraction of the various stationary distributions. Such results, describing (weak and strong) domains of attraction of π0\pi_{0} have been obtained in [32, 11, 6] and (weak) domains of attraction of πa\pi_{a} (a0a\neq 0) in [6]. Results analogous to [35, 32] for two-sided infinite Brownian systems have been obtained in [33].

1.2 Goals and results

Although the above results give us a good understanding of the local stability structure of the infinite Atlas model, the picture that one has is far from complete. A key obstacle here is that a full characterization of all extremal invariant distributions of the infinite Atlas model is currently an open problem. The goal of this work is to make some progress towards this goal and, moreover, provide some characterization of the structure of the set of invariant distributions. We will in fact consider the more general setting of the 𝒈{\boldsymbol{g}}-Atlas model where the drift vector 𝒈𝒟{\boldsymbol{g}}\in{\mathcal{D}}, with

𝒟{𝒈=(g0,g1,):i=0gi2<}.{\mathcal{D}}\doteq\Big{\{}{\boldsymbol{g}}=(g_{0},g_{1},\ldots)^{\prime}\in{\mathbb{R}}^{\infty}:\sum_{i=0}^{\infty}g_{i}^{2}<\infty\Big{\}}. (1.2)

For this setting it is known from the work of [35] that, once more, the process associated with the gap sequence of the ranked particle system has a continuum of stationary distributions given as

πa𝒈n=1Exp(n(2g¯n+a)),a>2infng¯n,\pi_{a}^{{\boldsymbol{g}}}\doteq\otimes_{n=1}^{\infty}\mbox{Exp}(n(2\bar{g}_{n}+a)),\;\;a>-2\inf_{n\in{\mathbb{N}}}\bar{g}_{n}, (1.3)

where g¯n1n(g0++gn1)\bar{g}_{n}\doteq\frac{1}{n}(g_{0}+\cdots+g_{n-1}). In the special case where 𝒈𝒟1{\boldsymbol{g}}\in{\mathcal{D}}_{1}, with

𝒟1{𝒈𝒟: there exist N1<N2< such that g¯k>g¯Nj,k=1,,Nj1, for all j1},{\mathcal{D}}_{1}\doteq\{{\boldsymbol{g}}\in{\mathcal{D}}:\text{ there exist }N_{1}<N_{2}<\dots\rightarrow\infty\text{ such that }\bar{g}_{k}>\bar{g}_{N_{j}},\ k=1,\dots,N_{j}-1,\text{ for all }j\geq 1\}, (1.4)

πa𝒈\pi_{a}^{{\boldsymbol{g}}} is also an invariant distribution for a=2infng¯n=2limjg¯Nja=-2\inf_{n\in{\mathbb{N}}}\bar{g}_{n}=-2\lim_{j\rightarrow\infty}\bar{g}_{N_{j}} (see [32, Section 4.2]). Note that 𝒈1𝒟1{\boldsymbol{g}}^{1}\in{\mathcal{D}}_{1} whereas the zero drift 𝒈=(0,0,){\boldsymbol{g}}=(0,0,\ldots)^{\prime} is in 𝒟{\mathcal{D}} but not in 𝒟1{\mathcal{D}}_{1}. Roughly speaking, a drift 𝒈{\boldsymbol{g}} lying in 𝒟1{\mathcal{D}}_{1} produces a ‘stabilizing interaction’ in the subsystem of the lowest NjN_{j} particles for any j1j\geq 1, due to which the gaps between them stabilize in time owing to stronger upward average drifts of the lower particles in this subsystem in comparison to the average drift of all the NjN_{j} particles. This intrinsic stabilizing influence from the drift of the particles leads to an additional stationary distribution, namely π2limjg¯Nj\pi_{-2\lim_{j\rightarrow\infty}\bar{g}_{N_{j}}} (in comparison to a 𝒈𝒟𝒟1{\boldsymbol{g}}\in{\mathcal{D}}\setminus{\mathcal{D}}_{1}). For 𝒈𝒟𝒟1{\boldsymbol{g}}\in{\mathcal{D}}\setminus{\mathcal{D}}_{1}, where such a mechanism is absent, local stability essentially arises from configurations with a rapidly increasing density of particles as one moves away from the lowest one and hence one only obtains “dense stationary distributions” corresponding to a>2infng¯na>-2\inf_{n\in{\mathbb{N}}}\bar{g}_{n}.

Using Kakutani’s theorem [22] it is easy to verify that, for different values of aa, the probability measures πa𝒈\pi_{a}^{{\boldsymbol{g}}} are mutually singular. These distributions are also special in that they have a product form structure. In particular, if the initial distribution of the gap process is chosen to be one of these distributions, then the laws of distinct gaps at any fixed time are independent despite these gaps having a highly correlated temporal evolution mechanism (see (2.3)-(2.4)). We now describe the two main results of this work.

First result: Extremality. In Theorem 3.3 we show that for each 𝒈𝒟{\boldsymbol{g}}\in{\mathcal{D}} and a>2infng¯na>-2\inf_{n\in{\mathbb{N}}}\bar{g}_{n}, πa𝒈\pi_{a}^{{\boldsymbol{g}}} is an extremal invariant distribution for the gap sequence process of the 𝒈{\boldsymbol{g}}-Atlas model. Further, if 𝒈𝒟1{\boldsymbol{g}}\in{\mathcal{D}}_{1}, πa𝒈\pi_{a}^{{\boldsymbol{g}}} is also an extremal invariant distribution for a=2infng¯n=2limjg¯Nja=-2\inf_{n\in{\mathbb{N}}}\bar{g}_{n}=-2\lim_{j\rightarrow\infty}\bar{g}_{N_{j}}; in particular for the infinite Atlas model πa=πa𝒈1\pi_{a}=\pi_{a}^{{\boldsymbol{g}}^{1}} is extremal for all a0a\geq 0. From equivalence between extremality and ergodicity (cf. Lemma 3.2) it then follows that all these invariant distributions are ergodic as well. This result also identifies non-trivial subsets in the weak domain of attraction of πa𝒈\pi_{a}^{{\boldsymbol{g}}} for each a>2infng¯na>-2\inf_{n\in{\mathbb{N}}}\bar{g}_{n} (a2infng¯na\geq-2\inf_{n\in{\mathbb{N}}}\bar{g}_{n} if 𝒈𝒟1{\boldsymbol{g}}\in{\mathcal{D}}_{1}); see Corollary 3.4.

Questions about extremality and ergodicity of stationary distributions have been addressed previously in the context of interacting particle systems on countably infinite graphs (see [26, 2, 37, 3] and references therein). However, in all these cases, the interactions are Poissonian, namely, the dynamics is given in terms of jumps of particles to neighboring vertices in a countably infinite graph at epochs of Poisson processes associated with edges or vertices. This enables one to use the (explicit) generator of the associated continuous-time jump-Markov processes in an effective way. The interactions in rank based diffusions are very ‘singular’ owing to the local time based dynamics (see (2.3)) and generator based methods seem to be less tractable. Furthermore, unlike previous works, the state space for the gap process (i.e. +\mathbb{R}_{+}^{\infty}) is not countable and has a non-smooth boundary, and the process has intricate interactions (oblique reflections) with the boundary. Hence, proving extremality requires new techniques. Our proofs are based on constructing appropriate couplings for these infinite dimensional diffusions which then allow us to prove suitable invariance properties (see e.g. (4.13)) and a certain ‘directional strong Feller property’ (see Proposition 4.3). Such coupling techniques, based on ‘mirror’ couplings of driving Brownian motions, are novel in the context of infinite rank based diffusions and provide a new method for establishing semigroup continuity properties for such processes. Moreover, the coupling approach introduced in the paper has the potential to be applicable to broader families of infinite-dimensional diffusion processes for which such directional strong Feller properties may be useful, e.g. in analysis of the ergodic behavior. Although our setting and methods are very different, at a high level, the approach we take, of proving extremality by showing the a.e. constancy of suitable invariant functions, is inspired by the papers [37, 3].

Second result: Characterization of product form stationary distributions. A natural question is whether there are any other product form stationary distributions of the 𝒈{\boldsymbol{g}}-Atlas model than the ones identified in [35]. Our second main result (Theorem 3.5) answers this question in the negative under certain conditions by showing that if 𝒈𝒟1{\boldsymbol{g}}\in{\mathcal{D}}_{1} and π\pi is a product form stationary distribution of the 𝒈{\boldsymbol{g}}-Atlas model satisfying a mild integrability condition (see (3.2)) then it must be πa𝒈\pi_{a}^{{\boldsymbol{g}}} for some a2limjg¯Nja\geq-2\lim_{j\rightarrow\infty}\bar{g}_{N_{j}}. Furthermore, this result gives a novel probabilistic interpretation to aa in terms of the resulting force acting on a tagged particle under the combined influence of hardcore interactions (collision local times) and soft potentials (drift terms). See Remark 3.6 for this interpretation and for a conjecture that is suggested by this interpretation.

1.3 Proof ideas

We now make some comments on proofs. The key step in proving the extremality of πa𝒈\pi_{a}^{{\boldsymbol{g}}} is to establish that any bounded measurable function ψ\psi on +{\mathbb{R}}_{+}^{\infty} that is πa𝒈\pi_{a}^{{\boldsymbol{g}}}-a.e. invariant, under the action of the semigroup of the Markov process corresponding to the 𝒈{\boldsymbol{g}}-Atlas gap sequence, is constant πa𝒈\pi_{a}^{{\boldsymbol{g}}}-a.e. If 𝒈=𝒈1{\boldsymbol{g}}={\boldsymbol{g}}^{1} and a=0a=0, we have that πa𝒈=i=1Exp(2)\pi_{a}^{{\boldsymbol{g}}}=\otimes_{i=1}^{\infty}\mbox{Exp}(2), and therefore the coordinate sequence {Zi}i=1\{Z_{i}\}_{i=1}^{\infty} is iid under πa𝒈\pi_{a}^{{\boldsymbol{g}}}. In this case, from the Hewitt-Savage zero-one law it suffices to show that ψ\psi is πa𝒈\pi_{a}^{{\boldsymbol{g}}}-a.e. invariant under all finite permutations of the coordinates of +{\mathbb{R}}_{+}^{\infty}. For this, in turn, it suffices to simply prove the above invariance property for transpositions of the ii-th and (i+1)(i+1)-th coordinates, for all ii\in{\mathbb{N}}. For a general πa𝒈\pi_{a}^{{\boldsymbol{g}}}, the situation is more involved as the coordinate sequence {Zi}i=1\{Z_{i}\}_{i=1}^{\infty} is not iid any more. Nevertheless, from the scaling properties of Exponential distributions it follows that, with cn2[n(2g¯n+a)]1c_{n}\doteq 2[n(2\bar{g}_{n}+a)]^{-1}, the sequence {Z~n}n1\{\tilde{Z}_{n}\}_{n\geq 1}, defined as Z~n=cn1Zn\tilde{Z}_{n}=c_{n}^{-1}Z_{n}, nn\in{\mathbb{N}}, is iid under πa𝒈\pi_{a}^{{\boldsymbol{g}}}. In this case, in order to invoke the Hewitt-Savage zero-one law, one needs to argue that for each ii the map ψ\psi is πa𝒈\pi_{a}^{{\boldsymbol{g}}}-a.e. invariant under the transformation that takes the (i,i+1)(i,i+1) coordinates (zi,zi+1)(z_{i},z_{i+1}) to (cici+1zi+1,ci+1cizi)(\frac{c_{i}}{c_{i+1}}z_{i+1},\frac{c_{i+1}}{c_{i}}z_{i}) and keeps the remaining coordinates the same. Establishing this property is at the heart of the proof of Theorem 3.3. A key technical idea in the proof is the construction of a mirror coupling of the first i+1i+1 Brownian motions, and synchronous coupling of the remaining Brownian motions, in the evolution of the ranked infinite 𝒈{\boldsymbol{g}}-Atlas model corresponding to a pair of nearby initial configurations. Estimates on the probability of a successful coupling, before any of the first ii-gap processes have hit zero or the lowest ii-particles have interacted with the higher ranked particles (in a suitable sense), are some of the important ingredients in the proof. We refer the reader to Section 4.1 for additional comments on the proof strategy.

The proof of Theorem 3.5 hinges on establishing a key identity for expectations, under the given product form invariant measure π\pi, of certain integrals involving the state process of the ii-th gap and the collision local time for the (j1)(j-1)-th and jj-th particle, for iji\neq j (see Lemma 5.4). This identity is a consequence of the product form structure of π\pi and basic results on local times of continuous semimartingales. One subtlety here is that although the product form structure of an invariant measure π\pi implies that the laws of the various gaps at any fixed time tt (when the process is initiated at π\pi) are independent, the laws of the paths of the various gap processes are not, and thus one cannot immediately deduce the independence beween the state of the ii-th gap at time tt and the local time of the jj-th gap at 0, at time tt, for iji\neq j. By using the form of the dynamics of the 𝒈{\boldsymbol{g}}-Atlas model, the identity in Lemma 5.4 allows us to obtain a recursive system of equations for the moment generating functions of the coordinate projections of π\pi that can then be solved explicitly from which it is then readily seen that π\pi must be πa𝒈\pi_{a}^{{\boldsymbol{g}}} for a suitable value of aa. See Section 5.1 for additional comments on the proof idea.

1.4 Open problem

Of course it is immediate to construct non-product form stationary distributions of the 𝒈{\boldsymbol{g}}-Atlas model by considering mixtures of the above product stationary distributions, however one can ask if these mixtures are all the invariant measures of the 𝒈{\boldsymbol{g}}-Atlas model. For the cases where 𝒈=(0,0,){\boldsymbol{g}}=(0,0,\ldots)^{\prime} this question was answered in the affirmative in [30, Theorem 4.2] under certain integrability constraints on the denseness of particle configurations. For a general 𝒈{\boldsymbol{g}} providing such a complete characterization is a challenging open problem.

In the context of interacting particle systems on countably infinite graphs, the analogous problem has been solved completely in a few cases such as the simple exclusion process [26] and the zero range process [2] where the extremal probability measures are fully characterized as an explicit collection of certain product form measures. However, in these models the particle density associated with distinct extremal measures are scalar multiples of each other owing to certain homogeneity properties in the dynamics (see, for example, [2, Theorem 1.10]). This, along with the Poissonian nature of the interactions, enables one to prove useful monotonicity properties of the ‘synchronously coupled’ dynamics (see [26, Section 2] and [2, Section 4]) using generator methods that are crucial to the above characterization results.

A key challenge in extending these methods to rank based diffusions of the form considered here is that, in addition to the singular local time interactions, the point process associated with the configuration of particles with gaps distributed as πa𝒈\pi_{a}^{{\boldsymbol{g}}} has an intensity function ρ(x)\rho(x), that grows exponentially as xx\to\infty when a>0a>0 and, due to a nonlinear dependence on aa, lacks the scalar multiple property for distinct values of aa. This is a direct consequence of the inhomogeneity of the topological interactions in our particle systems where the local stability in a certain region of the particle cloud is affected both by the density of particles in the neighborhood and their relative ranks in the full system. Moreover, unlike the above interacting particle systems, in rank based diffusions, when the initial gaps are given by a stationary distribution, the point process of particles is typically not stationary. This phenomenon, where the gaps are stationary while the associated point process is not, referred to as quasi-stationarity in [30], is technically challenging. We note that this latter paper studies one setting where the intensity function grows exponentially and the particle density lacks the scalar multiple property for distinct values of aa. However their setting, in the context of rank based diffusions, corresponds to the case 𝒈=(0,0,){\boldsymbol{g}}=(0,0,\ldots)^{\prime}, where the unordered particles behave like independent standard Brownian motions, and this fact is crucially exploited in [30].

1.5 Organization

Rest of the paper is organized as follows. We close this section by summarizing the main notation used in this work. In Section 2 we give the precise formulation of the model. In Section 3, we describe the questions of interest and give our main results. Finally Sections 4 and 5 give the proofs of our main results, namely Theorems 3.3 and 3.5, respectively.

1.6 Notation

The following notation will be used. Let 0{0,1,2,}{\mathbb{N}}_{0}\doteq\{0,1,2,\dots\}, {(x0,x1,):xi,i0}{\mathbb{R}}^{\infty}\doteq\{(x_{0},x_{1},\ldots)^{\prime}:x_{i}\in{\mathbb{R}},\,i\in{\mathbb{N}}_{0}\} and +{(x0,x1,):xi+,i0}{\mathbb{R}}_{+}^{\infty}\doteq\{(x_{0},x_{1},\ldots)^{\prime}:x_{i}\in{\mathbb{R}}_{+},\,i\in{\mathbb{N}}_{0}\}. We will equip +{\mathbb{R}}^{\infty}_{+} with the partial order ‘\leq’ under which 𝒙𝒚{\boldsymbol{x}}\leq{\boldsymbol{y}} for 𝒙=(xi)i0{\boldsymbol{x}}=(x_{i})_{i\in{\mathbb{N}}_{0}}, 𝒚=(yi)i0{\boldsymbol{y}}=(y_{i})_{i\in{\mathbb{N}}_{0}}, if xiyix_{i}\leq y_{i} for all i0i\in{\mathbb{N}}_{0}. Borel σ\sigma-fields on a metric space SS will be denoted as (S){\mathcal{B}}(S) and the space of probability measures on (S,(S))(S,{\mathcal{B}}(S)) will be denoted as 𝒫(S){\mathcal{P}}(S) which is equipped with the topology of weak convergence. We will denote by {\mathbb{Q}} the set of rationals. For a Polish space SS with a partial order ‘\leq’, we say for γ1,γ2𝒫(S)\gamma_{1},\gamma_{2}\in{\mathcal{P}}(S), γ1stγ2\gamma_{1}\leq_{\mbox{{\tiny st}}}\gamma_{2}, if there are SS-valued random variables 𝑿1,𝑿2{\boldsymbol{X}}_{1},{\boldsymbol{X}}_{2} given on a common probability space with 𝑿i{\boldsymbol{X}}_{i} distributed as γi\gamma_{i}, i=1,2i=1,2, and 𝑿1𝑿2{\boldsymbol{X}}_{1}\leq{\boldsymbol{X}}_{2} a.s. Let 𝕏𝒞([0,):+){\mathbb{X}}\doteq{\mathcal{C}}([0,\infty):{\mathbb{R}}_{+}^{\infty}) which is equipped with the topology of local uniform convergence (with +{\mathbb{R}}_{+}^{\infty} equipped with the product topology). For γ𝒫(+)\gamma\in{\mathcal{P}}({\mathbb{R}}_{+}^{\infty}), let L2(γ)L^{2}(\gamma) be the collection of all measurable ψ:+\psi:{\mathbb{R}}_{+}^{\infty}\to{\mathbb{R}} such that +|ψ(𝒛)|2γ(d𝒛)<\int_{{\mathbb{R}}_{+}^{\infty}}|\psi({\boldsymbol{z}})|^{2}\gamma(d{\boldsymbol{z}})<\infty. We denote the inner-product and the norm on L2(γ)L^{2}(\gamma) as ,\langle\cdot,\cdot\rangle and \|\cdot\| respectively.

2 Model Formulation

Let

𝒰{𝒙=(x0,x1,):i=0eα[(xi)+]2< for all α>0}.{\mathcal{U}}\doteq\Big{\{}{\boldsymbol{x}}=(x_{0},x_{1},\ldots)^{\prime}\in{\mathbb{R}}^{\infty}:\sum_{i=0}^{\infty}e^{-\alpha[(x_{i})_{+}]^{2}}<\infty\mbox{ for all }\alpha>0\Big{\}}.

Recall 𝒟{\mathcal{D}} defined in (1.2). Following [32], an 𝒙{\boldsymbol{x}}\in{\mathbb{R}}^{\infty} is called rankable if there is a one-to-one map 𝒑{\boldsymbol{p}} from 0{\mathbb{N}}_{0} onto itself such that x𝒑(i)x𝒑(j)x_{{\boldsymbol{p}}(i)}\leq x_{{\boldsymbol{p}}(j)} whenever i<ji<j, i,j0i,j\in{\mathbb{N}}_{0}. It is easily seen that any 𝒙𝒰{\boldsymbol{x}}\in{\mathcal{U}} is locally finite and hence rankable. For an 𝒙{\boldsymbol{x}} that is rankable we denote the unique permutation map 𝒑{\boldsymbol{p}} as above which breaks ties in the lexicographic order by 𝒑𝒙{\boldsymbol{p}}_{{\boldsymbol{x}}}. For a sequence {Wi}i0\{W_{i}\}_{i\in{\mathbb{N}}_{0}} of mutually independent standard Brownian motions given on some filtered probability space, and 𝒚=(y0,y1,)𝒰{\boldsymbol{y}}=(y_{0},y_{1},\ldots)^{\prime}\in{\mathcal{U}}, 𝒈=(g0,g1,)𝒟{\boldsymbol{g}}=(g_{0},g_{1},\ldots)^{\prime}\in{\mathcal{D}}, consider the following system of equations.

dYi(t)=[k=0𝟏(𝒑𝐘(t)(k)=i)gk]dt+dWi(t),dY_{i}(t)=\left[\sum_{k=0}^{\infty}{\boldsymbol{1}}({\boldsymbol{p}}_{{\mathbf{Y}}(t)}(k)=i)g_{k}\right]dt+dW_{i}(t), (2.1)

Yi(0)=yiY_{i}(0)=y_{i}, i0i\in{\mathbb{N}}_{0}, where for t0t\geq 0, 𝐘(t)=(Y0(t),Y1(t),){\mathbf{Y}}(t)=(Y_{0}(t),Y_{1}(t),\ldots)^{\prime}.

The following result is from [32] (see Theorems 3.2 and Lemma 3.4 therein).

Theorem 2.1 ([32]).

For every 𝐠𝒟{\boldsymbol{g}}\in{\mathcal{D}} and 𝐲𝒰{\boldsymbol{y}}\in{\mathcal{U}} there is a unique weak solution 𝐘(){\boldsymbol{Y}}(\cdot) to (2.1). Furthermore P(𝐘(t)𝒰 for all t0)=1P({\mathbf{Y}}(t)\in{\mathcal{U}}\mbox{ for all }t\geq 0)=1 and for any T<T<\infty and m(0,)m\in(0,\infty), the set {i0:inft[0,T]Yi(t)m}\{i\in{\mathbb{N}}_{0}:\inf_{t\in[0,T]}Y_{i}(t)\leq m\} is finite a.s.

When 𝒈=𝒈1=(1,0,0,){\boldsymbol{g}}={\boldsymbol{g}}^{1}=(1,0,0,\ldots)^{\prime} the process given by the above theorem is the well known (standard) infinite Atlas model. In general, the unique in law solution process given by Theorem 2.1 will be referred to as the (𝒈,𝒚)({\boldsymbol{g}},{\boldsymbol{y}})-infinite Atlas model. Since this solution process 𝐘(t){\mathbf{Y}}(t), under the conditions of the above theorem, is rankable a.s., we can define the ranked process {Y(i)(t),i0}t0\{Y_{(i)}(t),i\in{\mathbb{N}}_{0}\}_{t\geq 0} that gives the unique ranking of 𝐘(t){\mathbf{Y}}(t) (in which ties are broken in the lexicographic order) such that Y(0)(t)Y(1)(t)Y(2)(t)Y_{(0)}(t)\leq Y_{(1)}(t)\leq Y_{(2)}(t)\leq\cdots. From [32, Lemma 3.5], the processes defined by

Bi(t)j=00t𝟏(Yj(s)=Y(i)(s))𝑑Wj(s),i0,t0,B^{*}_{i}(t)\doteq\sum_{j=0}^{\infty}\int_{0}^{t}\mathbf{1}(Y_{j}(s)=Y_{(i)}(s))dW_{j}(s),\ i\in\mathbb{N}_{0},t\geq 0, (2.2)

are independent standard Brownian motions which can be used to write down the following stochastic differential equation (SDE) for {Y(i)():i0}\{Y_{(i)}(\cdot):i\in\mathbb{N}_{0}\}:

dY(i)(t)=gidt+dBi(t)12dLi+1(t)+12dLi(t),t0,Y(i)(0)=y(i),i0.dY_{(i)}(t)=g_{i}dt+dB^{*}_{i}(t)-\frac{1}{2}dL^{*}_{i+1}(t)+\frac{1}{2}dL^{*}_{i}(t),\ t\geq 0,\;\;Y_{(i)}(0)=y_{(i)},\,i\in\mathbb{N}_{0}. (2.3)

Here, L0()0L^{*}_{0}(\cdot)\equiv 0 and for ii\in\mathbb{N}, Li()L^{*}_{i}(\cdot) denotes the local time of collision between the (i1)(i-1)-th and ii-th particles, that is, the unique non-decreasing continuous process satisfying Li(0)=0L^{*}_{i}(0)=0 and Li(t)=0t𝟏(Y(i1)(s)=Y(i)(s))𝑑Li(s)L^{*}_{i}(t)=\int_{0}^{t}\mathbf{1}\left(Y_{(i-1)}(s)=Y_{(i)}(s)\right)dL^{*}_{i}(s) for all t0t\geq 0. The gap process for the (𝒈,𝒚)({\boldsymbol{g}},{\boldsymbol{y}})-infinite Atlas model is the +{\mathbb{R}}_{+}^{\infty}-valued process 𝒁()={\boldsymbol{Z}}(\cdot)= (Z1(),Z2(),)(Z_{1}(\cdot),Z_{2}(\cdot),\dots)^{\prime} defined by

Zi()Y(i)()Y(i1)(),i.Z_{i}(\cdot)\doteq Y_{(i)}(\cdot)-Y_{(i-1)}(\cdot),\ i\in\mathbb{N}. (2.4)

Let

𝒱{𝒛+: for some 𝒚𝒰,𝒛=(y(1)y(0),y(2)y(1),)}.{\mathcal{V}}\doteq\{{\boldsymbol{z}}\in{\mathbb{R}}_{+}^{\infty}:\mbox{ for some }{\boldsymbol{y}}\in{\mathcal{U}},{\boldsymbol{z}}=(y_{(1)}-y_{(0)},y_{(2)}-y_{(1)},\ldots)^{\prime}\}. (2.5)

Note that if 𝒛𝒱{\boldsymbol{z}}\in{\mathcal{V}} then 𝒚(𝒛)(0,z1,z1+z2,)𝒰{\boldsymbol{y}}({\boldsymbol{z}})\doteq(0,z_{1},z_{1}+z_{2},\ldots)^{\prime}\in{\mathcal{U}} and if 𝒚𝒰{\boldsymbol{y}}\in{\mathcal{U}} then 𝒛(𝒚)(y(1)y(0),y(2)y(2),)𝒱{\boldsymbol{z}}({\boldsymbol{y}})\doteq(y_{(1)}-y_{(0)},y_{(2)}-y_{(2)},\ldots)^{\prime}\in{\mathcal{V}}. Thus given 𝒛𝒱{\boldsymbol{z}}\in{\mathcal{V}} we can define a unique in law stochastic process {𝒁(t)}t0\{{\boldsymbol{Z}}(t)\}_{t\geq 0} with values in 𝒱{\mathcal{V}} that can be viewed as a 𝒱{\mathcal{V}}-valued Markov process referred to as the gap process of the 𝐠{\boldsymbol{g}}-Atlas model. The Markov property of 𝒁(){\boldsymbol{Z}}(\cdot) needs justification which we have sketched for completeness in Remark 2.3 at the end of the section.

