Extremal Invariant Distributions of Infinite Brownian Particle Systems with Rank Dependent Drifts
Abstract
Consider an infinite collection of particles on the real line moving according to independent Brownian motions and such that the -th particle from the left gets the drift . The case where and for all corresponds to the well studied infinite Atlas model. Under conditions on the drift vector 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 where 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 for some . 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 -th particle from the left gets a constant drift . The special case when and for is the well studied Atlas model. We refer to the general setting as the -Atlas model, where . 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 , 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 to this stationary distribution, as , occurs at a geometric rate [9]. Rates of convergence to stationarity, depending explicitly on the drift vector 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 , the process describing the gaps between the ranked particles in the system has a simple product form stationary distribution given as (here and later, for , denotes the Exponential distribution with mean ). 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
(1.1) |
As in the finite dimensional settings it is of interest to investigate convergence of the laws at time to the stationary distributions as . 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 have been obtained in [32, 11, 6] and (weak) domains of attraction of () 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 -Atlas model where the drift vector , with
(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
(1.3) |
where . In the special case where , with
(1.4) |
is also an invariant distribution for (see [32, Section 4.2]). Note that whereas the zero drift is in but not in . Roughly speaking, a drift lying in produces a ‘stabilizing interaction’ in the subsystem of the lowest particles for any , 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 particles. This intrinsic stabilizing influence from the drift of the particles leads to an additional stationary distribution, namely (in comparison to a ). For , 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 .
Using Kakutani’s theorem [22] it is easy to verify that, for different values of , the probability measures 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 and , is an extremal invariant distribution for the gap sequence process of the -Atlas model. Further, if , is also an extremal invariant distribution for ; in particular for the infinite Atlas model is extremal for all . 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 for each ( if ); 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. ) 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 -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 and is a product form stationary distribution of the -Atlas model satisfying a mild integrability condition (see (3.2)) then it must be for some . Furthermore, this result gives a novel probabilistic interpretation to 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 is to establish that any bounded measurable function on that is -a.e. invariant, under the action of the semigroup of the Markov process corresponding to the -Atlas gap sequence, is constant -a.e. If and , we have that , and therefore the coordinate sequence is iid under . In this case, from the Hewitt-Savage zero-one law it suffices to show that is -a.e. invariant under all finite permutations of the coordinates of . For this, in turn, it suffices to simply prove the above invariance property for transpositions of the -th and -th coordinates, for all . For a general , the situation is more involved as the coordinate sequence is not iid any more. Nevertheless, from the scaling properties of Exponential distributions it follows that, with , the sequence , defined as , , is iid under . In this case, in order to invoke the Hewitt-Savage zero-one law, one needs to argue that for each the map is -a.e. invariant under the transformation that takes the coordinates to 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 Brownian motions, and synchronous coupling of the remaining Brownian motions, in the evolution of the ranked infinite -Atlas model corresponding to a pair of nearby initial configurations. Estimates on the probability of a successful coupling, before any of the first -gap processes have hit zero or the lowest -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 , of certain integrals involving the state process of the -th gap and the collision local time for the -th and -th particle, for (see Lemma 5.4). This identity is a consequence of the product form structure of and basic results on local times of continuous semimartingales. One subtlety here is that although the product form structure of an invariant measure implies that the laws of the various gaps at any fixed time (when the process is initiated at ) 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 -th gap at time and the local time of the -th gap at , at time , for . By using the form of the dynamics of the -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 that can then be solved explicitly from which it is then readily seen that must be for a suitable value of . 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 -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 -Atlas model. For the cases where 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 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 has an intensity function , that grows exponentially as when and, due to a nonlinear dependence on , lacks the scalar multiple property for distinct values of . 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 . However their setting, in the context of rank based diffusions, corresponds to the case , 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 , and . We will equip with the partial order ‘’ under which for , , if for all . Borel -fields on a metric space will be denoted as and the space of probability measures on will be denoted as which is equipped with the topology of weak convergence. We will denote by the set of rationals. For a Polish space with a partial order ‘’, we say for , , if there are -valued random variables given on a common probability space with distributed as , , and a.s. Let which is equipped with the topology of local uniform convergence (with equipped with the product topology). For , let be the collection of all measurable such that . We denote the inner-product and the norm on as and respectively.
