Support theorem on stochastic flows with interaction

We prove an analogue of the Stroock–Varadhan theorem for stochastic flows describing a motion of interacting particles in a random media. A version of the Itˆo lemma for functions on a measure-valued process is obtained.

Gespeichert in:
Bibliographische Detailangaben
Datum:2006
1. Verfasser: Pilipenko, A.Yu.
Format: Artikel
Sprache:English
Veröffentlicht: Інститут математики НАН України 2006
Online Zugang:http://dspace.nbuv.gov.ua/handle/123456789/4448
Tags: Tag hinzufügen
Keine Tags, Fügen Sie den ersten Tag hinzu!
Назва журналу:Digital Library of Periodicals of National Academy of Sciences of Ukraine
Zitieren:Support theorem on stochastic flows with interaction / A.Yu. Pilipenko // Theory of Stochastic Processes. — 2006. — Т. 12 (28), № 1-2. — С. 127–141. — Бібліогр.: 13 назв.— англ.

Institution

Digital Library of Periodicals of National Academy of Sciences of Ukraine
id irk-123456789-4448
record_format dspace
spelling irk-123456789-44482009-11-11T12:00:39Z Support theorem on stochastic flows with interaction Pilipenko, A.Yu. We prove an analogue of the Stroock–Varadhan theorem for stochastic flows describing a motion of interacting particles in a random media. A version of the Itˆo lemma for functions on a measure-valued process is obtained. 2006 Article Support theorem on stochastic flows with interaction / A.Yu. Pilipenko // Theory of Stochastic Processes. — 2006. — Т. 12 (28), № 1-2. — С. 127–141. — Бібліогр.: 13 назв.— англ. 0321-3900 http://dspace.nbuv.gov.ua/handle/123456789/4448 519.21 en Інститут математики НАН України
institution Digital Library of Periodicals of National Academy of Sciences of Ukraine
collection DSpace DC
language English
description We prove an analogue of the Stroock–Varadhan theorem for stochastic flows describing a motion of interacting particles in a random media. A version of the Itˆo lemma for functions on a measure-valued process is obtained.
format Article
author Pilipenko, A.Yu.
spellingShingle Pilipenko, A.Yu.
Support theorem on stochastic flows with interaction
author_facet Pilipenko, A.Yu.
author_sort Pilipenko, A.Yu.
title Support theorem on stochastic flows with interaction
title_short Support theorem on stochastic flows with interaction
title_full Support theorem on stochastic flows with interaction
title_fullStr Support theorem on stochastic flows with interaction
title_full_unstemmed Support theorem on stochastic flows with interaction
title_sort support theorem on stochastic flows with interaction
publisher Інститут математики НАН України
publishDate 2006
url http://dspace.nbuv.gov.ua/handle/123456789/4448
citation_txt Support theorem on stochastic flows with interaction / A.Yu. Pilipenko // Theory of Stochastic Processes. — 2006. — Т. 12 (28), № 1-2. — С. 127–141. — Бібліогр.: 13 назв.— англ.
work_keys_str_mv AT pilipenkoayu supporttheoremonstochasticflowswithinteraction
first_indexed 2025-07-02T07:41:28Z
last_indexed 2025-07-02T07:41:28Z
_version_ 1836520139376295936
fulltext Theory of Stochastic Processes Vol. 12 (28), no. 1–2, 2006, pp. 127–141 UDC 519.21 A. YU. PILIPENKO SUPPORT THEOREM ON STOCHASTIC FLOWS WITH INTERACTION We prove an analogue of the Stroock–Varadhan theorem for stochastic flows describ- ing a motion of interacting particles in a random media. A version of the Itô lemma for functions on a measure-valued process is obtained. Introduction Let us consider a flow of interacting particles in a random media. Denote by xt(u) a position at an instant of time t of a particle starting from a point u ∈ R d. Let μt be a distribution of the mass of particles at the moment t. Suppose that each particle does not change its mass in time. So μt is the image of the measure μ0 under the mapping xt, i.e. μt = μ0 ◦ x−1 t . Assume that the motion of each particle depends not only on its position at the current time moment but also on the distribution of the total mass of particles and satisfies the following system of stochastic equations: (0.1) ⎧⎪⎨⎪⎩ dxt(u) = a(xt(u), μt)dt + ∑m k=1 bk(xt(u), μt)dwk(t), x0(u) = u, u ∈ R d, μt = μ0 ◦ x−1 t , t ∈ [0, T ], where a, bk : R d × P → R d, μ0 ∈ P ; {wk(t)}k=1,m are independent one-dimensional Wiener processes. Here, P is the space of probability measures on R d with a topology of weak