The following result identifies an important family of stationary distributions of this Markov process. The first statement in the theorem is from [35, Theorem 1.6]. The second statement is due to [32, Section 4.2] and [35, Remark 4]. Recall 𝒟1{\mathcal{D}}_{1} defined in (1.4).

Theorem 2.2 ([35, 32]).

Let 𝐠𝒟{\boldsymbol{g}}\in{\mathcal{D}}. Define for nn\in{\mathbb{N}}, g¯n1n(g0++gn1)\bar{g}_{n}\doteq\frac{1}{n}(g_{0}+\cdots+g_{n-1}). Then for each a>2infng¯na>-2\inf_{n\in{\mathbb{N}}}\bar{g}_{n}, the probability measure πa𝐠\pi_{a}^{{\boldsymbol{g}}} on +{\mathbb{R}}_{+}^{\infty} defined as

πa𝒈n=1Exp(n(2g¯n+a))\pi_{a}^{{\boldsymbol{g}}}\doteq\otimes_{n=1}^{\infty}\operatorname{Exp}(n(2\bar{g}_{n}+a))

is a stationary distribution for the gap process of the 𝐠{\boldsymbol{g}}-Atlas model. Furthermore, if 𝐠𝒟1{\boldsymbol{g}}\in{\mathcal{D}}_{1}, the above statement holds also for the case a=2infng¯n=2limjg¯Nja=-2\inf_{n\in{\mathbb{N}}}\bar{g}_{n}=-2\lim_{j\rightarrow\infty}\bar{g}_{N_{j}}. In particular, when 𝐠=𝐠1(1,0,){\boldsymbol{g}}={\boldsymbol{g}}^{1}\doteq(1,0,\ldots)^{\prime} (infinite Atlas model), πa𝐠1n=1Exp(2+na)\pi_{a}^{{\boldsymbol{g}}^{1}}\doteq\otimes_{n=1}^{\infty}\operatorname{Exp}(2+na) is a stationary distribution for the gap process for all a0a\geq 0.

The existence of the limit limjg¯Nj\lim_{j\rightarrow\infty}\bar{g}_{N_{j}} and the equality infng¯n=limjg¯Nj\inf_{n\in{\mathbb{N}}}\bar{g}_{n}=\lim_{j\rightarrow\infty}\bar{g}_{N_{j}} follow from the definition of 𝒟1{\mathcal{D}}_{1} (see first paragraph of Section 5).

As an immediate consequence of Kakutani’s theorem [22] we see that for any 𝒈𝒟{\boldsymbol{g}}\in{\mathcal{D}} and a,a>2infng¯na,a^{\prime}>-2\inf_{n\in{\mathbb{N}}}\bar{g}_{n} (and a,a2infng¯na,a^{\prime}\geq-2\inf_{n\in{\mathbb{N}}}\bar{g}_{n} when 𝒈𝒟1{\boldsymbol{g}}\in{\mathcal{D}}_{1}), aaa\neq a^{\prime}, the measures πa𝒈\pi_{a}^{{\boldsymbol{g}}} and πa𝒈\pi_{a^{\prime}}^{{\boldsymbol{g}}} are mutually singular.

Remark 2.3 (Markov property of the gap process).

The ranked process {Y(i)()}i0\{Y_{(i)}(\cdot)\}_{i\in{\mathbb{N}}_{0}} is formally constructed in [32] as an ‘approximative version’ using limits of finite-dimensional reflected diffusions. Namely, for any fixed k0k\in\mathbb{N}_{0}, {Y(i)()}0ik\{Y_{(i)}(\cdot)\}_{0\leq i\leq k} is obtained as an almost sure limit as mm\rightarrow\infty, uniformly over compact time intervals, of the first k+1k+1 coordinates of the reflected diffusion {Y(i)m()}0im\{Y_{(i)}^{m}(\cdot)\}_{0\leq i\leq m} given by the SDE (2.3) with Y(i)m(0)=y(i)Y_{(i)}^{m}(0)=y_{(i)}, Lm+1()0L^{*}_{m+1}(\cdot)\equiv 0, and a given collection {Bi}i0\{B_{i}^{*}\}_{i\in{\mathbb{N}}_{0}} of iid standard Brownian motions (see also Definition 4.4). Fix any time t0t\geq 0. Define for any m0m\in\mathbb{N}_{0} the process {Y(i)m,t(t+)}0im\{Y^{m,t}_{(i)}(t+\cdot)\}_{0\leq i\leq m} similarly by setting Y(i)m,t(t)=Y(i)(t)Y_{(i)}^{m,t}(t)=Y_{(i)}(t) and driven by the Brownian motions {Bi(t+)Bi(t)}0im\{B^{*}_{i}(t+\cdot)-B^{*}_{i}(t)\}_{0\leq i\leq m} via the SDE (2.3) (again with the local time for i=m+1i=m+1 set to zero). Define Zim,tZ^{m,t}_{i}, 1im1\leq i\leq m, the gap process sequence associated with Y(i)m,tY_{(i)}^{m,t}, 0im0\leq i\leq m. Fix k0k\in\mathbb{N}_{0}, s>0s>0, and any bounded continuous function f:kf:\mathbb{R}^{k}\rightarrow\mathbb{R}. Let 𝒢tσ{Zi(u):ut,i}{\mathcal{G}}_{\leq t}\doteq\sigma\{Z_{i}(u):u\leq t,i\in{\mathbb{N}}\} and 𝒢tσ{Zi(t):i}{\mathcal{G}}_{t}\doteq\sigma\{Z_{i}(t):i\in{\mathbb{N}}\}. For any mkm\geq k and t>0t>0, as {Bi(t+)Bi(t)}i0\{B^{*}_{i}(t+\cdot)-B^{*}_{i}(t)\}_{i\in\mathbb{N}_{0}} is independent of 𝒢t{\mathcal{G}}_{\leq t},

𝔼(f(Z1m,t(t+s),,Zkm,t(t+s))|𝒢t)=𝔼(f(Z1m,t(t+s),,Zkm,t(t+s))|𝒢t).\mathbb{E}\left(f\left(Z^{m,t}_{1}(t+s),\dots,Z^{m,t}_{k}(t+s)\right)\,\Big{|}\,{\mathcal{G}}_{\leq t}\right)=\mathbb{E}\left(f\left(Z^{m,t}_{1}(t+s),\dots,Z^{m,t}_{k}(t+s)\right)\,\Big{|}\,{\mathcal{G}}_{t}\right).

Thus, to deduce the Markov property, it suffices to show that (Z1m,t(t+s),,Zkm,t(t+s))\left(Z^{m,t}_{1}(t+s),\dots,Z^{m,t}_{k}(t+s)\right) converges almost surely to (Z1(t+s),,Zk(t+s))\left(Z_{1}(t+s),\dots,Z_{k}(t+s)\right) as mm\rightarrow\infty which will follow from the a.s. convergence of (Y(0)m,t(t+s),,Y(k)m,t(t+s))\left(Y^{m,t}_{(0)}(t+s),\dots,Y^{m,t}_{(k)}(t+s)\right) to (Y(0)(t+s),,Y(k)(t+s))\left(Y_{(0)}(t+s),\dots,Y_{(k)}(t+s)\right). The latter can be shown by exploiting monotonicity properties of rank-based diffusions [34, Theorem 3.2 and Corollary 3.9] as follows.

Fix ϵ>0\epsilon>0. By the construction of the process {Y(i)()}i0\{Y_{(i)}(\cdot)\}_{i\in{\mathbb{N}}_{0}} in [32], we can find (random) m00m_{0}\in\mathbb{N}_{0} such that, for any mm0m\geq m_{0}, |Y(i)m(t+s)Y(i)(t+s)|<ϵ|Y^{m}_{(i)}(t+s)-Y_{(i)}(t+s)|<\epsilon, for 0ik0\leq i\leq k. Now fix mm0m\geq m_{0}. Note that Y(i)m,t(t)Y(i)m(t)Y^{m,t}_{(i)}(t)\leq Y^{m}_{(i)}(t) for 0im0\leq i\leq m by [34, Corollary 3.9], and hence by [34, Theorem 3.2],

Y(i)m,t(t+s)Y(i)m(t+s)Y(i)(t+s)+ϵ for  0im.{Y^{m,t}_{(i)}(t+s)\leq Y^{m}_{(i)}(t+s)\leq Y_{(i)}(t+s)+\epsilon\ \text{ for }\ 0\leq i\leq m.}

To construct a lower bounding process for Y(i)m,t(t+s)Y^{m,t}_{(i)}(t+s), define for mmm^{\prime}\geq m the process {Y(i)m,m,t(t+)}0im\{Y^{m^{\prime},m,t}_{(i)}(t+\cdot)\}_{0\leq i\leq m} started with Y(i)m,m,t(t)=Y(i)m(t)Y^{m^{\prime},m,t}_{(i)}(t)=Y^{m^{\prime}}_{(i)}(t) and driven by the Brownian motions {Bi(t+)Bi(t)}0im\{B^{*}_{i}(t+\cdot)-B^{*}_{i}(t)\}_{0\leq i\leq m} via the SDE (2.3) (and again setting the local time for i=m+1i=m+1 to be zero). By the construction of the process {Y(i)()}i0\{Y_{(i)}(\cdot)\}_{i\in{\mathbb{N}}_{0}} in [32], one can choose m=m(m,ϵ)m^{\prime}=m^{\prime}(m,\epsilon) large enough so that Y(i)m(t)Y(i)(t)+ϵY^{m^{\prime}}_{(i)}(t)\leq Y_{(i)}(t)+\epsilon for 0im0\leq i\leq m. Moreover, if we consider the translation of the system {Y(i)m,t(t+)}0im\{Y^{m,t}_{(i)}(t+\cdot)\}_{0\leq i\leq m} by ϵ\epsilon, then the (m+1)(m+1)-particle process started from {Y(i)(t)+ϵ:0im}\{Y_{(i)}(t)+\epsilon:0\leq i\leq m\} at time tt and driven by the Brownian motions {Bi(t+)Bi(t)}0im\{B^{*}_{i}(t+\cdot)-B^{*}_{i}(t)\}_{0\leq i\leq m} has particle locations at time t+st+s given by {Y(i)m,t(t+s)+ϵ:0im}\{Y^{m,t}_{(i)}(t+s)+\epsilon:0\leq i\leq m\}. Hence, using [34, Theorem 3.2 and Corollary 3.9], we have

Y(i)(t+s)ϵY(i)m(t+s)Y(i)m,m,t(t+s)Y(i)m,t(t+s)+ϵ for  0im.Y_{(i)}(t+s)-\epsilon\leq Y^{m^{\prime}}_{(i)}(t+s)\leq Y^{m^{\prime},m,t}_{(i)}(t+s)\leq Y^{m,t}_{(i)}(t+s)+\epsilon\ \text{ for }\ 0\leq i\leq m. (2.6)

The first inequality above holds because mmm0m^{\prime}\geq m\geq m_{0}. We conclude from the above two displays that

Y(i)(t+s)2ϵY(i)m,t(t+s)Y(i)(t+s)+ϵ for  0ik.Y_{(i)}(t+s)-2\epsilon\leq Y^{m,t}_{(i)}(t+s)\leq Y_{(i)}(t+s)+\epsilon\ \text{ for }\ 0\leq i\leq k. (2.7)

This gives the desired almost sure convergence as ϵ>0\epsilon>0 is arbitrary.

3 Main Results

We are interested in the extremality properties of the probability measures πa𝒈\pi_{a}^{{\boldsymbol{g}}}. We also ask whether these are the only product form stationary distributions.

We begin with some notation. Recall 𝕏𝒞([0,):+){\mathbb{X}}\doteq{\mathcal{C}}([0,\infty):{\mathbb{R}}_{+}^{\infty}). Define measurable maps {θt}t0\{\theta_{t}\}_{t\geq 0} from (𝕏,(𝕏))({\mathbb{X}},{\mathcal{B}}({\mathbb{X}})) to itself as

θt(𝒁)(s)𝒁(t+s),t0,s0,𝒁𝕏.\theta_{t}({\boldsymbol{Z}})(s)\doteq{\boldsymbol{Z}}(t+s),\;t\geq 0,s\geq 0,\;{\boldsymbol{Z}}\in{\mathbb{X}}.

Given 𝒈𝒟{\boldsymbol{g}}\in{\mathcal{D}} and 𝒛𝒱{\boldsymbol{z}}\in{\mathcal{V}}, we denote the probability distribution of the gap process of the 𝒈{\boldsymbol{g}}-Atlas model on (𝕏,(𝕏))({\mathbb{X}},{\mathcal{B}}({\mathbb{X}})), with initial gap sequence 𝒛{\boldsymbol{z}}, by 𝒛𝒈{\mathbb{P}}^{{\boldsymbol{g}}}_{{\boldsymbol{z}}}. Also, for γ𝒫(+)\gamma\in{\mathcal{P}}({\mathbb{R}}_{+}^{\infty}) supported on 𝒱{\mathcal{V}} (namely, γ(𝒱)=1\gamma({\mathcal{V}})=1), let γ𝒈+𝒛𝒈γ(d𝒛){\mathbb{P}}^{{\boldsymbol{g}}}_{\gamma}\doteq\int_{{\mathbb{R}}_{+}^{\infty}}{\mathbb{P}}^{{\boldsymbol{g}}}_{{\boldsymbol{z}}}\;\gamma(d{\boldsymbol{z}}). The corresponding expectation operators will be denoted as 𝔼𝒛𝒈{\mathbb{E}}^{{\boldsymbol{g}}}_{{\boldsymbol{z}}} and 𝔼γ𝒈{\mathbb{E}}^{{\boldsymbol{g}}}_{\gamma} respectively. Denote by 𝒈{\mathcal{I}}^{{\boldsymbol{g}}} the collection of all invariant (stationary) probability measures of the gap process of the 𝒈{\boldsymbol{g}}-Atlas model supported on 𝒱{\mathcal{V}}, namely

𝒈{γ𝒫(𝕏):γ(𝒱c)=0, and γ𝒈θt1=γ𝒈 for all t0}.{\mathcal{I}}^{{\boldsymbol{g}}}\doteq\{\gamma\in{\mathcal{P}}({\mathbb{X}}):\gamma({\mathcal{V}}^{c})=0,\mbox{ and }{\mathbb{P}}^{{\boldsymbol{g}}}_{\gamma}\circ\theta_{t}^{-1}={\mathbb{P}}^{{\boldsymbol{g}}}_{\gamma}\mbox{ for all }t\geq 0\}.

Abusing notation, the canonical coordinate process on (𝕏,(𝕏))({\mathbb{X}},{\mathcal{B}}({\mathbb{X}})) will be denoted by {𝒁(t)}t0\{{\boldsymbol{Z}}(t)\}_{t\geq 0}. Let (+){\mathcal{M}}({\mathbb{R}}_{+}^{\infty}) be the collection of all real measurable maps on +{\mathbb{R}}_{+}^{\infty}. For f(+)f\in{\mathcal{M}}({\mathbb{R}}_{+}^{\infty}), 𝒛+{\boldsymbol{z}}\in{\mathbb{R}}_{+}^{\infty} and t0t\geq 0 such that 𝔼𝒛𝒈(|f(𝒁(t))|)<{\mathbb{E}}^{{\boldsymbol{g}}}_{{\boldsymbol{z}}}(|f({\boldsymbol{Z}}(t))|)<\infty we write

Tt𝒈f(𝒛)𝔼𝒛𝒈(f(𝒁(t))).T_{t}^{{\boldsymbol{g}}}f({\boldsymbol{z}})\doteq{\mathbb{E}}^{{\boldsymbol{g}}}_{{\boldsymbol{z}}}(f({\boldsymbol{Z}}(t))). (3.1)

Note that for 𝒈𝒟{\boldsymbol{g}}\in{\mathcal{D}}, γ𝒈\gamma\in{\mathcal{I}}^{{\boldsymbol{g}}}, and ψL2(γ)\psi\in L^{2}(\gamma), Tt𝒈ψT_{t}^{{\boldsymbol{g}}}\psi is γ\gamma a.e. well defined and belongs to L2(γ)L^{2}(\gamma). Furthermore, the collection {Tt𝒈}t0\{T_{t}^{{\boldsymbol{g}}}\}_{t\geq 0} defines a contraction semigroup on L2(γ)L^{2}(\gamma), namely

Tt𝒈Ts𝒈ψ=Tt+s𝒈ψ, and Tt𝒈ψψ for all s,t0 and ψL2(γ).T_{t}^{{\boldsymbol{g}}}T_{s}^{{\boldsymbol{g}}}\psi=T_{t+s}^{{\boldsymbol{g}}}\psi,\mbox{ and }\|T_{t}^{{\boldsymbol{g}}}\psi\|\leq\|\psi\|\mbox{ for all }s,t\geq 0\mbox{ and }\psi\in L^{2}(\gamma).

We now recall the definition of extremality and ergodicity. Let, for 𝒈{\boldsymbol{g}} as above and γ𝒈\gamma\in{\mathcal{I}}^{{\boldsymbol{g}}}, 𝕀γ𝒈{\mathbb{I}}_{\gamma}^{{\boldsymbol{g}}} be the collection of all Tt𝒈T_{t}^{{\boldsymbol{g}}}-invariant square integrable functions, namely,

𝕀γ𝒈{ψL2(γ):Tt𝒈ψ=ψ,γ a.s., for all t0}.{\mathbb{I}}_{\gamma}^{{\boldsymbol{g}}}\doteq\{\psi\in L^{2}(\gamma):T_{t}^{{\boldsymbol{g}}}\psi=\psi,\;\gamma\mbox{ a.s., for all }t\geq 0\}.

We denote the projection of a ψL2(γ)\psi\in L^{2}(\gamma) on to the closed subspace 𝕀γ𝒈{\mathbb{I}}_{\gamma}^{{\boldsymbol{g}}} as ψ^γ𝒈\hat{\psi}_{\gamma}^{{\boldsymbol{g}}}. Namely, ψ^γ𝒈\hat{\psi}_{\gamma}^{{\boldsymbol{g}}} is the unique element of 𝕀γ𝒈{\mathbb{I}}_{\gamma}^{{\boldsymbol{g}}} that satisfies

ψ,η=ψ^γ𝒈,η, for all η𝕀γ𝒈.\langle\psi,\eta\rangle=\langle\hat{\psi}_{\gamma}^{{\boldsymbol{g}}},\eta\rangle,\mbox{ for all }\eta\in{\mathbb{I}}_{\gamma}^{{\boldsymbol{g}}}.

This projection can be obtained as the limit of 1t0tTs𝒈ψ𝑑s\frac{1}{t}\int_{0}^{t}T_{s}^{{\boldsymbol{g}}}\psi\,ds in L2(γ)L^{2}(\gamma) as tt\rightarrow\infty (see (A.3)). Thus, for any ψL2(γ)\psi\in L^{2}(\gamma), ψ^γ𝒈()\hat{\psi}_{\gamma}^{{\boldsymbol{g}}}(\cdot) can be intuitively interpreted as the ‘long-time average’ of {𝔼𝒈(ψ(𝒁(t)):t0}\{{\mathbb{E}}^{{\boldsymbol{g}}}_{\cdot}(\psi({\boldsymbol{Z}}(t)):t\geq 0\}.

Definition 3.1.

Let 𝐠𝒟{\boldsymbol{g}}\in{\mathcal{D}}. A ν𝐠\nu\in{\mathcal{I}}^{{\boldsymbol{g}}} is said to be an extremal invariant distribution of the gap process of the 𝐠{\boldsymbol{g}}-Atlas model if, whenever for some ε(0,1){\varepsilon}\in(0,1) and ν1,ν2𝐠\nu_{1},\nu_{2}\in{\mathcal{I}}^{{\boldsymbol{g}}} we have ν=εν1+(1ε)ν2\nu={\varepsilon}\nu_{1}+(1-{\varepsilon})\nu_{2}, then ν1=ν2=ν\nu_{1}=\nu_{2}=\nu. We denote the collection of all such measures by e𝐠{\mathcal{I}}^{{\boldsymbol{g}}}_{e}.

We call ν𝐠\nu\in{\mathcal{I}}^{{\boldsymbol{g}}} an ergodic probability measure for the invariant distribution of the gap process of the 𝐠{\boldsymbol{g}}-Atlas model if for all ψL2(ν)\psi\in L^{2}(\nu), ψ^ν𝐠\hat{\psi}_{\nu}^{{\boldsymbol{g}}} is constant ν\nu-a.s. We denote the collection of all such measures by er𝐠{\mathcal{I}}^{{\boldsymbol{g}}}_{er}.

We note that (cf. proof of Lemma 3.2 below) if γer𝒈\gamma\in{\mathcal{I}}^{{\boldsymbol{g}}}_{er}, then for any ψL2(γ)\psi\in L^{2}(\gamma),

1t0tTs𝒈ψ()𝑑s+ψ(𝒙)γ(d𝒙), in L2(γ), as t.\frac{1}{t}\int_{0}^{t}T_{s}^{{\boldsymbol{g}}}\psi(\cdot)ds\to\int_{{\mathbb{R}}_{+}^{\infty}}\psi({\boldsymbol{x}})\gamma(d{\boldsymbol{x}}),\mbox{ in }L^{2}(\gamma),\mbox{ as }t\to\infty.

The following result, which says that extremal invariant measures and ergodic invariant measures are the same, is standard, however we provide a proof in the appendix for completeness.

Lemma 3.2.

Let 𝐠𝒟{\boldsymbol{g}}\in{\mathcal{D}}. Then e𝐠=er𝐠{\mathcal{I}}^{{\boldsymbol{g}}}_{e}={\mathcal{I}}^{{\boldsymbol{g}}}_{er}. Let γ𝐠\gamma\in{\mathcal{I}}^{{\boldsymbol{g}}} and suppose that for every bounded measurable ψ:+\psi:{\mathbb{R}}_{+}^{\infty}\to{\mathbb{R}}, ψ^γ𝐠\hat{\psi}_{\gamma}^{{\boldsymbol{g}}} is constant, γ\gamma a.s. Then γe𝐠\gamma\in{\mathcal{I}}^{{\boldsymbol{g}}}_{e}.

The following is the first main result of this work.

Theorem 3.3.

Let 𝐠𝒟{\boldsymbol{g}}\in{\mathcal{D}}. Then, for every a>2infng¯na>-2\inf_{n\in{\mathbb{N}}}\bar{g}_{n}, πa𝐠e𝐠=er𝐠\pi_{a}^{{\boldsymbol{g}}}\in{\mathcal{I}}^{{\boldsymbol{g}}}_{e}={\mathcal{I}}^{{\boldsymbol{g}}}_{er}. Furthermore, when 𝐠𝒟1{\boldsymbol{g}}\in{\mathcal{D}}_{1}, πa𝐠e𝐠=er𝐠\pi_{a}^{{\boldsymbol{g}}}\in{\mathcal{I}}^{{\boldsymbol{g}}}_{e}={\mathcal{I}}^{{\boldsymbol{g}}}_{er} also for a=2infng¯na=-2\inf_{n\in{\mathbb{N}}}\bar{g}_{n}.

The above theorem proves the extremality of the invariant measures πa𝒈\pi_{a}^{{\boldsymbol{g}}} for suitable values of aa. As an immediate consequence of this theorem one can identify natural collections of measures that are in the (weak) domain of attraction of a given πa𝒈\pi^{{\boldsymbol{g}}}_{a}, as noted in the corollary below. We recall that a measure γ𝒫(+)\gamma\in{\mathcal{P}}({\mathbb{R}}_{+}^{\infty}) is said to be in the weak (resp. strong) domain of attraction of πa𝒈\pi^{{\boldsymbol{g}}}_{a} if for any bounded continuous function ψ:+\psi:{\mathbb{R}}_{+}^{\infty}\rightarrow{\mathbb{R}},

1t0t𝔼γ𝒈(ψ(𝒁(s)))𝑑s+ψ(𝒙)πa𝒈(d𝒙),\frac{1}{t}\int_{0}^{t}{\mathbb{E}}^{{\boldsymbol{g}}}_{\gamma}(\psi({\boldsymbol{Z}}(s)))ds\to\int_{{\mathbb{R}}_{+}^{\infty}}\psi({\boldsymbol{x}})\pi^{{\boldsymbol{g}}}_{a}(d{\boldsymbol{x}}),

(resp. 𝔼γ𝒈(ψ(𝒁(t)))+ψ(𝒙)πa𝒈(d𝒙){\mathbb{E}}^{{\boldsymbol{g}}}_{\gamma}(\psi({\boldsymbol{Z}}(t)))\to\int_{{\mathbb{R}}_{+}^{\infty}}\psi({\boldsymbol{x}})\pi^{{\boldsymbol{g}}}_{a}(d{\boldsymbol{x}})), as tt\to\infty.