2 Model Formulation
Let
Recall defined in (1.2). Following [32], an is called rankable if there is a one-to-one map from onto itself such that whenever , . It is easily seen that any is locally finite and hence rankable. For an that is rankable we denote the unique permutation map as above which breaks ties in the lexicographic order by . For a sequence of mutually independent standard Brownian motions given on some filtered probability space, and , , consider the following system of equations.
(2.1) |
, , where for , .
The following result is from [32] (see Theorems 3.2 and Lemma 3.4 therein).
Theorem 2.1 ([32]).
For every and there is a unique weak solution to (2.1). Furthermore and for any and , the set is finite a.s.
When 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 -infinite Atlas model. Since this solution process , under the conditions of the above theorem, is rankable a.s., we can define the ranked process that gives the unique ranking of (in which ties are broken in the lexicographic order) such that . From [32, Lemma 3.5], the processes defined by
(2.2) |
are independent standard Brownian motions which can be used to write down the following stochastic differential equation (SDE) for :
(2.3) |
Here, and for , denotes the local time of collision between the -th and -th particles, that is, the unique non-decreasing continuous process satisfying and for all . The gap process for the -infinite Atlas model is the -valued process defined by
(2.4) |
Let
(2.5) |
Note that if then and if then . Thus given we can define a unique in law stochastic process with values in that can be viewed as a -valued Markov process referred to as the gap process of the -Atlas model. The Markov property of 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 defined in (1.4).
Theorem 2.2 ([35, 32]).
Let . Define for , . Then for each , the probability measure on defined as
is a stationary distribution for the gap process of the -Atlas model. Furthermore, if , the above statement holds also for the case . In particular, when (infinite Atlas model), is a stationary distribution for the gap process for all .
The existence of the limit and the equality follow from the definition of (see first paragraph of Section 5).
As an immediate consequence of Kakutani’s theorem [22] we see that for any and (and when ), , the measures and are mutually singular.
Remark 2.3 (Markov property of the gap process).
The ranked process is formally constructed in [32] as an ‘approximative version’ using limits of finite-dimensional reflected diffusions. Namely, for any fixed , is obtained as an almost sure limit as , uniformly over compact time intervals, of the first coordinates of the reflected diffusion given by the SDE (2.3) with , , and a given collection of iid standard Brownian motions (see also Definition 4.4). Fix any time . Define for any the process similarly by setting and driven by the Brownian motions via the SDE (2.3) (again with the local time for set to zero). Define , , the gap process sequence associated with , . Fix , , and any bounded continuous function . Let and . For any and , as is independent of ,
Thus, to deduce the Markov property, it suffices to show that converges almost surely to as which will follow from the a.s. convergence of to . The latter can be shown by exploiting monotonicity properties of rank-based diffusions [34, Theorem 3.2 and Corollary 3.9] as follows.
Fix . By the construction of the process in [32], we can find (random) such that, for any , , for . Now fix . Note that for by [34, Corollary 3.9], and hence by [34, Theorem 3.2],
To construct a lower bounding process for , define for the process started with and driven by the Brownian motions via the SDE (2.3) (and again setting the local time for to be zero). By the construction of the process in [32], one can choose large enough so that for . Moreover, if we consider the translation of the system by , then the -particle process started from at time and driven by the Brownian motions has particle locations at time given by . Hence, using [34, Theorem 3.2 and Corollary 3.9], we have
(2.6) |
The first inequality above holds because . We conclude from the above two displays that
(2.7) |
This gives the desired almost sure convergence as is arbitrary.
3 Main Results
We are interested in the extremality properties of the probability measures . We also ask whether these are the only product form stationary distributions.
We begin with some notation. Recall . Define measurable maps from to itself as
Given and , we denote the probability distribution of the gap process of the -Atlas model on , with initial gap sequence , by . Also, for supported on (namely, ), let . The corresponding expectation operators will be denoted as and respectively. Denote by the collection of all invariant (stationary) probability measures of the gap process of the -Atlas model supported on , namely
Abusing notation, the canonical coordinate process on will be denoted by . Let be the collection of all real measurable maps on . For , and such that we write
(3.1) |
Note that for , , and , is a.e. well defined and belongs to . Furthermore, the collection defines a contraction semigroup on , namely
We now recall the definition of extremality and ergodicity. Let, for as above and , be the collection of all -invariant square integrable functions, namely,
We denote the projection of a on to the closed subspace as . Namely, is the unique element of that satisfies
This projection can be obtained as the limit of in as (see (A.3)). Thus, for any , can be intuitively interpreted as the ‘long-time average’ of .
