A Degenerate Stochastic Partial Differential Equation for Superprocesses with Singular Interaction

Similar documents
9.2 Branching random walk and branching Brownian motions

Branching Processes II: Convergence of critical branching to Feller s CSB

Brownian Motion. 1 Definition Brownian Motion Wiener measure... 3

Regularity of the density for the stochastic heat equation

p 1 ( Y p dp) 1/p ( X p dp) 1 1 p

Learning Session on Genealogies of Interacting Particle Systems

On the martingales obtained by an extension due to Saisho, Tanemura and Yor of Pitman s theorem

Some Properties of NSFDEs

for all f satisfying E[ f(x) ] <.

Brownian motion. Samy Tindel. Purdue University. Probability Theory 2 - MA 539

The Continuity of SDE With Respect to Initial Value in the Total Variation

Kai Lai Chung

Convergence at first and second order of some approximations of stochastic integrals

Functional central limit theorem for super α-stable processes

LAN property for ergodic jump-diffusion processes with discrete observations

Martingale Problems. Abhay G. Bhatt Theoretical Statistics and Mathematics Unit Indian Statistical Institute, Delhi

Introduction to Random Diffusions

ON THE FIRST TIME THAT AN ITO PROCESS HITS A BARRIER

PROBABILITY: LIMIT THEOREMS II, SPRING HOMEWORK PROBLEMS

PROBABILITY: LIMIT THEOREMS II, SPRING HOMEWORK PROBLEMS

Cores for generators of some Markov semigroups

OPTIMAL SOLUTIONS TO STOCHASTIC DIFFERENTIAL INCLUSIONS

Lecture 12. F o s, (1.1) F t := s>t

Two viewpoints on measure valued processes

Stepping-Stone Model with Circular Brownian Migration

1. Stochastic Processes and filtrations

Harnack Inequalities and Applications for Stochastic Equations

ITÔ S ONE POINT EXTENSIONS OF MARKOV PROCESSES. Masatoshi Fukushima

Functional Limit theorems for the quadratic variation of a continuous time random walk and for certain stochastic integrals

CHENCHEN MOU AND YINGFEI YI

Brownian Motion and Stochastic Calculus

Properties of an infinite dimensional EDS system : the Muller s ratchet

Uniformly Uniformly-ergodic Markov chains and BSDEs

SMSTC (2007/08) Probability.

Measure-valued processes and related topics

Large Deviations for Perturbed Reflected Diffusion Processes

Fonctions on bounded variations in Hilbert spaces

Limit theorems for continuous time branching flows 1

Pathwise uniqueness for stochastic differential equations driven by pure jump processes

Stationary distribution and pathwise estimation of n-species mutualism system with stochastic perturbation

WEYL S LEMMA, ONE OF MANY. Daniel W. Stroock

STATISTICS 385: STOCHASTIC CALCULUS HOMEWORK ASSIGNMENT 4 DUE NOVEMBER 23, = (2n 1)(2n 3) 3 1.

{σ x >t}p x. (σ x >t)=e at.

Spatial diffusions with singular drifts: The construction of Super Brownian motion with immigration at unoccupied sites

Weak convergence and large deviation theory

Reflected Brownian Motion

Exponential martingales: uniform integrability results and applications to point processes

MATH 6605: SUMMARY LECTURE NOTES

THE SKOROKHOD OBLIQUE REFLECTION PROBLEM IN A CONVEX POLYHEDRON

Mathematical Methods for Neurosciences. ENS - Master MVA Paris 6 - Master Maths-Bio ( )

Filtrations, Markov Processes and Martingales. Lectures on Lévy Processes and Stochastic Calculus, Braunschweig, Lecture 3: The Lévy-Itô Decomposition

Introduction to self-similar growth-fragmentations

BSDEs and PDEs with discontinuous coecients Applications to homogenization K. Bahlali, A. Elouain, E. Pardoux. Jena, March 2009

Simulation methods for stochastic models in chemistry

Some SDEs with distributional drift Part I : General calculus. Flandoli, Franco; Russo, Francesco; Wolf, Jochen

A REPRESENTATION FOR THE KANTOROVICH RUBINSTEIN DISTANCE DEFINED BY THE CAMERON MARTIN NORM OF A GAUSSIAN MEASURE ON A BANACH SPACE

n E(X t T n = lim X s Tn = X s

THE MALLIAVIN CALCULUS FOR SDE WITH JUMPS AND THE PARTIALLY HYPOELLIPTIC PROBLEM

A NEW PROOF OF THE WIENER HOPF FACTORIZATION VIA BASU S THEOREM

SOLUTIONS OF SEMILINEAR WAVE EQUATION VIA STOCHASTIC CASCADES

Spatial Ergodicity of the Harris Flows

EULER MARUYAMA APPROXIMATION FOR SDES WITH JUMPS AND NON-LIPSCHITZ COEFFICIENTS

Feller Processes and Semigroups

Exercises. T 2T. e ita φ(t)dt.

