914
Views
0
CrossRef citations to date
0
Altmetric
Review

Weak martingale solution of stochastic critical Oldroyd-B type models perturbed by pure jump noise

ORCID Icon &
Pages 657-690 | Received 23 Sep 2020, Accepted 16 Jun 2021, Published online: 24 Aug 2021

Abstract

We investigate the existence of a weak martingale solution for a two-dimensional critical viscoelastic flow of the Oldroyd type driven by pure jump Lévy noise. Due to the viscoelastic nature, noise in the equation modeling stress tensor is considered in the Marcus canonical form. Owing to the lack of dissipation and taking into account of the structure of the non-linear terms, the proof requires higher order estimates.

2010 Mathematics Subject Classification:

1. Introduction

1.1. Description of the deterministic model

It is well-known that the full classical Oldroyd type models for viscoelastic fluids (see Oldroyd [Citation1]) in R2 can be written as (1.1a) vt+(v·)vνΔv+p=ν1Div τinR2×(0,T),(1.1a) (1.1b) τt+(v·)τκΔτ+aτ+Q(τ,v)=ν2D(v)inR2×(0,T),(1.1b) (1.1c) ·v=0   inR2×(0,T),(1.1c) (1.1d) v(·,0)=v0,  τ(·,0)=τ0 inR2,(1.1d) where v is the velocity vector field which is assumed to be divergence free, τ is the non-Newtonian part of the stress tensor (i.e., τ(x,t) is a (2, 2) symmetric matrix), p is the pressure of the fluid, which is an unknown scalar, i.e., part of the solution. The parameters ν (the viscosity of the fluid), a (the reciprocal of the relaxation time), ν1 and ν2 (determined by the dynamical viscosity of the fluid, the retardation time and a) are assumed to be non-negative. D(v) is called the deformation tensor and is the symmetric part of the velocity gradient D(v)=12(v+tv). Q is a quadratic form in (τ,v) and for Oldroyd fluids one usually chooses Q(τ,v)=τW(v)W(v)τb(D(v)τ+τD(v)), where b[1,1] is a constant and W(v)=12(vtv) is the vorticity tensor, and is the skew-symmetric part of velocity gradient. We impose an idealized condition (or far-field condition) on the fluid, i.e., v(x,t)0andτ(x,t)0as|x|.

The case κ = 0 in Equation(1.1) is a classical issue and analysis of the model under this situation is quiet challenging. For more review of such problems, we refer to [Citation2, Citation3] and the references cited therein. Authors in [Citation4] have studied the two dimensional model when there is a lack of dissipation in the velocity equation only, i.e., when ν=0,κ>0, and when Q has a special structure (that is, in the corotational case). As an improvization of [Citation4], the authors in [Citation5] have exploited the method of transferring dissipation and have obtained the existence of a unique global solution in three dimensional space provided the initial data is small in the Hs Sobolev-norm with s>52. It is worth noting that even when Q=0, the coupling is critical with respect to the smoothing effect provided by the partial parabolic regularization (see, for instance, [Citation4] for details). When the diffusion is present in both the equations i.e., ν,κ>0, the problem is subcritical in dimension two, and global well-posedness of strong solutions can be obtained in [Citation6] due to the smoothing effect provided by the fully parabolic system. One can clearly observe that Equation(1.1a) is similar to Navier-Stokes equations (NSE) if ν1=0. We have fairly good understanding of two dimensional NSE in both deterministic and stochastic set-ups. Although in three dimensional NSE settings, there are still open questions on global solvability and only partial results are available in this direction. In this regard, we restrict ourselves to the two dimensional problem only.

Let us now briefly describe the connection between some deterministic systems with critical coupling which are close to our model. In particular, a critical coupling for the Boussinesq system has been studied in Hmidi et al. [Citation7, Citation8], Manna and Panda [Citation9], for the MHD system in Caflisch et al. [Citation10], for the liquid crystal model in Lin et al. [Citation11], to name a few. Many other interesting results are devoted to the Oldroyd-B model and related models and are available in [Citation12–16] and the references therein.