Definition 3.1.
Let . A is said to be an extremal invariant distribution of the gap process of the -Atlas model if, whenever for some and we have , then . We denote the collection of all such measures by .
We call an ergodic probability measure for the invariant distribution of the gap process of the -Atlas model if for all , is constant -a.s. We denote the collection of all such measures by .
We note that (cf. proof of Lemma 3.2 below) if , then for any ,
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 . Then . Let and suppose that for every bounded measurable , is constant, a.s. Then .
The following is the first main result of this work.
Theorem 3.3.
Let . Then, for every , . Furthermore, when , also for .
The above theorem proves the extremality of the invariant measures for suitable values of . 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 , as noted in the corollary below. We recall that a measure is said to be in the weak (resp. strong) domain of attraction of if for any bounded continuous function ,
(resp. ), as .
Corollary 3.4.
Let . Fix any . Let be absolutely continuous with respect to . Then lies in the weak domain of attraction of . The assertion holds for any if .
Sufficient conditions for a probability measure to be in the strong domain of attraction of were obtained in [32, 11] whereas weak domain of attraction results for , , have been obtained in [6]. The above corollary provides a weak domain of attraction result for a general class of -Atlas models.
One can ask whether these are the only extremal invariant measures of the gap process of the -Atlas model supported on . As noted in the Introduction, the answer to this question when 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 (in fact even for ) this is currently a challenging open problem. However, we make partial progress towards this goal in the next result by showing that for any (and under a mild integrability condition), the collection exhausts all the extremal product form invariant distributions. In fact we prove the substantially stronger statement that the measures 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 and let be a product measure, i.e. for some , . With , suppose that
(3.2) |
Then, for some , . Moreover, has the representation
(3.3) |
for any , where denote the collision local times in (2.3).
Recall that defined in (2.5) consists of for which for all . In comparison, condition (3.2) requires a finite expectation of when is distributed as . 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 ).
The equalities (3.3) give a probabilistic interpretation to . By stationarity of , can be thought of as the expected rate of change of the local time . Hence, is intuitively the expected rate at which the bottom particle is ‘pushed down’ by the particle above it during collisions and denotes its upward drift in time. Thus, 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 , one obtains a similar interpretation as follows. Consider the subsystem consisting of the lowest particles viewed as a rank based diffusion , where gets upward drift if its rank in the subsystem is , and it is reflected downwards when it collides with the minimum of the particles outside the subsystem. It can be deduced that each particle 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 . Hence, and respectively quantify the effect of reflection and drift on each particle among the lowest particles, and captures the difference between these effects. The positivity of implies that the hardcore interactions dominate in the long term. Indeed, the results of [41] show that when , under , for any , almost surely as . We conjecture that the same result is true for any .
Remark 3.7.
When , 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 , 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 satisfies and there exist such that for all . If there is a stationary approximative version of the infinite-dimensional gap process with initial (invariant) distribution supported on , and if is a product measure that satisfies the integrability property in (3.2), then for some .
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 for some . Then one may conjecture that, for each ,
One may also ask the following ‘domain of attraction’ question. Given , identify the set such that for , for each ,
In this case, we conjecture that any supported on is in the strong domain of attraction of . We leave the study of these questions for future work.
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 .
Definition 4.1.
Let and be such that . We say that a probability measure on defines a coupling of the gap process of the -Atlas model with initial distributions , if, denoting the coordinate processes on as and , namely
we have
Define the coupling time
where if the above set is empty. When and for some , we write .
Since will be fixed throughout the section, we will frequently suppress it from the notation.