Basic Definitions: Indexed Collections and Random Functions

Stability of Stochastic Differential Equations

The Pedestrian s Guide to Local Time

Stochastic optimal control with rough paths

4.6 Example of non-uniqueness.

Random Process Lecture 1. Fundamentals of Probability

Generalized Gaussian Bridges of Prediction-Invertible Processes

Homework # , Spring Due 14 May Convergence of the empirical CDF, uniform samples

Stochastic Calculus February 11, / 33

Richard F. Bass Krzysztof Burdzy University of Washington

Bayesian quickest detection problems for some diffusion processes

A Central Limit Theorem for Fleming-Viot Particle Systems Application to the Adaptive Multilevel Splitting Algorithm

L -uniqueness of Schrödinger operators on a Riemannian manifold

Simulation of diffusion. processes with discontinuous coefficients. Antoine Lejay Projet TOSCA, INRIA Nancy Grand-Est, Institut Élie Cartan

Homework #6 : final examination Due on March 22nd : individual work

ON THE POLICY IMPROVEMENT ALGORITHM IN CONTINUOUS TIME

NON-UNIQUENESS FOR NON-NEGATIVE SOLUTIONS OF PARABOLIC STOCHASTIC PARTIAL DIFFERENTIAL EQUATIONS

13 The martingale problem

Large time behavior of reaction-diffusion equations with Bessel generators

A REVERSIBILITY PROBLEM FOR FLEMING-VIOT PROCESSES

Citation Osaka Journal of Mathematics. 41(4)

Example 4.1 Let X be a random variable and f(t) a given function of time. Then. Y (t) = f(t)x. Y (t) = X sin(ωt + δ)

Stochastic Differential Equations.

Lecture 21 Representations of Martingales

A NOTE ON STOCHASTIC INTEGRALS AS L 2 -CURVES

4th Preparation Sheet - Solutions

Stochastic Integration.

On pathwise stochastic integration

ERRATA: Probabilistic Techniques in Analysis

A Concise Course on Stochastic Partial Differential Equations

Latent voter model on random regular graphs

Stochastic Processes

Obstacle problems for nonlocal operators

The Chaotic Character of the Stochastic Heat Equation

Applications of Ito s Formula

Fast-slow systems with chaotic noise

Transcription:

Published in: Probability Theory and Related Fields 13 (24), 1: 1 17. A Degenerate Stochastic Partial Differential Equation for Superprocesses with Singular Interaction Zenghu Li 1 Department of Mathematics, Beijing Normal University, Beijing 1875, P.R. China lizh@bnu.edu.cn Hao Wang 2 Department of Mathematics, University of Oregon, Eugene OR 9743-1222, U.S.A. haowang@math.uoregon.edu Jie Xiong 3 Department of Mathematics, University of Tennessee Knoxville, TN 37996-13, U.S.A. jxiong@math.utk.edu and Mathematical and Statistical Sciences, University of Alberta Edmonton, Alberta T6G 2G1, Canada. jxiong@math.ualberta.ca Abstract The scaling limit for a class of interacting superprocesses and the associated singular, degenerate stochastic partial differential equation (SDSPDE) are investigated. It is proved that the scaling limit is a coalescing, purely-atomic-measurevalued process which is the unique strong solution of a reconstructed, associated SDSPDE. AMS Subject Classifications: Primary 6G57, 6H15; secondary 6J8. Key words and phrases: coalescing Brownian motion, scaling limit, purely atomic superprocess, interaction, stochastic partial differential equation, strong solution, pathwise uniqueness. 1 Research supported by NNSF (No. 11314 and No. 112111). 2 Research supported partially by the research grant of UO. 2 Corresponding author. 3 Research supported partially by NSA, NSERC, PIms, Lockheed Martin Naval Electronics and Surveillance Systems, Lockheed Martin Canada and VisionSmart through a MITACS center of excellence entitled Prediction in Interacting Systems. 1

1 Introduction Currently interactions and stochastic partial differential equations are two of the hot topics in the field of Superprocesses. We have seen that diverse interactions have been introduced into the models of superprocesses such as population interaction in the mutually catalytic model, interactive branching mechanism, inter-particle interaction, particle random medium interaction, mean-field interaction, and so on. (See the survey article [3] and the references therein). In Dawson et al. [5], a particle random medium interaction model with location dependent branching, which generalized the model introduced in Wang [14] [13], is introduced and the limiting superprocesses, which will be called superprocesses with dependent spatial motion and branching (SDSM s), are constructed and characterized. An open problem left from the recent work of Dawson et al. [6] and Wang [16] on this model has especially raised our interest in the investigation and consideration presented in the present paper. In order to describe the question clearly, we introduce necessary notations and the model first. For fixed natural integers k, m 1, let C k (R m ) be the set of functions on R m having continuous derivatives of order k and C k(rm ) be the set of functions in C k (R m ) which, together with their derivatives up to order k, can be extended continuously to R m := R m { }, the one point compactification of R m. C k(rm ) denotes the subset of C k(rm ) of functions that, together with their derivatives up to order k, vanish at infinity. Let M(R m ) be the space of finite Borel measures on R m equipped with the topology of weak convergence. We denote by C b (R m ) the set of bounded continuous functions on R m, and by C (R m ) its subset of continuous functions vanishing at infinity. The subsets of non-negative elements of C b (R m ) and C (R m ) are denoted by C b (R m ) + and C (R m ) +, respectively. S(R) stands for the space of all infinitely differentiable functions which, together with all their derivatives, are rapidly decreasing at infinity. Let B(R) (resp. C(R)) be the collection of all Borel (resp. continuous) functions on R. For f B(R) and µ M(R), set f, µ = R fdµ. Suppose that {W (x, t) : x R, t } is a Brownian sheet (see Walsh [12]) and {B i (t) : t },i N, is a family of independent standard Brownian motions which are independent of {W (x, t) : x R, t }. For each natural number n, which serves as a control parameter for our finite branching particle systems, we consider a system of particles (initially, there are m n particles) which move, die and produce offspring in a random medium on R. The diffusive part of such a branching particle system has the form dx n i (t) = c(x n i (t))db i (t) + h(y x n i (t))w (dy, dt), t, (1.1) R where c C b (R) is a Lipschitz function and h C 2 (R) is a square-integrable function. By Lemma 3.1 of Dawson et al. [5], for any initial conditions x n i () = x i R, the stochastic equations (1.1) have unique strong solution {x n i (t) : t } and, for each integer m 1, {(x n 1 (t),, xn m(t)) : t } is an m-dimensional diffusion process which is generated by the differential operator G m := 1 2 m i=1 a(x i ) 2 x 2 i + 1 2 m i,j=1,i j 2 ρ(x i x j ). (1.2) x i x j In particular, {x n i (t) : t } is a one-dimensional diffusion process with generator G := 2