Corollary 3.4.

Let 𝐠𝒟{\boldsymbol{g}}\in{\mathcal{D}}. Fix any a>2infng¯na>-2\inf_{n\in{\mathbb{N}}}\bar{g}_{n}. Let γ𝒫(+)\gamma\in{\mathcal{P}}({\mathbb{R}}_{+}^{\infty}) be absolutely continuous with respect to πa𝐠\pi^{{\boldsymbol{g}}}_{a}. Then γ\gamma lies in the weak domain of attraction of πa𝐠\pi^{{\boldsymbol{g}}}_{a}. The assertion holds for any a2infng¯na\geq-2\inf_{n\in{\mathbb{N}}}\bar{g}_{n} if 𝐠𝒟1{\boldsymbol{g}}\in{\mathcal{D}}_{1}.

Sufficient conditions for a probability measure to be in the strong domain of attraction of π0𝒈1\pi^{{\boldsymbol{g}}^{1}}_{0} were obtained in [32, 11] whereas weak domain of attraction results for πa𝒈1\pi^{{\boldsymbol{g}}^{1}}_{a}, a0a\geq 0, have been obtained in [6]. The above corollary provides a weak domain of attraction result for a general class of 𝒈{\boldsymbol{g}}-Atlas models.

One can ask whether these are the only extremal invariant measures of the gap process of the 𝒈{\boldsymbol{g}}-Atlas model supported on 𝒱{\mathcal{V}}. As noted in the Introduction, the answer to this question when 𝒈=𝟎{\boldsymbol{g}}={\boldsymbol{0}} is affirmative from results of [30, Theorem 4.2] , if one restricts to extremal measures satisfying certain integrability constraints on the denseness of particle configurations. For a general 𝒈𝒟{\boldsymbol{g}}\in{\mathcal{D}} (in fact even for 𝒈=𝒈1{\boldsymbol{g}}={\boldsymbol{g}}^{1}) this is currently a challenging open problem. However, we make partial progress towards this goal in the next result by showing that for any 𝒈𝒟1{\boldsymbol{g}}\in{\mathcal{D}}_{1} (and under a mild integrability condition), the collection {πa𝒈}\{\pi_{a}^{{\boldsymbol{g}}}\} exhausts all the extremal product form invariant distributions. In fact we prove the substantially stronger statement that the measures πa𝒈\pi_{a}^{{\boldsymbol{g}}} are the only product form (extremal or not) invariant distributions under a mild integrability condition. Qualitatively, this result says that these measures are the only invariant distributions that preserve independence of the marginal laws of the gaps in time.

Theorem 3.5.

Let 𝐠𝒟1{\boldsymbol{g}}\in{\mathcal{D}}_{1} and let π𝐠\pi\in{\mathcal{I}}^{{\boldsymbol{g}}} be a product measure, i.e. π=i=1πi\pi=\otimes_{i=1}^{\infty}\pi_{i} for some πi𝒫(+)\pi_{i}\in{\mathcal{P}}({\mathbb{R}}_{+}), ii\in{\mathbb{N}}. With F(𝐳)j=1e14(l=1jzl)2,𝐳+F({\boldsymbol{z}})\doteq\sum_{j=1}^{\infty}e^{-\frac{1}{4}(\sum_{l=1}^{j}z_{l})^{2}},\,{\boldsymbol{z}}\in{\mathbb{R}}_{+}^{\infty}, suppose that

+F(𝒛)π(d𝒛)<.\int_{{\mathbb{R}}_{+}^{\infty}}F({\boldsymbol{z}})\pi(d{\boldsymbol{z}})<\infty. (3.2)

Then, for some a2infng¯na\geq-2\inf_{n\in{\mathbb{N}}}\bar{g}_{n}, π=πa𝐠\pi=\pi_{a}^{{\boldsymbol{g}}}. Moreover, aa has the representation

a=𝔼π𝒈(L1(1))2g0=𝔼π𝒈(Lk(1))k2k(g0++gk1),a={\mathbb{E}}^{{\boldsymbol{g}}}_{\pi}(L^{*}_{1}(1))-2g_{0}=\frac{{\mathbb{E}}^{{\boldsymbol{g}}}_{\pi}(L^{*}_{k}(1))}{k}-\frac{2}{k}(g_{0}+\dots+g_{k-1}), (3.3)

for any kk\in{\mathbb{N}}, where {Li}i\{L^{*}_{i}\}_{i\in{\mathbb{N}}} denote the collision local times in (2.3).

Recall that 𝒱{\mathcal{V}} defined in (2.5) consists of 𝒛+{\boldsymbol{z}}\in{\mathbb{R}}_{+}^{\infty} for which j=1eα(l=1jzl)2<\sum_{j=1}^{\infty}e^{-\alpha(\sum_{l=1}^{j}z_{l})^{2}}<\infty for all α>0\alpha>0. In comparison, condition (3.2) requires a finite expectation of j=1e14(l=1jzl)2\sum_{j=1}^{\infty}e^{-\frac{1}{4}(\sum_{l=1}^{j}z_{l})^{2}} when 𝒛{\boldsymbol{z}} is distributed as π\pi. Roughly speaking, condition (3.2) puts a restriction on the rate of increase of the density of particles as one moves away from the lowest ranked particle.

Several remarks are in order.

Remark 3.6 (Probabilistic interpretation of aa).

The equalities (3.3) give a probabilistic interpretation to aa. By stationarity of π\pi, 𝔼π𝐠(L1(1)){\mathbb{E}}^{{\boldsymbol{g}}}_{\pi}(L^{*}_{1}(1)) can be thought of as the expected rate of change of the local time L1L^{*}_{1}. Hence, 𝔼π𝐠(L1(1))2\frac{{\mathbb{E}}^{{\boldsymbol{g}}}_{\pi}(L^{*}_{1}(1))}{2} is intuitively the expected rate at which the bottom particle is ‘pushed down’ by the particle above it during collisions and g0g_{0} denotes its upward drift in time. Thus, a/2a/2 captures the difference between two kinds of forces acting on the bottom particle: the hardcore interactions due to collisions and the soft potential corresponding to the drift. For k>1k>1, one obtains a similar interpretation as follows. Consider the subsystem consisting of the kk lowest particles viewed as a rank based diffusion (Y0k,,Yk1k)(Y^{k}_{0},\dots,Y^{k}_{k-1}), where YikY^{k}_{i} gets upward drift gjg_{j} if its rank in the subsystem is jj, and it is reflected downwards when it collides with the minimum of the particles outside the subsystem. It can be deduced that each particle YikY^{k}_{i} accrues roughly the same proportion of local time due to downward reflection as time grows. Moreover, it asymptotically spends an equal proportion of time at each rank j{0,,k1}j\in\{0,\dots,k-1\}. Hence, 𝔼π𝐠(Lk(1))2k\frac{{\mathbb{E}}^{{\boldsymbol{g}}}_{\pi}(L^{*}_{k}(1))}{2k} and 1k(g0++gk1)\frac{1}{k}(g_{0}+\dots+g_{k-1}) respectively quantify the effect of reflection and drift on each particle among the lowest kk particles, and a/2a/2 captures the difference between these effects. The positivity of aa implies that the hardcore interactions dominate in the long term. Indeed, the results of [41] show that when 𝐠=𝐠1{\boldsymbol{g}}={\boldsymbol{g}}^{1}, under πa𝐠1\pi^{{\boldsymbol{g}}^{1}}_{a}, for any kk\in{\mathbb{N}}, Y(k)(t)/ta/2Y_{(k)}(t)/t\rightarrow-a/2 almost surely as tt\rightarrow\infty. We conjecture that the same result is true for any 𝐠𝒟1{\boldsymbol{g}}\in{\mathcal{D}}_{1}.

Remark 3.7.

When 𝐠𝒟{\boldsymbol{g}}\notin{\mathcal{D}}, uniqueness in law for the infinite-dimensional gap process is currently an open problem, and therefore what one means by a stationary distribution is not clear. However, under conditions, for 𝐠𝒟{\boldsymbol{g}}\notin{\mathcal{D}}, one can still construct stationary ‘approximative’ versions of this gap process by taking ‘limits’ of finite-dimensional processes [32, Definition 7] (see also Definition 4.4 below). See [32, Theorem 4.4, Lemma 4.5 and Section 4.2] and [35, Remark 3] for some situations where such versions can be constructed. Theorem 3.5 can be extended to such settings as follows, as is clear from an inspection of the proof. Suppose 𝐠{\boldsymbol{g}} satisfies infn1g¯n>\inf_{n\geq 1}\bar{g}_{n}>-\infty and there exist N1<N2<N_{1}<N_{2}<\dots\rightarrow\infty such that g¯k>g¯Nj\bar{g}_{k}>\bar{g}_{N_{j}} for all k=1,,Nj1,j1k=1,\dots,N_{j}-1,\ j\geq 1. If there is a stationary approximative version of the infinite-dimensional gap process with initial (invariant) distribution π\pi supported on 𝒱{\mathcal{V}}, and if π\pi is a product measure that satisfies the integrability property in (3.2), then π=πa𝐠\pi=\pi_{a}^{{\boldsymbol{g}}} for some a2limjg¯Nja\geq-2\lim_{j\rightarrow\infty}\bar{g}_{N_{j}}.

Remark 3.8.

Stationary distributions for finite dimensional reflected Brownian motions (with drift) in the positive orthant, have been studied in [16, 15]. In particular, the paper [15] shows that the unique stationary distribution can be characterized through an identity, holding true for all suitable smooth test functions, referred to as the Basic Adjoint Relationship (BAR) (see [15, Section 8]). Using this characterization it is shown in [15, Section 9] that if the stationary distribution is product form then it must necessarily be a product of Exponential distributions. The proof relies on using a suitable class of exponential test functions in the BAR to characterize the moment generating function of the stationary distribution. In the infinite dimensional setting considered here, although similar test functions are useful, we do not know of a similar BAR characterization for all stationary distributions. To circumvent this, we show in Lemma 5.4 that for any product form stationary distribution, one can obtain ‘local’ descriptions for the expectations of certain path functionals of the process. This result is key and essentially plays the role of BAR in our context in obtaining a recursive system of equations for the moment generating functions of the marginal distributions that can then be solved explicitly to prove Theorem 3.5.

Remark 3.9.

A referee has proposed the following interesting direction of investigation. Suppose that γπb𝐠\gamma\ll\pi_{b}^{{\boldsymbol{g}}} for some b2infng¯nb\geq-2\inf_{n\in{\mathbb{N}}}\bar{g}_{n}. Then one may conjecture that, for each kk\in{\mathbb{N}},

𝔼γ𝒈(Lk(t))kt2k(g0++gk1)b as t.\frac{{\mathbb{E}}^{{\boldsymbol{g}}}_{\gamma}(L^{*}_{k}(t))}{kt}-\frac{2}{k}(g_{0}+\dots+g_{k-1})\to b\mbox{ as }t\to\infty.

One may also ask the following ‘domain of attraction’ question. Given b2infng¯nb\geq-2\inf_{n\in{\mathbb{N}}}\bar{g}_{n}, identify the set 𝒱b𝒱{\mathcal{V}}_{b}\subset{\mathcal{V}} such that for z𝒱bz\in{\mathcal{V}}_{b}, for each kk\in{\mathbb{N}},

𝔼z𝒈(Lk(t))kt2k(g0++gk1)b as t.\frac{{\mathbb{E}}^{{\boldsymbol{g}}}_{z}(L^{*}_{k}(t))}{kt}-\frac{2}{k}(g_{0}+\dots+g_{k-1})\to b\mbox{ as }t\to\infty.

In this case, we conjecture that any γ𝒫(+)\gamma\in{\mathcal{P}}({\mathbb{R}}_{+}^{\infty}) supported on 𝒱b{\mathcal{V}}_{b} is in the strong domain of attraction of πb𝐠\pi_{b}^{{\boldsymbol{g}}}. We leave the study of these questions for future work.

Rest of the paper is devoted to the proofs of Theorem 3.3 and Theorem 3.5. Proof of Lemma 3.2 is given in the Appendix for completeness.

4 Proof of Theorem 3.3

We will only prove the first statement in Theorem 3.3. The proof of the second is similar and is therefore omitted.

We begin with the following definition. Let 𝕐𝒞([0,):+×+){\mathbb{Y}}\doteq{\mathcal{C}}([0,\infty):{\mathbb{R}}_{+}^{\infty}\times{\mathbb{R}}_{+}^{\infty}).

Definition 4.1.

Let 𝐠𝒟{\boldsymbol{g}}\in{\mathcal{D}} and γ,γ𝒫(+)\gamma,\gamma^{\prime}\in{\mathcal{P}}({\mathbb{R}}_{+}^{\infty}) be such that γ(𝒱)=γ(𝒱)=1\gamma({\mathcal{V}})=\gamma^{\prime}({\mathcal{V}})=1. We say that a probability measure γ,γ𝐠{\mathbb{P}}^{{\boldsymbol{g}}}_{\gamma,\gamma^{\prime}} on (𝕐,(𝕐))({\mathbb{Y}},{\mathcal{B}}({\mathbb{Y}})) defines a coupling of the gap process of the 𝐠{\boldsymbol{g}}-Atlas model with initial distributions γ,γ\gamma,\gamma^{\prime}, if, denoting the coordinate processes on 𝕐{\mathbb{Y}} as 𝐙(1){\boldsymbol{Z}}^{(1)} and 𝐙(2){\boldsymbol{Z}}^{(2)}, namely

𝒁(1)(ω)(t)ω(1)(t),𝒁(2)(ω)(t)ω(2)(t),ω=(ω(1),ω(2))𝕐,t0,{\boldsymbol{Z}}^{(1)}(\omega)(t)\doteq\omega^{(1)}(t),\;\;{\boldsymbol{Z}}^{(2)}(\omega)(t)\doteq\omega^{(2)}(t),\;\omega=(\omega^{(1)},\omega^{(2)})\in{\mathbb{Y}},\;t\geq 0,

we have

γ,γ𝒈(𝒁(1))1=γ𝒈,γ,γ𝒈(𝒁(2))1=γ𝒈.{\mathbb{P}}^{{\boldsymbol{g}}}_{\gamma,\gamma^{\prime}}\circ({\boldsymbol{Z}}^{(1)})^{-1}={\mathbb{P}}^{{\boldsymbol{g}}}_{\gamma},\;\;{\mathbb{P}}^{{\boldsymbol{g}}}_{\gamma,\gamma^{\prime}}\circ({\boldsymbol{Z}}^{(2)})^{-1}={\mathbb{P}}^{{\boldsymbol{g}}}_{\gamma^{\prime}}.

Define the coupling time

τcinf{t0:𝒁(1)(s)=𝒁(2)(s) for all st},\tau_{c}\doteq\inf\{t\geq 0:{\boldsymbol{Z}}^{(1)}(s)={\boldsymbol{Z}}^{(2)}(s)\text{ for all }s\geq t\},

where τc\tau_{c}\doteq\infty if the above set is empty. When γ=δ𝐳\gamma=\delta_{{\boldsymbol{z}}} and γ=δ𝐳\gamma^{\prime}=\delta_{{\boldsymbol{z}}^{\prime}} for some 𝐳,𝐳𝒱{\boldsymbol{z}},{\boldsymbol{z}}^{\prime}\in{\mathcal{V}}, we write γ,γ𝐠=𝐳,𝐳𝐠{\mathbb{P}}^{{\boldsymbol{g}}}_{\gamma,\gamma^{\prime}}={\mathbb{P}}^{{\boldsymbol{g}}}_{{\boldsymbol{z}},{\boldsymbol{z}}^{\prime}}.

Since 𝒈𝒟{\boldsymbol{g}}\in{\mathcal{D}} will be fixed throughout the section, we will frequently suppress it from the notation.

Consider now γ=πa𝒈\gamma=\pi_{a}^{{\boldsymbol{g}}}, where aa is as in the statement of the theorem, and a bounded measurable map ψ0:+\psi_{0}:{\mathbb{R}}_{+}^{\infty}\to{\mathbb{R}} such that

Ttψ0=ψ0,γ a.s. for every t0,T_{t}\psi_{0}=\psi_{0},\;\gamma\mbox{ a.s. for every }t\geq 0, (4.1)

where TtT_{t} is defined as in (3.1), namely, Ttψ0(𝒛)T_{t}\psi_{0}({\boldsymbol{z}}) is defined for γ\gamma a.e. zz as Ttψ0(𝒛)=𝔼𝒛(ψ0(𝒁(t)))T_{t}\psi_{0}({\boldsymbol{z}})={\mathbb{E}}_{{\boldsymbol{z}}}(\psi_{0}({\boldsymbol{Z}}(t))), t0t\geq 0. In order to prove Theorem 3.3 it suffices, in view of Lemma 3.2, to show that ψ0\psi_{0} is γ\gamma-a.e. constant. This, in view of (4.1), is equivalent to showing that for some fixed t0>0t_{0}>0, ψTt0ψ0\psi\doteq T_{t_{0}}\psi_{0} is γ\gamma-a.e. constant. For the rest of the section we will fix a t0>0t_{0}>0 and consider ψ\psi defined as above. Note that (4.1) holds with ψ0\psi_{0} replaced by ψ\psi.

4.1 Proof overview

Before we proceed to the details, we give a brief overview of the proof strategy for showing that ψ\psi is γ\gamma-a.e. constant. The first step is to show using the TtT_{t}-invariance of ψ\psi that for any t0t\geq 0, ψ(𝒁(t))=ψ(𝒛)\psi({\boldsymbol{Z}}(t))=\psi({\boldsymbol{z}}) for γ\gamma-a.e. 𝒛{\boldsymbol{z}}. Moreover, using the product form of γ\gamma, the same conclusion is seen to hold for the process 𝒁(){\boldsymbol{Z}}(\cdot) started from a ‘perturbed’ initial point obtained by changing any two co-ordinates of γ\gamma-a.e. 𝒛{\boldsymbol{z}} by given numbers (see (4.11)). Up to this point, we only use quite general arguments not involving the specific dynamics of the 𝒈{\boldsymbol{g}}-Atlas model. However, the dynamics comes into play crucially in the subsequent steps, which involve construction of a coupling of two 𝒈{\boldsymbol{g}}-Atlas models started from initial points that differ at a finite number of co-ordinates. This is achieved by a combination of ‘mirror’ and synchronous couplings of the infinite collection of driving Brownian motions (see (4.20) and (4.21)). The coupling is utilized in two ways. First, it is shown that for any s>0s>0, the coupled 𝒈{\boldsymbol{g}}-Atlas models coalesce with positive probability by time ss (Proposition 4.2). Using this and (4.11), it follows that the value of ψ\psi remains unchanged upon changing any pair of coordinates by rational numbers (see (4.13)). To extend this to perturbation by real numbers (see (4.14) and (4.15)), we need a key ‘directional strong Feller property’ described in Proposition 4.3, which is once again established using the coupling. The equality of ψ\psi under pairwise perturbations is then extended to perturbation by any finite permutation via straightforward algebraic manipulations. The proof of γ\gamma almost sure constancy of ψ\psi, and hence of Theorem 3.3, is finally achieved by an application of the Hewitt-Savage zero-one law.

4.2 Preliminary Results

Now, we proceed to the details. We begin by noting that, from (4.1) (with ψ0\psi_{0} replaced by ψ\psi),

0\displaystyle 0 =𝔼γ(ψ(𝒁(t))2)𝔼γ(ψ(𝒁(0))2)=𝔼γ(ψ(𝒁(t))2)+𝔼γ(ψ(𝒁(0))2)2𝔼γ(ψ(𝒁(0))ψ(𝒁(t)))\displaystyle={\mathbb{E}}_{\gamma}\left(\psi({\boldsymbol{Z}}(t))^{2}\right)-{\mathbb{E}}_{\gamma}\left(\psi({\boldsymbol{Z}}(0))^{2}\right)={\mathbb{E}}_{\gamma}\left(\psi({\boldsymbol{Z}}(t))^{2}\right)+{\mathbb{E}}_{\gamma}\left(\psi({\boldsymbol{Z}}(0))^{2}\right)-2{\mathbb{E}}_{\gamma}\left(\psi({\boldsymbol{Z}}(0))\psi({\boldsymbol{Z}}(t))\right)
=𝔼γ(ψ(𝒁(t))ψ(𝒁(0)))2.\displaystyle={\mathbb{E}}_{\gamma}\left(\psi({\boldsymbol{Z}}(t))-\psi({\boldsymbol{Z}}(0))\right)^{2}.

This says

ψ(𝒁(t))=ψ(𝒁(0)),γ a.s., for every t0.\psi({\boldsymbol{Z}}(t))=\psi({\boldsymbol{Z}}(0)),\;{\mathbb{P}}_{\gamma}\mbox{ a.s., for every }t\geq 0. (4.2)

Next let

H(𝒛,t)𝔼𝒛|ψ(𝒁(t))ψ(𝒁(0))|,𝒛𝒱,t0.H({\boldsymbol{z}},t)\doteq{\mathbb{E}}_{{\boldsymbol{z}}}|\psi({\boldsymbol{Z}}(t))-\psi({\boldsymbol{Z}}(0))|,\;{\boldsymbol{z}}\in{\mathcal{V}},\,t\geq 0. (4.3)

For ii\in{\mathbb{N}}, x,y+x,y\in{\mathbb{R}}_{+}, and γ¯𝒫(𝒱)\bar{\gamma}\in{\mathcal{P}}({\mathcal{V}}) of the form γ¯=i=1γ¯i\bar{\gamma}=\otimes_{i=1}^{\infty}\bar{\gamma}_{i} for some γ¯i𝒫(+)\bar{\gamma}_{i}\in{\mathcal{P}}({\mathbb{R}}_{+}), ii\in{\mathbb{N}}, define

ηiγ¯(x,y,t)+H(z1,,zi1,x,y,zi+2,,t)j{i,i+1}γ¯j(dzj).\eta_{i}^{\bar{\gamma}}(x,y,t)\doteq\int_{{\mathbb{R}}_{+}^{\infty}}H(z_{1},\ldots,z_{i-1},x,y,z_{i+2},\ldots,t)\prod_{j\in{\mathbb{N}}\setminus\{i,i+1\}}\bar{\gamma}_{j}(dz_{j}). (4.4)

We have from the Markov property that

𝔼γ¯|ψ(𝒁(t))ψ(𝒁(0))|=+2ηiγ¯(x,y,t)γ¯i(dx)γ¯i+1(dy).{\mathbb{E}}_{\bar{\gamma}}\left|\psi({\boldsymbol{Z}}(t))-\psi({\boldsymbol{Z}}(0))\right|=\int_{{\mathbb{R}}_{+}^{2}}\eta_{i}^{\bar{\gamma}}(x,y,t)\bar{\gamma}_{i}(dx)\bar{\gamma}_{i+1}(dy). (4.5)

Now take γ¯=γ=πa𝒈=i=1γi\bar{\gamma}=\gamma=\pi_{a}^{{\boldsymbol{g}}}=\otimes_{i=1}^{\infty}\gamma_{i}. Then, from (4.2),

0\displaystyle 0 =𝔼γ|ψ(𝒁(t))ψ(𝒁(0))|=+2ηiγ(x,y,t)γi(dx)γi+1(dy).\displaystyle={\mathbb{E}}_{\gamma}\left|\psi({\boldsymbol{Z}}(t))-\psi({\boldsymbol{Z}}(0))\right|=\int_{{\mathbb{R}}_{+}^{2}}\eta_{i}^{\gamma}(x,y,t)\gamma_{i}(dx)\gamma_{i+1}(dy). (4.6)

Recall that, for each ii\in{\mathbb{N}}, γi\gamma_{i} is an Exponential distribution and thus is mutually absolutely continuous with respect to the Lebesgue measure λ\lambda on +{\mathbb{R}}_{+}. Since ηiγ\eta_{i}^{\gamma} is nonnegative, we have from this mutual absolute continuity property that, for any ii\in{\mathbb{N}},

ηiγ(x,y,t)=0,λλ a.e. (x,y)+2, for every t0.\eta_{i}^{\gamma}(x,y,t)=0,\;\lambda\otimes\lambda\mbox{ a.e. }(x,y)\in{\mathbb{R}}_{+}^{2},\,\mbox{ for every }t\geq 0. (4.7)

Fix δ1,δ2>0\delta_{1},\delta_{2}>0, ii\in{\mathbb{N}}, t0t\geq 0. For y0y\geq 0, let ςδ2(y)δ21y>δ2\varsigma_{\delta_{2}}(y)\doteq\delta_{2}1_{y>\delta_{2}}. Letting 𝒁{\boldsymbol{Z}} be a +{\mathbb{R}}_{+}^{\infty}-valued random variable distributed as γ\gamma, denote by γ~𝒫(+)\tilde{\gamma}\in{\mathcal{P}}({\mathbb{R}}_{+}^{\infty}) the probability distribution of 𝒁~𝒁+δ1𝒆iςδ2(Zi+1)𝒆i+1\tilde{\boldsymbol{Z}}\doteq{\boldsymbol{Z}}+\delta_{1}{\boldsymbol{e}}_{i}-\varsigma_{\delta_{2}}(Z_{i+1}){\boldsymbol{e}}_{i+1} where 𝒆i{\boldsymbol{e}}_{i} is the unit vector in +{\mathbb{R}}_{+}^{\infty} with 11 at the ii-th coordinate. Note that by definition in (4.4), ηiγ=ηiγ~\eta_{i}^{\gamma}=\eta_{i}^{\tilde{\gamma}}, and in particular, from (4.7),