Consider now , where is as in the statement of the theorem, and a bounded measurable map such that
(4.1) |
where is defined as in (3.1), namely, is defined for a.e. as , . In order to prove Theorem 3.3 it suffices, in view of Lemma 3.2, to show that is -a.e. constant. This, in view of (4.1), is equivalent to showing that for some fixed , is -a.e. constant. For the rest of the section we will fix a and consider defined as above. Note that (4.1) holds with replaced by .
4.1 Proof overview
Before we proceed to the details, we give a brief overview of the proof strategy for showing that is -a.e. constant. The first step is to show using the -invariance of that for any , for -a.e. . Moreover, using the product form of , the same conclusion is seen to hold for the process started from a ‘perturbed’ initial point obtained by changing any two co-ordinates of -a.e. by given numbers (see (4.11)). Up to this point, we only use quite general arguments not involving the specific dynamics of the -Atlas model. However, the dynamics comes into play crucially in the subsequent steps, which involve construction of a coupling of two -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 , the coupled -Atlas models coalesce with positive probability by time (Proposition 4.2). Using this and (4.11), it follows that the value of 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 under pairwise perturbations is then extended to perturbation by any finite permutation via straightforward algebraic manipulations. The proof of almost sure constancy of , 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 replaced by ),
This says
(4.2) |
Next let
(4.3) |
For , , and of the form for some , , define
(4.4) |
We have from the Markov property that
(4.5) |
Now take . Then, from (4.2),
(4.6) |
Recall that, for each , is an Exponential distribution and thus is mutually absolutely continuous with respect to the Lebesgue measure on . Since is nonnegative, we have from this mutual absolute continuity property that, for any ,
(4.7) |
Fix , , . For , let . Letting be a -valued random variable distributed as , denote by the probability distribution of where is the unit vector in with at the -th coordinate. Note that by definition in (4.4), , and in particular, from (4.7),
(4.8) |
Thus, in view of (4.5),
(4.9) |
where we have used the fact that is mutually absolutely continuous with respect to . For , let . Then, combining (4.6) and (4.9), we have
(4.10) |
Since when , we get
(4.11) |
We will need the following proposition. The proof is given in Section 4.4.
Proposition 4.2.
Fix , with , and . Then there exists a coupling of the gap process of the -Atlas model with initial distributions and such that, for any ,
Now, for and ,
where are as given by Proposition 4.2 and the second equality follows from (4.11). Hence, from Proposition 4.2, for every and ,
Thus we have shown that
(4.12) |
This implies that
(4.13) |
To see this, let denote the event on the left side of (4.13). Then if and , then for some , and , which shows that is in the event on the left side of (4.12) which in view of (4.12) says that , proving the statement in (4.13). The following proposition enables us to extend (4.13) to all . The proof is given in Section 4.4.
Proposition 4.3.
For each with and , the map is right continuous on . That is, if and and with as , then as .
We remark that our proof shows that in the above proposition can be taken to be for any real bounded measurable function on and , namely it need not be -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 a.e. ,
(4.14) |
A similar argument shows that, for a.e. ,
(4.15) |
We now proceed to the proof of the first statement in Theorem 3.3. Recall that , where satisfies the condition in Theorem 3.3. For notational simplicity, let , . Let be defined as
if . For all other , set . We denote as for simplicity. Observe that, for any ,
(4.16) |
Consider the set with on which the two statements in (4.14) and (4.15) hold. Then for any such that for all , we have,
(4.17) |
Indeed, if , then the statement follows from (4.14) on taking and , and if , the statement follows from (4.15) on taking and . Since , we have that the probability on the right side of (4.3) is and so,
(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 ,
(4.19) |
4.4 Proofs of Propositions 4.2 and 4.3
Recall that we fix . Also, throughout this section we fix such that and . Define for and
Observe that, with the above choice of starting points of the particles, the corresponding gaps are and .
Let be a sequence of mutually independent standard Brownian motions on some probability space . Consider the -dimensional diffusion process
where , , , and is an matrix with for all , for all and otherwise. The process will be used to analyze the evolution of the first gaps before any of them hit zero or the lowest particles interact with the higher ranked particles (in an appropriate sense), as stated more precisely later.