(a(x)/2), where is the Laplacian operator, ρ(x) := h(y x)h(y)dy, (1.3) R and a(x) := c 2 (x) + ρ() for x R. The function ρ is twice continuously differentiable with ρ and ρ bounded since h is square-integrable and twice continuously differentiable with h and h bounded. The quadratic variational process for the system given by (1.1) is x n i (t), x n j (t) = ρ(x n i (s) x n j (s))ds + δ {i=j} c 2 (x i (s)) ds, (1.4) where we set δ {i=j} = 1 or according as i = j or i j, where i, j N. Here x n i (t) is the location of the i th particle. We assume that each particle has mass 1/θ n and branches at rate γθ n, where γ and θ 2 are fixed constants. We assume that when a particle 1 θ δ n x, which has location at x, dies, it produces k particles with probability p k (x); x R, k N {}. This means that the branching mechanism depends on the spatial location. The offspring distribution is assumed to satisfy: p 1 (x) =, kp k (x) = 1, and m 2 (x) := k 2 p k (x) < for all x R. (1.5) k= k= The second condition indicates that we are solely interested in the critical case. After branching, the resulting set of particles evolve in the same way as their parents and they start off from the parent particle s branching site. Let m n t denote the total number of particles at time t. Denote the empirical measure process by µ n t ( ) := 1 θ n m n t δ x n i (t)( ). (1.6) In order to obtain measure-valued processes by use of an appropriate rescaling, we assume that there is a positive constant ξ > such that m n /θn ξ for all n and that weak convergence of the initial laws µ n µ holds, for some finite measure µ. As for the convergence from branching particle systems to a SDSM, the reader is referred to Wang [14] and Dawson et al. [5]. Let E := M(R) be the Polish space of all finite Radon measures on R with the weak topology defined by µ n µ if and only if f, µ n f, µ for f C b (R). By Ito s formula and the conditional independence of motions and branching, we can obtain the following formal generators (usually called pregenerators) for the limiting measure-valued processes: L c,σ F (µ) := A c F (µ) + B σ F (µ), (1.7) where B σ F (µ) := 1 2 R i=1 σ(x) δ2 F (µ) µ(dx), (1.8) δµ(x) 2 3

and A c F (µ) := 1 2 R a(x)( d2 (µ) )δf dx2 δµ(x) µ(dx) (1.9) + 1 2 R R ρ(x y)( d dx )( d dy ) δ 2 F (µ) δµ(x)δµ(y) µ(dx)µ(dy) for F (µ) D(L c,σ ) C(E), where σ(x) := γ(m 2 (x) 1) for any x R, the variational derivative is defined by δf (µ) δµ(x) := lim h F (µ + hδ x ) F (µ), (1.1) h and D(L c,σ ) is the domain of the pregenerator L c,σ. Especially, we denote L,σ = A + B σ for L c,σ = A c +B σ with c(x). Let B(R) + be the space of all non-negative, bounded, measurable functions on R. We cite one theorem proved in Dawson et al. [5]. Theorem 1.1 Let c C b (R) be a Lipschitz function, h C 2 (R) be a square-integrable function on R, and σ(x) B(R) +. Then, for any µ E, (L c,σ, δ µ )-martingale problem (MP) has a unique solution which is denoted by X t with sample paths in D([, ), M(R)). Then X t is a diffusion process. Proof: For the proof of this theorem, the reader is referred to the section 5 of Dawson et al. [5]. X t is often called the high density limit of the branching particle systems we discussed. Similar to Konno-Shiga s famous results for super-brownian motion (See [9]), for this interactive model it was proved by Wang [13] that X t is absolutely continuous. Also, Dawson et al [8] derived a stochastic partial differential equation (SPDE) for the density for the case of c( ) = ɛ >. An interesting case is due to Wang ([13], [15]) who proved that when c( ), X t is a purely atomic measure valued process. In addition, Dawson et al [6] derived a degenerated SPDE φ, X t = φ, X t + h(y )φ, X u W (dydu) for + 1 2 ρ() + i I(t ) t R t φ, X u du (1.11) t φ(x i (u)) σ(x i (u))a i (u)db i (u), φ S(R), t t > X t = i I(t) a i (t)δ xi (t) and proved that the above degenerate SPDE has a pathwise or strong unique solution. On the other hand, Wang [16] recently proved that if c( ) ɛ >, and h is a singular function (See Wang [16] for the precise definition) which can be roughly defined as the square root of the Dirac delta function, the high density limit X t is just the super-brownian motion. Wang [16] 4