ηiγ~(x+δ1,yςδ2(y),t)=ηiγ(x+δ1,yςδ2(y),t)=0,λλ a.e. (x,y)+2.\eta_{i}^{\tilde{\gamma}}(x+\delta_{1},y-\varsigma_{\delta_{2}}(y),t)=\eta_{i}^{\gamma}(x+\delta_{1},y-\varsigma_{\delta_{2}}(y),t)=0,\;\lambda\otimes\lambda\mbox{ a.e. }(x,y)\in{\mathbb{R}}_{+}^{2}. (4.8)

Thus, in view of (4.5),

𝔼γ~|ψ(𝒁(t))ψ(𝒁(0))|=+2ηiγ~(x,y,t)γ~i(dx)γ~i+1(dy)=+2ηiγ~(x+δ1,yςδ2(y),t)γi(dx)γi+1(dy)=0,{\mathbb{E}}_{\tilde{\gamma}}\left|\psi({\boldsymbol{Z}}(t))-\psi({\boldsymbol{Z}}(0))\right|=\int_{{\mathbb{R}}_{+}^{2}}\eta_{i}^{\tilde{\gamma}}(x,y,t)\tilde{\gamma}_{i}(dx)\tilde{\gamma}_{i+1}(dy)=\int_{{\mathbb{R}}_{+}^{2}}\eta_{i}^{\tilde{\gamma}}(x+\delta_{1},y-\varsigma_{\delta_{2}}(y),t)\gamma_{i}(dx)\gamma_{i+1}(dy)=0, (4.9)

where we have used the fact that γiγi+1\gamma_{i}\otimes\gamma_{i+1} is mutually absolutely continuous with respect to λλ\lambda\otimes\lambda. For 𝒛𝒱{\boldsymbol{z}}\in{\mathcal{V}}, let βδ1,δ2(𝒛)𝒛+δ1𝒆iςδ2(zi+1)𝒆i+1\beta_{\delta_{1},\delta_{2}}({\boldsymbol{z}})\doteq{\boldsymbol{z}}+\delta_{1}{\boldsymbol{e}}_{i}-\varsigma_{\delta_{2}}(z_{i+1}){\boldsymbol{e}}_{i+1}. Then, combining (4.6) and (4.9), we have

𝔼𝒛|ψ(𝒁(t))ψ(𝒁(0))|=𝔼βδ1,δ2(𝒛)|ψ(𝒁(t))ψ(𝒁(0))|=0,γ a.e. 𝒛, for every t0.{\mathbb{E}}_{{\boldsymbol{z}}}\left|\psi({\boldsymbol{Z}}(t))-\psi({\boldsymbol{Z}}(0))\right|={\mathbb{E}}_{\beta_{\delta_{1},\delta_{2}}({\boldsymbol{z}})}\left|\psi({\boldsymbol{Z}}(t))-\psi({\boldsymbol{Z}}(0))\right|=0,\;\gamma\mbox{ a.e. }{\boldsymbol{z}},\mbox{ for every }t\geq 0. (4.10)

Since βδ1,δ2(𝒛)=𝒛+δ1𝒆iδ2𝒆i+1𝒛δ1,δ2,i\beta_{\delta_{1},\delta_{2}}({\boldsymbol{z}})={\boldsymbol{z}}+\delta_{1}{\boldsymbol{e}}_{i}-\delta_{2}{\boldsymbol{e}}_{i+1}\doteq{\boldsymbol{z}}^{\delta_{1},\delta_{2},i} when zi+1>δ2z_{i+1}>\delta_{2}, we get

ψ(𝒁(t))=ψ(𝒛),𝒛 a.e., ψ(𝒁(t))=ψ(𝒛δ1,δ2,i),𝒛δ1,δ2,i a.e., for γ a.e. 𝒛 with zi+1>δ2, for every t0.\psi({\boldsymbol{Z}}(t))=\psi({\boldsymbol{z}}),\,{\mathbb{P}}_{{\boldsymbol{z}}}\mbox{ a.e., }\psi({\boldsymbol{Z}}(t))=\psi({\boldsymbol{z}}^{\delta_{1},\delta_{2},i}),\,{\mathbb{P}}_{{\boldsymbol{z}}^{\delta_{1},\delta_{2},i}}\mbox{ a.e., for }\;\gamma\mbox{ a.e. }{\boldsymbol{z}}\mbox{ with }z_{i+1}>\delta_{2},\mbox{ for every }t\geq 0. (4.11)

We will need the following proposition. The proof is given in Section 4.4.

Proposition 4.2.

Fix ii\in{\mathbb{N}}, 𝐳𝒱{\boldsymbol{z}}\in{\mathcal{V}} with 𝐳>0{\boldsymbol{z}}>0, and δ1>0,δ2(0,zi+1)\delta_{1}>0,\delta_{2}\in(0,z_{i+1}). Then there exists a coupling 𝐳,δ1,δ2,i{\mathbb{P}}_{{\boldsymbol{z}},\delta_{1},\delta_{2},i} of the gap process of the 𝐠{\boldsymbol{g}}-Atlas model with initial distributions δ𝐳\delta_{{\boldsymbol{z}}} and δ𝐳δ1,δ2,i\delta_{{\boldsymbol{z}}^{\delta_{1},\delta_{2},i}} such that, for any s>0s>0,

𝒛,δ1,δ2,i(𝒁(1)(s)=𝒁(2)(s))>0.{\mathbb{P}}_{{\boldsymbol{z}},\delta_{1},\delta_{2},i}({\boldsymbol{Z}}^{(1)}(s)={\boldsymbol{Z}}^{(2)}(s))>0.

Now, for δ1,δ2>0\delta_{1},\delta_{2}>0 and s>0s>0,

{𝒛:zi+1>δ2}|ψ(𝒛δ1,δ2,i)ψ(𝒛)|𝒛,δ1,δ2,i(𝒁(1)(s)=𝒁(2)(s))γ(d𝒛)\displaystyle\int_{\{{\boldsymbol{z}}:z_{i+1}>\delta_{2}\}}|\psi({\boldsymbol{z}}^{\delta_{1},\delta_{2},i})-\psi({\boldsymbol{z}})|{\mathbb{P}}_{{\boldsymbol{z}},\delta_{1},\delta_{2},i}({\boldsymbol{Z}}^{(1)}(s)={\boldsymbol{Z}}^{(2)}(s))\gamma(d{\boldsymbol{z}})
={𝒛:zi+1>δ2}𝔼𝒛,δ1,δ2,i[|ψ(𝒛δ1,δ2,i)ψ(𝒛)|1{𝒁(1)(s)=𝒁(2)(s)}]γ(d𝒛)\displaystyle=\int_{\{{\boldsymbol{z}}:z_{i+1}>\delta_{2}\}}{\mathbb{E}}_{{\boldsymbol{z}},\delta_{1},\delta_{2},i}\left[|\psi({\boldsymbol{z}}^{\delta_{1},\delta_{2},i})-\psi({\boldsymbol{z}})|1_{\{{\boldsymbol{Z}}^{(1)}(s)={\boldsymbol{Z}}^{(2)}(s)\}}\right]\gamma(d{\boldsymbol{z}})
={𝒛:zi+1>δ2}𝔼𝒛,δ1,δ2,i[|ψ(𝒁(2)(s))ψ(𝒁(1)(s))|1{𝒁(1)(s)=𝒁(2)(s)}]γ(d𝒛)=0,\displaystyle=\int_{\{{\boldsymbol{z}}:z_{i+1}>\delta_{2}\}}{\mathbb{E}}_{{\boldsymbol{z}},\delta_{1},\delta_{2},i}\left[|\psi({\boldsymbol{Z}}^{(2)}(s))-\psi({\boldsymbol{Z}}^{(1)}(s))|1_{\{{\boldsymbol{Z}}^{(1)}(s)={\boldsymbol{Z}}^{(2)}(s)\}}\right]\gamma(d{\boldsymbol{z}})=0,

where 𝒁(i){\boldsymbol{Z}}^{(i)} are as given by Proposition 4.2 and the second equality follows from (4.11). Hence, from Proposition 4.2, for every δ1,δ2>0\delta_{1},\delta_{2}>0 and ii\in{\mathbb{N}},

ψ(𝒛+δ1𝒆iδ2𝒆i+1)=ψ(𝒛) for γ a.e. 𝒛 with zi+1>δ2.\psi({\boldsymbol{z}}+\delta_{1}{\boldsymbol{e}}_{i}-\delta_{2}{\boldsymbol{e}}_{i+1})=\psi({\boldsymbol{z}})\mbox{ for }\gamma\mbox{ a.e. }{\boldsymbol{z}}\mbox{ with }z_{i+1}>\delta_{2}.

Thus we have shown that

γ(δ1,δ2(0,){𝒛:zi+1>δ2, and ψ(𝒛+δ1𝒆iδ2𝒆i+1)ψ(𝒛)})=0.\gamma(\cup_{\delta_{1},\delta_{2}\in{\mathbb{Q}}\cap(0,\infty)}\{{\boldsymbol{z}}:z_{i+1}>\delta_{2},\mbox{ and }\psi({\boldsymbol{z}}+\delta_{1}{\boldsymbol{e}}_{i}-\delta_{2}{\boldsymbol{e}}_{i+1})\neq\psi({\boldsymbol{z}})\})=0. (4.12)

This implies that

γ(𝒛:ψ(𝒛+δ1𝒆iδ2𝒆i+1)=ψ(𝒛) for all δ1(0,),δ2(0,zi+1))=1.\gamma({\boldsymbol{z}}:\psi({\boldsymbol{z}}+\delta_{1}{\boldsymbol{e}}_{i}-\delta_{2}{\boldsymbol{e}}_{i+1})=\psi({\boldsymbol{z}})\mbox{ for all }\delta_{1}\in(0,\infty)\cap{\mathbb{Q}},\ \delta_{2}\in(0,z_{i+1})\cap{\mathbb{Q}})=1. (4.13)

To see this, let BB denote the event on the left side of (4.13). Then if 𝒛Bc{\boldsymbol{z}}\in B^{c} and zi+1>0z_{i+1}>0, then for some δ1,δ2(0,)\delta_{1},\delta_{2}\in(0,\infty)\cap{\mathbb{Q}}, zi+1>δ2z_{i+1}>\delta_{2} and ψ(𝒛+δ1𝒆iδ2𝒆i+1)ψ(𝒛)\psi({\boldsymbol{z}}+\delta_{1}{\boldsymbol{e}}_{i}-\delta_{2}{\boldsymbol{e}}_{i+1})\neq\psi({\boldsymbol{z}}), which shows that 𝒛{\boldsymbol{z}} is in the event on the left side of (4.12) which in view of (4.12) says that γ(Bc)=0\gamma(B^{c})=0, proving the statement in (4.13). The following proposition enables us to extend (4.13) to all (δ1,δ2)(0,)×(0,zi+1)(\delta_{1},\delta_{2})\in(0,\infty)\times(0,z_{i+1}). The proof is given in Section 4.4.

Proposition 4.3.

For each 𝐳𝒱{\boldsymbol{z}}\in{\mathcal{V}} with 𝐳>0{\boldsymbol{z}}>0 and ii\in{\mathbb{N}}, the map (δ1,δ2)ψ(𝐳+δ1𝐞iδ2𝐞i+1)(\delta_{1},\delta_{2})\mapsto\psi({\boldsymbol{z}}+\delta_{1}{\boldsymbol{e}}_{i}-\delta_{2}{\boldsymbol{e}}_{i+1}) is right continuous on [0,)×[0,zi+1)[0,\infty)\times[0,z_{i+1}). That is, if (δ1,δ2)[0,)×[0,zi+1)(\delta_{1},\delta_{2})\in[0,\infty)\times[0,z_{i+1}) and δ1,nδ1\delta_{1,n}\downarrow\delta_{1} and δ2,n[0,zi+1)\delta_{2,n}\in[0,z_{i+1}) with δ2,nδ2\delta_{2,n}\downarrow\delta_{2} as nn\rightarrow\infty, then ψ(𝐳+δ1,n𝐞iδ2,n𝐞i+1)ψ(𝐳+δ1𝐞iδ2𝐞i+1)\psi({\boldsymbol{z}}+\delta_{1,n}{\boldsymbol{e}}_{i}-\delta_{2,n}{\boldsymbol{e}}_{i+1})\rightarrow\psi({\boldsymbol{z}}+\delta_{1}{\boldsymbol{e}}_{i}-\delta_{2}{\boldsymbol{e}}_{i+1}) as nn\rightarrow\infty.

We remark that our proof shows that in the above proposition ψ\psi can be taken to be Tt0ψ0T_{t_{0}}\psi_{0} for any real bounded measurable function ψ0\psi_{0} on +{\mathbb{R}}_{+}^{\infty} and t0>0t_{0}>0, namely it need not be {Tt}\{T_{t}\}-invariant. Thus the property established in the above proposition can be viewed as a certain type of ‘directional strong Feller property’.

4.3 Completing the proof of Theorem 3.3

As an immediate consequence of the above proposition and (4.13) we have that for γ\gamma a.e. 𝒛{\boldsymbol{z}},

ψ(𝒛+δ1𝒆iδ2𝒆i+1)=ψ(𝒛) for all δ1(0,),δ2(0,zi+1),i.\psi({\boldsymbol{z}}+\delta_{1}{\boldsymbol{e}}_{i}-\delta_{2}{\boldsymbol{e}}_{i+1})=\psi({\boldsymbol{z}})\mbox{ for all }\delta_{1}\in(0,\infty),\delta_{2}\in(0,z_{i+1}),\;i\in{\mathbb{N}}. (4.14)

A similar argument shows that, for γ\gamma a.e. 𝒛{\boldsymbol{z}},

ψ(𝒛δ1𝒆i+δ2𝒆i+1)=ψ(𝒛) for all δ1(0,zi),δ2(0,),i.\psi({\boldsymbol{z}}-\delta_{1}{\boldsymbol{e}}_{i}+\delta_{2}{\boldsymbol{e}}_{i+1})=\psi({\boldsymbol{z}})\mbox{ for all }\delta_{1}\in(0,z_{i}),\delta_{2}\in(0,\infty),\;i\in{\mathbb{N}}. (4.15)

We now proceed to the proof of the first statement in Theorem 3.3. Recall that γ=πa𝒈n=1Exp(n(2g¯n+a))\gamma=\pi_{a}^{{\boldsymbol{g}}}\doteq\otimes_{n=1}^{\infty}\mbox{Exp}(n(2\bar{g}_{n}+a)), where aa satisfies the condition in Theorem 3.3. For notational simplicity, let cn2[n(2g¯n+a)]1c_{n}\doteq 2[n(2\bar{g}_{n}+a)]^{-1}, nn\in{\mathbb{N}}. Let ψ~:𝒱\tilde{\psi}:{\mathcal{V}}\to{\mathbb{R}} be defined as

ψ~(z1,z2,z3,)ψ(c1z1,c2z2,c3z3,),𝒛=(z1,z2,)𝒱\tilde{\psi}(z_{1},z_{2},z_{3},\ldots)\doteq\psi(c_{1}z_{1},c_{2}z_{2},c_{3}z_{3},\ldots),\;{\boldsymbol{z}}=(z_{1},z_{2},\ldots)\in{\mathcal{V}}

if (c1z1,c2z2,c3z3,)𝒱(c_{1}z_{1},c_{2}z_{2},c_{3}z_{3},\ldots)\in{\mathcal{V}}. For all other 𝒛𝒱{\boldsymbol{z}}\in{\mathcal{V}}, set ψ~(𝒛)=0\tilde{\psi}({\boldsymbol{z}})=0. We denote π0𝒈1n=1Exp(2)\pi_{0}^{{\boldsymbol{g}}^{1}}\doteq\otimes_{n=1}^{\infty}\mbox{Exp}(2) as π0\pi_{0} for simplicity. Observe that, for any i0i\in\mathbb{N}_{0},

π0(𝒛:ψ~(z1,z2,,zi1,zi,zi+1,zi+2,)=ψ~(z1,z2,,zi1,zi+1,zi,zi+2,))\displaystyle\pi_{0}({\boldsymbol{z}}:\tilde{\psi}(z_{1},z_{2},\ldots,z_{i-1},z_{i},z_{i+1},z_{i+2},\ldots)=\tilde{\psi}(z_{1},z_{2},\ldots,z_{i-1},z_{i+1},z_{i},z_{i+2},\ldots))
=πa𝒈(𝒛:ψ(z1,z2,,zi1,zi,zi+1,zi+2,)=ψ(z1,z2,,zi1,cici+1zi+1,ci+1cizi,zi+2,)).\displaystyle=\pi_{a}^{{\boldsymbol{g}}}\left({\boldsymbol{z}}:\psi(z_{1},z_{2},\ldots,z_{i-1},z_{i},z_{i+1},z_{i+2},\ldots)=\psi\left(z_{1},z_{2},\ldots,z_{i-1},\frac{c_{i}}{c_{i+1}}z_{i+1},\frac{c_{i+1}}{c_{i}}z_{i},z_{i+2},\ldots\right)\right). (4.16)

Consider the set C(𝕏)C\in{\mathcal{B}}({\mathbb{X}}) with πa𝒈(C)=1\pi_{a}^{{\boldsymbol{g}}}(C)=1 on which the two statements in (4.14) and (4.15) hold. Then for any 𝒛C{\boldsymbol{z}}\in C such that zi>0z_{i}>0 for all ii\in{\mathbb{N}}, we have,

ψ(z1,z2,,zi1,zi,zi+1,zi+2,)=ψ(z1,z2,,zi1,cici+1zi+1,ci+1cizi,zi+2,).\psi(z_{1},z_{2},\ldots,z_{i-1},z_{i},z_{i+1},z_{i+2},\ldots)=\psi\left(z_{1},z_{2},\ldots,z_{i-1},\frac{c_{i}}{c_{i+1}}z_{i+1},\frac{c_{i+1}}{c_{i}}z_{i},z_{i+2},\ldots\right). (4.17)

Indeed, if cici+1zi+1zi>0\frac{c_{i}}{c_{i+1}}z_{i+1}-z_{i}>0, then the statement follows from (4.14) on taking δ1=cici+1zi+1zi\delta_{1}=\frac{c_{i}}{c_{i+1}}z_{i+1}-z_{i} and δ2=ci+1ciδ1\delta_{2}=\frac{c_{i+1}}{c_{i}}\delta_{1}, and if zicici+1zi+1>0z_{i}-\frac{c_{i}}{c_{i+1}}z_{i+1}>0, the statement follows from (4.15) on taking δ1=zicici+1zi+1\delta_{1}=z_{i}-\frac{c_{i}}{c_{i+1}}z_{i+1} and δ2=ci+1ciδ1\delta_{2}=\frac{c_{i+1}}{c_{i}}\delta_{1}. Since πa𝒈(C)=1\pi_{a}^{{\boldsymbol{g}}}(C)=1, we have that the probability on the right side of (4.3) is 11 and so,

π0(𝒛:ψ~(z1,z2,,zi1,zi,zi+1,zi+2,)=ψ~(z1,z2,,zi1,zi+1,zi,zi+2,))=1.\pi_{0}({\boldsymbol{z}}:\tilde{\psi}(z_{1},z_{2},\ldots,z_{i-1},z_{i},z_{i+1},z_{i+2},\ldots)=\tilde{\psi}(z_{1},z_{2},\ldots,z_{i-1},z_{i+1},z_{i},z_{i+2},\ldots))=1. (4.18)

As any finite permutation can be obtained as a composition of finitely many adjacent transpositions, it now follows that, in fact for any finite permutation ρ:\rho:{\mathbb{N}}\to{\mathbb{N}},

π0(𝒛:ψ~(z1,z2,)=ψ~(zρ(1),zρ(2),))=1.\pi_{0}({\boldsymbol{z}}:\tilde{\psi}(z_{1},z_{2},\ldots)=\tilde{\psi}(z_{\rho(1)},z_{\rho(2)},\ldots))=1. (4.19)

Now using the Hewitt-Savage zero-one law (cf. [23, Theorem 2.15]), ψ~\tilde{\psi} is π0\pi_{0} a.e. constant, namely, there is a α\alpha\in{\mathbb{R}} such that π0(𝒛:ψ~(𝒛)=α)=1\pi_{0}({\boldsymbol{z}}:\tilde{\psi}({\boldsymbol{z}})=\alpha)=1. This shows that

πa𝒈(𝒛:ψ(𝒛)=α)=π0(𝒛:ψ~(𝒛)=α)=1.\pi_{a}^{{\boldsymbol{g}}}({\boldsymbol{z}}:\psi({\boldsymbol{z}})=\alpha)=\pi_{0}({\boldsymbol{z}}:\tilde{\psi}({\boldsymbol{z}})=\alpha)=1.

Hence, ψ\psi is constant πa𝒈\pi_{a}^{{\boldsymbol{g}}} a.s. Appealing to Lemma 3.2, this completes the proof of the first statement in Theorem 3.3. The second statement follows similarly. ∎

4.4 Proofs of Propositions 4.2 and 4.3

Recall that we fix 𝒈𝒟{\boldsymbol{g}}\in{\mathcal{D}}. Also, throughout this section we fix 𝒛𝒱{\boldsymbol{z}}\in{\mathcal{V}} such that 𝒛>0{\boldsymbol{z}}>0 and ii\in{\mathbb{N}}. Define for δ2(0,zi+1)\delta_{2}\in(0,z_{i+1}) and δ1>0\delta_{1}>0

𝒚(0,z1,z1+z2,),𝒚δ1,δ2𝒚+j=0i1(δ2δ1)𝒆j+δ2𝒆i.{\boldsymbol{y}}\doteq(0,z_{1},z_{1}+z_{2},\ldots)^{\prime},\;\;{\boldsymbol{y}}^{\delta_{1},\delta_{2}}\doteq{\boldsymbol{y}}+\sum_{j=0}^{i-1}(\delta_{2}-\delta_{1}){\boldsymbol{e}}_{j}+\delta_{2}{\boldsymbol{e}}_{i}.

Observe that, with the above choice of starting points of the particles, the corresponding gaps are 𝒛(𝒚)=𝒛{\boldsymbol{z}}({\boldsymbol{y}})={\boldsymbol{z}} and 𝒛(𝒚δ1,δ2)=𝒛δ1,δ2{\boldsymbol{z}}({\boldsymbol{y}}^{\delta_{1},\delta_{2}})={\boldsymbol{z}}^{\delta_{1},\delta_{2}}.

Let B0,B1,B_{0},B_{1},\ldots be a sequence of mutually independent standard Brownian motions on some probability space (Ω,,)(\Omega,{\mathcal{F}},{\mathbb{P}}). Consider the (i+1)(i+1)-dimensional diffusion process

𝚿(t)=𝚿(0)+D𝑩(i)(t)+𝒃t,t0,\mathbf{\Psi}(t)=\mathbf{\Psi}(0)+D{\boldsymbol{B}}^{(i)}(t)+{\boldsymbol{b}}t,\ t\geq 0,

where 𝚿(0)=(z1,,zi,(zi+1+δ2)/2)\mathbf{\Psi}(0)=(z_{1},\dots,z_{i},(z_{i+1}+\delta_{2})/2)^{\prime}, 𝑩(i)()=(B0(),B1(),,Bi()){\boldsymbol{B}}^{(i)}(\cdot)=(B_{0}(\cdot),B_{1}(\cdot),\dots,B_{i}(\cdot))^{\prime}, 𝒃=(g1g0,,gigi1,gi){\boldsymbol{b}}=(g_{1}-g_{0},\dots,g_{i}-g_{i-1},-g_{i})^{\prime}, and DD is an (i+1)×(i+1)(i+1)\times(i+1) matrix with Djj=1D_{jj}=-1 for all 1ji+11\leq j\leq i+1, Dj(j+1)=1D_{j(j+1)}=1 for all 1ji1\leq j\leq i and Djl=0D_{jl}=0 otherwise. The process {𝚿(t):t0}\{\mathbf{\Psi}(t):\,t\geq 0\} will be used to analyze the evolution of the first i+1i+1 gaps before any of them hit zero or the lowest ii particles interact with the higher ranked particles (in an appropriate sense), as stated more precisely later.