It can be checked that is non-singular and for all and otherwise. Let and define . Also define the stopping time
where denotes the standard Euclidean norm. Define the mirror coupled -dimensional Brownian motion by
(4.20) |
Since is a unitary matrix, it follows from the strong Markov property that is indeed a Brownian motion. can be thought of as the reflection of the Brownian motion in a hyperplane perpendicular to the vector till the first time when hits this hyperplane (which is also the first meeting time of and ), and then coalescing with . Using , define a coupled version of the process by
Extend to an infinite collection of standard Brownian motions .
Our arguments will involve a coupling of two copies of the infinite ordered -Atlas model started from and and respectively driven by the Brownian motions and . For the finite particle -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 -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 and consider a collection of independent standard Brownian motions . Consider for fixed , the system of SDE in (2.3) for , with starting configuration , , local times of collision between the -th and -th particles denoted by , and . Denote by and the unique strong solution to this finite-dimensional system of reflected SDE with driving Brownian motions .
Then (see [32, Definition 7 and Theorem 3.7], [32, Lemma 6.4] and the discussion following it), there exist continuous -valued processes , , adapted to , such that, a.s. satisfies (2.3) with associated local times given by and for any ,
We will call the ‘infinite ordered -Atlas model’ with driving Brownian motions started from .
We denote the infinite ordered -Atlas model defined on with initial condition and driving Brownian motions as . Similarly, denote the infinite ordered -Atlas model defined on with initial condition and driving Brownian motions as . Denote the gap processes associated with and as and respectively, namely
It then follows that
(4.21) |
is a coupling of the gap process of the -Atlas model with initial distributions and . Moreover, the process gives the evolution of before any of the co-ordinates of hit zero or hits the level (note that guarantees that ). This can be seen from (2.3). Similarly, gives the evolution of before any of the co-ordinates of hit zero or hits the level from above.
We will now construct tractable events of positive probability under which the ‘mirror coupled’ processes and will successfully couple before any of their co-ordinates hit zero or (equivalently, ) hits the level . Towards this end, observe that is a polyhedral convex domain contained in the nonpositive orthant of . Let denote the line segment joining and . By convexity of , and since ,
(4.22) |
where denotes the boundary of and denotes Euclidean distance of a point from this set. Also define the processes
where the last equality can be verified from (4.20). Observe that is a standard Brownian motion and moreover, and are independent Gaussian processes. From a geometric point of view, denotes the component of along the vector and denotes the ‘synchronously’ coupled projections of and along the hyperplane perpendicular to . Define the stopping time
Consider any . Note that, if , then, using ,
Furthermore,
Hence, by definition of , . If, on the other hand, , then recalling ,
again implying . Hence, we conclude that, on the event , for all . A similar argument gives for all on the same event.
Since , the sequence is bounded. Let be such that
Fix any
(4.23) |
For and , define the following events in :
Let . For notational convenience, we suppress the dependence of on .
We claim that . To see this, observe that on the event , the ordered Atlas particles and stay above the level by time . Moreover, on , which, by the previous discussion, implies and for all , that is, none of the co-ordinates of or hit zero by time . Hence, for all , and . Further, by the mirror coupling dynamics, and thus, under ,
for all . As under , we conclude that the above equality holds for all . Finally, as and (also, and ) do not meet by time , for all , for all , and hence for all . These observations imply on .
The following lemma gives a key estimate on the probabilities of this event. Recall for which .
Lemma 4.5.
For any , there is a and such that for all .
Proof.
Note that the inequality is immediate from the discussion above the lemma. Now fix . Constants appearing in this proof may depend on and this dependence is not noted explicitly. By a union bound and properties of Brownian motion we see that for any and ,
(4.25) |
where for ,
(4.26) |
Note that, for , . By our condition on in (4.23) and using ,
Thus, from (4.25) it follows that
(4.27) | ||||
(4.28) |
where the first inequality uses for , the second uses the fact that , and the third uses the inequality for . Thus, choosing such that
(here we use ), we obtain
(4.29) |
Without loss of generality we assume that . Next, note that, by the independence of and , for any ,
(4.30) |
Recall defined in (4.22). Writing to highlight its dependence on , it follows from (4.22) and the explicit form of and that there exists and such that
Hence, we obtain such that
(4.31) |
Recall . As and depends only on and not on , we can obtain depending only on such that
Hence, we obtain depending on such that
(4.32) |
From (4.31) and (4.32), for all ,
(4.33) |
Using (4.31) and (4.33) in (4.30), we obtain for all ,
(4.34) |
From (4.29) and (4.34), we conclude
which proves the lemma. ∎
Proof of Proposition 4.3. In order to prove the proposition it suffices to show the right continuity of at . Recall that . Let be arbitrary and let . Let be as in (4.23) and, for this chosen , let and be as in Lemma 4.5. For , , let be as in (4.21) and let be the corresponding expectation operator. For any ,
where the fourth inequality uses Lemma 4.5.