Stochastic analysis of the above critical coupled systems influenced by random forcing are very limited and quite recent (e.g. see Yamazaki [Citation17] for Boussinesq system with zero dissipation, Manna et al. [Citation18] for the non-resistive MHD system, Brzeźniak et al. [Citation19] for the liquid crystal model). Existence and asymptotic behavior of a linear visoelastic fluid equation perturbed by additive or multiplicative Wiener processes has been studied, in Barbu et al. [Citation20] and Razafimandimby [Citation21].

In this context, it is worthy to mention that some recent works have been devoted to understand the fluid behaviors in the micro-macro regimes. In order to build a micro-macro model, one needs to go down to the microscopic scale and make use of the kinetic theory to obtain a mathematical model for the evolution of the microstructures of the fluid (e.g. configurations of the polymer chains in the case of polymeric fluid). For a quick survey and more details, we refer to our earlier paper [Citation2, Citation3] and references therein.

1.2. The stochastic model

In the present work, we are interested in the mathematical analysis of a stochastic version of (1.1a)–(1.1d). Our work is motivated by the importance of external perturbation on the dynamics of the velocity field for fluids with memory. The viscoelastic property demands that the material must return to its original shape after any deforming external force has been removed (i.e., it will show an elastic response) even though it may take time to do so. Hence the equation modeling stress tensor (i.e. the equation for τ) is invariant under coordinate transformation. Therefore the question of how to incorporate a suitable perturbation modeling the stress tensor without destroying its invariance property is a delicate one.

Hence for a complete knowledge of the effect of fluctuating forcing field on the behavior of the viscoelastic fluids, one needs to take into account the dynamics of v and τ. To initiate this kind of investigation, we propose a mathematical study of the following system of equations which basically describes an approximation of the system governing the viscoelastic fluids under the influence of fluctuating external forces.

1.2.1. Physical motivation for the choice of noise

In continuum mechanics, it is well-known that in a solid material, the elastic component of the stress can be attributed to the deformation of the bonds between the tiniest particles of the material. As a subject of interest, in a fluid, the elastic stress can be assigned to the change in the mean spacing of the particles. This phenomena in the fluid affects the particles’ interaction or collision rate and hence the transfer of momentum across the fluid takes place. Therefore, it is related to the microscopic thermal random component of the particles’ motion, which motivates us to establish more realistic models by incorporating stochastic processes. The randomness leads to a variate of new phenomena and may have highly non-trivial impact on the behavior of the solution. As per the existing literature, the viscous component of the stress arises from the macroscopic mean velocity of the particles. It can be attributed to friction or particle diffusion between adjacent parcels of the medium that have different mean velocities, which inspires us to incorporate noise (both continuous and jump) for both the equations (see [Citation22]) for further details.

1.3. Model problem

In this work we consider the following two dimensional stochastic viscoelastic equation (ν=0,κ>0 and Q=0) with pure jump noise (in Itô-Lévy sense for the velocity equation and Marcus canonical sense for the stress tensor equation) (1.2a) dv(t)+[(v(t)·)v(t)+p]dt=ν1Div τ(t)dt+ZF(v(t),z)N˜1(dt,dz),(1.2a) (1.2b) dτ(t)+[(v(t)·)τ(t)+aτ(t)κΔτ(t)]dt=ν2D(v(t))dt+i=1k(hiτ(t))dLi(t),(1.2b) (1.2c) ·v=0,  v(·,0)=v0,  τ(·,0)=τ0.(1.2c) Each hi;i=1,,k is a bounded function in R2×2,L(t):=(L1(t),,Lk(t)) is a Rk-valued pure jump Lévy process with the intensity measure λ2 such that suppλ2B, where B:=B(0,1)\{0}Rk; lRk i.e., L(t)=0tBl N˜2(ds,dl) where N˜2 represents time homogeneous compensated Poisson random measure. Here N˜1 represents time homogeneous compensated Poisson random measure on a certain measurable space (Z,Z). Precise definition of will be stated later. The tensor product ⊗ denotes usual matrix multiplication. Properties of F have been stated in Assumption 2.3.