also pointed out by example that if c( ), the conclusion is not clear when h is a singular function due to coalescence. This naturally raised a challenging question: Can we identify the high density limit as well as its associated degenerate SPDE when c( ) and h is a singular function? We now outline the rough idea for approaching the problem. Since the square root of the Dirac delta function cannot be defined in the sense of distribution, we have to find a way such that it makes sense. The first idea is to seek a method to handle it as a sequential limit. The scaling limit argument automatically becomes a good candidate for us. From the related literature, we found that Dawson and Fleischmann [2] studied the clumping property of the classical super-brownian motion (SBM) using the scaling limit. Mimicking this paper, we defined Xt K (B) = K 1 X Kt (KB), B B(R). (1.12) We found that the limit of X K is the same as that of [2]. The external term is simply too weak to carry over at the end of the scaling. However, this does not give all that we want. When we carefully checked the proportions of scaling for different coefficients, we realized that we must adjust the scaling proportion of the random medium term. This adjustment also matches the real world situation. In fact, h and W characterize the outside force applying to the whole system. It should use a much larger scale compared to the motion of each individual in the system. Imagine our own movement and that of the earth! Therefore, we replaced the original h(x) by Kh(x) and applied the scaling (1.12). We denote the resulted process by Zt K. Assume that the scaled initial measures converge. We then prove that the limit Z t of Zt K exists and is characterized as follows: At time t >, Z t is a Poisson random measure with intensity t 1 µ. The particles move according to coalescing Brownian motion, whose mechanism is determined by the external term ρ, until its mass, governed by independent Feller s branching diffusions, reaches. Note that when h =, the particles do not move and we get the same result as Dawson and Fleischmann [2]. From this procedure, we derive a singular, degenerated SPDE for the limit process, where the motion dynamic is driven by a sequence of coalescing Brownian motions. Nevertheless, the singular, degenerated SPDE does not have strong uniqueness due to coalescence. After we replace the coalescing Brownian motions by killing Brownian motions, the strong uniqueness of the singular, degenerated SPDE is recovered and we get our expected results. A similar problem is studied by Dawson et al. [7] from a different point of view. Since the paper is not yet published, we briefly list here the contents of that paper. The main purpose of the paper is to give a proof of the observation of Dawson et al. [5]. Section 2 gives characterizations for a coalescing Brownian motion flow and shows that the flow is actually the scaling limit of the interacting Brownian flow that serves as the carrier of the purely atomic SDSM in the excursion representation given in Dawson and Li [4]. Section 3 constructs the limiting superprocess in terms of one-dimensional excursions using the coalescing Brownian flow as a carrier. Section 4 derives the scaling limit of the SDSM from that of the interacting Brownian flow and the excursion representations. The major differences between [7] and our paper are as follows: First, we consider the usual scaling (1.12) and compare with the results of Dawson and Fleischmann for classical SBM. In [7], the scaling K 2 X Kt (KB) is used. Secondly, our scaling limit is for a general superprocess over a stochastic flow, namely, we derive the degenerate limit from random fields. In [7], it started from the degenerated process (i.e. c = ) so that the scaling limit is essentially for finite-dimensional processes. Thirdly, our proof makes 5

use of a new representation theorem which provides an easy approach to Wang s (cf. [15], [13]) result for the non-singular case. Finally, [7] has discussed the excursion representation of the limiting superprocess. In the current paper we consider an SPDE for the superprocess which is degenerated as well as singular. These two topics distinguish the two papers emphases. A related model was studied by Skoulakis and Adler [11], and its properties were investigated by Xiong [18], [17]. This article is organized as follows: In section 2, we prove the weak convergence of X K and characterize the limit by a martingale problem. In section 3, we discuss the weak convergence of Z K with h replaced by Kh and prove that the limit martingale problem coincides with that studied by Dawson et al [7] which arises from a system of coalescing Brownian motions. In section 4, we show the nonuniqueness for the solution to the degenerated SPDE which is a natural extension of (1.11) for the singular case. Finally, We modify the driving Brownian motions and prove strong uniqueness for the modified SPDE. 2 Weak convergence under clumping scaling To accommodate a larger class of processes, we consider tempered measures. Let φ λ (x) = y <1 ( dye λ x y exp 1 ) / ( 1 y 2 dy exp 1 ) y <1 1 y 2. Note that to each λ R and m there are positive constants c λ,m and c λ,m such that c λ,m φ λ (x) dm dx m φ λ (x) c λ,m φ λ (x), x R, (cf. (2.1) of Mitoma [1]). We define M tem (R) to be the collection of all measures µ such that φ λ, µ <, λ >. Let C rap (R) be the collection of all functions f such that for all λ >, there exists c λ such that f(x) c λ φ λ (x) for all x R. Lemma 2.1 Assume that a, σ and ρ are bounded and sup φλ, µ K <, λ > K where µ K is define by the same fashion as in (1.12). Then for any α 2, λ, T >, we have sup K E sup t T φλ, X K t α <. Proof: Denote a K (x) = K 1 a(kx). σ K and ρ K are defined similarly. Note that X K satisfies the following martingale problem: φ C rap (R), M K t (φ) φ, X K t φ, µ K 1 ak φ, Xu K 2 du (2.1) 6