Since is arbitrary, the result follows. ∎
Proof of Proposition 4.2. Let and be as in the statement of the proposition and let be as in (4.21). Recall the event from (4.4) (defined for any ) and consider the following modification of :
Let . From the definition of the coupling it is easily seen by an argument similar to that given before Lemma 4.5 that
Note that, as is given in terms of Brownian motions and is defined in terms of , and are independent. Thus, for any , and it suffices to show that each term in the product is positive.
As , as . Hence, we can obtain depending on such that for all and . Thus,
(4.35) |
Moreover, from standard Brownian motion estimates,
(4.36) |
From the independence of the events considered in (4.4) and (4.36), we conclude that . Finally, from standard Brownian motion estimates and the explicit form of and in terms of ,
The result follows. ∎
5 Proof of Theorem 3.5
Recall from (1.4). Observe that for , for all . Moreover, as , . Thus, exists, is finite, and . As adding the same drift to each ordered particle in (2.3) keeps the gaps unchanged, we can assume without loss of generality that . In particular, for all .
Fix and let be as in the statement of Theorem 3.5. The assumption (3.2) on will be taken to hold throughout the section. From Theorem 2.1 we can construct a filtered probability space equipped with mutually independent real -Brownian motions and continuous processes that solve the SDE (2.3), where are as introduced below (2.3), such that with defined as in (2.4), the process has the distribution . Furthermore, without loss of generality, we can assume that . We will write and respectively for the probability and expectation under the law of this -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 ; 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 for any , which is later used in showing the existence of and to identify this as ( being the local time at zero of the -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 -th gap process against the -th local time process for . The aforementioned uniform integrability is crucially used here. For any , the m.g.f. of the -th gap at time 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 for some 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 and , we have .
Proof.
Recall the sequence associated with . Fix any such that and consider the dimensional -Atlas model defined by replacing with 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 , the associated gap sequence defined by (2.4), where the processes are defined by (2.3), for , has a unique stationary distribution (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 on given as the first marginal distribution of satisfies . In particular, for all . Since as , we can find a , with , such that . Then
which completes the proof. ∎
The next three lemmas concern the collection described above. We remind the reader that in these lemmas the probability measure and the expectation correspond to and respectively, where and are as fixed at the beginning of the section.
Lemma 5.2.
For every , .
Proof.
From (2.3) it follows that, for , and ,
(5.1) |
where for , and . This says that
From this and Lemma 5.1 it follows that if for some , then as well. Thus it suffices to show that . From (2.3), and recalling that , we see that
which says that
Thus to prove the lemma it suffices to show that
(5.2) |
where solve the system of equations in (2.1) with and the vector
distributed as .
Note that for ,
(5.3) |
The next lemma will allow us to interchange expectations and limits as .
Lemma 5.3.
The family is uniformly integrable.
Proof.
For , define as
(5.4) |
Note that is a function with an absolutely continuous derivative. It then follows from Itô’s formula (see [25, Problem 3.7.3]) applied to given by (5.1), that
(5.5) |
Note that, for all ,
Also,
Combining these, and dividing by in (5.5), we have
The desired uniform integrability now follows from Lemmas 5.1 and 5.2 and the observation that
∎
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 at zero for each , as described in the lemma. We note that the product form structure of is crucially exploited here.
Lemma 5.4.
For any , the limit exists and . Furthermore, for any measurable and , ,
(5.6) |
Proof.