1.3.1. Why Marcus type noise?

The choice of the noise modeling the stress tensor for the Stratonovich type noise (for continuous noise) and Lévy noise (for jump noise) in the Marcus canonical form are due to its special property of invariance under diffeomorphism. We refer to our earlier paper [Citation2] where the EquationEquation (1.1a) for velocity field is perturbed by a very general noise of multiplicative Lévy type, whereas EquationEquation (1.1b) for the stress tensor is perturbed by a Wiener noise in the Stratonovich form. However, interested readers may look into the recent manuscript [Citation23] of the first author and his collaborator, where for the first time a certain kind of SPDE driven by Lévy noise in the Marcus canonical form has been studied. Such invariance property is known in the literature for Stratonovich integrals (see e.g. Brzeźniak and Elworthy [Citation24, Theorem 3.3 and Proposition 4.4]). However, one needs an analogue of the Stratonovich integral in the case of stochastic integral with respect to compensated Poisson random measure. The work of Marcus [Citation25], developed later by Applebaum and Kunita, (see e.g. Section 6.10 of [Citation26] and [Citation27]; see also [Citation28]) provides a framework to resolve this technical issue. In other words, one needs a result of the following type (which is framed in SDE set-up with notations as above). Authors of the current paper have studied the critically coupled system in [Citation2] (when ν>0, κ=0, Q0) under the influence of a Wiener process in Stratonovich form and in [Citation3] in presence of Marcus noise (when ν>0, κ>0, Q=0) in R2.

1.3.2. Novelty of our work

We construct a martingale solution of the stochastic critical Oldroyd-B system Equation(1.2) with zero dissipation (in the velocity equation) and a convection diffusion (in the equation for τ) in dimension two. The proof is based on a variant of the Faedo-Galerkin method. Furthermore, we use tightness criteria (see Appendix A.2 of the current manuscript and the references cited therein) in certain spaces of cádlág functions and Jakubowski’s generalization of the Skorokhod-Theorem to nonmetric spaces; we refer to [Citation29, Citation30], Proposition 9 in [Citation31], Proposition 3.7 in Step IV of the current manuscript.

Since the original system Equation(3.1) lacks dissipation, it behaves more like an incompressible Euler equation. Hence mere L2-estimate on the velocity is not enough to talk about well-posedness of such problems and one needs to consider more regular initial data to obtain an enstrophy estimate. Besides, the rigorous analysis involved in this paper owing to the critical coupling of the random system in presence of pure jump noise is quiet technical and delicate compared to the deterministic ones, see [Citation4, Citation5].

Proposition 1.1.

Let L(t) be a Rk-valued pure jump Lévy process with the intensity measure λ such that suppλB:=B(0,1)Rk; lRk i.e., L(t)=0tBl N˜(ds,dl) where N˜ represents time homogeneous compensated Poisson random measure. Let u,w1,,wk:R2R2 be complete C1-vector fields. Define w:R2L(Rk,R2) such that w(y)(h):=j=1kwj(y)hj,hRk,yR2. Consider the following “Marcus” stochastic differential equation (1.3) dY(t)=u(Y(t)) dt+w(Y(t))dL(t)=u(Y(t)) dt+j=1kwj(Y(t))dLj(t),(1.3) which is defined in the integral form as follows Y(t)=Y0+0tu(Y(s)) ds+0tB[Φ(1,l,Y(s))Y(s)]N˜(ds,dl)+0tB[Φ(1,l,Y(s))Y(s)j=1kljwj(Y(s))]λ(dl)ds, where z(t):=Φ(t,l,z0) solves dzdt=j=1kljwj(z),with initial condition z(0)=z0. Suppose, Y, an R2-valued process, is a solution to Equation(1.3). Let φ:R2R2 be a C1 diffeomorphism. Define for each j=0,1,,k, the “Push-forward” of the vector fields u and wj by φ as û:R2yφ(φ1(y))(u(φ1(y)))R2,ŵj:R2yφ(φ1(y))(wj(φ1(y)))R2. Let ŵ:R2L(Rk,R2) be such that ŵ(y)(h):=j=1kŵj(y)hj, hRk.