convergence. Systems of stochastic equations of the type (0.1) for flows of interacting particles were introduced by A.A. Dorogovtsev and P. Kotelenez [1]. Measure-valued processes which are solutions of (0.1) were initially obtained in [2] as some weak limits of the finite number of systems of interacting particles. However, there was not considered any flow describing the individual behavior of particles for the limit process. The theorem on uniqueness and existence can be proved under some natural conditions on the coefficients of (0.1) [3]. Moreover, if a, bk are smooth enough, then a process xt(u) possesses a modification which is differentiable with respect to u (the corresponding number of times), and its derivatives ∂kxt(u) ∂uk are continuous in (t, u). So, we can consider the process xt(·), t ≥ 0 as a continuous stochastic process with values in Ck(Rd, Rd). Here, a space Ck(Rd, Rd) of k times continuously differentiable functions is provided by the topology of uniform convergence of functions and their derivatives on compact sets. On the other hand, x = xt(u) can be considered as a random element in the space Ck = C([0, T ]; Ck(Rd, Rd)). The aim of this article is a characterization of the support of a stochastic flow xt(u). 2000 AMS Mathematics Subject Classification. Primary 60K35, 60G57. Key words and phrases. Support of stochastic flow, systems of interacting particles. 127 128 A. YU. PILIPENKO The corresponding result for the usual stochastic differential equation dξt = a(ξt)dt + m∑ k=1 bk(ξt) ◦ dwk(t) is given by the well-known Stroock—Varadhan theorem [4]. It asserts that, under some smoothness assumptions on the coefficients of Eq. (0.1), the support of ξ(t) distribution is equal to the closure in C([0, T ], Rd) of the set {xψ, ψ is piecewise smooth}, where xψ is a solution of the non-random equation dxψ(t) = a(xψ(t))dt + m∑ k=1 bk(xψ(t))dψ(t). A similar result for the stochastic flow xt(u) generated by s.d.e. depending on the initial condition (0.2) { dxt(u) = a(xt(u))dt + ∑ k bk(xt(u)) ◦ dwk(t), x0(u) = u. was obtained by H.Kunita [5]. Stochastic flows generated by the equation (0.3) dxt(u) = ∫ Rd α(xt(u), xt(v))μ(dv)dt + m∑ k=1 (∫ Rd βk(xt(u), xt(v))μ(dv) ) ◦ dwk(t) with interaction were considered in [6, 7]. Note that Eq. (0.3) is a partial case of Eq. (0.1) with a(u, ν) = ∫ Rd α(u, v)ν(dv), bk(u, ν) = ∫ Rd βk(u, v)ν(dv). Usually, it is convenient to formulate a Stroock—Varadhan-type theorem for stochastic equations written in the Stratonovich form. Moreover, it is always supposed that the coefficients of equations are differentiable more than one time. The coefficients of (0.1) depends on the measure-valued process μt. So we need some version of the Itô formula for functions depending on a measure-valued process. We prove the corresponding result in §2. The rest of the proof of the Stroock—Varadhan theorem for a stochastic flow xt(u) is quite standard. We estimate the “small-balls” probability in §3 and prove that, for each ε > 0 and a piecewise smooth function ψ, the conditional probability P ( sup |u|<R sup t∈[0,T ] |∇(xt(u) − xψ t (u))|/ sup t∈[0;T ] |w(t) − ψ(t)| ) converges to zero as δ → 0 + . This gives us the inclusion {xψ, ψ is piecewise smooth} ⊂ supp Px. The support is a closed set, so it contains the closure of {xψ, ψ is piecewise smooth}, Section 4 is devoted to the proof of the approximation theorem. Particularly, we prove that if a sequence of processes {ξε k(t), ε > 0} is such that wε k(t) = ∫ t 0 ξε k(s)ds converges as ε → 0 to a Wiener process wk(t) in some sense, then a solution of Eq. (0.1) with wε(t) instead of w(t) converges to the solution of (0.1), but may be with corrected coefficients. It is worth to mention one principal difference between Eqs. (0.1) and (0.2). SUPPORT THEOREM ON STOCHASTIC FLOWS WITH INTERACTION 129 While studying Eq. (0.2), we can consider different starting points independently. Then the differentiability with respect to u is just a differentiability with respect to a parameter. However, all the equations in (0.1) are coupled, and (0.1) should be consid- ered as an infinite (even uncountable) system of stochastic equations or as a stochastic equation in the functional space. 