It can be checked that DD is non-singular and (D1)jl=1(D^{-1})_{jl}=-1 for all 1jli+11\leq j\leq l\leq i+1 and (D1)jl=0(D^{-1})_{jl}=0 otherwise. Let 𝚿~(0)(z1,,zi1,zi+δ1,(zi+1δ2)/2)\tilde{\mathbf{\Psi}}(0)\doteq(z_{1},\dots,z_{i-1},z_{i}+\delta_{1},(z_{i+1}-\delta_{2})/2)^{\prime} and define 𝒗D1(𝚿~(0)𝚿(0))=(δ2δ1,,δ2δ1,δ2){\boldsymbol{v}}\doteq D^{-1}(\tilde{\mathbf{\Psi}}(0)-\mathbf{\Psi}(0))=(\delta_{2}-\delta_{1},\dots,\delta_{2}-\delta_{1},\delta_{2})^{\prime}. Also define the stopping time

σinf{t0:𝒗𝑩(i)(t)=𝒗2/2},\sigma\doteq\inf\{t\geq 0:\,{\boldsymbol{v}}^{\prime}{\boldsymbol{B}}^{(i)}(t)=\|{\boldsymbol{v}}\|^{2}/2\},

where \|\cdot\| denotes the standard Euclidean norm. Define the mirror coupled (i+1)(i+1)-dimensional Brownian motion 𝑩~(i)\tilde{\boldsymbol{B}}^{(i)} by

𝑩~(i)(t)=(I2𝒗𝒗𝒗2)𝑩(i)(t)1[tσ]+(𝑩(i)(t)𝒗)1[t>σ],t0.\tilde{\boldsymbol{B}}^{(i)}(t)=\left(I-\frac{2{\boldsymbol{v}}{\boldsymbol{v}}^{\prime}}{\|{\boldsymbol{v}}\|^{2}}\right){\boldsymbol{B}}^{(i)}(t)1_{[t\leq\sigma]}+({\boldsymbol{B}}^{(i)}(t)-{\boldsymbol{v}})1_{[t>\sigma]},\ t\geq 0. (4.20)

Since (I2𝒗𝒗𝒗2)\left(I-\frac{2{\boldsymbol{v}}{\boldsymbol{v}}^{\prime}}{\|{\boldsymbol{v}}\|^{2}}\right) is a unitary matrix, it follows from the strong Markov property that 𝑩~(i)\tilde{\boldsymbol{B}}^{(i)} is indeed a Brownian motion. 𝑩~(i)\tilde{\boldsymbol{B}}^{(i)} can be thought of as the reflection of the Brownian motion 𝑩(i){\boldsymbol{B}}^{(i)} in a hyperplane perpendicular to the vector 𝒗{\boldsymbol{v}} till the first time σ\sigma when 𝑩(i){\boldsymbol{B}}^{(i)} hits this hyperplane (which is also the first meeting time of 𝑩(i){\boldsymbol{B}}^{(i)} and 𝑩~(i)\tilde{\boldsymbol{B}}^{(i)}), and then coalescing with 𝑩(i){\boldsymbol{B}}^{(i)}. Using 𝑩~(i)\tilde{\boldsymbol{B}}^{(i)}, define a coupled version of the process 𝚿\mathbf{\Psi} by

𝚿~(t)=𝚿~(0)+D𝑩~(i)(t)+𝒃t,t0.\tilde{\mathbf{\Psi}}(t)=\tilde{\mathbf{\Psi}}(0)+D\tilde{\boldsymbol{B}}^{(i)}(t)+{\boldsymbol{b}}t,\ t\geq 0.

Extend 𝑩~(i)\tilde{\boldsymbol{B}}^{(i)} to an infinite collection of standard Brownian motions 𝑩~=(B~0,B~1,)(𝑩~(i),Bi+1,)\tilde{\boldsymbol{B}}=(\tilde{B}_{0},\tilde{B}_{1},\dots)^{\prime}\doteq(\tilde{\boldsymbol{B}}^{(i)},B_{i+1},\dots)^{\prime}.

Our arguments will involve a coupling of two copies of the infinite ordered 𝒈{\boldsymbol{g}}-Atlas model started from 𝒚{\boldsymbol{y}} and 𝒚δ1,δ2{\boldsymbol{y}}^{\delta_{1},\delta_{2}} and respectively driven by the Brownian motions {Bj}j0\{B_{j}\}_{j\in{\mathbb{N}}_{0}} and {B~j}j0\{\tilde{B}_{j}\}_{j\in{\mathbb{N}}_{0}}. For the finite particle 𝒈{\boldsymbol{g}}-Atlas model, this coupling can be directly constructed using the existence of a strong solution to the finite version of the SDE (2.3). However, for the infinite 𝒈{\boldsymbol{g}}-Atlas model, this is a delicate issue. We will use the recipe of approximative versions of [32], which we now introduce.

Definition 4.4.

Suppose 𝐱𝒰{\boldsymbol{x}}\in\mathcal{U} and consider a collection of independent standard Brownian motions {Bj}j0\{B^{*}_{j}\}_{j\in{\mathbb{N}}_{0}}. Consider for fixed mm\in{\mathbb{N}}, the system of SDE in (2.3) for i=0,1,,mi=0,1,\ldots,m, with starting configuration Xi(0)=xiX^{*}_{i}(0)=x_{i}, 0im0\leq i\leq m, local times of collision between the (i1)(i-1)-th and ii-th particles denoted by Li,1imL^{*}_{i},1\leq i\leq m, and L0()Lm+1()0L^{*}_{0}(\cdot)\equiv L^{*}_{m+1}(\cdot)\equiv 0. Denote by 𝐗,(m)()=(X0,(m)(),,Xm,(m)()){\boldsymbol{X}}^{*,(m)}(\cdot)=(X^{*,(m)}_{0}(\cdot),\dots,X^{*,(m)}_{m}(\cdot))^{\prime} and 𝐋,(m)()=(L0,(m)(),,Lm,(m)()){\boldsymbol{L}}^{*,(m)}(\cdot)=(L^{*,(m)}_{0}(\cdot),\dots,L^{*,(m)}_{m}(\cdot))^{\prime} the unique strong solution to this finite-dimensional system of reflected SDE with driving Brownian motions (B0(),,Bm())(B^{*}_{0}(\cdot),\dots,B^{*}_{m}(\cdot))^{\prime}.

Then (see [32, Definition 7 and Theorem 3.7], [32, Lemma 6.4] and the discussion following it), there exist continuous {\mathbb{R}}^{\infty}-valued processes 𝐗()(Xi():i0){\boldsymbol{X}}^{*}(\cdot)\doteq(X^{*}_{i}(\cdot):i\in\mathbb{N}_{0})^{\prime}, 𝐋()(Li():i0){\boldsymbol{L}}^{*}(\cdot)\doteq(L^{*}_{i}(\cdot):i\in\mathbb{N}_{0})^{\prime}, adapted to tσ{Bi(s):st,i0}{\mathcal{F}}_{t}\doteq\sigma\{B^{*}_{i}(s):s\leq t,i\in{\mathbb{N}}_{0}\}, such that, a.s. 𝐗{\boldsymbol{X}}^{*} satisfies (2.3) with associated local times given by 𝐋{\boldsymbol{L}}^{*} and for any T(0,)T\in(0,\infty),

limmsupt[0,T][|Xi,(m)(t)Xi(t)|+|Li,(m)(t)Li(t)|]=0 a.s., for all i0.\lim_{m\rightarrow\infty}\sup_{t\in[0,T]}\left[\left|X^{*,(m)}_{i}(t)-X^{*}_{i}(t)\right|+\left|L^{*,(m)}_{i}(t)-L^{*}_{i}(t)\right|\right]=0\ \text{ a.s., for all }i\in\mathbb{N}_{0}.

We will call 𝐗(){\boldsymbol{X}}^{*}(\cdot) the ‘infinite ordered 𝐠{\boldsymbol{g}}-Atlas model’ with driving Brownian motions {Bj}j0\{B^{*}_{j}\}_{j\in{\mathbb{N}}_{0}} started from 𝐱=(x0,x1,x2){\boldsymbol{x}}=(x_{0},x_{1},x_{2}\dots)^{\prime}.

We denote the infinite ordered 𝒈{\boldsymbol{g}}-Atlas model defined on (Ω,,)(\Omega,{\mathcal{F}},{\mathbb{P}}) with initial condition 𝒚{\boldsymbol{y}} and driving Brownian motions {Bj}j0\{B_{j}\}_{j\in{\mathbb{N}}_{0}} as 𝑿=(X0,X1,){\boldsymbol{X}}=(X_{0},X_{1},\ldots)^{\prime}. Similarly, denote the infinite ordered 𝒈{\boldsymbol{g}}-Atlas model defined on (Ω,,)(\Omega,{\mathcal{F}},{\mathbb{P}}) with initial condition 𝒚δ1,δ2{\boldsymbol{y}}^{\delta_{1},\delta_{2}} and driving Brownian motions {B~j}j0\{\tilde{B}_{j}\}_{j\in{\mathbb{N}}_{0}} as 𝑿~=(X~0,X~1,)\tilde{\boldsymbol{X}}=(\tilde{X}_{0},\tilde{X}_{1},\ldots)^{\prime}. Denote the gap processes associated with 𝑿{\boldsymbol{X}} and 𝑿~\tilde{\boldsymbol{X}} as 𝒁{\boldsymbol{Z}} and 𝒁~\tilde{\boldsymbol{Z}} respectively, namely

ZiXiXi1,Z~iX~iX~i1,;i.Z_{i}\doteq X_{i}-X_{i-1},\;\;\tilde{Z}_{i}\doteq\tilde{X}_{i}-\tilde{X}_{i-1},\;;i\in{\mathbb{N}}.

It then follows that

𝒛,δ1,δ2,i(𝒁,𝒁~)1{\mathbb{P}}_{{\boldsymbol{z}},\delta_{1},\delta_{2},i}\doteq{\mathbb{P}}\circ({\boldsymbol{Z}},\tilde{\boldsymbol{Z}})^{-1} (4.21)

is a coupling of the gap process of the 𝒈{\boldsymbol{g}}-Atlas model with initial distributions δ𝒛\delta_{{\boldsymbol{z}}} and δ𝒛δ1,δ2,i\delta_{{\boldsymbol{z}}^{\delta_{1},\delta_{2},i}}. Moreover, the process {𝚿(t):t0}\{\mathbf{\Psi}(t):\,t\geq 0\} gives the evolution of {(Z1(t),,Zi(t),yi+yi+1+δ22Xi(t)):t0}\{\left(Z_{1}(t),\dots,Z_{i}(t),\frac{y_{i}+y_{i+1}+\delta_{2}}{2}-X_{i}(t)\right)^{\prime}:\,t\geq 0\} before any of the co-ordinates of 𝚿()\mathbf{\Psi}(\cdot) hit zero or Xi+1X_{i+1} hits the level yi+yi+1+δ22\frac{y_{i}+y_{i+1}+\delta_{2}}{2} (note that zi+1>δ2z_{i+1}>\delta_{2} guarantees that Xi+1(0)>yi+yi+1+δ22X_{i+1}(0)>\frac{y_{i}+y_{i+1}+\delta_{2}}{2}). This can be seen from (2.3). Similarly, {𝚿~(t):t0}\{\tilde{\mathbf{\Psi}}(t):\,t\geq 0\} gives the evolution of {(Z~1(t),,Z~i(t),yi+yi+1+δ22X~i(t)):t0}\{\left(\tilde{Z}_{1}(t),\dots,\tilde{Z}_{i}(t),\frac{y_{i}+y_{i+1}+\delta_{2}}{2}-\tilde{X}_{i}(t)\right)^{\prime}:\,t\geq 0\} before any of the co-ordinates of 𝚿~()\tilde{\mathbf{\Psi}}(\cdot) hit zero or X~i+1\tilde{X}_{i+1} hits the level yi+yi+1+δ22\frac{y_{i}+y_{i+1}+\delta_{2}}{2} from above.

We will now construct tractable events of positive probability under which the ‘mirror coupled’ processes 𝚿\mathbf{\Psi} and 𝚿~\tilde{\mathbf{\Psi}} will successfully couple before any of their co-ordinates hit zero or Xi+1X_{i+1} (equivalently, X~i+1\tilde{X}_{i+1}) hits the level yi+yi+1+δ22\frac{y_{i}+y_{i+1}+\delta_{2}}{2}. Towards this end, observe that 𝒟{D1𝒙:𝒙+i+1}\mathcal{D}\doteq\{D^{-1}{\boldsymbol{x}}:{\boldsymbol{x}}\in\mathbb{R}^{i+1}_{+}\} is a polyhedral convex domain contained in the nonpositive orthant of i+1\mathbb{R}^{i+1}. Let {\mathcal{L}} denote the line segment joining D1𝚿(0)D^{-1}\mathbf{\Psi}(0) and D1𝚿~(0)D^{-1}\tilde{\mathbf{\Psi}}(0). By convexity of 𝒟\mathcal{D}, and since 𝚿(0),𝚿~(0)>0\mathbf{\Psi}(0),\tilde{\mathbf{\Psi}}(0)>0,

rinf𝒖dist(𝒖,𝒟)>0,r\doteq\inf_{{\boldsymbol{u}}\in{\mathcal{L}}}\operatorname{dist}({\boldsymbol{u}},\partial\mathcal{D})>0, (4.22)

where 𝒟\partial\mathcal{D} denotes the boundary of 𝒟\mathcal{D} and dist\operatorname{dist} denotes Euclidean distance of a point from this set. Also define the processes

M(t)𝒗𝒗𝑩(i)(t),𝑴(t)(I𝒗𝒗𝒗2)𝑩(i)(t)=(I𝒗𝒗𝒗2)𝑩~(i)(t),t0,\displaystyle M(t)\doteq\frac{{\boldsymbol{v}}^{\prime}}{\|{\boldsymbol{v}}\|}{\boldsymbol{B}}^{(i)}(t),\ {\boldsymbol{M}}^{\perp}(t)\doteq\left(I-\frac{{\boldsymbol{v}}{\boldsymbol{v}}^{\prime}}{\|{\boldsymbol{v}}\|^{2}}\right){\boldsymbol{B}}^{(i)}(t)=\left(I-\frac{{\boldsymbol{v}}{\boldsymbol{v}}^{\prime}}{\|{\boldsymbol{v}}\|^{2}}\right)\tilde{\boldsymbol{B}}^{(i)}(t),\ t\geq 0,

where the last equality can be verified from (4.20). Observe that MM is a standard Brownian motion and moreover, MM and 𝑴{\boldsymbol{M}}^{\perp} are independent Gaussian processes. From a geometric point of view, 2M2M denotes the component of 𝑩~(i)𝑩(i)\tilde{\boldsymbol{B}}^{(i)}-{\boldsymbol{B}}^{(i)} along the vector 𝒗{\boldsymbol{v}} and 𝑴{\boldsymbol{M}}^{\perp} denotes the ‘synchronously’ coupled projections of 𝑩~(i)\tilde{\boldsymbol{B}}^{(i)} and 𝑩(i){\boldsymbol{B}}^{(i)} along the hyperplane perpendicular to 𝒗{\boldsymbol{v}}. Define the stopping time

τinf{t0:M(t)=r/4 or 𝑴(t)=r/4}.\tau^{*}\doteq\inf\{t\geq 0:M(t)=-r/4\text{ or }\|{\boldsymbol{M}}^{\perp}(t)\|=r/4\}.

Consider any tστr8D1𝒃+1t\leq\sigma\wedge\tau^{*}\wedge\frac{r}{8\|D^{-1}{\boldsymbol{b}}\|+1}. Note that, if M(t)0M(t)\geq 0, then, using tσt\leq\sigma,

𝒖D1𝚿(0)+M(t)𝒗𝒗=(1𝒗𝑩(i)(t)𝒗2)D1𝚿(0)+𝒗𝑩(i)(t)𝒗2D1𝚿~(0).{\boldsymbol{u}}\doteq D^{-1}\mathbf{\Psi}(0)+M(t)\frac{{\boldsymbol{v}}}{\|{\boldsymbol{v}}\|}=\left(1-\frac{{\boldsymbol{v}}^{\prime}{\boldsymbol{B}}^{(i)}(t)}{\|{\boldsymbol{v}}\|^{2}}\right)D^{-1}\mathbf{\Psi}(0)+\frac{{\boldsymbol{v}}^{\prime}{\boldsymbol{B}}^{(i)}(t)}{\|{\boldsymbol{v}}\|^{2}}D^{-1}\tilde{\mathbf{\Psi}}(0)\in{\mathcal{L}}.

Furthermore,

D1𝚿(t)𝒖=𝑴(t)+D1𝒃t𝑴(t)+D1𝒃t3r8.\|D^{-1}\mathbf{\Psi}(t)-{\boldsymbol{u}}\|=\|{\boldsymbol{M}}^{\perp}(t)+D^{-1}{\boldsymbol{b}}t\|\leq\|{\boldsymbol{M}}^{\perp}(t)\|+\|D^{-1}{\boldsymbol{b}}\|t\leq\frac{3r}{8}.

Hence, by definition of rr, D1𝚿(t)𝒟D^{-1}\mathbf{\Psi}(t)\notin\partial\mathcal{D}. If, on the other hand, M(t)<0M(t)<0, then recalling tτt\leq\tau^{*},

D1𝚿(t)D1𝚿(0)\displaystyle\|D^{-1}\mathbf{\Psi}(t)-D^{-1}\mathbf{\Psi}(0)\| =𝑩(i)(t)+D1𝒃t=M(t)𝒗𝒗+𝑴(t)+D1𝒃t\displaystyle=\|{\boldsymbol{B}}^{(i)}(t)+D^{-1}{\boldsymbol{b}}t\|=\|M(t)\frac{{\boldsymbol{v}}}{\|{\boldsymbol{v}}\|}+{\boldsymbol{M}}^{\perp}(t)+D^{-1}{\boldsymbol{b}}t\|
|M(t)|+𝑴(t)+D1𝒃t5r8,\displaystyle\leq|M(t)|+\|{\boldsymbol{M}}^{\perp}(t)\|+\|D^{-1}{\boldsymbol{b}}\|t\leq\frac{5r}{8},

again implying D1𝚿(t)𝒟D^{-1}\mathbf{\Psi}(t)\notin\partial\mathcal{D}. Hence, we conclude that, on the event {στr8D1𝒃+1}\{\sigma\leq\tau^{*}\wedge\frac{r}{8\|D^{-1}{\boldsymbol{b}}\|+1}\}, D1𝚿(t)𝒟D^{-1}\mathbf{\Psi}(t)\notin\partial\mathcal{D} for all tσt\leq\sigma. A similar argument gives D1𝚿~(t)𝒟D^{-1}\tilde{\mathbf{\Psi}}(t)\notin\partial\mathcal{D} for all tσt\leq\sigma on the same event.

Since 𝒈𝒟{\boldsymbol{g}}\in{\mathcal{D}}, the sequence {gi}\{g_{i}\} is bounded. Let gl[1,)g^{l}\in[1,\infty) be such that

gjgl for all j0.g_{j}\geq-g^{l}\mbox{ for all }j\in{\mathbb{N}}_{0}.

Fix any

δ¯(0,zi+1/(2+4gl)).\underline{\delta}\in(0,z_{i+1}/(2+4g^{l})). (4.23)

For s(0,δ¯]s\in(0,\underline{\delta}] and δ1,δ2(0,δ¯)\delta_{1},\delta_{2}\in(0,\underline{\delta}), define the following events in {\mathcal{F}}:

1(s)\displaystyle{\mathcal{E}}^{1}(s) {infji+1inf0ts(yj+Bj(r))>yi+yi+1+δ22+glδ¯},\displaystyle\doteq\left\{\inf_{j\geq i+1}\inf_{0\leq t\leq s}(y_{j}+B_{j}(r))>\frac{y_{i}+y_{i+1}+\delta_{2}}{2}+g^{l}\underline{\delta}\right\},
2(s)\displaystyle{\mathcal{E}}^{2}(s) {inftsr8D1𝒃+1M(t)>r/4,suptsr8D1𝒃+1M(t)𝒗/2,suptsr8D1𝒃+1𝑴(t)<r/4}.\displaystyle\doteq\left\{\inf_{t\leq s\wedge\frac{r}{8\|D^{-1}{\boldsymbol{b}}\|+1}}M(t)>-r/4,\ \sup_{t\leq s\wedge\frac{r}{8\|D^{-1}{\boldsymbol{b}}\|+1}}M(t)\geq\|{\boldsymbol{v}}\|/2,\ \sup_{t\leq s\wedge\frac{r}{8\|D^{-1}{\boldsymbol{b}}\|+1}}\|{\boldsymbol{M}}^{\perp}(t)\|<r/4\right\}.

Let (s)1(s)2(s){\mathcal{E}}(s)\doteq{\mathcal{E}}^{1}(s)\cap{\mathcal{E}}^{2}(s). For notational convenience, we suppress the dependence of 1(s),2(s),(s){\mathcal{E}}^{1}(s),{\mathcal{E}}^{2}(s),{\mathcal{E}}(s) on δ1,δ2\delta_{1},\delta_{2}.

We claim that (s){τcs}{\mathcal{E}}(s)\subseteq\{\tau_{c}\leq s\}. To see this, observe that on the event 1(s){\mathcal{E}}^{1}(s), the ordered Atlas particles Xi+1X_{i+1} and X~i+1\tilde{X}_{i+1} stay above the level yi+yi+1+δ22\frac{y_{i}+y_{i+1}+\delta_{2}}{2} by time ss. Moreover, on 2(s){\mathcal{E}}^{2}(s), {στr8D1𝒃+1s}\{\sigma\leq\tau^{*}\wedge\frac{r}{8\|D^{-1}{\boldsymbol{b}}\|+1}\wedge s\} which, by the previous discussion, implies D1𝚿(t)𝒟D^{-1}\mathbf{\Psi}(t)\notin\partial\mathcal{D} and D1𝚿~(t)𝒟D^{-1}\tilde{\mathbf{\Psi}}(t)\notin\partial\mathcal{D} for all tσt\leq\sigma, that is, none of the co-ordinates of 𝚿()\mathbf{\Psi}(\cdot) or 𝚿~()\tilde{\mathbf{\Psi}}(\cdot) hit zero by time σ\sigma. Hence, for all tσt\leq\sigma, 𝚿(t)=(Z1(t),,Zi(t),yi+yi+1+δ22Xi(t))\mathbf{\Psi}(t)=\left(Z_{1}(t),\dots,Z_{i}(t),\frac{y_{i}+y_{i+1}+\delta_{2}}{2}-X_{i}(t)\right)^{\prime} and 𝚿~(t)=(Z~1(t),,Z~i(t),yi+yi+1+δ22X~i(t))\tilde{\mathbf{\Psi}}(t)=\left(\tilde{Z}_{1}(t),\dots,\tilde{Z}_{i}(t),\frac{y_{i}+y_{i+1}+\delta_{2}}{2}-\tilde{X}_{i}(t)\right)^{\prime}. Further, by the mirror coupling dynamics, 𝚿(σ)=𝚿~(σ)\mathbf{\Psi}(\sigma)=\tilde{\mathbf{\Psi}}(\sigma) and thus, under (s){\mathcal{E}}(s),

(Z1(t),,Zi(t),yi+yi+1+δ22Xi(t))=(Z~1(t),,Z~i(t),yi+yi+1+δ22X~i(t))\left(Z_{1}(t),\dots,Z_{i}(t),\frac{y_{i}+y_{i+1}+\delta_{2}}{2}-X_{i}(t)\right)=\left(\tilde{Z}_{1}(t),\dots,\tilde{Z}_{i}(t),\frac{y_{i}+y_{i+1}+\delta_{2}}{2}-\tilde{X}_{i}(t)\right)

for all tσt\geq\sigma. As σs\sigma\leq s under (s){\mathcal{E}}(s), we conclude that the above equality holds for all tst\geq s. Finally, as XiX_{i} and Xi+1X_{i+1} (also, X~i\tilde{X}_{i} and X~i+1\tilde{X}_{i+1}) do not meet by time σ\sigma, Xj(t)=X~j(t)X_{j}(t)=\tilde{X}_{j}(t) for all ji+1j\geq i+1, for all tσt\leq\sigma, and hence for all t0t\geq 0. These observations imply τcs\tau_{c}\leq s on (s){\mathcal{E}}(s).

The following lemma gives a key estimate on the probabilities of this event. Recall t0>0t_{0}>0 for which ψ=Tt0ψ0\psi=T_{t_{0}}\psi_{0}.

Lemma 4.5.

For any η>0\eta>0, there is a δ0(0,δ¯)\delta_{0}\in(0,\underline{\delta}) and t1(0,t0δ¯)t_{1}\in(0,t_{0}\wedge\underline{\delta}) such that 𝐳,δ1,δ2,i(τc>t1)((t1)c)η{\mathbb{P}}_{{\boldsymbol{z}},\delta_{1},\delta_{2},i}(\tau_{c}>t_{1})\leq{\mathbb{P}}({\mathcal{E}}(t_{1})^{c})\leq\eta for all δ1,δ2(0,δ0)\delta_{1},\delta_{2}\in(0,\delta_{0}).

Proof.

Note that the inequality 𝒛,δ1,δ2,i(τc>t1)((t1)c){\mathbb{P}}_{{\boldsymbol{z}},\delta_{1},\delta_{2},i}(\tau_{c}>t_{1})\leq{\mathbb{P}}({\mathcal{E}}(t_{1})^{c}) is immediate from the discussion above the lemma. Now fix η(0,1)\eta\in(0,1). Constants appearing in this proof may depend on 𝒛{\boldsymbol{z}} and this dependence is not noted explicitly. By a union bound and properties of Brownian motion we see that for any δ2(0,δ¯)\delta_{2}\in(0,\underline{\delta}) and s(0,1)s\in(0,1),

((1(s))c)2j=i+1Φ¯(yjyi+1+12(zi+1δ2)glδ¯s),{\mathbb{P}}(({\mathcal{E}}^{1}(s))^{c})\leq 2\sum_{j=i+1}^{\infty}\bar{\Phi}\left(\frac{y_{j}-y_{i+1}+\frac{1}{2}(z_{i+1}-\delta_{2})-g^{l}\underline{\delta}}{\sqrt{s}}\right), (4.25)

where for uu\in{\mathbb{R}},

Φ¯(u)=12πuev2/2𝑑v.\bar{\Phi}(u)=\frac{1}{\sqrt{2\pi}}\int_{u}^{\infty}e^{-v^{2}/2}dv. (4.26)

Note that, for u0u\geq 0, Φ¯(u)2eu2/4\bar{\Phi}(u)\leq\sqrt{2}e^{-u^{2}/4}. By our condition on δ¯\underline{\delta} in (4.23) and using δ2(0,δ¯)\delta_{2}\in(0,\underline{\delta}),

12(zi+1δ2)glδ¯12(zi+1(2gl+1)δ¯)c1>0.\frac{1}{2}(z_{i+1}-\delta_{2})-g^{l}\underline{\delta}\geq\frac{1}{2}(z_{i+1}-(2g^{l}+1)\underline{\delta})\doteq c_{1}>0.