Then Y is a solution to Equation(1.3) if and only if Z(t):=φ(Y(t)) is a solution to dZ(t)=û(Z(t)) dt+ŵ(Z(t))dL(t),Z0=φ(Y0).

Proof of the above result can be borrowed from [Citation32, Theorem B.1], where Brzeźniak and the first named author of this article have obtained the result for the pure jump case.

Let us briefly describe the content of the paper. In Section 2, we present detailed description of the functional spaces, the Marcus map and assumptions of this paper. In Section 3, we discuss statement of the main result and its proof.

Appendix A is split into two parts. The very first subsection has been presented by series of technical lemmas which are useful during the course of analysis made throughout the paper. The second subsection has been devoted to certain useful compactness and tightness criterion for cádlag functions and predictable processes.

2. Functional settings, assumptions and the Marcus map

We denote Cc(R2;R2) as the space of all R2valued compactly supported C functions in R2, V:={vCc(R2;R2):Div v=0},  V˜:={τCc(R2;R4):τT=τ}. We define H,V,H˜ and V˜ be the closure of V in L2(R2;R2),H1(R2;R2),L2(R2;R4) and H1(R2;R4) respectively. For s>0, let Vs and Vs˜ be the closure of V and V˜ in Hs(R2;R2) and Hs(R2;R4) respectively. It is clear from the definitions that V1=V and V˜1=V˜.

Notation 2.1.

Throughout this paper, we denote X as the dual of X where X is a Hilbert space and C as a generic constant.

We note the following:

  1. H is a Hilbert space with scalar product (u,v)H:=(u,v)L2,   u,vH and V is a Hilbert space with scalar product (u,v)V:=(u,v)L2+(u,v)L2,  u,vV.

  2. H˜ is a Hilbert space with scalar product (u,v)H˜:=(u,v)L2   u,vH˜ and V˜ is a Hilbert space with scalar product (u,v)V˜:=(u,v)L2+(u,v)L2,u,vV˜.

2.1. Trilinear operators b and b˜

We define the trilinear operators b:Lp(R2;R2)×W1,q(R2;R2)×Lr(R2;R2)R and b˜:Lp(R2;R2)×W1,q(R2;R4)×Lr(R2;R4)R by b(u,v,w):=i,j=12R2u(i)xiv(j)w(j)dx,uLp(R2;R2), vW1,q(R2;R2), wLr(R2;R2)b˜(v,τ,θ):=i,j,k=12R2v(i)xiτ(jk)θ(jk)dx,vLp(R2;R2), τW1,q(R2;R4), θLr(R2;R4) with 1p+1q+1r=1 and p,q,r[1,]. Let us define bilinear maps B:V×VV and B˜:V×V˜V˜ by B(u,v)=b(u,v,·) and B˜(v,τ)=b˜(v,τ,·). Then we can show that B:V×VV and B˜:V×V˜V˜ are linear and continuous.