From results on local times of continuous semimartingales (see e.g. [29, Corollary VI.1.9]) it follows that, for all and , converges a.s. to -times the semimartingale local time of at (as defined in [29, VI.1.2]) as . Furthermore, one has (see [29, Exercise VI.1.16 (3)]) that
where the first equality on the last line follows from the facts that (which follows from ), and that the Atlas model does not have triple collisions a.s. (see [32, Theorem 5.1]). It then follows that, as ,
(5.7) |
Combining this with Lemma 5.3 and using the fact that is a stationary process, we now have that, as ,
(5.8) |
This proves the first statement in the lemma. In order to prove (5.6) it suffices to consider the case where 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 is a continuous and bounded function. From (5.7) we can find such that and for all
(5.9) |
and for all . In particular this says that, for every and , the collection of measures defined as
is relatively compact in the weak convergence topology. From this and using (5.9) again we now see that, for every and , converges weakly to defined as
From the sample path continuity of we can assume without loss of generality that for every and , is a continuous map. From the above weak convergence it then follows that, for and ,
Using Lemma 5.3 and the fact that is a stationary process with a product form stationary distribution, we now see that for , ,
∎
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 for , we see that, for all ,
(5.10) |
Applying the above identity for , and setting , we have
Proceeding by induction, suppose that for some ,
(5.11) |
Then from (5.10),
Thus it follows that, for every ,
(5.12) |
Fix and . Then by Itô’s formula applied to given by (5.1),
(5.13) |
Since , from Lemma 5.1, and consequently the stochastic integral in the above display has mean . Moreover, note that
Thus taking expectations in (5.3) and using Lemma 5.4, we have,
Rearranging terms, we have
where the second equality in the above display follows from (5.10). Thus we have shown that for all and ,
Thus, since , by uniqueness of Laplace transforms, we must have that for each . Finally note that, since by Lemma 5.1, for all , we must have that . Thus, in view of (5.12), we have shown that, for some , for all and so for some .
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 will be fixed in the proof we suppress it from the notation. Let .
Define
We claim that , where denotes the orthogonal complement of the closure of the linear subspace generated by in . Indeed, if , then for all and , . Taking ,
(A.1) |
Using this, and the contraction property of , for all ,
where the first equality on the second line follows from (A.1). This says that (as elements of ) for all and so and shows the claim . Any can be written as where and .
Next, for and , define as
Then . By definition . Also, if then for some and . Then from the contraction property of it follows that
Similarly, if , there is a such that
Finally, let . Then, given , there is a such that . It follows, using again the contraction property, that for all ,
Thus . Since is arbitrary, we obtain
(A.2) |
From these observations we obtain, for any ,
(A.3) |
The above convergence in fact shows that and consequently is the projection of on to . To see this, recall that it was argued above that . Now consider the reverse inclusion and let . Then we can write where and . Thus, for ,
As , we have from (A.2) that, , which says that . This proves the inclusion and we have the claimed statement .
Now we proceed to the proof of the statements in the lemma. We first consider the second statement in the lemma. Fix . Suppose that for every bounded measurable map , is constant a.s. We will now show that this implies . Suppose there is a and such that . Note that since from (A.3) and is invariant, we must have , and so from (A.3) it follows that
(A.4) |
Also, from definition,
(A.5) |
Thus in . Also, since , and consequently, . Since is an arbitrary bounded measurable function, we must have . This proves that . We have thus shown the second statement in the lemma and in fact also shown that .
Finally we argue that . Suppose and that . Then there is a such that is not a.s. constant under . Thus there is a such that, with , . Note that by definition
(A.6) |
We refer to this property as is harmonic (with respect to the semigroup ). We claim that this implies that is harmonic as well, namely
(A.7) |
To see this, note that, from the definition of , if is harmonic, then
(A.8) |
This, together with the fact that is a contraction, says that
which in view of (A.8) shows that is harmonic. From the linearity of it then follows that is harmonic as well. This implies that if are harmonic then and are harmonic as well. Recalling tha is harmonic, we now have that is harmonic for every . The property in (A.7) is now immediate from this on observing that a.s. and dominated convergence theorem. This proves the claim.
Consider the probability measures on defined as
Using (A.7) it is easily seen that . Indeed, if and ,
where the third and fourth equalities follow from (A.7) and the fifth uses the invariance of . This shows the invariance of from which (together with the fact that is invariant) the invariance of follows immediately.
Finally note that , and by definition . This contradicts the fact that and thus we must have . We have thus shown that 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]