is a martingale with quadratic variation process M K (φ) t = KσK φ 2, Xu K t du + du ρ K (y z)φ (y)φ (z)xu K (dy)xu K (dz). (2.2) R 2 By Burkholder s inequality, it is easy to see that E sup φλ, Xt K t s ( s ) α c + ce φλ, Xt K α dt ( s ) +ce φ 2 λ, Xt K α/2 ( s ) dt + ce φλ, Xt K α/2 2 dt. Since φ 2 λ φ λ, using x 1 + x 2 and Hölder s inequality, we can continue with c + c s E φ λ, X K t α dt. The conclusion of the lemma then follows from Gronwall s inequality. Theorem 2.1 Under the conditions of Lemma 2.1, {X K : K 1} is tight in C(R +, M tem (R)). Proof: It is well known that we only need to prove the tightness of { φ, X K : K 1} in C(R +, R) for each fixed φ C rap (R). Note that by the martingale problem (2.1,2.2) and Lemma 2.1, we have E φ, Xt K φ, X K s α c t s α/2. Take α > 2; the tightness then follows from Kolmogorov s criterion. Theorem 2.2 Suppose that σ( ) = lim x σ(x) and µ = lim K µ K exist and is not an atom of µ. Under the conditions of Lemma 2.1, X K converges in law to the unique solution of the following martingale problem: φ C rap (R), is a martingale with quadratic variation process M t (φ) φ, X t φ, µ (2.3) M (φ) t = σ( )φ 2, X u du. (2.4) Proof: Note that for any ɛ > fixed, Kσ K (y) converges to σ( ) as K uniformly for y Sɛ c where S ɛ = ( ɛ, ɛ). On the other hand, if we choose f ɛ C b (R) such that 1 Sɛ f ɛ 1 S2ɛ, then lim sup E φ 2 1 Sɛ, Xu K lim sup µ K (dx) f ɛ (y)φ 2 (y)p K u (x, dy) K K = µ (dx)f ɛ (x)φ 2 (x), where p K u (x, dy) is the transition probability of the Markov process generated by L K φ = 1 2 a Kφ. Let ɛ ; we have µ (dx)f ɛ (x)φ 2 (x). By (2.1), (2.2), it is then easy to see that every limit point of X K solves the martingale problem (2.3), (2.4). The uniqueness of this martingale problem follows from [2]. The conclusion of the theorem then follows easily. 7

3 Weak convergence under strong interaction In this section, we consider strong interaction, namely, replace h by Kh and then apply clumping scaling discussed in the previous section. For any φ C rap (R), we have that U K t (φ) φ, Z K t φ, µ K 1 (K 1 c 2 (Kx) + ρ())φ, Z K u du 2 is a martingale with quadratic variation process U K (φ) t = KσK φ 2, Zu K t du + du ρ(k(y z))φ (y)φ (z)zu K (dy)zu K (dz). R 2 We shall prove that Z K is a tight sequence and characterize the limit Z. Note that Z K is a measure-valued process with density and, as we will show, Z is a purely atomic measure-valued process. Therefore, it is not easy to derive the limit martingale problem of type (3.8) below from the martingale problem for Z K studied in Dawson et al. [5]. With a complicated argument as that in Xiong and Zhou [19], we believe that it can be proved that Z satisfies the martingale problem (3.6), (3.7). However, that martingale problem is not well-posed. To determine the distribution of Z uniquely, we start with the dual relation between Z K and (Y K, M) and prove the convergence of the latter; then the distribution of Z is determined by the limit of (Y K, M). Finally, we construct a process which is clearly a Markov process, since it is the unique solution to the martingale problem (3.8) and has the same distribution as Z. First we need the following lemma. Lemma 3.1 Suppose that lim x ρ(x) = and η K is governed by the following SDE: dη K (t) = ρ() ρ(kη K (t))db t. Then η K η which is a Brownian motion with as an absorbing boundary. Proof: It is easy to see that η K is a tight sequence of diffusions on [, ), each with as an absorbing boundary. Let D = {f C 2 ([, )) : f () = }. Then for any f D, 1 f(η K (t)) 2 (ρ() ρ(kη K(s)))f (η K (s))ds is a martingale. Let η be a limit point. Then for any f D, f(η(t)) 1 2 ρ()f (η(s))ds is a martingale. This proves the conclusion of the lemma. To characterize the limit of Z K, we need the concept of the coalescing Brownian motion (CBM) which was first introduced by R. Arratia [1], where the coalescing Brownian motion is constructed from a system of discrete random walks. An interesting thing here is that we can construct the coalescing Brownian motion simply from a given Brownian sheet. 8

Definition 3.1 (x 1 (t),, x m (t)) is a CBM if the components move as independent Brownian motions until any pair, say x i (t) and x j (t), (i < j), meet. Starting from the meeting time, x j (t) assumes the values of x i (t), x i (t) disappears, and the system continues to evolve in the same fashion. Theorem 3.1 Suppose lim x ρ(x) = and the conditions of Theorem 2.2 hold. Then, Z K is tight and its limit finite marginal distribution is determined by the following duality relation: for all f C rap (R m ), [ Yt E f, (Z t ) m = E m,f, µ Mt { 1 exp 2 }] M s (M s 1)ds (3.1) and recursive relation: for t 1 < t 2 < < t j+1, ( ) E Π j+1 i=1 f i, (Z ti ) m i ( = E E Y m j+1,f j+1 [ Y tj+1 t j, (Z tj ) M t j+1 t j exp Π j i=1 f i, (Z ti ) m i ) { 1 tj+1 t j }] M s (M s 1)ds 2 (3.2) where M t is Kingman s coalescent process starting at m with jumping time = τ < τ 1 < < τ m =, and where Y t, starting at f, is a function-valued process defined by Y t = P Mτ k t τ k Γ k P Mτ 1 τ 2 τ 1 Γ 1 P Mτ τ 1 Y, t [τ k, τ k+1 ), k < m (3.3) where Pt m is the semigroup of the m-dimensional coalescing Brownian motion, Γ k is taking one of the Φ ij randomly and Φ ij f(x 1,, x m 1 ) = σ( )f(x 1,, x m 1,, x m 1,, x m 2 ), (3.4) where x m 1 is in the places of the ith and jth variables. Proof: The tightness follows from an argument similar to that in section 2. By [5], we have E f, (Zt K ) m [ Y K = E m,f t, µ Mt { 1 exp 2 }] M s (M s 1)ds, where Y K starting from f is defined similarly as in (3.3)-(3.4) with Φ ij, σ( ) and Pt m by Φ K ij f(x 1,, x m 1 ) = σ(kx m 1 )f(x 1,, x m 1,, x m 1,, x m 2 ), σ(k ), and P m,k t, respectively. P m,k t G m,k = 1 2 m i=1 K 1 c 2 (Kx i ) 2 x 2 i is the semigroup with generator + 1 2 1 i,j m 2 ρ(k(x i x j )). x i x j replaced 9