2.2. Some additional operators and notations

If s>2, then by the Sobolev embedding theorem Hs1(R2;R2)Cb(R2;R2)L(R2;R2) where Cb(R2;R2) denotes the space of continuous and bounded R2 valued functions defined on R2. We have the following conclusions: (2.1) VsjsVjHHjVjsVs,andVs˜j˜sV˜j˜H˜H˜j˜V˜j˜sVs˜.(2.1) The following embeddings j,js,j˜,j˜s are continuous and dense. By Lemma C.1 in Appendix C of [Citation23], there exists a Hilbert space U such that UVs and (2.2) the natural embedding ls:UVs is dense and compact.(2.2) Combining Equation(2.1) and Equation(2.2) we have UlsVsjsVjHHjVjsVslsU. Let us consider the adjoint of the following maps j,js,ls denoted by j*,js*,ls* respectively and are defined as: j*:HV,  js*:VVs,  ls*:VsU. Since all the embeddings j,js,ls are dense so j*,js*,ls* are one-one. Let us define the map i:UHbyi:=j°js°ls. Note that i is compact and range of i is dense in H. We consider the adjoint of i defined as i*:HU, which is one-one. In the similar manner, one can have U˜l˜sVs˜j˜sV˜j˜H˜H˜j˜V˜j˜sVs˜l˜sU˜ and i˜:U˜H˜byi˜:=j˜°j˜s°l˜s. Let us define (A,D(A)) by D(A):=j*(H)V, Av:=(j*)1v,vD(A). In the similar manner, we can define (As,D(As)),(A,D(A)),(As,D(As)),(A˜,D(A˜)),(A˜s,D(A˜s)),(A˜,D(A˜)) and (A˜s,D(A˜s)), corresponding to maps (js*)1, (i*)1,(is*)1,((j˜)*)1,((j˜s)*)1,((i˜)*)1 and ((i˜s)*)1 respectively. We note that A=A°As°As and A˜=A˜°A˜s°A˜s. Note that for each vV, using the continuity of u(v,u)H, one can conclude existence of an element AvV such that VAv,uV=(v,u)H,v,uV. Similarly we can define A˜:V˜V˜ by V˜A˜τ,wV˜=(τ,w)H,τ,wV˜.

Remark 1.

In view of [Citation23], one can observe that both A,A˜ are self adjoint and A1,(A˜)1 are compact.

Lemma 2.2.

For vD(A),uV,τD(A˜),wV˜, the following results hold:

  1. ((AI)v,u)H=(v,u)H=VAv,uV,  and  |Av|V|Avv|H.

  2. ((A˜I)τ,w)H˜=(τ,w)H˜=V˜A˜τ,wV˜,  and  |A˜τ|V˜|A˜ττ|H˜.

For proof see [Citation23].

Let (S,S) be a measurable space and let MN̂(S) be the set of all N̂ valued measures on (S,S), where N̂=N{0,}. On the set MN̂(S), we consider the σfield MN̂(S) defined as the smallest σfield for all AS: the map iA:MN̂(S)λjλj(A)N̂ is measurable.

Let (Z,Z) be a measurable space. Let N1 on (Z,Z) over a filtered probability space (Ω,F,F,P) be the time homogeneous Poisson random measure (for definition, see e.g. [Citation33]), which is known to be a measurable function N1:(Ω,F)(MN̂(R+×Z),MN̂(R+×Z)). The formula λ1(A):=E[N1((0,1]×A)],AZ defines a measure on (Z,Z), called an intensity measure of N1. Moreover, for all T< and all AZ such that E[N1((0,T]×A)]<, the R-valued process {N1((0,t]×A)tλ1(A)}t(0,T] is an integrable martingale on (Ω,F,F,P). The random measure Lebλ1 on B(R+)Z is called the compensator of N1 and N˜1:=N1Lebλ1 is called the compensated time homogeneous Poisson random measure.

Similarly, one can define the time homogeneous Poisson random measure N2.

Assumption 2.3.

Below we state the assumptions made throughout this paper.

(A.1) L=(Ω,F,F,P) is a filtered probability space, where F=(Ft)t0 is the filtration, satisfying the usual conditions, i.e.,

  1. P is complete on (Ω,F),

  2. for each t0, Ft contains all (F,P)-null sets,

  3. the filtration Ft is right-continuous.

(A.2) N1 is a time homogeneous Poisson random measure on a measurable space (Z,Z) over the above probability space with intensity measure λ1 (and compensator Lebλ1).