1. Preliminaries Let P be a set of probability measures on R d with a topology of weak convergence. Definition 1.1. A pair (xt, μt) is said to be a solution to Eq. (0.1) if 1) mapping x = xt(u, ω) : [0, T ] × R d × Ω → R d is measurable in a triple of its arguments; 2) processes xt(u), μt are adapted to a filtration generated by the Wiener processes wk(t), k = 1, m; 3) for each u ∈ R d with probability one, the equality xt(u) = u + ∫ t 0 a(xs(u), μs)ds + m∑ k=1 ∫ t 0 bk(xs(u), μs)dwk(s), t ∈ [0, T ], where μt = μ0 ◦ x−1 t is the image of the initial measure μ0 under the random mapping xt, is satisfied. Let us introduce the Wasserstein metric on the space P : γ(μ1, μ2) := inf κ∈Q(μ1,μ2) ∫ Rd ∫ Rd |u − v| ∧ 1κ(du, dv), where Q(μ1, μ2) is a set of probability measures on R d × R d with marginals μ1 and μ2. It is well known that the metric space (P , γ) is complete and separable [8], and the convergence in metric γ is equivalent to the weak convergence of measures. Let us formulate the theorem of existence and uniqueness for the solution of Eq. (0.1). Theorem 1.1 [3]. Assume that ∃L > 0 ∀u1, u2 ∈ R d ∀μ1, μ2 ∈ P : (1.1) |a(u1, μ1) − a(u2, μ2)| + m∑ k=1 |bk(u1, μ1) − bk(u2, μ2)| ≤ ≤ L(|u1 − u2| + γ(μ1, μ2)). Then there exists a unique solution for Eq. (0.1). Further, we always assume that the conditions of Theorem 1.1 are satisfied. Put ã(x, t) := a(x, μt(ω)), b̃k(x, t) := bk(x, μt(ω)). Observe that if xt, μt satisfy (0.1) then xt satisfies the following Itô equation: (1.2) { dxt(u) = ã(xt(u), t)dt + ∑m k=1 b̃k(xt(u), t)dwk(t), t ≥ 0, x0(u) = u. The functions ã, b̃k are continuous and satisfy the Lipschitz condition in x. So, the flow xt(u) has a modification which is continuous in (t, u) (cf.[5]). Moreover, if the coefficients a, bk are (n + 1) times continuously differentiable with respect to u and have bounded derivatives, then the partial derivatives ∂jxt(u) ∂uj , j = 1, n exist and are continuous in (t, u). Therefore, we can consider the process xt(·) as a continuous process with values in Cn(Rd, Rd) or as a random element in Cn = C([0, T ], Cn(Rd, Rd)), where the space of 130 A. YU. PILIPENKO functions Cn(Rd, Rd) differentiable n-times has a topology of uniform convergence of the functions and their derivatives on compact sets. Usually, to study the support of a solution for some stochastic equation, one needs a stronger condition of smoothness than the Lipschitz condition. That’s why we need the following definition of a derivative for functions depending on a measure-valued argument [9]. Let M be a space of finite measures on R d with a topology of weak convergence, and let F be a continuous function from M to R. Definition 1.2. Assume that, for each μ ∈ M, u ∈ R d, there exists the limit (1.3) δF (μ) δμ (u) := lim ε→0+ F (μ + εδu) − F (μ) ε which is continuous in (μ, u), where δu is a unit measure concentrated at the point u ∈ R d. Then function F is said to be continuous differentiable, and δF δμ is its derivative. Higher order derivatives δnF δμn (u1, . . . , un) are defined iteratively. Remark 1. Let μt be a process from (0.1). Then the full mass of μt is not changing in time, μt(Rd) = μ0(Rd) = 1. So, in general, in order to consider Eq. (0.1), it is natural to know the coefficients in the space R d ×P only, but not in the larger space R d ×M. Note also that if μ ∈ P , then μ+εδu /∈ P as ε �= 0. We can replace the definition of continuous differentiability (1.3) remaining in a class P by using Newton—Leibnitz formula: ∃ δF δμ ∈ C(P × R d) ∀μ1, μ2 ∈ P : (1.4) F (μ2) − F (μ1) = ∫ 1 0 ∫ Rd δF (τμ2 + (1 − τ)μ1) δμ (u)(μ2 − μ1)(du)dτ. The proof of (1.4) can be done for discrete measures at first. Then it is not difficult to extend (1.4) for an arbitrary measure by continuity (recall that δF δμ is continuous in (μ, u)). However, formula (1.4) seems to be less natural than (1.3). Moreover, if the function F is defined on P and satisfies (1.4) for all μ1, μ2 ∈ P , then its extension F̃ (μ) := F ( μ μ(Rd) ) to the space of all finite measures M is continuous differentiable in the sense of Definition 1.2. The following statement will be useful in our investigations. Lemma 1.1. Assume that F : P → R satisfies (1.4), where the function δF δμ is bounded and satisfies the Lipschitz condition in u : ∃L1 ∀μ ∈ P ∀u ∈ R d : ∣∣∣∣δF (μ) δμ (u) ∣∣∣∣ ≤ L1; ∃L2 ∀μ ∈ P ∀u1, u2 ∈ R d : ∣∣∣∣δF (μ) δμ (u1) − δF (μ) δμ (u2) ∣∣∣∣ ≤ L2|u1 − u2|. Let ν ∈ P be a probability measure. Then there exists a constant L = L(F ) such that, for all measurable functions f1, f2 : R d → R d, the following inequality holds: (1.5) |F (ν2) − F (ν1)| ≤ L ∫ Rd |f1(u) − f2(u)| ∧ 1ν(du), where νi = ν ◦ f−1 i , i = 1, 2. Proof. Put μτ = τν2 + (1 − τ)ν1. Then, due to (1.4), we have |F (ν2) − F (ν1)| ≤ ∫ 1 0 ∣∣∣∣∫ Rd δF (μτ ) δμ (u)(ν2(du) − ν1(du)) ∣∣∣∣ dτ = SUPPORT THEOREM ON STOCHASTIC FLOWS WITH INTERACTION 131 = ∫ 1 0 ∣∣∣∣∫ Rd ( δF (μτ ) δμ (f2(u)) − δF (μτ ) δμ (f1(u)) ) ν(du) ∣∣∣∣ dτ ≤ ≤ ∫ 1 0 ∫ Rd (L2|f2(u) − f1(u)|) ∧ 2L1ν(du)dτ ≤ ≤ (L2 + 2L1) ∫ Rd |f2(u) − f1(u)| ∧ 1ν(du). Lemma 1.1 is proved. Remark 2. While studying carefully the proof of Theorem 1.1 (see [3]), one can notice that it is sufficient to demand a condition of type (1.5) instead of (1.1). So, the existence of the bounded derivatives ∂a(x,μ) ∂x , ∂ ∂u δa(x,μ) δμ (u), ∂bk(x,μ) ∂x , and ∂ ∂u δbk(x,μ) δμ (u) is enough for Theorem 1.1. This condition will always be supposed in §2, §3. However, it seems to the author that the formulation of Theorem 1.1 is methodologically more useful than a similar theorem with the condition ∃L > 0 ∀x, u ∈ R d ∀μ :∣∣∣∣∂a(x, μ) ∂x ∣∣∣∣ + ∣∣∣∣ ∂ ∂u δa(x, μ) δμ (u) ∣∣∣∣ + m∑ k=1 (∣∣∣∣∂bk(x, μ) ∂x ∣∣∣∣ + ∣∣∣∣ ∂ ∂u δbk(x, μ) δμ (u) ∣∣∣∣) ≤ L instead of (1.1). 2. Itô formula for functions of measure-valued argument In this section, we prove a version of the Itô formula for a process F (ξt, μt), where μt is defined in (0.1), and the process ξt has a stochastic differential (2.1) dξt = αtdt + m∑ j=1 βj t dwj(t). The main result is as follows: Theorem 2.1. Assume that a function F : R k × M → R is such that the following derivatives exist, bounded, and continuous in all arguments: ∂F (x, μ) ∂x , ∂2F (x, μ) ∂x2 , δF (x, μ) δμ (u), δ2F (x, μ) δμ2 (u, v), ∂ ∂u δF (x, μ) δμ (u), ∂2 ∂u2 δF (x, μ) δμ (u), ∂2 ∂x∂u δF (x, μ) δμ (u), ∂2 ∂u∂v δ2F (x, μ) δμ2 (u, v). Suppose that a process ξt = (ξ1 t , . . . , ξk t ) has the stochastic differential (2.1), and the functions a, bk satisfy the assumptions of Theorem 1.1 and are bounded. Let us introduce the following notations: 〈f, μ〉 = ∫ fdμ, σ2 ξ (t) = m∑ l=1 βl t (βl t) ∗, σ2 ξμ(t, x, μ) = m∑ l=1 βl tb ∗ l (x, μ), σ2 μμ(x, y, μ) = m∑ l=1 bl(x, μ)b∗l (y, μ), σ2 μ(x, μ) = σ2 μμ(x, x, μ), where ∗ is the transposition operator of a matrix. Set L1(t)F (x, μ) = F ′ x(x, μ)αt + 1 2 sp(F ′′ xx(x, μ)σ2 ξ,ξ(t)), 132 A. YU. PILIPENKO L2F (x, μ) = ∫ Rd ( ∂ ∂u δF (x, μ) δμ (u)a(u, μ) + 1 2 sp ∂2 ∂u2 δF (x, μ) δμ (u)σ2 μ(u, μ) ) μ(du)+ (2.2) + 1 2 ∫ Rd ∫ Rd sp ∂2 ∂u∂v δ2F (x, μ) δμ2 (u, v)σ2 μμ(u, v, μ)μ(du)μ(dv). Then dF (ξt, μt) = L1(t)F (ξt, μt)dt + F ′ 1(ξt, μt) m∑ l=1 βl tdwl(t)+ +L2F (ξt, μt)dt + m∑ l=1 ∫ Rd ∂ ∂u δF (ξt, μt) δμ (u)bl(u, μt)μt(du)dwl(t)+ (2.3) + 1 2 sp ∫ Rd ∂2 ∂ξ∂u δF (ξt, μt) δμ (u)σ2 ξμ(t, u, μt)μ(du)dt. Proof. Let us verify formula (2.3) for a function F (x, μ) = F (μ) depending on the second argument only. The general case can be proved similarly but with additional routine calculations. Let μn = n∑ j=1 cj,nδuj,n , n ∈ N, cj,n ≥ 0, ∑ j cj,n = 1 be a sequence of discrete measures which converges weakly to μ0 as n → ∞. Denote, by xn t (u), a solution of Eq. (0.1) with initial measure μn instead of μ0, μn t := μn ◦ (xn t )−1. The plan of the proof is to obtain the Itô formula for μn t at first, and then to pass to a limit as n → ∞. Remark 3. The Itô formula for superprocesses with interaction was obtained in [10]. There, a function F (μ) was approximated by polynomials in μ, and the proof is a hard technical work. It is worth mentioning that the process μt is neither a superprocess nor the Fleming—Viot process. This fact can be easily checked if we write down and compare their generators [3,9]. Observe that the measure μn t is also discrete, μn t = ∑n j=1 cj,nδxn t (uj,n). To simplify notations, we will write cj , uj instead of cj,n, uj,n. The processes xn t (uk), k = 1, n satisfy a finite system of stochastic equations dxn t (uk) = a(xn t (uk), n∑ j=1 cjδxn t (uj))dt+ + m∑ l=1 bl(xn t (uk), n∑ j=1 cjδxn t (uj))dwl(t), k = 1, n. Consider a function f : R nd → R defined as follows: f(v1, . . . , vn) = F ( n∑ j=1 cjδvj ). Then F (μn t ) = f(xn t (u1), . . . , xn t (un)). Let us check that f is twice continuously differ- entiable. Assume for simplicity