Thus, from (4.25) it follows that

((1(s))c)\displaystyle{\mathbb{P}}(({\mathcal{E}}^{1}(s))^{c}) 8ec12/4sj=i+1e(yjyi+1)2/4s\displaystyle\leq\sqrt{8}e^{-c_{1}^{2}/4s}\sum_{j=i+1}^{\infty}e^{-(y_{j}-y_{i+1})^{2}/4s} (4.27)
8ec12/4sj=i+1e(yjyi+1)2/48ec12/4seyi+12/4j=i+1eyj2/8,\displaystyle\leq\sqrt{8}e^{-c_{1}^{2}/4s}\sum_{j=i+1}^{\infty}e^{-(y_{j}-y_{i+1})^{2}/4}\leq\sqrt{8}e^{-c_{1}^{2}/4s}e^{y_{i+1}^{2}/4}\sum_{j=i+1}^{\infty}e^{-y_{j}^{2}/8}, (4.28)

where the first inequality uses (a+b)2a2+b2(a+b)^{2}\geq a^{2}+b^{2} for a,b0a,b\geq 0, the second uses the fact that s(0,1)s\in(0,1), and the third uses the inequality (ab)2a2/2b2(a-b)^{2}\geq a^{2}/2-b^{2} for a,ba,b\in{\mathbb{R}}. Thus, choosing t(0,1)t^{\prime}\in(0,1) such that

8ec12/4teyi+12/4j=i+1eyj2/8η/4,\sqrt{8}e^{-c_{1}^{2}/4t^{\prime}}e^{y_{i+1}^{2}/4}\sum_{j=i+1}^{\infty}e^{-y_{j}^{2}/8}\leq\eta/4,

(here we use 𝒚𝒰{\boldsymbol{y}}\in{\mathcal{U}}), we obtain

((1(s))c)η/4 for all δ2(0,δ¯) and all st.{\mathbb{P}}(({\mathcal{E}}^{1}(s))^{c})\leq\eta/4\text{ for all }\delta_{2}\in(0,\underline{\delta})\text{ and all }s\leq t^{\prime}. (4.29)

Without loss of generality we assume that t<t0δ¯t^{\prime}<t_{0}\wedge\underline{\delta}. Next, note that, by the independence of MM and 𝑴{\boldsymbol{M}}^{\perp}, for any s(0,r8D1𝒃+1]s\in(0,\frac{r}{8\|D^{-1}{\boldsymbol{b}}\|+1}],

((2(s))c)=1(inftsM(t)>r/4,suptsM(t)𝒗/2)(supts𝑴(t)<r/4).{\mathbb{P}}(({\mathcal{E}}^{2}(s))^{c})=1-{\mathbb{P}}\left(\inf_{t\leq s}M(t)>-r/4,\ \sup_{t\leq s}M(t)\geq\|{\boldsymbol{v}}\|/2\right){\mathbb{P}}\left(\sup_{t\leq s}\|{\boldsymbol{M}}^{\perp}(t)\|<r/4\right). (4.30)

Recall rr defined in (4.22). Writing r=r(δ1,δ2)r=r(\delta_{1},\delta_{2}) to highlight its dependence on δ1,δ2\delta_{1},\delta_{2}, it follows from (4.22) and the explicit form of 𝚿(0)\mathbf{\Psi}(0) and 𝚿~(0)\tilde{\mathbf{\Psi}}(0) that there exists δ(0,δ¯)\delta^{\prime}\in(0,\underline{\delta}) and c2>0c_{2}>0 such that

infδ1,δ2(0,δ)r(δ1,δ2)c2>0.\inf_{\delta_{1},\delta_{2}\in(0,\delta^{\prime})}r(\delta_{1},\delta_{2})\doteq c_{2}>0.

Hence, we obtain t1(0,tr8D1𝒃+1)t_{1}\in(0,t^{\prime}\wedge\frac{r}{8\|D^{-1}{\boldsymbol{b}}\|+1}) such that

(inftt1M(t)>r/4)1η/4,(suptt1𝑴(t)<r/4)1η/4, for all δ1,δ2(0,δ).{\mathbb{P}}\left(\inf_{t\leq t_{1}}M(t)>-r/4\right)\geq 1-\eta/4,\ \ {\mathbb{P}}\left(\sup_{t\leq t_{1}}\|{\boldsymbol{M}}^{\perp}(t)\|<r/4\right)\geq 1-\eta/4,\text{ for all }\delta_{1},\delta_{2}\in(0,\delta^{\prime}). (4.31)

Recall 𝒗=D1(𝚿~(0)𝚿(0)){\boldsymbol{v}}=D^{-1}(\tilde{\mathbf{\Psi}}(0)-\mathbf{\Psi}(0)). As 𝚿~(0)𝚿(0)=(0,,0,δ1,δ2)\tilde{\mathbf{\Psi}}(0)-\mathbf{\Psi}(0)=(0,\dots,0,\delta_{1},-\delta_{2}) and D1D^{-1} depends only on ii and not on δ1,δ2\delta_{1},\delta_{2}, we can obtain c3>0c_{3}>0 depending only on ii such that

𝒗c3δ12+δ22 for all δ1,δ2>0.\|{\boldsymbol{v}}\|\leq c_{3}\sqrt{\delta_{1}^{2}+\delta_{2}^{2}}\text{ for all }\delta_{1},\delta_{2}>0.

Hence, we obtain δ0(0,δ)\delta_{0}\in(0,\delta^{\prime}) depending on t1t_{1} such that

(suptt1M(t)𝒗/2)1η/4 for all δ1,δ2(0,δ0).{\mathbb{P}}\left(\sup_{t\leq t_{1}}M(t)\geq\|{\boldsymbol{v}}\|/2\right)\geq 1-\eta/4\text{ for all }\delta_{1},\delta_{2}\in(0,\delta_{0}). (4.32)

From (4.31) and (4.32), for all δ1,δ2(0,δ0)\delta_{1},\delta_{2}\in(0,\delta_{0}),

(inftt1M(t)>r/4,suptt1M(t)𝒗/2)(inftt1M(t)>r/4)(suptt1M(t)<𝒗/2)1η/2.{\mathbb{P}}\left(\inf_{t\leq t_{1}}M(t)>-r/4,\ \sup_{t\leq t_{1}}M(t)\geq\|{\boldsymbol{v}}\|/2\right)\geq{\mathbb{P}}\left(\inf_{t\leq t_{1}}M(t)>-r/4\right)-{\mathbb{P}}\left(\sup_{t\leq t_{1}}M(t)<\|{\boldsymbol{v}}\|/2\right)\geq 1-\eta/2. (4.33)

Using (4.31) and (4.33) in (4.30), we obtain for all δ1,δ2(0,δ0)\delta_{1},\delta_{2}\in(0,\delta_{0}),

((2(t1))c)1(1η/2)(1η/4)3η/4.{\mathbb{P}}(({\mathcal{E}}^{2}(t_{1}))^{c})\leq 1-(1-\eta/2)(1-\eta/4)\leq 3\eta/4. (4.34)

From (4.29) and (4.34), we conclude

(((t1))c)((1(t1))c)+((2(t1))c)η for all δ1,δ2(0,δ0),{\mathbb{P}}(({\mathcal{E}}(t_{1}))^{c})\leq{\mathbb{P}}(({\mathcal{E}}^{1}(t_{1}))^{c})+{\mathbb{P}}(({\mathcal{E}}^{2}(t_{1}))^{c})\leq\eta\text{ for all }\delta_{1},\delta_{2}\in(0,\delta_{0}),

which proves the lemma. ∎

Proof of Proposition 4.3. In order to prove the proposition it suffices to show the right continuity of (δ1,δ2)ψ(𝒛+δ1𝒆iδ2𝒆i+1)(\delta_{1},\delta_{2})\mapsto\psi({\boldsymbol{z}}+\delta_{1}{\boldsymbol{e}}_{i}-\delta_{2}{\boldsymbol{e}}_{i+1}) at (δ1,δ2)=(0,0)(\delta_{1},\delta_{2})=(0,0). Recall that ψ=Tt0ψ0\psi=T_{t_{0}}\psi_{0}. Let ε>0{\varepsilon}>0 be arbitrary and let ηε/(2ψ0)\eta\doteq{\varepsilon}/(2\|\psi_{0}\|_{\infty}). Let δ¯\underline{\delta} be as in (4.23) and, for this chosen η\eta, let δ0(0,δ¯)\delta_{0}\in(0,\underline{\delta}) and t1(0,t0δ¯)t_{1}\in(0,t_{0}\wedge\underline{\delta}) be as in Lemma 4.5. For δ1>0\delta_{1}>0, δ2(0,zi+1)\delta_{2}\in(0,z_{i+1}), let 𝒛,δ1,δ2,i{\mathbb{P}}_{{\boldsymbol{z}},\delta_{1},\delta_{2},i} be as in (4.21) and let 𝔼𝒛,δ1,δ2,i{\mathbb{E}}_{{\boldsymbol{z}},\delta_{1},\delta_{2},i} be the corresponding expectation operator. For any δ1,δ2(0,δ0)\delta_{1},\delta_{2}\in(0,\delta_{0}),

|ψ(𝒛+δ1𝒆iδ2𝒆i+1)ψ(𝒛)|\displaystyle|\psi({\boldsymbol{z}}+\delta_{1}{\boldsymbol{e}}_{i}-\delta_{2}{\boldsymbol{e}}_{i+1})-\psi({\boldsymbol{z}})| =|𝔼𝒛δ1,δ2,iψ0(𝒁(t0))𝔼𝒛ψ0(𝒁(t0))|\displaystyle=|{\mathbb{E}}_{{\boldsymbol{z}}^{\delta_{1},\delta_{2},i}}\psi_{0}({\boldsymbol{Z}}(t_{0}))-{\mathbb{E}}_{{\boldsymbol{z}}}\psi_{0}({\boldsymbol{Z}}(t_{0}))|
𝔼𝒛,δ1,δ2,i[|ψ0(𝒁(1)(t0))ψ0(𝒁(2)(t0))|1{τc>t1}]\displaystyle\leq{\mathbb{E}}_{{\boldsymbol{z}},\delta_{1},\delta_{2},i}\left[|\psi_{0}({\boldsymbol{Z}}^{(1)}(t_{0}))-\psi_{0}({\boldsymbol{Z}}^{(2)}(t_{0}))|1_{\{\tau_{c}>t_{1}\}}\right]
2ψ0𝒛,δ1,δ2,i(τc>t1)\displaystyle\leq 2\|\psi_{0}\|_{\infty}{\mathbb{P}}_{{\boldsymbol{z}},\delta_{1},\delta_{2},i}(\tau_{c}>t_{1})
2ψ0((t1)c)2ψ0η=ε,\displaystyle\leq 2\|\psi_{0}\|_{\infty}{\mathbb{P}}({\mathcal{E}}(t_{1})^{c})\leq 2\|\psi_{0}\|_{\infty}\eta={\varepsilon},

where the fourth inequality uses Lemma 4.5. Since ε>0{\varepsilon}>0 is arbitrary, the result follows. ∎

Proof of Proposition 4.2. Let 𝒛{\boldsymbol{z}} and δ1,δ2\delta_{1},\delta_{2} be as in the statement of the proposition and let 𝒛,δ1,δ2,i{\mathbb{P}}_{{\boldsymbol{z}},\delta_{1},\delta_{2},i} be as in (4.21). Recall the event 2(s){\mathcal{E}}^{2}(s) from (4.4) (defined for any s>0s>0) and consider the following modification of 1(s){\mathcal{E}}^{1}(s):

~1(s){infji+1inf0ts(yj+Bj(t)tgl)>yi+yi+1+δ22},s>0.\displaystyle\tilde{\mathcal{E}}^{1}(s)\doteq\left\{\inf_{j\geq i+1}\inf_{0\leq t\leq s}(y_{j}+B_{j}(t)-tg^{l})>\frac{y_{i}+y_{i+1}+\delta_{2}}{2}\right\},\ s>0.

Let ~(s)~1(s)2(s)\tilde{\mathcal{E}}(s)\doteq\tilde{\mathcal{E}}^{1}(s)\cap{\mathcal{E}}^{2}(s). From the definition of the coupling 𝒛,δ1,δ2,i{\mathbb{P}}_{{\boldsymbol{z}},\delta_{1},\delta_{2},i} it is easily seen by an argument similar to that given before Lemma 4.5 that

𝒛,δ1,δ2,i(𝒁(1)(s)=𝒁(2)(s))(~(s)).{\mathbb{P}}_{{\boldsymbol{z}},\delta_{1},\delta_{2},i}({\boldsymbol{Z}}^{(1)}(s)={\boldsymbol{Z}}^{(2)}(s))\geq{\mathbb{P}}(\tilde{\mathcal{E}}(s)).

Note that, as ~1(s)\tilde{\mathcal{E}}^{1}(s) is given in terms of Brownian motions {Bj}ji+1\{B_{j}\}_{j\geq i+1} and 2(s){\mathcal{E}}^{2}(s) is defined in terms of {Bj}ji\{B_{j}\}_{j\leq i}, ~1(s)\tilde{\mathcal{E}}^{1}(s) and 2(s){\mathcal{E}}^{2}(s) are independent. Thus, for any s>0s>0, (~(s))=(~1(s))(2(s)){\mathbb{P}}(\tilde{\mathcal{E}}(s))={\mathbb{P}}(\tilde{\mathcal{E}}^{1}(s)){\mathbb{P}}({\mathcal{E}}^{2}(s)) and it suffices to show that each term in the product is positive.

As 𝒚𝒰{\boldsymbol{y}}\in\mathcal{U}, yjy_{j}\rightarrow\infty as jj\rightarrow\infty. Hence, we can obtain j0i+2j_{0}\geq i+2 depending on s,is,i such that yj2sgl+2(yi+yi+1+δ2)y_{j}\geq 2sg^{l}+2(y_{i}+y_{i+1}+\delta_{2}) for all jj0j\geq j_{0} and j=j0Φ¯(yj4s)1/4\sum_{j=j_{0}}^{\infty}\bar{\Phi}\left(\frac{y_{j}}{4\sqrt{s}}\right)\leq 1/4. Thus,

(infjj0inf0ts(yj+Bj(t)tgl)>yi+yi+1+δ22)\displaystyle{\mathbb{P}}\left(\inf_{j\geq j_{0}}\inf_{0\leq t\leq s}(y_{j}+B_{j}(t)-tg^{l})>\frac{y_{i}+y_{i+1}+\delta_{2}}{2}\right) (infjj0inf0ts(yj2+Bj(t))>yi+yi+1+δ22)\displaystyle\geq{\mathbb{P}}\left(\inf_{j\geq j_{0}}\inf_{0\leq t\leq s}(\frac{y_{j}}{2}+B_{j}(t))>\frac{y_{i}+y_{i+1}+\delta_{2}}{2}\right)
12j=j0Φ¯(yj4s)1/2.\displaystyle\geq 1-2\sum_{j=j_{0}}^{\infty}\bar{\Phi}\left(\frac{y_{j}}{4\sqrt{s}}\right)\geq 1/2. (4.35)

Moreover, from standard Brownian motion estimates,

(infi+1j<j0inf0ts(yj+Bj(t)tgl)>yi+yi+1+δ22)>0.\displaystyle{\mathbb{P}}\left(\inf_{i+1\leq j<j_{0}}\inf_{0\leq t\leq s}(y_{j}+B_{j}(t)-tg^{l})>\frac{y_{i}+y_{i+1}+\delta_{2}}{2}\right)>0. (4.36)

From the independence of the events considered in (4.4) and (4.36), we conclude that (~1(s))>0{\mathbb{P}}(\tilde{\mathcal{E}}^{1}(s))>0. Finally, from standard Brownian motion estimates and the explicit form of M()M(\cdot) and 𝑴{\boldsymbol{M}}^{\perp} in terms of 𝑩(i)(){\boldsymbol{B}}^{(i)}(\cdot),

(inftsr8D1𝒃+1M(t)r/4,suptsr8D1𝒃+1M(t)𝒗/2,suptsr8D1𝒃+1𝑴(t)r/4)=(inftsr8D1𝒃+1M(t)r/4,suptsr8D1𝒃+1M(t)𝒗/2)(suptsr8D1𝒃+1𝑴(t)r/4)>0.{\mathbb{P}}\left(\inf_{t\leq s\wedge\frac{r}{8\|D^{-1}{\boldsymbol{b}}\|+1}}M(t)\geq-r/4,\ \sup_{t\leq s\wedge\frac{r}{8\|D^{-1}{\boldsymbol{b}}\|+1}}M(t)\geq\|{\boldsymbol{v}}\|/2,\ \sup_{t\leq s\wedge\frac{r}{8\|D^{-1}{\boldsymbol{b}}\|+1}}\|{\boldsymbol{M}}^{\perp}(t)\|\leq r/4\right)\\ ={\mathbb{P}}\left(\inf_{t\leq s\wedge\frac{r}{8\|D^{-1}{\boldsymbol{b}}\|+1}}M(t)\geq-r/4,\ \sup_{t\leq s\wedge\frac{r}{8\|D^{-1}{\boldsymbol{b}}\|+1}}M(t)\geq\|{\boldsymbol{v}}\|/2\right){\mathbb{P}}\left(\sup_{t\leq s\wedge\frac{r}{8\|D^{-1}{\boldsymbol{b}}\|+1}}\|{\boldsymbol{M}}^{\perp}(t)\|\leq r/4\right)>0.

The result follows. ∎

5 Proof of Theorem 3.5

Recall 𝒟1{\mathcal{D}}_{1} from (1.4). Observe that for 𝒈𝒟1{\boldsymbol{g}}\in{\mathcal{D}}_{1}, g¯Nj+1<g¯Nj\bar{g}_{N_{j+1}}<\bar{g}_{N_{j}} for all j1j\geq 1. Moreover, as 𝒈𝒟{\boldsymbol{g}}\in{\mathcal{D}}, infng¯ninfngn>\inf_{n\in{\mathbb{N}}}\bar{g}_{n}\geq\inf_{n\in{\mathbb{N}}}g_{n}>-\infty. Thus, g¯limjg¯Nj\bar{g}_{\infty}\doteq\lim_{j\rightarrow\infty}\bar{g}_{N_{j}} exists, is finite, and limjg¯Nj=infng¯n\lim_{j\rightarrow\infty}\bar{g}_{N_{j}}=\inf_{n\in{\mathbb{N}}}\bar{g}_{n}. As adding the same drift g¯dt-\bar{g}_{\infty}dt to each ordered particle in (2.3) keeps the gaps unchanged, we can assume without loss of generality that infngn=g¯=0\inf_{n\in{\mathbb{N}}}g_{n}=\bar{g}_{\infty}=0. In particular, g¯n>0\bar{g}_{n}>0 for all nn\in\mathbb{N}.

Fix 𝒈𝒟1{\boldsymbol{g}}\in{\mathcal{D}}_{1} and let π=i=1πi\pi=\otimes_{i=1}^{\infty}\pi_{i} be as in the statement of Theorem 3.5. The assumption (3.2) on π\pi will be taken to hold throughout the section. From Theorem 2.1 we can construct a filtered probability space (Ω,,,{t})(\Omega,{\mathcal{F}},{\mathbb{P}},\{{\mathcal{F}}_{t}\}) equipped with mutually independent real {t}t0\{{\mathcal{F}}_{t}\}_{t\geq 0}-Brownian motions {Bi}i0\{B^{*}_{i}\}_{i\in{\mathbb{N}}_{0}} and continuous processes {Y(i),i0}\{Y_{(i)},i\in{\mathbb{N}}_{0}\} that solve the SDE (2.3), where {Li}i0\{L^{*}_{i}\}_{i\in{\mathbb{N}}_{0}} are as introduced below (2.3), such that with {Zi}i\{Z_{i}\}_{i\in{\mathbb{N}}} defined as in (2.4), the process 𝒁=(Z1,Z2,){\boldsymbol{Z}}=(Z_{1},Z_{2},\ldots)^{\prime} has the distribution π𝒈{\mathbb{P}}^{{\boldsymbol{g}}}_{\pi}. Furthermore, without loss of generality, we can assume that Y(0)(0)=0Y_{(0)}(0)=0. We will write {\mathbb{P}} and 𝔼{\mathbb{E}} respectively for the probability and expectation under the law of this {\mathbb{R}}^{\infty}-valued process.

5.1 Proof overview

First we give an overview of the approach. We will use moment generating functions (m.g.f) to identify the marginals of π\pi; so the first step is to establish finiteness of the m.g.f. of any fixed gap in a positive interval around zero. This is achieved in Lemma 5.1 by using comparison techniques between the gap processes of infinite and finite versions of the model, the latter having a unique invariant distribution that is a product of Exponential distributions. Lemmas 5.2 and 5.3 together establish the uniform integrability of {1ε011{0Zi(s)ε}𝑑s,ε(0,1/2)}\{\frac{1}{{\varepsilon}}\int_{0}^{1}1_{\{0\leq Z_{i}(s)\leq{\varepsilon}\}}ds,\;{\varepsilon}\in(0,1/2)\} for any ii\in{\mathbb{N}}, which is later used in showing the existence of limε01επi[0,ε]\lim_{{\varepsilon}\downarrow 0}\frac{1}{{\varepsilon}}\pi_{i}[0,{\varepsilon}] and to identify this as 𝔼(Li(1)){\mathbb{E}}(L^{*}_{i}(1)) (LiL^{*}_{i} being the local time at zero of the ii-th gap in (2.3)). Lemma 5.4, which is key to the proof of Theorem 3.5, gives an explicit representation for the expectation of the integral of a function of the ii-th gap process against the jj-th local time process for iji\neq j. The aforementioned uniform integrability is crucially used here. For any ii\in{\mathbb{N}}, the m.g.f. of the ii-th gap at time 11 is then identified by an application of Itô’s formula to exponential functions of the gap and using the representation in Lemma 5.4 to evaluate the local time terms. The obtained m.g.f. corresponds to that of an exponential random variable. The associated rates are then shown to agree with that of πa𝒈\pi^{{\boldsymbol{g}}}_{a} for some a2infng¯na\geq-2\inf_{n\in{\mathbb{N}}}\bar{g}_{n} via a recursive relation resulting from taking expectations in (5.1). The representation (3.3) is obtained as a by-product of our computations.

5.2 Preliminary results

We begin with some preliminary results.

Lemma 5.1.

For any ii\in{\mathbb{N}} and λ<2k=0i1gk\lambda<2\sum_{k=0}^{i-1}g_{k}, we have +eλzπi(dz)<\int_{{\mathbb{R}}_{+}}e^{\lambda z}\pi_{i}(dz)<\infty.

Proof.

Recall the sequence {Nj}\{N_{j}\} associated with 𝒈𝒟1{\boldsymbol{g}}\in{\mathcal{D}}_{1}. Fix any dd\in\mathbb{N} such that Nd>iN_{d}>i and consider the NdN_{d} dimensional (𝒈,𝒚)({\boldsymbol{g}},{\boldsymbol{y}})-Atlas model defined by replacing \infty with Nd1N_{d}-1 in equation (2.1). This model has been studied extensively in previous works (see e.g. [32, 34]) and it is well known that, since 𝒈𝒟1{\boldsymbol{g}}\in{\mathcal{D}}_{1}, the associated gap sequence {Zj}j=1Nd1\{Z_{j}\}_{j=1}^{N_{d}-1} defined by (2.4), where the processes Y(j)Y_{(j)} are defined by (2.3), for j=0,1,,Nd1j=0,1,\ldots,N_{d}-1, has a unique stationary distribution π~(Nd1)l=1Nd1Exp(2l(g¯lg¯Nd))\tilde{\pi}^{(N_{d}-1)}\doteq\otimes_{l=1}^{N_{d}-1}\mbox{Exp}(2l(\bar{g}_{l}-\bar{g}_{N_{d}})) (see [32, Proposition 2.2(4)]). Using monotonocity and comparison estimates for finite and infinite Atlas models (cf. [32, Corollary 3.14] and [34, equations (58)-(60)]) it now follows that the probability measure π|Nd1\pi|_{N_{d}-1} on +Nd1{\mathbb{R}}_{+}^{N_{d}-1} given as the first Nd1N_{d}-1 marginal distribution of π\pi satisfies π|Nd1stπ~(Nd1)\pi|_{N_{d}-1}\leq_{\mbox{{\tiny st}}}\tilde{\pi}^{(N_{d}-1)}. In particular, πistExp(2i(g¯ig¯Nd))\pi_{i}\leq_{\mbox{{\tiny st}}}\mbox{Exp}(2i(\bar{g}_{i}-\bar{g}_{N_{d}})) for all Nd>iN_{d}>i. Since g¯Nd0\bar{g}_{N_{d}}\to 0 as dd\to\infty, we can find a dd\in{\mathbb{N}}, with Nd>iN_{d}>i, such that λ<2i(g¯ig¯Nd)\lambda<2i(\bar{g}_{i}-\bar{g}_{N_{d}}). Then

+eλzπi(dz)+eλzExp(2i(g¯ig¯Nd))(dz)<\int_{{\mathbb{R}}_{+}}e^{\lambda z}\pi_{i}(dz)\leq\int_{{\mathbb{R}}_{+}}e^{\lambda z}\mbox{Exp}(2i(\bar{g}_{i}-\bar{g}_{N_{d}}))(dz)<\infty

which completes the proof. ∎

The next three lemmas concern the collection {Zi,Li}\{Z_{i},L^{*}_{i}\} described above. We remind the reader that in these lemmas the probability measure \mathbb{P} and the expectation 𝔼\mathbb{E} correspond to π𝒈{\mathbb{P}}^{{\boldsymbol{g}}}_{\pi} and 𝔼π𝒈{\mathbb{E}}^{{\boldsymbol{g}}}_{\pi} respectively, where 𝒈{\boldsymbol{g}} and π\pi are as fixed at the beginning of the section.