Let We define G m = 1 2 D 1 = {f C 2 b (Rm ) : 1 i,j m 2 ρ()1 {xi =x j }. x i x j 2 f x i x j = if x i = x j, for some i j}. Let x K (t) be the process generated by G m,k. It is easy to see that x K is tight as an m- dimensional process. Let x be a limit point. For f D 1, we have G m,k f G m f and hence f(x(t)) f(x()) G m f(x(s))ds is a martingale. From this, it is easy to see that x(t) is an m-dimensional Brownian motion before it reaches the set {x : i < j, x i = x j }. To study its behavior after it reaches that set, we consider x K i (t) xk j (t). Note that d x K dt i x K j t = K 1 c 2 (Kx K i (t)) + K 1 c 2 (Kx K j (t)) + (ρ() ρ(k(x K i (t) x K j (t))). The first and the second terms converge to uniformly. As in Lemma 3.1, it is then easy to show that x i (t) x j (t) is a Brownian motion with as absorbing boundary. This implies that x(t) is the coalescing Brownian motion. Hence, P m,k t converges to Pt m, and hence Y K converges to Y. As f C rap (R m ), we have f(x 1,, x m ) c λ φ λ (x 1 ) φ λ (x m ) c λ φ λ (x). It is easy to verify that P m,k t φ(x) c m λ,t φ λ(x). Therefore, Yt K, µ Mt c Mt λ,t φ λ, µ Mt. (3.1) follows from the dominated convergence theorem. The same argument as in [5] shows that (3.1) determines the distribution of Z t uniquely. Thus, for any fixed m and n, we have that f, (Z K t ) m n is integrable uniformly in K since E ( f, (Z K t ) m 2n ) = E f 2n, (Z K t ) 2mn. Finally we prove (3.2). Note that ( E Π j+1 i=1 fi, (Zt K i ) m ) ( i = E E ( where f K i = E ( ) f j+1, (Zt K j+1 ) m j+1 Ft K j Π j i=1 fi, (Zt K i ) m ) i E Y m j+1,f j+1 [ Y K t j+1 t j, (Z K t j ) M t j+1 t j { 1 exp 2 j+1 t j ( E Π j i=1 f K i, (Zt K i ) m ) i, }] M s (M s 1)ds Π j i=1 fi, (Zt K i ) m ) i = f i for i < j and ( { 1 tj+1 fj K = f j E Y m j+1,f j+1 Yt K t j }) j+1 t j exp M s (M s 1)ds. 2 Note that f K j c λ φ λ. Letting K in (3.5), we have (3.2). 1 (3.5)

Next we construct Z t from another point of view. Let f, Z t n = 1 n n ξ i (t)f(x i (t)), i=1 where {ξ i } are independent Feller s branching diffusions with branching rate σ( ) and (x i (t),, x n (t)) is the n-dimensional coalescent Brownian motion with diffusion coefficient ρ(). Here we consider the limit of Z n by adapting the method of Xiong and Zhou [19]. Theorem 3.2 Under the conditions of Theorem 3.1, Zn is tight and its limit Z solves the following martingale problem: U t (φ) φ, Z t φ, µ 1 2 ρ() is a martingale with quadratic variation process U (φ) t = where = {(x, x) : x R}. σ( )φ 2, Z u du + Proof: Applying Itô s formula, it is easy to show that U n t (φ) φ, Z n t is a martingale with quadratic variation process U n (φ) t = 1 n = n i=1 + ρ() n 2 φ, µ 1 2 ρ() σ( )ξ i (s)φ(x i (s)) 2 ds n i,j=1 τ ij t σ( )φ 2, Z u n du + φ, Z u du (3.6) du ρ()φ (y)φ (z) Z u (dy) Z u (dz), (3.7) φ, Z u n du ξ i (s)ξ j (s)φ (x i (s))φ (x j (s))ds du ρ()φ (y)φ (z) Z u n (dy) Z u n (dz), where τ ij is the first time x i (t) and x j (t) meet. It is easy to prove the tightness of {( Z n, U n (φ))}. Denote the limit by ( Z, Λ). By arguments similar to those in [19], we see that Λ(t) = σ( )φ 2, Z u du + du ρ()φ (y)φ (z) Z u (dy) Z u (dz) and hence Z solves the martingale problem (3.6)-(3.7). Remark 3.3 The solution to the martingale problem (3.6)-(3.7) is not unique. For example, if we replace (x 1 (t),, x n (t)) by n-dimensional ordinary Brownian motion, the limit will provide another example. The ordinary SBM with diffusion coefficient ρ() and branching rate σ( ) is a third example of the solution to the martingale problem (3.6)-(3.7). 11