that n = 2 and calculate 〈 ∂f ∂v1 , e〉Rd , where e ∈ R d (recall that v1 ∈ R d). SUPPORT THEOREM ON STOCHASTIC FLOWS WITH INTERACTION 133 Set ν = c1δv1 + c2δv2 , ντ = (1− τ)(c1δv1 + c2δv2) + τ(c1δv1+εe + c2δv2). Formula (1.4) implies lim ε→0 f(v1 + εe, v2) − f(v1, v2) ε = lim ε→0 1 ε ∫ 1 0 (∫ Rd δF (ντ ) δμ (u)(c1δv1+εe + c2δv2)(du) − ∫ Rd δF (ντ ) δμ (u)(c1δv1 + c2δv2)(du) ) = lim ε→0 1 ε ∫ 1 0 c1 ( δF (ντ ) δμ (v1 + εe) − δF (ντ ) δμ (v1) ) dτ = c1 ( ∂ ∂v1 δF (ν) δμ , e ) Rd . Here, while grounding the passage to the limit, we use the continuous differentiability of δF (μ) δμ (u) with respect to the parameter u and the boundedness of its derivative. So, we have verified that (2.4) δf ∂vj = cj∇δF (ν) δμ (vj), where ∇ is the derivative with respect to the argument vj . Analogously, (2.5) ∂2f ∂vi∂vj = cicj∇1∇2 δ2F (ν) δμ2 (vi, vj) + cj∇2 δF (ν) δμ (vj)δij , where δij is the Kronecker symbol. Let us apply the usual Itô lemma to the process ηn t = f(xn t (u1), . . . , xn t (un)) = F ( n∑ j=1 cjδxn t (uj)) = F (μn t ) and use (2.4), (2.5): dηn t = ∑ i ci∇δF (μn t ) δμ (xn t (ui)) ( a(xn t (ui), μn t )dt + ∑ l bl(xn t (ui), μn t )dwl(t) ) + + 1 2 ∑ i,j cicjsp ( ∇1∇2 δ2F (μn t ) δμ2 (xn t (ui), xn t (uj))σ2 μμ(xn t (ui), xn t (uj), μn t ) ) dt+ (2.6) + 1 2 ∑ i cisp ( ∇2 δF (μn t ) δμ (xn t (ui))σ2 μ(xn t (ui), μn t ) ) dt. Observe that, for each g = g(u), h = h(u, v), we have ∑ i cig(xn t (ui)) = ∫ Rd g(u)μn t (du), ∑ i,j cicjh(xn t (ui), xn t (uj)) = ∫ Rd ∫ Rd h(u, v)μn t (du)μn t (dv). 134 A. YU. PILIPENKO So, the expression in (2.6) is equal to (2.7) dηn t = L2F (μn t )dt + m∑ l=1 ∫ Rd ∂ ∂u δF (μn t ) δμ (u)bl(u, μn t )μn t (du)dwl(t), where the operator L2 is defined in (2.2). The initial distributions μn are weakly convergent to μ0. Therefore [3], for each t ≥ 0, we have the convergence μn t to μt in probability. Taking a subsequence, if it is needed, it can be assumed without loss of generality that, for almost all ω ∈ Ω and almost all t ≥ 0, μn t (ω) ⇒ μt(ω), n → ∞. All functions in the differentiated expression in (2.7) are bounded and continuous. The proof that we can pass to a limit under the integral sign in (2.7) follows from the next lemma. Lemma 2.1. Let {νn, n ≥ 1} ⊂ P be a non-random sequence of probability measures which converges weakly to ν0. Assume that the function g : R d × P → R is continuous and bounded. Then ∫ Rd g(u, νn)νn(du) → ∫ Rd g(u, ν0)ν0(du), n → ∞. Proof. ∣∣∣∣∫ Rd g(u, νn)νn(du) − ∫ Rd g(u, ν0)ν0(du) ∣∣∣∣ ≤ ≤ ∫ Rd |g(u, νn) − g(u, ν0)|νn(du) + ∣∣∣∣∫ Rd g(u, ν0)νn(du) − ∫ Rd g(u, ν0)ν0(du) ∣∣∣∣ . The second item converges to zero due to the weak convergence νn ⇒ ν0, n → ∞. To estimate the first item, observe that the set {νn, n ≥ 0} is a weak compact. By the Prokhorov theorem, for each ε > 0, there exists a compact Kε ⊂ R d such that νn(Rd \ Kε) < ε for all n ≥ 0. The function g is continuous on a compact Kε × {νn; n ≥ 0} ⊂ R d ×P and, hence, it is uniformly continuous. Thus, sup u∈Kε |g(u, νn) − g(u, ν0)| → 0, n → ∞. Therefore, lim n→∞ ∫ Rd |g(u, νn) − g(u, ν0)|νn(du) ≤ ≤ lim n→∞ ∫ Kε |g(u, νn) − g(u, ν0)|νn(du) + lim n→∞ ∫ Rd\Kε (|g(u, νn)| + |g(u, ν0)|)νn(du) ≤ ≤ 2 sup u∈Rd sup ν∈P |g(u, ν)|ε. The number ε > 0 is arbitrary. Thus, Lemma 2.1 and also Theorem 2.1 are proved. 3. Lower estimate for support of xt(u) Let ψ = (ψ1, . . . , ψm) : [0, T ] → R m be a piecewise smooth function. Denote, by xψ , a solution of the following (deterministic) equation (3.1) ⎧⎪⎨⎪⎩ dxψ t (u) = ã(xψ t (u), μψ t )dt + ∑m k=1 bk(xψ t (u), μψ t )ψk(t), u ∈ R d, μψ t = μ0 ◦ (xψ t )−1, t ∈ [0, T ], xψ 0 (u) = u, SUPPORT THEOREM ON STOCHASTIC FLOWS WITH INTERACTION 135 where (3.2) ã(x, μ) = a(x, μ) − 1 2 m∑ k=1 ( ∂bk(x, μ) ∂x bk(x, μ) + ∫ Rd ∂ ∂v δbk(x, μ) δμ (v)bk(v, μ)μ(dv) ) is a corrected coefficient. The aim of this section is to show that the support suppx of the flow xt(u) contains the set S = {xψ : ψ is piecewise smooth}. Note that if we replace the integral with respect to ψk(t) by the Stratonovich integral w.r.t. wk(t), then we obtain Eq. (0.1). Theorem 3.1. Assume that the coefficients a, bk are such that the derivatives ∂α ∂ūα ∂β ∂xβ δja(x, μ) δμj (ū) and ∂α ∂ūα ∂β ∂xβ δjbk(x, μ) δμj (ū) are bounded and continuous in all their arguments, where j = 0, n + 3, ū = (u1, . . . , uj), α = (α1, . . . , αj), αi ≥ 0, β = (β1, . . . , βd), βi ≥ 0, α1 + · · · + αj + β1 + · · ·+ βd ≤ n + 3. Then the support of xt considered as a random element in C([0, T ], Cn(Rd, Rd)) contains the set S. Proof. We restrict ourselves only to the case n = 0 and ψ ≡ 0. The case of arbitrary ψ can be considered with the use of the Girsanov theorem (see the reasoning of Theorem [11]). The reasoning for arbitrary n is similar to the case n = 0 (see [6] for details). Lemma 3.1. Assume that the conditions of Theorem 3.1 are satisfied. Then, for each ball U ⊂ R d and each ε > 0, δ > 0, we have the following convergence of conditional probabilities: lim δ→0+ P ⎛⎜⎝ sup u∈U t∈[0,T ] ∣∣∣∣∫ t 0 bk(xs(u), μs) ◦ dwk(s) ∣∣∣∣ > ε/‖w‖ < δ ⎞⎟⎠ = 0, where | · | is a norm in R d, ‖w‖ = supt∈[0,T ] maxi=1,m |wi(t)| is a norm in C([0, T ], Rm). Proof. Due to the Sobolev embedding theorems [12], it is enough to verify that lim δ→0+ P ( sup t∈[0,T ] ‖ ∫ t 0 bk(xs(·), μs) ◦ dwk(s)‖W 1 p (U) ≥ ε/‖w‖ < δ ) = 0, where p > d, ‖f‖W 1 p (U) = (∫ U (|f |p + |∇f |p)dx )1/p . The proof of the corresponding fact can be done similarly [7]. We need to use the Itô formula (2.3), boundedness of a, bk and their derivatives, and some estimates of the moments for xt(u) and ∂xt(u) ∂u (to estimate the moments, one can apply results in [5], §4.5, 4.6, to (1.2)). Let us show now that x0 ∈ suppx. To verify this, it is sufficient to check that, for every ball U ⊂ R d and ε > 0, P ( sup t∈[0,T ] sup u∈U |xt(u) − x0 t (u)| < ε ) > 0. 136 A. YU. PILIPENKO Denote ∑m k=1 ∫ t 0 bk(xs(u), μs) ◦ dwk(s) by ot(u). Then xt(u) = u + ∫ t 0 ã(xs(u), μs)ds + ot(u), x0 t (u) = u + ∫ t 0 ã(x0 s(u), μ0 s)ds, where ã is defined in (3.2), x0 t (u) is a solution of (3.1) with ψ ≡ 0, μ0 t = μ0 ◦ (x0 t ) −1. Let us use inequality (1.5) to estimate the difference |xt(u) − x0 t (u)|, u ∈ U : |xt(u) − x0 t (u)| ≤ L ∫ t 0 ( |xs(u) − x0 s(u)| + ∫ Rd |xs(v) − x0 s(v)| ∧ 1μ0(dv) ) ds + |ot(u)| ≤ ≤ 2L ∫ t 0 (sup v∈U |xs(v) − x0 s(v)| + μ(Rd \ U))ds + sup s∈[0,t] v∈U |os(v)|, where L > 0 is a constant. So, by the Gronwall lemma, we get the estimate sup v∈U |xt(v) − x0 t (v)| ≤ ⎛⎜⎝ sup v∈U s∈[0,t] |os(v)| + μ(Rd \ U) ⎞⎟⎠ e2LT . Let ε > 0, and let U ⊂ R d be fixed. Choose a ball Ũ ⊃ U such that μ(Rd \ Ũ)e2LT < ε 2 . Then we choose δ > 0 such that P ⎛⎜⎝ sup u∈U t∈[0,T ] |ot(u)| < e−2LT ε 2 / ‖w‖ < δ ⎞⎟⎠ > 0. Therefore, P ⎛⎜⎝ sup u∈U t∈[0,T ] |xt(u) − x0 t (u)| ≤ ε ⎞⎟⎠ ≥ P ⎛⎜⎝ sup u∈U t∈[0,T ] |xt(u) − x0 t (u)| ≤ ε ⎞⎟⎠ ≥ ≥ P ⎛⎜⎝( sup u∈U t∈[0,T ] |ot(u)| + μ(Rd \ Ũ))e2LT ≤ ε / ‖w‖ < δ ⎞⎟⎠ · P (‖w‖ < δ) > 0. Moreover, it is easy to verify that P ⎛⎜⎝ sup u∈U t∈[0,T ] |xt(u) − x0 t (u)| ≤ ε / ‖w‖ < δ ⎞⎟⎠ → 0, δ → 0 + . Theorem 3.1 is proved. SUPPORT THEOREM ON STOCHASTIC FLOWS WITH INTERACTION 137 4. Approximation theorem. Let us consider the system⎧⎪⎨⎪⎩ d dtx ε t (u) = a(xε t (u), με t ) + ∑m k=1 bk(xε t (u), με t )ξ ε k(t), xε 0(u) = u, u ∈ R d, με t = μ0 ◦ (xε t ) −1, t ∈ [0, T ], where the stochastic processes ξε k(t) are such that wε k(t) = ∫ t 0 ξε k(s)ds converges as ε → 0 to a Wiener process wk(t) in some sense. In this section, we give some sufficient conditions that ensure the convergence of (xε t , μ ε t ) to (xt, μt), where (xt, μt) is a solution of (0.1). Theorem 4.1. Assume that the functions a, bk, k = 1, m satisfy the conditions of The- orem 3.1 and A1. The collection of processes Kε(s) = N∑ l,j=1 (∫ T s |E (ξε l (r)/Gε s) |dr ) ( 1 + |ξε j (s)|) , ε > 0 is uniformly Lp-bounded for each p ≥ 1: ∀p ≥ 1 sup ε sup s EKp ε (s) = Kp < ∞ and is uniformly exponentially bounded in mean: ∀λ > 0 sup ε E exp{λ ∫ T 0 Kε(s) ds} < ∞, where Gε t := σ(ξε k(z) : 0 ≤ z ≤ t, k = 1, . . . , N). A2. For each s ∈ [0, T ], the following convergence holds in L2 :∫ T s |E( ξε l (z)/Gε s )| dz → 0, ε → 0. A3. There exist the deterministic bounded functions σlm(t), 1 ≤ l, m ≤ N such that, for all s < t, we have the following convergence in L1 : E (∫ t s ξε l (τ)dτ ∫ τ s ξε m(v)dv/Gε s ) −→ ε→0 ∫ t s σlm(z)dz. Then the distribution in the space Cn × C([0; T ], Rm) × C([0; T ],P) of the triple (xε t , w ε(t), με t ) converges weakly as ε → 0 to a limit measure such that: 1) the distribution of the second coordinate w(t) = (w1(t), . . . , wm(t)) is a Wiener process with the covariation matrix ‖ ∫ t 0 (σij(s) + σji(s))ds‖; 138 A. YU. PILIPENKO 2) the processes xt(u), w(t) and