(A.3) (L(t))t0 is a Rk-valued, (Ft)-adapted Lévy process of pure jump type defined on the above probability space with drift 0 and the corresponding time homogenous Poisson random measure N2.

(A.4) The intensity measure λ2 of N2 (with compensator Lebλ2) is such that suppλ2B and λ2(B)<, where B:=B(0,1)\{0}Rk.

(A.5) N1 and N2 are mutually independent.

(A.6) hiW1,, for each i=1,2,k.

(A.7) F:H×ZH is a measurable function such that there exists a positive constant L such that (2.3) Z|F(v1,z)F(v2,z)|H2 λ1(dz)L|v1v2|H2,v1,v2H,(2.3) and for each p1 there exists a positive constant Cp such that (2.4) Z|F(v,z)|Hp λ1(dz)Cp(1+|v|Hp),vH.(2.4)

(A.8) F:V×ZV is a measurable function such that there exists a positive constant Cp, depending upon p2, such that Z|Curl F(v,z)|Hpλ1(dz)Cp(1+|v|Hp),   vV.

Assumption (A.8) is required for the gradient and curl estimates in Lemma 3.4 and Corollary 3.5.

2.3. The Marcus map

For general discussion about this topic, we refer to the Appendix A. For hiW1,, we consider the bounded linear map gi:H˜H˜bygi(τ)=hiτ. It is easy to observe that gi:V˜V˜ is also a continuous linear map. We define a generalized Marcus mapping Φ:R+×Rk×H˜H˜ such that for each fixed lRk,  τ0H˜, the function tΦ(t,l,τ0) is the continuously differentiable solution of the ordinary differential equation dy(t)dt=i=1kligi(y(t))t0,y(0)=τ0. Therefore, we can write Φ(t,l,τ0)=Φ(0,l,τ0)+i=1k0tligi(Φ(s,l,τ0))ds,t0.

Notation.

Let us fix t = 1 and denote Φ(l,·)=Φ(1,l,·).

EquationEquation (1.2) with notation is defined in the integral form as following: (2.5) τ(t)=τ00t[(v(s)·)τ(s)+aτ(s)+κΔτ(s)ν2D(v(s))]ds+0tB[Φ(l,τ(s))τ(s)]N˜2(ds,dl)+0tB{Φ(l,τ(s))τ(s)i=1kligi(τ(s))}λ2(dl)ds,(2.5) where τ0 is a F0-measurable random variable.

For zH˜, we denote G(l,z):=Φ(l,z)z, K(l,z):=Φ(l,z)zi=1kligi(z),and b(z):=BK(l,z) λ2(dl). With these above notations, Equation(2.5) can be written as: τ(t)=τ00t[(v(s)·)τ(s)+aτ(s)+κΔτ(s)ν2D(v(s))]ds+0tBG(l,τ(s))N˜2(ds,dl)+0tb(τ(s))ds. Now, define a linear operator R:H˜τi=1kligi(τ)H˜. Then RL(H˜)|l|RkgL(H˜).

Thus, if we denote y(t):=Φ(t,l,x), then y satisfies dydt=Ry,y(0)=x.

Hence y(t)=etRx=j=0tjj!Rjx.

We then have the following useful result. We postpone the proof to Appendix A.

Lemma 2.4.

Let ψ:H˜R be defined by ψ(τ)=|τ|H˜p, p1. If N:=01esRpds, then

  1. |ψ(Φ(l,τ))ψ(τ)|N p |l|RkgL(H˜)|τ|H˜p.

  2. |ψ(Φ(l,τ))ψ(τ)ψ(τ)Rτ|N p2 |l|Rk2gL(H˜)2|τ|H˜p.