Finally, we prove Z and Z have the same distribution. Theorem 3.4 Under the conditions of Theorem 3.1, Z = Z in distribution. Therefore, Z t is a Markov process. Proof: We only need to prove (3.1) holds with Z replaced by Z. For f Cb 2(Rm ), let F m,f (µ) = f, µ m. Define LF m,f (µ) = F m,g m f (µ) + 1 F m 1,Φij f (µ). 2 1 i j m By Itô s formula, it is easy to see that Z n t solves the following martingale problem: f C 2 b (Rm ), F m,f ( Z n t ) F m,f ( Z n ) LF m,f ( Z n s )ds is a martingale. Let n, we see that Z t satisfies: f C 2 b (Rm ), F m,f ( Z t ) F m,f (µ) LF m,f ( Z s )ds (3.8) is a martingale. Mimicking the proof of Theorem 2.1 in [5], we see that (3.1) holds with Z t replaced by Z t. Since Z t is a Markov process, it is easy to verify that (3.2) holds with Z t replaced by Z t. Therefore, Z and Z have the same distribution and Z is a Markov process. 4 SPDE In this section, we first show that Z t is of purely atomic type. Then we characterize Z t as the unique strong solution to an SPDE. Since we are only interested in establishing the property for Z t, we may and will assume that c = and σ = σ( ) are constants. Let h n converges to the square root of the delta function so that ρ n (x) converges to when x and ρ n () ρ() >. Fix h n as in [5], we first reprove a theorem of Wang [13] (see also [15]) by a new representation of Z n t. This representation in a more general setup will involve Perkins historical calculus and will be developed in another paper. Throughout this section, we assume that the conditions of Theorem 3.1 remain in force. Theorem 4.1 Let Xt n (x, W ) be the strong solution of Xt n = x + h n (y Xs n )W (dsdy). Let ζ t be the superprocess with branching rate σ and spatial motion-free. Define Z n t ( ) = ζ t {x : X n t (x, W ) }. Then U n t (φ) = φ, Z n t φ, Z n 1 2 ρ n()φ, Zu n du 12

is a martingale with quadratic variation process U n (φ) t = σφ 2, Zu n t du + du ρ n (y z)φ (y)φ (z)zu n (dy)zu n (dz). R 2 Proof: Applying Itô s formula, we have d φ, Z n t = d φ(x n t (, W )), ζ t = dφ(xt n (, W )), ζ t + φ(xt n (, W )), dζ t = ρ n() φ, Zt n φ dt + (X n t (, W ))h n (y ), ζ t W (dtdy) 2 + φ(xt n (, W )), dζ t. Hence, U n t (φ) is a martingale with quadratic variation process U n (φ) t = = φ (Xs n t 2 (, W ))h n (y ), ζ s dsdy + σφ 2 (X n s (, W )), ζ s ds du ρ n (y z)φ (y)φ (z)zu n (dy)zu n (dz) + σφ 2, Z n u du. R 2 Theorem 4.2 For all t >, Z t is of purely-atomic type. Proof: It is well-known that ζ t = i I(t) ξi tδ x i is of purely-atomic type (cf. [2] and [13]). Hence Zt n = i I(t) ξi tδ X n t (x i,w ). Taking a limit, it is clear that Z t is of purely-atomic type. Mimicking [6], we consider the following SPDE φ, µ t = φ, µ t + + 1 2 ρ() + i I(t ) i I(t ) ρ() t φ, µ u du t φ (x i (u))ξ i (u)dw i (u) (4.1) t φ(x i (u)) σ( )ξ i (u)db i (u), where (W 1, W 2, ) are coalescing Brownian motions independent of {B i : i 1}. If x i (t) = ρ()w i (t) and ξ i (t) is the Feller s branching diffusion generated by B i, it is easy to see that µ t = ξ i (t)δ xi (t) (4.2) satisfies (4.1). It is easy to verify that the solution of (4.1) satisfies the martingale problem (3.8). Therefore, weak uniqueness holds for the solution to (4.1). The next theorem shows the non-strong-uniqueness of the solution. Theorem 4.3 Let t = and µ = i I ξ i()δ xi (), where I is a countable set. pathwise uniqueness for the SPDE (4.1) does not hold. Then, the 13

Proof: Let τ 1 = inf{t : i j, W i (t) = W j (t)}. Define τ 2 similarly. Suppose W i and W j (i < j) meet at time τ 1. For t τ 1, we define ξ k (t) = ξ k (t) for all k. For < τ 1 < t τ 2, if k i, j we define ξ k (t) = ξ k (t); otherwise, we define that ξ j (t) = and let ξ i (t) be the unique solution of the following SDE ξ i (t) = ξ i (τ 1 ) + ξ j (τ 1 ) + σ( ) τ 1 ξ i (u)db i (u). Then µ t = ξi (t)δ xi (t) (4.3) is another solution to (4.1). Therefore, the pathwise uniqueness for the SPDE (4.1) does not hold. To derive an SPDE with strong uniqueness, we need to modify the Brownian driving system in (4.1). Definition 4.1 W = { W k } is a system of killing Brownian motions (KBM) if each starts with an independent Brownian motion until a pair of them meet; at that time, the process with higher index will be killed and the indexes of the other higher indexed processes is lowered by 1. The system continues to evolve in this fashion. B = { B k } is a system of adjoint (to W ) killing Brownian motions (AKBM) if each starts with an independent Brownian motion. A member will be killed when the corresponding member in W is killed. Now we modify (4.1) and consider φ, µ t = φ, µ t + + 1 2 ρ() + i I(t ) i I(t ) ρ() t φ, µ u du t φ (x i (u))ξ i (u)d W i (u) (4.4) t φ(x i (u)) σ( )ξ i (u)d B i (u). If we construct W and B in an obvious way, it is then clear that µ t defined by (4.3) is a solution to (4.4). Theorem 4.4 Let t = and µ = i I ξ i()δ xi (), where I is a countable set. SPDE (4.4) has a pathwise unique solution. Then, the Proof: Let µ t be a solution and τ 1 = inf{t : i j, W i (t) = W j (t)}. 14