Lemma 5.2.

For every ii\in{\mathbb{N}}, 𝔼(Li(1))2<{\mathbb{E}}(L^{*}_{i}(1))^{2}<\infty.

Proof.

From (2.3) it follows that, for ii\in{\mathbb{N}}, and 0t10\leq t\leq 1,

Zi(t)=Zi(0)+hit+Wi(t)12Li1(t)12Li+1(t)+Li(t),Z_{i}(t)=Z_{i}(0)+h_{i}t+W_{i}^{*}(t)-\frac{1}{2}L^{*}_{i-1}(t)-\frac{1}{2}L^{*}_{i+1}(t)+L^{*}_{i}(t), (5.1)

where for ii\in{\mathbb{N}}, hi=gigi1h_{i}=g_{i}-g_{i-1} and Wi=BiBi1W_{i}^{*}=B_{i}^{*}-B_{i-1}^{*}. This says that

Li+1(1)=2(Zi(0)Zi(1)+hi+Wi(1)+Li(1))Li1(1).L^{*}_{i+1}(1)=2(Z_{i}(0)-Z_{i}(1)+h_{i}+W_{i}^{*}(1)+L^{*}_{i}(1))-L^{*}_{i-1}(1).

From this and Lemma 5.1 it follows that if for some ii\in{\mathbb{N}} 𝔼(Li(1))2<{\mathbb{E}}(L^{*}_{i}(1))^{2}<\infty, then 𝔼(Li+1(1))2<{\mathbb{E}}(L^{*}_{i+1}(1))^{2}<\infty as well. Thus it suffices to show that 𝔼(L1(1))2<{\mathbb{E}}(L^{*}_{1}(1))^{2}<\infty. From (2.3), and recalling that Y(0)(0)=0Y_{(0)}(0)=0, we see that

Y(0)(1)=g0+B1(1)12L1(1)Y_{(0)}(1)=g_{0}+B^{*}_{1}(1)-\frac{1}{2}L^{*}_{1}(1)

which says that

L1(1)=2(Y(0)(1)+g0+B1(1)).L^{*}_{1}(1)=2(-Y_{(0)}(1)+g_{0}+B^{*}_{1}(1)).

Thus to prove the lemma it suffices to show that

𝔼(infj0Yj(1))2<,{\mathbb{E}}\left(\inf_{j\in{\mathbb{N}}_{0}}Y_{j}(1)\right)^{2}<\infty, (5.2)

where {Yj}\{Y_{j}\} solve the system of equations in (2.1) with Y0(0)=0Y_{0}(0)=0 and the vector

(Y1(0),Y2(0)Y1(0),Y3(0)Y2(0),)(Y_{1}(0),Y_{2}(0)-Y_{1}(0),Y_{3}(0)-Y_{2}(0),\ldots)

distributed as π\pi.

Note that for x>g0x>g_{0},

(infj0Yj(1)>x)(g0+B1(1)>x)2e(xg0)2/4.{\mathbb{P}}(\inf_{j\in{\mathbb{N}}_{0}}Y_{j}(1)>x)\leq{\mathbb{P}}(g_{0}+B^{*}_{1}(1)>x)\leq\sqrt{2}e^{-(x-g_{0})^{2}/4}. (5.3)

Next, as 𝒈𝒟{\boldsymbol{g}}\in{\mathcal{D}}, there exists gl>0g^{l}>0 such that giglg_{i}\geq-g^{l} for all i0i\in\mathbb{N}_{0}. Thus, we have, for x0x\geq 0,

(infj0Yj(1)<x)\displaystyle{\mathbb{P}}(\inf_{j\in{\mathbb{N}}_{0}}Y_{j}(1)<-x) (infj0(Yj(0)+Wj(1))<glx)\displaystyle\leq{\mathbb{P}}(\inf_{j\in{\mathbb{N}}_{0}}(Y_{j}(0)+W_{j}(1))<g^{l}-x)
j=0(Wj(1)glxYj(0))=j=0𝔼Φ¯(x+Yj(0)gl),\displaystyle\leq\sum_{j=0}^{\infty}{\mathbb{P}}(W_{j}(1)\leq g^{l}-x-Y_{j}(0))=\sum_{j=0}^{\infty}{\mathbb{E}}\bar{\Phi}(x+Y_{j}(0)-g^{l}),

where Φ¯(z)\bar{\Phi}(z) was defined in (4.26). Thus, for xglx\geq g^{l},

(infj0Yj(1)x)\displaystyle{\mathbb{P}}(\inf_{j\in{\mathbb{N}}_{0}}Y_{j}(1)\leq-x) 2j=0𝔼e(x+Yj(0)gl)2/42e(xgl)2/4j=0𝔼e(Yj(0))2/4.\displaystyle\leq\sqrt{2}\sum_{j=0}^{\infty}{\mathbb{E}}e^{-(x+Y_{j}(0)-g^{l})^{2}/4}\leq\sqrt{2}e^{-(x-g^{l})^{2}/4}\sum_{j=0}^{\infty}{\mathbb{E}}e^{-(Y_{j}(0))^{2}/4}.

The desired square integrability in (5.2) is now immediate from (5.3) and the above on observing that

j=0𝔼e(Yj(0))2/4=1++(j=1e14(l=1jzl)2)π(d𝒛)<\sum_{j=0}^{\infty}{\mathbb{E}}e^{-(Y_{j}(0))^{2}/4}=1+\int_{{\mathbb{R}}_{+}^{\infty}}\left(\sum_{j=1}^{\infty}e^{-\frac{1}{4}(\sum_{l=1}^{j}z_{l})^{2}}\right)\pi(d{\boldsymbol{z}})<\infty

by our assumption. ∎

The next lemma will allow us to interchange expectations and limits as ε0{\varepsilon}\to 0.

Lemma 5.3.

The family {1ε011{0Zi(s)ε}𝑑s,ε(0,1/2)}\{\frac{1}{{\varepsilon}}\int_{0}^{1}1_{\{0\leq Z_{i}(s)\leq{\varepsilon}\}}ds,\;{\varepsilon}\in(0,1/2)\} is uniformly integrable.

Proof.

For ε(0,1/2){\varepsilon}\in(0,1/2), define ψε:\psi_{{\varepsilon}}:{\mathbb{R}}\to{\mathbb{R}} as

ψε(z){z22 if 0zεε22+(zε)ε if z>ε.\psi_{{\varepsilon}}(z)\doteq\begin{cases}\frac{z^{2}}{2}&\mbox{ if }0\leq z\leq{\varepsilon}\\ \frac{{\varepsilon}^{2}}{2}+(z-{\varepsilon}){\varepsilon}&\mbox{ if }z>{\varepsilon}.\end{cases} (5.4)

Note that ψε\psi_{{\varepsilon}} is a C1C^{1} function with an absolutely continuous derivative. It then follows from Itô’s formula (see [25, Problem 3.7.3]) applied to ZiZ_{i} given by (5.1), that

ψε(Zi(1))\displaystyle\psi_{{\varepsilon}}(Z_{i}(1)) =ψε(Zi(0))+01ψε(Zi(s))hi𝑑s+01ψε(Zi(s))𝑑Wi(s)\displaystyle=\psi_{{\varepsilon}}(Z_{i}(0))+\int_{0}^{1}\psi^{\prime}_{{\varepsilon}}(Z_{i}(s))h_{i}ds+\int_{0}^{1}\psi^{\prime}_{{\varepsilon}}(Z_{i}(s))dW^{*}_{i}(s)
1201ψε(Zi(s))𝑑Li+1(s)1201ψε(Zi(s))𝑑Li1(s)\displaystyle\quad-\frac{1}{2}\int_{0}^{1}\psi^{\prime}_{{\varepsilon}}(Z_{i}(s))dL^{*}_{i+1}(s)-\frac{1}{2}\int_{0}^{1}\psi^{\prime}_{{\varepsilon}}(Z_{i}(s))dL^{*}_{i-1}(s)
+01ψε(Zi(s))𝑑Li(s)+01ψε′′(Zi(s))𝑑s.\displaystyle\quad+\int_{0}^{1}\psi^{\prime}_{{\varepsilon}}(Z_{i}(s))dL^{*}_{i}(s)+\int_{0}^{1}\psi^{\prime\prime}_{{\varepsilon}}(Z_{i}(s))ds. (5.5)

Note that, for all z+z\in{\mathbb{R}}_{+},

0ψε(z)zε,  0ψε(z)ε,ψε(0)=0.0\leq\psi_{{\varepsilon}}(z)\leq z{\varepsilon},\;\;0\leq\psi^{\prime}_{{\varepsilon}}(z)\leq{\varepsilon},\;\;\psi^{\prime}_{{\varepsilon}}(0)=0.

Also,

ψε′′(z)={1 if 0z<ε0 if z>ε.\psi^{\prime\prime}_{{\varepsilon}}(z)=\begin{cases}1&\mbox{ if }0\leq z<{\varepsilon}\\ 0&\mbox{ if }z>{\varepsilon}.\end{cases}

Combining these, and dividing by ε{\varepsilon} in (5.5), we have

1ε011{0Zi(s)ε}𝑑s\displaystyle\frac{1}{{\varepsilon}}\int_{0}^{1}1_{\{0\leq Z_{i}(s)\leq{\varepsilon}\}}ds Zi(1)1ε01ψε(Zi(s))𝑑Wi(s)+|hi|+12Li1(1)+12Li+1(1).\displaystyle\leq Z_{i}(1)-\frac{1}{{\varepsilon}}\int_{0}^{1}\psi^{\prime}_{{\varepsilon}}(Z_{i}(s))dW^{*}_{i}(s)+|h_{i}|+\frac{1}{2}L^{*}_{i-1}(1)+\frac{1}{2}L^{*}_{i+1}(1).

The desired uniform integrability now follows from Lemmas 5.1 and 5.2 and the observation that

𝔼(1ε01ψε(Zi(s))𝑑Wi(s))22.{\mathbb{E}}\left(\frac{1}{{\varepsilon}}\int_{0}^{1}\psi^{\prime}_{{\varepsilon}}(Z_{i}(s))dW^{*}_{i}(s)\right)^{2}\leq 2.

The following lemma will be key to proving Theorem 3.5. It will be used to represent expectations of integrals of nonnegative measurable functions with respect to local time in terms of stationary integrals and the ‘density’ of πi\pi_{i} at zero for each ii, as described in the lemma. We note that the product form structure of π\pi is crucially exploited here.

Lemma 5.4.

For any ii\in{\mathbb{N}}, the limit νilimε01επi[0,ε]\nu_{i}\doteq\lim_{{\varepsilon}\downarrow 0}\frac{1}{{\varepsilon}}\pi_{i}[0,{\varepsilon}] exists and νi=𝔼(Li(1))\nu_{i}={\mathbb{E}}(L^{*}_{i}(1)). Furthermore, for any measurable f:++f:{\mathbb{R}}_{+}\to{\mathbb{R}}_{+} and i,ji,j\in{\mathbb{N}}, iji\neq j,

𝔼01f(Zi(s))𝑑Lj(s)=νj+f(z)πi(dz).{\mathbb{E}}\int_{0}^{1}f(Z_{i}(s))dL^{*}_{j}(s)=\nu_{j}\int_{{\mathbb{R}}_{+}}f(z)\pi_{i}(dz). (5.6)
Proof.

From results on local times of continuous semimartingales (see e.g. [29, Corollary VI.1.9]) it follows that, for all t[0,1]t\in[0,1] and ii\in{\mathbb{N}}, 1ε0t1{0Zi(s)ε}𝑑s\frac{1}{{\varepsilon}}\int_{0}^{t}1_{\{0\leq Z_{i}(s)\leq{\varepsilon}\}}ds converges a.s. to 1/21/2-times the semimartingale local time Λi(t)\Lambda_{i}(t) of ZiZ_{i} at 0 (as defined in [29, VI.1.2]) as ε0{\varepsilon}\downarrow 0. Furthermore, one has (see [29, Exercise VI.1.16 (3)]) that

Λi(t)\displaystyle\Lambda_{i}(t) =2[hi0t1{Zi(s)=0}ds120t1{Zi(s)=0}dLi1(s)\displaystyle=2\Big{[}h_{i}\int_{0}^{t}1_{\{Z_{i}(s)=0\}}ds-\frac{1}{2}\int_{0}^{t}1_{\{Z_{i}(s)=0\}}dL^{*}_{i-1}(s)
120t1{Zi(s)=0}dLi+1(s)+0t1{Zi(s)=0}dLi(s)]\displaystyle\quad-\frac{1}{2}\int_{0}^{t}1_{\{Z_{i}(s)=0\}}dL^{*}_{i+1}(s)+\int_{0}^{t}1_{\{Z_{i}(s)=0\}}dL^{*}_{i}(s)\Big{]}
=20t1{Zi(s)=0}𝑑Li(s)=2Li(t),\displaystyle=2\int_{0}^{t}1_{\{Z_{i}(s)=0\}}dL^{*}_{i}(s)=2L^{*}_{i}(t),

where the first equality on the last line follows from the facts that 0t1{Zi(s)=0}𝑑s=0\int_{0}^{t}1_{\{Z_{i}(s)=0\}}ds=0 (which follows from Λi(t)<\Lambda_{i}(t)<\infty), and that the Atlas model does not have triple collisions a.s. (see [32, Theorem 5.1]). It then follows that, as ε0{\varepsilon}\downarrow 0,

1ε0t1{0Zi(s)ε}𝑑sLi(t), a.s. for every t[0,1] and i.\frac{1}{{\varepsilon}}\int_{0}^{t}1_{\{0\leq Z_{i}(s)\leq{\varepsilon}\}}ds\to L^{*}_{i}(t),\mbox{ a.s. for every }t\in[0,1]\mbox{ and }i\in{\mathbb{N}}. (5.7)

Combining this with Lemma 5.3 and using the fact that 𝒁{\boldsymbol{Z}} is a stationary process, we now have that, as ε0{\varepsilon}\downarrow 0,

1επi[0,ε]=𝔼1ε011{0Zi(s)ε}𝑑s𝔼(Li(1)), for all i.\frac{1}{{\varepsilon}}\pi_{i}[0,{\varepsilon}]={\mathbb{E}}\frac{1}{{\varepsilon}}\int_{0}^{1}1_{\{0\leq Z_{i}(s)\leq{\varepsilon}\}}ds\to{\mathbb{E}}(L^{*}_{i}(1)),\mbox{ for all }i\in{\mathbb{N}}. (5.8)

This proves the first statement in the lemma. In order to prove (5.6) it suffices to consider the case where ff is bounded (as the general case can be then recovered by monotone convergence theorem). In fact by appealing to the monotone class theorem (cf. [28, Theorem I.8]) we can assume without loss of generality that ff is a continuous and bounded function. From (5.7) we can find Ω0\Omega_{0}\in{\mathcal{F}} such that (Ω0)=1{\mathbb{P}}(\Omega_{0})=1 and for all ωΩ0\omega\in\Omega_{0}

1ε0t1{0Zi(s,ω)ε}𝑑sLi(t,ω), for every t[0,1],\frac{1}{{\varepsilon}}\int_{0}^{t}1_{\{0\leq Z_{i}(s,\omega)\leq{\varepsilon}\}}ds\to L^{*}_{i}(t,\omega),\mbox{ for every }t\in[0,1]\cap{\mathbb{Q}}, (5.9)

and Li(1,ω)<L^{*}_{i}(1,\omega)<\infty for all ii\in{\mathbb{N}}. In particular this says that, for every ωΩ0\omega\in\Omega_{0} and ii\in{\mathbb{N}}, the collection of measures {Πiε,ω,ε(0,1/2)}\{\Pi^{{\varepsilon},\omega}_{i},{\varepsilon}\in(0,1/2)\} defined as

Πiε,ω[a,b]1εab1{0Zi(s,ω)ε}𝑑s, 0ab1\Pi^{{\varepsilon},\omega}_{i}[a,b]\doteq\frac{1}{{\varepsilon}}\int_{a}^{b}1_{\{0\leq Z_{i}(s,\omega)\leq{\varepsilon}\}}ds,\;0\leq a\leq b\leq 1

is relatively compact in the weak convergence topology. From this and using (5.9) again we now see that, for every ωΩ0\omega\in\Omega_{0} and ii\in{\mathbb{N}}, Πiε,ω\Pi^{{\varepsilon},\omega}_{i} converges weakly to Πiω\Pi^{\omega}_{i} defined as

Πiω[a,b]Li(b,ω)Li(a,ω), 0ab1.\Pi^{\omega}_{i}[a,b]\doteq L^{*}_{i}(b,\omega)-L^{*}_{i}(a,\omega),\;0\leq a\leq b\leq 1.

From the sample path continuity of ZiZ_{i} we can assume without loss of generality that for every ωΩ0\omega\in\Omega_{0} and ii\in{\mathbb{N}}, sf(Zi(s,ω))s\mapsto f(Z_{i}(s,\omega)) is a continuous map. From the above weak convergence it then follows that, for ωΩ0\omega\in\Omega_{0} and i,ji,j\in{\mathbb{N}},

01f(Zi(s,ω))𝑑Lj(s,ω)\displaystyle\int_{0}^{1}f(Z_{i}(s,\omega))dL^{*}_{j}(s,\omega) =01f(Zi(s,ω))𝑑Πjω(s)\displaystyle=\int_{0}^{1}f(Z_{i}(s,\omega))d\Pi^{\omega}_{j}(s)
=limε001f(Zi(s,ω))𝑑Πjε,ω(s)=limε01ε01f(Zi(s,ω))1{0Zj(s,ω)ε}𝑑s.\displaystyle=\lim_{{\varepsilon}\to 0}\int_{0}^{1}f(Z_{i}(s,\omega))d\Pi^{{\varepsilon},\omega}_{j}(s)=\lim_{{\varepsilon}\to 0}\frac{1}{{\varepsilon}}\int_{0}^{1}f(Z_{i}(s,\omega))1_{\{0\leq Z_{j}(s,\omega)\leq{\varepsilon}\}}ds.

Using Lemma 5.3 and the fact that 𝒁{\boldsymbol{Z}} is a stationary process with a product form stationary distribution, we now see that for i,ji,j\in{\mathbb{N}}, iji\neq j,

𝔼01f(Zi(s))𝑑Lj(s)\displaystyle{\mathbb{E}}\int_{0}^{1}f(Z_{i}(s))dL^{*}_{j}(s) =limε01ε𝔼01f(Zi(s))1{0Zj(s)ε}𝑑s\displaystyle=\lim_{{\varepsilon}\to 0}\frac{1}{{\varepsilon}}{\mathbb{E}}\int_{0}^{1}f(Z_{i}(s))1_{\{0\leq Z_{j}(s)\leq{\varepsilon}\}}ds
=limε01επj[0,ε]f(z)πi(dz)=νjf(z)πi(dz).\displaystyle=\lim_{{\varepsilon}\to 0}\frac{1}{{\varepsilon}}\pi_{j}[0,{\varepsilon}]\int f(z)\pi_{i}(dz)=\nu_{j}\int f(z)\pi_{i}(dz).

5.3 Proof of Theorem 3.5

We now complete the proof of the theorem. Recalling Lemmas 5.1 and 5.2), taking expectations in (5.1), and using the identity νi=𝔼(Li(1))\nu_{i}={\mathbb{E}}(L^{*}_{i}(1)) for ii\in{\mathbb{N}}, we see that, for all ii\in{\mathbb{N}},

hi+νi12νi+112νi1=0.h_{i}+\nu_{i}-\frac{1}{2}\nu_{i+1}-\frac{1}{2}\nu_{i-1}=0. (5.10)

Applying the above identity for i=1i=1, and setting aν12g0a\doteq\nu_{1}-2g_{0}, we have

ν2=2ν1+2h1=2(g0+g1)+2(ν12g0)=2(g0+g1)+2a.\nu_{2}=2\nu_{1}+2h_{1}=2(g_{0}+g_{1})+2(\nu_{1}-2g_{0})=2(g_{0}+g_{1})+2a.

Proceeding by induction, suppose that for some k2k\geq 2,

νi=ia+2(g0++gi1), for all 1ik.\nu_{i}=ia+2(g_{0}+\cdots+g_{i-1}),\mbox{ for all }1\leq i\leq k. (5.11)

Then from (5.10),

νk+1\displaystyle\nu_{k+1} =2(νk12νk1+hk)\displaystyle=2(\nu_{k}-\frac{1}{2}\nu_{k-1}+h_{k})
=2ak+4(g0++gk1)a(k1)2(g0++gk2)+2(gkgk1)\displaystyle=2ak+4(g_{0}+\cdots+g_{k-1})-a(k-1)-2(g_{0}+\cdots+g_{k-2})+2(g_{k}-g_{k-1})
=(k+1)a+2(g0++gk).\displaystyle=(k+1)a+2(g_{0}+\cdots+g_{k}).

Thus it follows that, for every k1k\geq 1,

νk=ka+2(g0++gk1)=k(a+2g¯k).\nu_{k}=ka+2(g_{0}+\cdots+g_{k-1})=k(a+2\bar{g}_{k}). (5.12)

Fix ii\in{\mathbb{N}} and λ<k=0i1gk\lambda<\sum_{k=0}^{i-1}g_{k}. Then by Itô’s formula applied to ZiZ_{i} given by (5.1),

eλZi(1)\displaystyle e^{\lambda Z_{i}(1)} =eλZi(0)+λhi01eλZi(s)𝑑s+λ01eλZi(s)𝑑Wi(s)λ201eλZi(s)𝑑Li1(s)\displaystyle=e^{\lambda Z_{i}(0)}+\lambda h_{i}\int_{0}^{1}e^{\lambda Z_{i}(s)}ds+\lambda\int_{0}^{1}e^{\lambda Z_{i}(s)}dW_{i}^{*}(s)-\frac{\lambda}{2}\int_{0}^{1}e^{\lambda Z_{i}(s)}dL^{*}_{i-1}(s)
λ201eλZi(s)𝑑Li+1(s)+λ01eλZi(s)𝑑Li(s)+λ201eλZi(s)𝑑s.\displaystyle\quad-\frac{\lambda}{2}\int_{0}^{1}e^{\lambda Z_{i}(s)}dL^{*}_{i+1}(s)+\lambda\int_{0}^{1}e^{\lambda Z_{i}(s)}dL^{*}_{i}(s)+\lambda^{2}\int_{0}^{1}e^{\lambda Z_{i}(s)}ds. (5.13)

Since λ<k=0i1gk\lambda<\sum_{k=0}^{i-1}g_{k}, from Lemma 5.1, 01𝔼e2λZi(s)𝑑s=+e2λzπi(dz)<\int_{0}^{1}{\mathbb{E}}e^{2\lambda Z_{i}(s)}ds=\int_{{\mathbb{R}}_{+}}e^{2\lambda z}\pi_{i}(dz)<\infty and consequently the stochastic integral in the above display has mean 0. Moreover, note that

𝔼01eλZi(s)𝑑Li(s)=𝔼01𝑑Li(s)=𝔼(Li(1))=νi.{\mathbb{E}}\int_{0}^{1}e^{\lambda Z_{i}(s)}dL^{*}_{i}(s)={\mathbb{E}}\int_{0}^{1}dL^{*}_{i}(s)={\mathbb{E}}(L^{*}_{i}(1))=\nu_{i}.

Thus taking expectations in (5.3) and using Lemma 5.4, we have,

+eλzπi(dz)\displaystyle\int_{{\mathbb{R}}_{+}}e^{\lambda z}\pi_{i}(dz) =+eλzπi(dz)+λhi+eλzπi(dz)λ2νi1+eλzπi(dz)\displaystyle=\int_{{\mathbb{R}}_{+}}e^{\lambda z}\pi_{i}(dz)+\lambda h_{i}\int_{{\mathbb{R}}_{+}}e^{\lambda z}\pi_{i}(dz)-\frac{\lambda}{2}\nu_{i-1}\int_{{\mathbb{R}}_{+}}e^{\lambda z}\pi_{i}(dz)
λ2νi+1+eλzπi(dz)+λνi+λ2+eλzπi(dz).\displaystyle\quad-\frac{\lambda}{2}\nu_{i+1}\int_{{\mathbb{R}}_{+}}e^{\lambda z}\pi_{i}(dz)+\lambda\nu_{i}+\lambda^{2}\int_{{\mathbb{R}}_{+}}e^{\lambda z}\pi_{i}(dz).

Rearranging terms, we have

λνi=(λhi+λ2νi1+λ2νi+1λ2)+eλzπi(dz)=(λνiλ2)+eλzπi(dz),\lambda\nu_{i}=\left(-\lambda h_{i}+\frac{\lambda}{2}\nu_{i-1}+\frac{\lambda}{2}\nu_{i+1}-\lambda^{2}\right)\int_{{\mathbb{R}}_{+}}e^{\lambda z}\pi_{i}(dz)=(\lambda\nu_{i}-\lambda^{2})\int_{{\mathbb{R}}_{+}}e^{\lambda z}\pi_{i}(dz),

where the second equality in the above display follows from (5.10). Thus we have shown that for all ii\in{\mathbb{N}} and λ<k=0i1gk\lambda<\sum_{k=0}^{i-1}g_{k},

+eλzπi(dz)=νiνiλ.\int_{{\mathbb{R}}_{+}}e^{\lambda z}\pi_{i}(dz)=\frac{\nu_{i}}{\nu_{i}-\lambda}.