3. Main result and key ideas of the proof

We now consider the following stochastic critical Oldroyd-B system (with zero dissipation in the velocity equation and a convection diffusion equation for τ) in dimension two (3.1a) dv(t)+[B(v(t))ν1Div τ(t)]dt=ZF(v(t),z)N˜1(dt,dz),t0(3.1a) (3.1b) dτ(t)+[κA˜τ(t)+B˜(v(t),τ(t))+aτ(t)ν2D(v(t))]dt=BG(l,τ(t))N˜2(dt,dl)+b(τ(t))dt,t0(3.1b) (3.1c) v(0)=v0;  τ(0)=τ0.(3.1c) We first introduce the following functional spaces endowed with the respective topologies: D([0,T];U):=the space of ca`dla`g functions v:[0,T]U with the topology T1 induced by the Skorokhod metric δT,U; Lw2(0,T;V):=the space L2(0,T;V) with the weak topology T2; L2(0,T;Hloc):=the space of measurable functions v:[0,T]H such that for all R>0, pT,R(v):=vL2(0,T;HBR):=(0TBR|v(x,t)|H2 dx dt)1/2<, with the topology T3 generated by the seminorms (pT,R)R>0, where BR is a closed ball of radius R; Hw:=the Hilbert space H endowed with the weak topology; D([0,T];Hw):=the space of all weakly ca`dla`g functions v:[0,T] H with the weakest topology T4 such that for all hH the mappings D([0,T];Hw)v(v(·),h)HD([0,T];R) are continuous. In particular, vnv in D([0,T];Hw) iff for all hH (vn(·),h)H(v(·),h)H in D([0,T];R). Similarly, one can define the spaces D([0,T];U˜), Lw2(0,T;V˜), L2(0,T;H˜loc) and D([0,T];H˜w).

Now let us consider the spaces Zi;  i=1,2 defined by (3.2) {Z1:=D([0,T];U)D([0,T];Hw)Lw2(0,T;V)L2(0,T;Hloc),Z2:=D([0,T];U˜)D([0,T];H˜w)Lw2(0,T;V˜)L2(0,T;H˜loc).(3.2) and let Ki be the supremumFootnote1 of the corresponding topologies in the spaces Zi,i=1,2. Let Z:=Z1×Z2 and T be the product topology on Z.

We now recall the definition of a weak martingale solution.

Definition 3.1.

A weak martingale solution of Equation(3.1) is a system (L̂,v̂,τ̂,N̂1,N̂2), where

  1. L̂=(Ω̂,F̂,F̂,P̂) is a filtered probability space with a filtration F̂=(F̂t)t0,

  2. N̂1 is a time homogeneous Poisson random measure on (Z,Z) over L̂ with the intensity measure λ1 and N̂2 is a time homogeneous Poisson random measure on (B,B(B)) with the intensity measure λ2,

  3. v0L2(Ω̂;V) and τ0L2(Ω̂;V˜),

  4. v̂:[0,T]×Ω̂H is an F̂-progressively measurable process with P̂-a.e. paths v̂(·,ω)D([0,T];Hw)L2(0,T;V) such that for all t[0,T] and all ϕV (v̂(t),ϕ)H+0tB(v̂(s)),ϕ ds=(v0,ϕ)H+ν10tDiv τ̂(s),ϕ ds+0tZ(F(v̂(s),z),ϕ)HN̂˜1(ds,dz),P̂a.s. and satisfies the inequality Ê(sup0tT|v̂(t)|H2+sup0tT|×v̂(t)|H2)<, where Ê denotes the expectation with respect to P̂.

  5. τ̂:[0,T]×Ω̂H˜ is an F̂-progressively measurable process with P̂-a.e. paths τ̂(·,ω)D([0,T];H˜w)L2(0,T;V˜) such that for all t[0,T] and all ψV˜, (τ̂(t),ψ)H˜+κ0tA˜τ̂(s),ψ ds+0tB˜(v̂(s),τ̂(s)),ψ ds+a0tτ̂(s),ψ ds=(τ̂(0),ψ)H˜+ν20tD(v̂(s)),ψ ds+0tb(τ̂(s)),ψ ds+0tB(G(l,τ̂(s)),ψ)H˜ N̂˜2(ds,dl),P̂a.s., and satisfies the inequality Ê(sup0tT|τ̂(t)|H˜2+2κ0T|τ̂(t)|V˜2 dt)<.