Now we prove the uniqueness of the solution before time τ 1 by adapting the technique of [6] to the present setup. Let ɛ = inf{ x i () x j () : i j I} and η 1 = inf{t [, τ 1 ) : x i (t) x i () 1 3 ɛ for some i I}. Then η 1 is a stopping time. Take φ such that its support is within 2 3 ɛ of x i (). Then ξ i (t)φ(x i (t)) = ξ i ()φ(x i ()) + ρ() φ (x i (u))ξ i (u)d W i (u) + 1 2 ρ() ξ i (u)φ (x i (u))du + σ( ) Also, take φ(x) = 1 for x within 1 3 ɛ of x i (). Then for t η 1 Applying Itô s formula, we then have ξ i (t) = ξ i () + σ( ) ξi (u)d B i (u). dφ(x i (t)) = ρ()φ (x i (t))d W i (t) + 1 2 ρ()φ (x i (t))dt. φ(x i (u)) ξ i (u)d B i (u). This implies that x i (t) = ρ() W i (t). By the definition of η 1, we have {x i (η 1 ) : i I} are all distinct; hence we may start from η 1 and define η 2 accordingly. As in [6], we then can prove that η n converges to τ 1 and get the uniqueness for t τ 1. Continuing this procedure, we get the uniqueness for all t. The argument for countable I is the same as that at the end of the proof of Theorem 4.1 in [6]. Acknowledgment The authors would like to express their gratitude to Professor Steven N. Evans and the anonymous referee for their helpful suggestions. In particular, the authors thank Professor Kenneth A. Ross for his comments and suggestions. References [1] Arratia, R. (1979). Coalescing Brownian motion on the line. Ph.D. thesis, University of Wisconsin, Madison, 1979. [2] Dawson, D. A. and Fleischmann, K. (1988). Strong clumping of critical space-time branching models in subcritical dimension. Stochastic Process. Appl., 3:193 28, 1988. [3] Dawson, D. A. and Perkins, E. A. (1999). Measure-valued processes and renormalization of branching particle systems. Stochastic Partial Differential Equations: Six Perspectives, Mathematical Surveys and Monographs,Vol. 64,Amer. Math. Soc.,Providence:45 16, 1999. [4] Dawson, D. A.; Li, Z. (23). Construction of immigration superprocesses with dependent spatial motion from one-dimensional excursions. Probab. Th. Rel. Fields, 127,1:37 61, 23. 15

[5] Dawson, D. A.; Li, Z. and Wang, H. (21). Superprocesses with dependent spatial motion and general branching densities. Electron. J. Probab., V6, 25:1 33, 21. [6] Dawson, D. A.; Li, Z. and Wang, H. (23). A degenerate stochastic partial differential equation for the purely atomic superprocess with dependent spatial motion. To appear in Infinite Dimensional Analysis, Quantum Probability and Related Topics, 23. [7] Dawson, D. A.; Li, Z. and Zhou, X. W. (22). Superprocesses with coalescing Brownian spatial motion as large scale limits. Submitted, 22. [8] Dawson, D. A.; Vaillancourt, J. and Wang, H. (2). Stochastic partial differential equations for a class of interacting measure-valued diffusions. Ann. Inst. Henri Poincaré, Probabilités et Statistiques, 36,2:167 18, 2. [9] Konno, N. and Shiga, T. (1988). Stochastic partial differential equations for some measurevalued diffusions. Probab. Th. Rel. Fields, 79:21 225, 1988. [1] Mitoma, I. (1985). An -dimensional inhomogeneous Langevin s equation. Journal of Functional Analysis, 61:342 359, 1983. [11] Skoulakis, G. and Adler, R.J. (21). Superprocesses over a stochastic flow. Ann. Appl. Probab., 11,2:488 543, 21. [12] Walsh, J. B. (1986). An introduction to stochastic partial differential equations. Lecture Notes in Math., 118:265 439, 1986. [13] Wang, H. (1997). State classification for a class of measure-valued branching diffusions in a Brownian medium. Probab. Th. Rel. Fields, 19:39 55, 1997. [14] Wang, H. (1998). A class of measure-valued branching diffusions in a random medium. Stochastic Anal. Appl., 16(4):753 786, 1998. [15] Wang, H. (22). State classification for a class of interacting superprocesses with location dependent branching. Electron. Comm. Probab., 7:157 167, 22. [16] Wang, H. (23). Singular spacetime Itô integral and a class of singular interacting branching particle systems. Infin. Dimens. Anal. Quantum Probab. Relat. Top., Vol.6, N 2:321 335, 23. [17] Xiong, J. (22). Long-term behavior for superprocesses over a stochastic flow. Submitted, 22. [18] Xiong, J. (22). A stochastic log-laplace equation. Submitted, 22. [19] Xiong, J., Zhou, X. W. (22). On the duality between coalescing Brownian motions. Submitted, 22. 16