Thus, since k=0i1gk=ig¯i>0\sum_{k=0}^{i-1}g_{k}=i\bar{g}_{i}>0, by uniqueness of Laplace transforms, we must have that πi=Exp(νi)\pi_{i}=\mbox{Exp}(\nu_{i}) for each ii\in{\mathbb{N}}. Finally note that, since by Lemma 5.1, +eλzπ1(dz)<\int_{{\mathbb{R}}_{+}}e^{\lambda z}\pi_{1}(dz)<\infty for all λ<2g0\lambda<2g_{0}, we must have that a=ν12g00a=\nu_{1}-2g_{0}\geq 0. Thus, in view of (5.12), we have shown that, for some a0a\geq 0, πi=Exp(νi)=Exp(i(2g¯i+a))\pi_{i}=\mbox{Exp}(\nu_{i})=\mbox{Exp}(i(2\bar{g}_{i}+a)) for all ii\in{\mathbb{N}} and so π=πa𝒈\pi=\pi_{a}^{{\boldsymbol{g}}} for some a0a\geq 0.

The assertion (3.3) follows from (5.12) upon recalling that νk=𝔼(Lk(1))\nu_{k}={\mathbb{E}}(L^{*}_{k}(1)) for kk\in{\mathbb{N}}.

Appendix A Proof of Lemma 3.2

We will like to acknowledge the lecture notes of Sethuraman [36] which are used at several steps in the proof below.

Since 𝒈𝒟{\boldsymbol{g}}\in{\mathcal{D}} will be fixed in the proof we suppress it from the notation. Let γ\gamma\in{\mathcal{I}}.

Define

𝔾{ηTtη:ηL2(γ),t0}.{\mathbb{G}}\doteq\{\eta-T_{t}\eta:\eta\in L^{2}(\gamma),\,t\geq 0\}.

We claim that [(span(𝔾))cl]𝕀γ[(\mbox{span}({\mathbb{G}}))^{cl}]^{\perp}\subset{\mathbb{I}}_{\gamma}, where [(span(𝔾))cl][(\mbox{span}({\mathbb{G}}))^{cl}]^{\perp} denotes the orthogonal complement of the closure of the linear subspace generated by 𝔾{\mathbb{G}} in L2(γ)L^{2}(\gamma). Indeed, if ψ[(span(𝔾))cl]\psi\in[(\mbox{span}({\mathbb{G}}))^{cl}]^{\perp}, then for all ηL2(γ)\eta\in L^{2}(\gamma) and t0t\geq 0, ψ,ηTtη=0\langle\psi,\eta-T_{t}\eta\rangle=0. Taking η=ψ\eta=\psi,

ψ,ψ=ψ,Ttψ for all t0.\langle\psi,\psi\rangle=\langle\psi,T_{t}\psi\rangle\mbox{ for all }t\geq 0. (A.1)

Using this, and the contraction property of TtT_{t}, for all t0t\geq 0,

0\displaystyle 0 Ttψψ,Ttψψ=Ttψ,Ttψ+ψ,ψ2ψ,Ttψ\displaystyle\leq\langle T_{t}\psi-\psi,T_{t}\psi-\psi\rangle=\langle T_{t}\psi,T_{t}\psi\rangle+\langle\psi,\psi\rangle-2\langle\psi,T_{t}\psi\rangle
=Ttψ,Ttψ+ψ,ψ2ψ,ψ=Ttψ,Ttψψ,ψ0,\displaystyle=\langle T_{t}\psi,T_{t}\psi\rangle+\langle\psi,\psi\rangle-2\langle\psi,\psi\rangle=\langle T_{t}\psi,T_{t}\psi\rangle-\langle\psi,\psi\rangle\leq 0,

where the first equality on the second line follows from (A.1). This says that Ttψ=ψT_{t}\psi=\psi (as elements of L2(γ)L^{2}(\gamma)) for all t0t\geq 0 and so ψ𝕀γ\psi\in{\mathbb{I}}_{\gamma} and shows the claim [(span(𝔾))cl]𝕀γ[(\mbox{span}({\mathbb{G}}))^{cl}]^{\perp}\subset{\mathbb{I}}_{\gamma}. Any ψL2(γ)\psi\in L^{2}(\gamma) can be written as ψ=ψ^γ+ψ~γ\psi=\hat{\psi}_{\gamma}+\tilde{\psi}_{\gamma} where ψ^γ[(span(𝔾))cl]𝕀γ\hat{\psi}_{\gamma}\in[(\mbox{span}({\mathbb{G}}))^{cl}]^{\perp}\subset{\mathbb{I}}_{\gamma} and ψ~γ(span(𝔾))cl\tilde{\psi}_{\gamma}\in(\mbox{span}({\mathbb{G}}))^{cl}.

Next, for ψL2(γ)\psi\in L^{2}(\gamma) and t>0t>0, define AtψL2(γ)A_{t}\psi\in L^{2}(\gamma) as

Atψ1t0tTsψ𝑑s.A_{t}\psi\doteq\frac{1}{t}\int_{0}^{t}T_{s}\psi\,ds.

Then Atψ=Atψ^γ+Atψ~γA_{t}\psi=A_{t}\hat{\psi}_{\gamma}+A_{t}\tilde{\psi}_{\gamma}. By definition Atψ^γ=ψ^γA_{t}\hat{\psi}_{\gamma}=\hat{\psi}_{\gamma}. Also, if ϕ𝔾\phi\in{\mathbb{G}} then for some t00t_{0}\geq 0 and ηL2(γ)\eta\in L^{2}(\gamma) ϕ=ηTt0η\phi=\eta-T_{t_{0}}\eta. Then from the contraction property of TtT_{t} it follows that

Atϕ2ηt0t0 as t.\|A_{t}\phi\|\leq\frac{2\|\eta\|t_{0}}{t}\to 0\mbox{ as }t\to\infty.

Similarly, if ϕspan(𝔾)\phi\in\mbox{span}({\mathbb{G}}), there is a c(ϕ)(0,)c(\phi)\in(0,\infty) such that

Atϕc(ϕ)t0 as t.\|A_{t}\phi\|\leq\frac{c(\phi)}{t}\to 0\mbox{ as }t\to\infty.

Finally, let ϕ(span(𝔾))cl\phi\in(\mbox{span}({\mathbb{G}}))^{cl}. Then, given ε>0{\varepsilon}>0, there is a ϕεspan(𝔾)\phi^{{\varepsilon}}\in\mbox{span}({\mathbb{G}}) such that ϕϕεε\|\phi-\phi^{{\varepsilon}}\|\leq{\varepsilon}. It follows, using again the contraction property, that for all t>0t>0,

AtϕAt(ϕϕε)+Atϕεϕϕε+c(ϕε)tε+c(ϕε)t.\|A_{t}\phi\|\leq\|A_{t}(\phi-\phi^{{\varepsilon}})\|+\|A_{t}\phi^{{\varepsilon}}\|\leq\|\phi-\phi^{{\varepsilon}}\|+\frac{c(\phi^{{\varepsilon}})}{t}\leq{\varepsilon}+\frac{c(\phi^{{\varepsilon}})}{t}.

Thus lim suptAtϕε\limsup_{t\to\infty}\|A_{t}\phi\|\leq{\varepsilon}. Since ε>0{\varepsilon}>0 is arbitrary, we obtain

limtAtϕ=0 for all ϕ(span(𝔾))cl.\lim_{t\to\infty}\|A_{t}\phi\|=0\mbox{ for all }\phi\in(\mbox{span}({\mathbb{G}}))^{cl}. (A.2)

From these observations we obtain, for any ψL2(γ)\psi\in L^{2}(\gamma),

limtAtψ=limt(Atψ^γ+Atψ~γ)=limt(ψ^γ+Atψ~γ)=ψ^γ.\lim_{t\to\infty}A_{t}\psi=\lim_{t\to\infty}(A_{t}\hat{\psi}_{\gamma}+A_{t}\tilde{\psi}_{\gamma})=\lim_{t\to\infty}(\hat{\psi}_{\gamma}+A_{t}\tilde{\psi}_{\gamma})=\hat{\psi}_{\gamma}. (A.3)

The above convergence in fact shows that [(span(𝔾))cl]=𝕀γ[(\mbox{span}({\mathbb{G}}))^{cl}]^{\perp}={\mathbb{I}}_{\gamma} and consequently ψ^γ\hat{\psi}_{\gamma} is the projection of ψ\psi on to 𝕀γ{\mathbb{I}}_{\gamma}. To see this, recall that it was argued above that [(span(𝔾))cl]𝕀γ[(\mbox{span}({\mathbb{G}}))^{cl}]^{\perp}\subset{\mathbb{I}}_{\gamma}. Now consider the reverse inclusion and let φ𝕀γ\varphi\in{\mathbb{I}}_{\gamma}. Then we can write φ=φ1+φ2\varphi=\varphi_{1}+\varphi_{2} where φ1(span(𝔾))cl\varphi_{1}\in(\mbox{span}({\mathbb{G}}))^{cl} and φ2[(span(𝔾))cl]𝕀γ\varphi_{2}\in[(\mbox{span}({\mathbb{G}}))^{cl}]^{\perp}\subset{\mathbb{I}}_{\gamma}. Thus, for t0t\geq 0,

φ=Atφ=Atφ1+Atφ2=Atφ1+φ2.\varphi=A_{t}\varphi=A_{t}\varphi_{1}+A_{t}\varphi_{2}=A_{t}\varphi_{1}+\varphi_{2}.

As tt\to\infty, we have from (A.2) that, Atφ10A_{t}\varphi_{1}\to 0, which says that φ=φ2\varphi=\varphi_{2}. This proves the inclusion 𝕀γ[(span(𝔾))cl]{\mathbb{I}}_{\gamma}\subset[(\mbox{span}({\mathbb{G}}))^{cl}]^{\perp} and we have the claimed statement [(span(𝔾))cl]=𝕀γ[(\mbox{span}({\mathbb{G}}))^{cl}]^{\perp}={\mathbb{I}}_{\gamma}.

Now we proceed to the proof of the statements in the lemma. We first consider the second statement in the lemma. Fix γ\gamma\in{\mathcal{I}}. Suppose that for every bounded measurable map ψ:+\psi:{\mathbb{R}}_{+}^{\infty}\to{\mathbb{R}}, ψ^γ\hat{\psi}_{\gamma} is constant γ\gamma a.s. We will now show that this implies γe\gamma\in{\mathcal{I}}_{e}. Suppose there is a ε(0,1){\varepsilon}\in(0,1) and γ1,γ2\gamma_{1},\gamma_{2}\in{\mathcal{I}} such that γ=εγ1+(1ε)γ2\gamma={\varepsilon}\gamma_{1}+(1-{\varepsilon})\gamma_{2}. Note that since from (A.3) ψ^γ=limtAtψ\hat{\psi}_{\gamma}=\lim_{t\to\infty}A_{t}\psi and γ\gamma is invariant, we must have ψ^γ=+ψ𝑑γ\hat{\psi}_{\gamma}=\int_{{\mathbb{R}}_{+}^{\infty}}\psi\,d\gamma, and so from (A.3) it follows that

+(Atψψ𝑑γ)2𝑑γ=Atψψ^γ20 as t.\int_{{\mathbb{R}}_{+}^{\infty}}\left(A_{t}\psi-\int\psi d\gamma\right)^{2}d\gamma=\|A_{t}\psi-\hat{\psi}_{\gamma}\|^{2}\to 0\mbox{ as }t\to\infty. (A.4)

Also, from definition,

lim supt+(Atψψ𝑑γ)2𝑑γ1lim suptε1+(Atψψ𝑑γ)2𝑑γ=0.\limsup_{t\to\infty}\int_{{\mathbb{R}}_{+}^{\infty}}\left(A_{t}\psi-\int\psi d\gamma\right)^{2}d\gamma_{1}\leq\limsup_{t\to\infty}{\varepsilon}^{-1}\int_{{\mathbb{R}}_{+}^{\infty}}\left(A_{t}\psi-\int\psi d\gamma\right)^{2}d\gamma=0. (A.5)

Thus Atψψ𝑑γA_{t}\psi\to\int\psi d\gamma in L2(γ1)L^{2}(\gamma_{1}). Also, since γ1\gamma_{1}\in{\mathcal{I}}, Atψ𝑑γ1=ψ𝑑γ1\int A_{t}\psi d\gamma_{1}=\int\psi d\gamma_{1} and consequently, ψ𝑑γ1=ψ𝑑γ\int\psi d\gamma_{1}=\int\psi d\gamma. Since ψ\psi is an arbitrary bounded measurable function, we must have γ=γ1\gamma=\gamma_{1}. This proves that γe\gamma\in{\mathcal{I}}_{e}. We have thus shown the second statement in the lemma and in fact also shown that ere{\mathcal{I}}_{er}\subset{\mathcal{I}}_{e}.

Finally we argue that eer{\mathcal{I}}_{e}\subset{\mathcal{I}}_{er}. Suppose γe\gamma\in{\mathcal{I}}_{e} and that γer\gamma\not\in{\mathcal{I}}_{er}. Then there is a ψL2(γ)\psi\in L^{2}(\gamma) such that ψ^γ\hat{\psi}_{\gamma} is not a.s. constant under γ\gamma. Thus there is a cc\in{\mathbb{R}} such that, with A={ψ^γ>c}A=\{\hat{\psi}_{\gamma}>c\}, γ(A)ε(0,1)\gamma(A)\doteq{\varepsilon}\in(0,1). Note that by definition

Ttψ^γ=ψ^γ,γ a.s.  for all t0.T_{t}\hat{\psi}_{\gamma}=\hat{\psi}_{\gamma},\;\gamma\mbox{ a.s. }\mbox{ for all }t\geq 0. (A.6)

We refer to this property as ψ^γ\hat{\psi}_{\gamma} is harmonic (with respect to the semigroup {Tt}\{T_{t}\}). We claim that this implies that 1A1_{A} is harmonic as well, namely

Tt1A=1A,γ a.s.  for all t0.T_{t}1_{A}=1_{A},\,\gamma\mbox{ a.s. }\mbox{ for all }t\geq 0. (A.7)

To see this, note that, from the definition of TtT_{t}, if fL2(γ)f\in L^{2}(\gamma) is harmonic, then

|f|=|Ttf|Tt|f|,γ a.s. |f|=|T_{t}f|\leq T_{t}|f|,\;\gamma\mbox{ a.s. } (A.8)

This, together with the fact that TtT_{t} is a contraction, says that

fTt|f|f\|f\|\leq\|T_{t}|f|\|\leq\|f\|

which in view of (A.8) shows that |f||f| is harmonic. From the linearity of TtT_{t} it then follows that f0=12(f+|f|)f\vee 0=\frac{1}{2}(f+|f|) is harmonic as well. This implies that if f1,f2f_{1},f_{2} are harmonic then f1f2f_{1}\vee f_{2} and f1f2f_{1}\wedge f_{2} are harmonic as well. Recalling tha ψ^γ\hat{\psi}_{\gamma} is harmonic, we now have that gn(n(ψ^γc)+1)g_{n}\doteq\left(n\left(\hat{\psi}_{\gamma}-c\right)^{+}\wedge 1\right) is harmonic for every nn\in\mathbb{N}. The property in (A.7) is now immediate from this on observing that gn1Ag_{n}\to 1_{A} a.s. and dominated convergence theorem. This proves the claim.

Consider the probability measures γ1,γ2\gamma_{1},\gamma_{2} on (+,(+))({\mathbb{R}}_{+}^{\infty},{\mathcal{B}}({\mathbb{R}}_{+}^{\infty})) defined as

γ1(B)ε1γ(BA),γ2(B)(1ε)1γ(BAc),B(+).\gamma_{1}(B)\doteq{\varepsilon}^{-1}\gamma(B\cap A),\;\;\gamma_{2}(B)\doteq(1-{\varepsilon})^{-1}\gamma(B\cap A^{c}),\;\;B\in{\mathcal{B}}({\mathbb{R}}_{+}^{\infty}).

Using (A.7) it is easily seen that γ1,γ2\gamma_{1},\gamma_{2}\in{\mathcal{I}}. Indeed, if B(+)B\in{\mathcal{B}}({\mathbb{R}}_{+}^{\infty}) and t0t\geq 0,

+Tt1B𝑑γ1\displaystyle\int_{{\mathbb{R}}_{+}^{\infty}}T_{t}1_{B}\,d\gamma_{1} =ε1ATt1B𝑑γ=ε1ATt1AB𝑑γ+ε1ATt1AcB𝑑γ\displaystyle={\varepsilon}^{-1}\int_{A}T_{t}1_{B}\,d\gamma={\varepsilon}^{-1}\int_{A}T_{t}1_{AB}\,d\gamma+{\varepsilon}^{-1}\int_{A}T_{t}1_{A^{c}B}\,d\gamma
=ε1ATt1AB𝑑γ=ε1+Tt1AB𝑑γ=ε1γ(BA)=γ1(B),\displaystyle={\varepsilon}^{-1}\int_{A}T_{t}1_{AB}\,d\gamma={\varepsilon}^{-1}\int_{{\mathbb{R}}_{+}^{\infty}}T_{t}1_{AB}\,d\gamma={\varepsilon}^{-1}\gamma(B\cap A)=\gamma_{1}(B),

where the third and fourth equalities follow from (A.7) and the fifth uses the invariance of γ\gamma. This shows the invariance of γ1\gamma_{1} from which (together with the fact that γ\gamma is invariant) the invariance of γ2\gamma_{2} follows immediately.

Finally note that γ1γ2\gamma_{1}\neq\gamma_{2}, and by definition γ=εγ1+(1ε)γ2\gamma={\varepsilon}\gamma_{1}+(1-{\varepsilon})\gamma_{2}. This contradicts the fact that γe\gamma\in{\mathcal{I}}_{e} and thus we must have γer\gamma\in{\mathcal{I}}_{er}. We have thus shown that eer{\mathcal{I}}_{e}\subset{\mathcal{I}}_{er} which completes the proof. ∎

Acknowledgements: Research supported in part by the RTG award (DMS-2134107) from the NSF. SB was supported in part by the NSF-CAREER award (DMS-2141621). AB was supported in part by the NSF (DMS-2152577). We thank two anonymous referees whose valuable inputs significantly improved the article.

Data availability statement: This manuscript has no associated data.

References

  • [1] David Aldous. “Up the river” game story. 2002. Download available from: http://www. stat. berkeley. edu/~ aldous/Research/OP/river. pdf, 41:42–47.
  • [2] Enrique Daniel Andjel. Invariant measures for the zero range process. The Annals of Probability, 10(3):525–547, 1982.
  • [3] Márton Balázs, Firas Rassoul-Agha, Timo Seppäläinen, and S. Sethuraman. Existence of the zero range process and a deposition model with superlinear growth rates. The Annals of Probability, 35(4):1201–1249, 2007.
  • [4] Sayan Banerjee and Brendan Brown. Dimension-free local convergence and perturbations for reflected Brownian motions. Annals of Applied Probability (to appear), 2022.
  • [5] Sayan Banerjee and Amarjit Budhiraja. Parameter and dimension dependence of convergence rates to stationarity for reflecting Brownian motions. Annals of Applied Probability, 30(5):2005–2029, 2020.
  • [6] Sayan Banerjee and Amarjit Budhiraja. Domains of attraction of invariant distributions of the infinite Atlas model. The Annals of Probability, 50(4):1610–1646, 2022.
  • [7] Adrian D Banner, Robert Fernholz, and Ioannis Karatzas. Atlas models of equity markets. The Annals of Applied Probability, 15(4):2296–2330, 2005.
  • [8] Richard F Bass and Etienne Pardoux. Uniqueness for diffusions with piecewise constant coefficients. Probability Theory and Related Fields, 76(4):557–572, 1987.
  • [9] Amarjit Budhiraja and Chihoon Lee. Long time asymptotics for constrained diffusions in polyhedral domains. Stochastic processes and their applications, 117(8):1014–1036, 2007.
  • [10] Manuel Cabezas, Amir Dembo, Andrey Sarantsev, and Vladas Sidoravicius. Brownian particles with rank-dependent drifts: Out-of-equilibrium behavior. Communications on Pure and Applied Mathematics, 72(7):1424–1458, 2019.
  • [11] Amir Dembo, Milton Jara, and Stefano Olla. The infinite Atlas process: Convergence to equilibrium. Annales de l’Institut Henri Poincaré, Probabilités et Statistiques, 55(2):607–619, 2019.
  • [12] Amir Dembo, Mykhaylo Shkolnikov, SR Srinivasa Varadhan, and Ofer Zeitouni. Large deviations for diffusions interacting through their ranks. Communications on Pure and Applied Mathematics, 69(7):1259–1313, 2016.
  • [13] E Robert Fernholz. Stochastic portfolio theory. In Stochastic portfolio theory, pages 1–24. Springer, 2002.
  • [14] Robert Fernholz and Ioannis Karatzas. Stochastic portfolio theory: an overview. Handbook of numerical analysis, 15:89–167, 2009.
  • [15] J Michael Harrison and Ruth J Williams. Brownian models of open queueing networks with homogeneous customer populations. Stochastics: An International Journal of Probability and Stochastic Processes, 22(2):77–115, 1987.
  • [16] J Michael Harrison and Ruth J Williams. Multidimensional reflected Brownian motions having exponential stationary distributions. The Annals of Probability, pages 115–137, 1987.
  • [17] Tomoyuki Ichiba and Ioannis Karatzas. On collisions of Brownian particles. Annals of Applied Probability, 20(3):951–977, 2010.
  • [18] Tomoyuki Ichiba, Ioannis Karatzas, and Mykhaylo Shkolnikov. Strong solutions of stochastic equations with rank-based coefficients. Probability Theory and Related Fields, 156(1-2):229–248, 2013.
  • [19] Tomoyuki Ichiba, Soumik Pal, and Mykhaylo Shkolnikov. Convergence rates for rank-based models with applications to portfolio theory. Probability Theory and Related Fields, 156(1-2):415–448, 2013.
  • [20] Tomoyuki Ichiba and Andrey Sarantsev. Yet another condition for absence of collisions for competing Brownian particles. Electronic Communications in Probability, 22, 2017.
  • [21] Benjamin Jourdain and Florent Malrieu. Propagation of chaos and Poincaré inequalities for a system of particles interacting through their CDF. Annals of Applied Probability, 18(5):1706–1736, 2008.
  • [22] Shizuo Kakutani. On equivalence of infinite product measures. Annals of Mathematics, pages 214–224, 1948.
  • [23] Olav Kallenberg. Foundations of modern probability, volume 2. Springer, 1997.
  • [24] Ioannis Karatzas, Soumik Pal, and Mykhaylo Shkolnikov. Systems of Brownian particles with asymmetric collisions. Annales de l’Institut Henri Poincaré, Probabilités et Statistiques, 52(1):323–354, 2016.
  • [25] Ioannis Karatzas and Steven Shreve. Brownian Motion and Stochastic Calculus, volume 113. Springer Science & Business Media, 2012.
  • [26] Thomas M Liggett. Coupling the simple exclusion process. The Annals of Probability, pages 339–356, 1976.
  • [27] Soumik Pal and Jim Pitman. One-dimensional Brownian particle systems with rank-dependent drifts. The Annals of Applied Probability, 18(6):2179–2207, 2008.
  • [28] Philip E Protter. Stochastic Integration and Differential Equations. Springer, 2005.
  • [29] Daniel Revuz and Marc Yor. Continuous Martingales and Brownian Motion, volume 293. Springer Science & Business Media, 2013.
  • [30] Anastasia Ruzmaikina and Michael Aizenman. Characterization of invariant measures at the leading edge for competing particle systems. The Annals of Probability, 33(1):82–113, 2005.
  • [31] Andrey Sarantsev. Triple and simultaneous collisions of competing Brownian particles. Electronic journal of probability, 20, 2015.
  • [32] Andrey Sarantsev. Infinite systems of competing Brownian particles. Annales de l’Institut Henri Poincaré, Probabilités et Statistiques, 53(4):2279–2315, 2017.
  • [33] Andrey Sarantsev. Two-sided infinite systems of competing brownian particles. ESAIM: Probability and Statistics, 21:317–349, 2017.
  • [34] Andrey Sarantsev. Comparison techniques for competing Brownian particles. Journal of Theoretical Probability, 32(2):545–585, 2019.
  • [35] Andrey Sarantsev and Li-Cheng Tsai. Stationary gap distributions for infinite systems of competing Brownian particles. Electronic Journal of Probability, 22, 2017.
  • [36] S. Sethuraman. Lecture notes: Invariant measures: Ergodicity and extremality.
  • [37] Sunder Sethuraman. On extremal measures for conservative particle systems. Annales de l’Institut Henri Poincaré, Probabilités et Statistiques, 37(2):139–154, 2001.
  • [38] Mykhaylo Shkolnikov. Competing particle systems evolving by interacting Lévy processes. The Annals of Applied Probability, 21(5):1911–1932, 2011.
  • [39] Mykhaylo Shkolnikov. Large systems of diffusions interacting through their ranks. Stochastic Processes and their Applications, 122(4):1730–1747, 2012.
  • [40] Wenpin Tang and Li-Cheng Tsai. Optimal surviving strategy for drifted Brownian motions with absorption. The Annals of Probability, 46(3):1597–1650, 2018.
  • [41] Li-Cheng Tsai. Stationary distributions of the Atlas model. Electronic Communications in Probability, 23:1 – 10, 2018.

S. Banerjee and A. Budhiraja,
Department of Statistics and Operations Research
University of North Carolina
Chapel Hill, NC 27599, USA
email: [email protected]
email: [email protected]