μt are connected by the system (4.1) ⎧⎪⎪⎪⎪⎪⎪⎪⎪⎪⎨⎪⎪⎪⎪⎪⎪⎪⎪⎪⎩ dxt(u) = [ a(xt(u), μt) + ∑m j,k=1 ( ∂bk(xt(u),μt) ∂x bj(xt(u), μt)+ + ∫ Rd ∂ ∂v δbk(xt(u),μt) δμ (v)bj(v, μt)μt(dv) )] σkj(t)dt + ∑m k=1 bk(xt(u), μt)dwk(t), x0(u) = u, u ∈ R d, μt = μ0 ◦ x−1 t , t ∈ [0, T ]. If, in addition, the processes wε(t) and w(t) are given on the same probability space and wε(t) → w(t), ε → 0 in probability for all t ∈ [0; T ], then (4.2) ∀p ≥ 1 ∀R > 0 E sup |u|≤R sup t∈[0;T ] n∑ k=0 |∇k(xε t (u) − xt(u))|p → 0, ε → 0. An example of the sequence wε(t) with σij = { 0, i �= j 0.5, i = j is a polygonal approxi- mation: ξj ε(t) = ε−1(wj((k + 1)ε) − wj(kε)) if t ∈ [kε; (k + 1)ε). Theorem 4.1 and this example imply that the support of the xt(u) distribution in the space Cn is contained in a set {xψ : ψ is a piecewise linear function, xψ satisfies (3.1)}. It is easy to show that if {ψn} is a sequence of piecewise smooth functions such that sup t∈[0;T ] |ψn(t) − ψ0(t)| → 0, n → ∞ ess supt∈[0;T ]|ψ′ n(t) − ψ′ 0(t)| → 0, n → ∞, then we have the convergence xψn → xψ0 , n → ∞ in Cn. So, the closure of {xψ : ψ is piecewise linear function, xψ satisfies (3.1)}. contains a set {xψ : ψ is piecewise smooth function, xψ satisfies (3.1)}. Therefore, we have obtained the following support theorem. Theorem 4.2. Assume that the conditions of Theorems 3.1, 4.1 are satisfied. Then the support of the xt(u) distribution in Cn is equal to the closure in Cn of the set {xψ : ψ is a piecewise smooth function, xψ satisfies (3.1)}. Proof of the Theorem 4.1. We follow the ideas of [5] Ch.5, see also [7] for some details related to the equations with interaction. Note, at first, that, under the conditions of Theorem 4.1, the processes wε converge weakly to a Wiener process w with required covariation [5] §5.7. SUPPORT THEOREM ON STOCHASTIC FLOWS WITH INTERACTION 139 Lemma 4.1. ∀m ∈ N ∀R > 0 ∃C = CR,m > 0 : sup ε E‖xε(·, t)‖2m W k 2m(BR) ≤ C, ∀t′, t′′ : sup ε E‖xε(·, t′) − xε(·, t′′)‖2m W k 2m(BR) ≤ C|t′ − t′′|2− 1 m , where BR is a ball in R d with the center at zero and with radius R. The course of the proof is similar to [5] §5.2 (see also [7] for the equations with inter- action), but we have to use Theorem 2.1 instead of the formula of the usual integration by parts. It is well known that the Sobolev space Wn+1 p,loc is compactly embedded into the space Cn(Rd, Rd) if p > d. Thus, by Theorem 1.4.7 [5], we have the weak relative compactness of the family xε, wε in Cn × C([0; T ], Rm). The next lemma implies that if a sequence of couples (xεk , wεk) converges weakly to some limit (x, w), then a sequence of triples (xεk . (·), wεk (·), μεk . ) converges weakly to (x.(·), w(·), μ.), where μt = μ0 ◦ (xt)−1. Lemma 4.2. Assume that a sequence of random elements yn = yn t (u) converges weakly to y0 = y0 t (u) in C([0; T ], C(Rd, Rd)). Then, for each probability measure μ, we have the weak convergence of measure-valued processes μn t = μ ◦ (yn t )−1 → μ0 t = μ ◦ (y0 t )−1 in the space C([0; T ],P). Proof. By the Skorokhod theorem [13], we can assume that all the elements yn, n ≥ 0 are given on the same probability space and with probability one: ∀R > 0 : sup t∈[0;T ] sup |u|<R |yn t (u) − y0 t (u)| → 0, n → ∞. Let ω be from the corresponding set of full probability. Denote, by κ̃t, the image of the measure μ w.r.t. mapping (yn t , y0 t ). Then sup t∈[0;T ] γ(μn t , μ0 t ) = sup t∈[0;T ] sup κ∈Q(μn t ;μ0 t ) ∫ ∫ |u − v| ∧ 1κ(du, dv) ≤ (4.3) ≤ sup t∈[0;T ] ∫ ∫ |u − v| ∧ 1κ̃t(du, dv) = sup t∈[0;T ] ∫ |yn t (u) − y0 t (u)| ∧ 1μ(du). Let ε > 0 be fixed. Choose R > 0, n0 ≥ 1 such that μ(u : |u| ≥ R) < ε and ∀n ≥ n0 : sup t∈[0;T ] sup |u|<R |yn t (u) − y0 t (u)| < ε. Then the right-hand side of (4.3) is less than 2ε. Lemma 4.2 is proved. If we verify that the limit triple (xt, wt, μt) satisfies Eq. (4.1), then, by uniqueness of the solution, we get the desired convergence (xε t , w ε t , μ ε t ) → (xt, wt, μt). To check (4.1) for a limit, it is sufficient to verify that M(u, t) = x(u, t) − u − ∫ t 0 [ a(xz(u), μz) + m∑ j,k=1 (∂bk(xz(u), μz) ∂x bj(xz(u), μz)+ (4.4) + ∫ Rd ∂ ∂v δbk(xz(u), μt) δμ (v)bj(v, μz)μz(dv) ) σkj(z) ] dz 140 A. YU. PILIPENKO is a continuous L2-martingale with respect to Ft = σ ( w(s), xs(u), s ∈ [0, t], u ∈ R d ) , t ∈ [0, T ], with the square characteristics (4.5) 〈M (i)(u, t), M (j)(v, t)〉 = ∑ k,l ∫ t 0 ∫ b (i) k (x(u, s), μs)b (j) l (x(v, s), μs))(σkl(s) + σlk(s))ds (4.6) 〈M (i)(u, t), wj(t)〉 = ∑ k ∫ t 0 ∫ b (i) k (xs(u)μs)(σkj(s) + σjk(s))ds. The process bk(xε t (u), με t ) is differentiable with respect to t by Theorem 2.1 and ∂ ∂t (bk(xε t (u), με t )) = = ∂ ∂x (bk(xε t )(u), με t )) ⎛⎝a(xε t (u), με t ) + m∑ j=1 bj(xε t (u), με t )ξ ε j (t) ⎞⎠ + + ∫ Rd ∂ ∂v δbk(xε t (u), με t ) δμ (v) ⎛⎝a(v, με t ) + m∑ j=1 bj(v, με t )ξ ε j (t) ⎞⎠ με t (dv). Then xε t (u) − xε s(u) − ∫ t s a(xε z(u), με z)dz = m∑ k=1 ∫ t s bk(xε z(u), με z)ξ ε k(z)dz = = m∑ k=1 bk(xε s(u), με s) ∫ t s ξk z dz + m∑ k=1 ∫ t s ∫ z s ∂ ∂r (bk(xε r(u), με r)) drξε k(z)dz = = m∑ k=1 bk(xε s(u), με s) ∫ t s ξε k(z)dz+ + m∑ k=1 ∫ t s ∫ z s [ ∂ ∂x (bk(xε r)(u), μr)) ⎛⎝a(xε r(u), με r) + m∑ j=1 bj(xε r(u), με r)ξ ε j (r) ⎞⎠ + + ∫ Rd ∂ ∂v δbk(xε r(u), με r) δμ (v) ⎛⎝a(v, με r) + m∑ j=1 bj(v, με r)ξ ε j (r) ⎞⎠] με r(dv)drξε k(z)dz. Let s ∈ [0; T ], sj ∈ [0; s), uj ∈ R d, l ∈ N. Put Φε = f(xε s1 (u1), . . . , xε sl (ul), wε(s1), . . . , wε(sl)), where f is some continuous bounded function, Φ = f(xs1(u1), . . . , xsl (ul), w(s1), . . . , w(sl)). We recall that {εn} is such that the sequence of triples (xεn . (·), wεn (·), μεn . ) converges weakly to some limit (x.(·), w(·), μ.) The following statement can be proved similarly to Lemma 5 [7]; see also the reasoning of Lemma 4.2. SUPPORT THEOREM ON STOCHASTIC FLOWS WITH INTERACTION 141 Lemma 4.3. Let α = α(x, μ) be a bounded continuous function. Then E ∫ t s α(xεn r (u), μεn r )drΦεn → E ∫ t s α(xr(u), μr)drΦ, n → ∞; E ∫ t s α(xεn r (u), μεn r )ξεn k (r)drΦεn → 0, n → ∞; E ∫ t s ∫ z s α(xεn r (u), μεn r )ξεn k (r)ξεn j (z)drdzΦεn → E ∫ t s α(xr(u), μr)σkj(r)drΦ, n → ∞. As a corollary of Lemma 4.3, we have a fact that M(u, t) is Ft−martingale. The reasoning for (4.5), (4.6) is similar. That is, the limit triple satisfies (4.1). So we have proved the desired weak convergence. The proof of (4.2) is quite classical and may be roughly formulated as follows. Let a sequence of solutions of SDEs given on the same probability space be weakly relatively compact, and let the uniqueness for the limit equation hold. Then the sequence converges strongly (cf. Theorem 5.2.8 [5]). Bibliography 1. Dorogovtsev A.A. and Kotelenez P., Smooth Stationary Solutions of Quasilinear Stochastic Partial Differential Equations: 1. Finite Mass, Preprint No. 97-145, Cleveland (1997). 2. Kotelenez P., A class of quasilinear stochastic partial differential equations of McKean-Vlasov type with mass conservation., Probab. Theory Related Fields 102 (1995), no. 2, 159–188. 3. Dorogovtsev A.A., Kotelenez P., Stochastic flows with interaction and random measures. (to appear) 4. Stroock D.W., Varadhan S.R.S., On the support of diffusion processes with applications to the strong maximum principle., Proceedings of the Sixth Berkeley Symposium on Mathematical Statistics and Probability (Univ. California, Berkeley, Calif., 1970/1971), Vol. III: Probability theory, pp. 333–359. 5. Kunita H., Stochastic Flows and Stochastic Differential Equations, vol. 24, Cambridge studies in advanced mathematics, 1990, pp. 346. 6. Pilipenko A. Yu., Stroock-Varadhan theorem for flows generated by stochastic differential equa- tions with interaction, Ukrainian Math. J. 54 (2002), no. 2, 227–236. 7. Pilipenko A.Yu., Approximation theorem for stochastic differential equations with interaction, Random Oper. and Stoch. Equ. 11 (2003), no. 3,, 213–228. 8. Dudley R. M., Real Analysis and Probability, Pacific Grove, Calif.: Wodsworth & Brooks/Cole Pub. Co., 1989. 9. Dawson D.A., Measure-Valued Markov processes. Ecole d’Ete de Probabilites de Saint-Flour XXI—1991, Lecture Notes in Math., vol. 1541, Springer, Berlin, 1993, pp. 1–260. 10. Jacka S., Tribe R., Comparisons for measure valued processes with interactions., Ann. Probab. 31 (2003), no. 3, 1679–1712. 11. Ikeda N., Watanabe S., Stochastic Differential Equations and Diffusion Processes, North- Holland Mathematical Library, 24. North-Holland Publishing Co., Amsterdam-New York; Ko- dansha, Ltd., Tokyo, 1981. 12. Adams R.A., Sobolev Spaces, Academic Press, New York, 1975. 13. Ethier S.N, Kurtz T.G., Markov Processes: Characterization and Convergence, Wiley, New York, 1986. E-mail : apilip@imath.kiev.ua