We now present our main result.

Theorem 3.1.

For the sake of brevity, we take ν1=ν2=κ=1. Let Assumption 2.3 hold and let v0L2(Ω̂;V) and τ0L2(Ω̂;V˜). Then there exists a weak martingale solution (L̂,v̂,τ̂,N̂1,N̂2) to the system Equation(3.1) satisfying Ê[sup0sT(|v̂(s)|H2+|τ̂(s)|H˜2)]+2Ê[0T|τ̂(s)|V˜2ds]<,Ê(sup0sT|×v̂(s)|H2)<,

3.1. Main ideas of the proof

Due to lack of dissipation in the velocity equation Equation(3.1), we first incorporate an artificial dissipative term. In other words, for every 1>ν0ν>0, ν0 being fixed, we consider the following approximating Oldroyd-B system: (3.3a) dvν(t)+[νAvν(t)+B(vν(t))ν1Div τν(t)]dt=ZF(vν(t),z)N˜1(dt,dz),t0(3.3a) (3.3b) dτν(t)+[κA˜τν(t)+B˜(vν(t),τν(t))+aτν(t)ν2D(vν(t))]dt=BG(l,τν(t))N˜2(dt,dl)+b(τν(t))dt,t0(3.3b) (3.3c) vν(0)=v0;  τν(0)=τ0.(3.3c) We now consider an Faedo-Galerkin approximation of the above system and let the Galerkin solution be (vnν,τnν). Our strategy is the following: we prove that the system (L̂,v̂,τ̂,N̂1,N̂2) is a weak martingale solution to Equation(3.1), which is obtained as a limit (as n or ν0) of a random variable (v̂nν,τ̂nν,N̂1n,N̂2n) on the new probability space L̂ and this random variable has the same laws as the random variable (vnν,τnν,N1n,N2n) on the original probability space L. The proof is split into a few steps.

Firstly, our aim is to have uniform energy bounds for the Galerkin solution (vnν,τnν), which are also valid bounds for the Galerkin solution of the original system Equation(3.1). Therefore the bounds need to be free from the choice of ν (and n). Toward this, we obtain a basic energy estimate for the Galerkin solution (vnν,τnν), which reads, for p2, E[sup0sT(|vnν(s)|Hp+|τnν(s)|H˜p)]+2E[0T|τnν(s)|V˜2|τnν(s)|H˜p2ds]C where C depends on the parameter of the problem but is independent of n and ν. Since the original system (3.1) lacks dissipation, it behaves more like an incompressible Euler equation. Hence mere L2-estimate on the velocity is not enough to talk about well-posedness of such problems and one needs to consider more regular initial data to obtain an enstrophy estimate. Toward this goal our aim is to get the L2-estimate on the finite dimensional projection on the vorticity, i.e. for Curl vnν:=wnν. Drawing motivation from [Citation4], we define the singular integral operator R, that operates on matrices in the following way: R(·):=(Δ)1Curl Div. Now applying Curl(:=) and R to (3.3), we get for t[0,T], dwnν(t)+[(vnν(t)·)wnν(t)νΔwnν(t)]dt=Div τnν(t)dt+ZFn(vnν(t),z)N˜1(dt,dz)dR(τnν)(t)+[RΔτnν(t)+R(vnν(t)·)τnν(t)+aR(τnν)(t)]dt=R(D(vnν)(t))dt+BRGn(l,τnν(t))N˜2(dt,dl)+Rbn(τnν(t))dt. Define Γnν:=wnνCνR(τnν), where Cν:=11ν>0, for ν(0,ν0],1>ν0 being fixed.

Using the commutator property R(vnν·)τnν=(vnν·)R(τnν)[vnν·,R]τnν and a crucial simplification due to the identity RD(vnν)=wnν, we see that Γnν satisfies the following equation for t[0,T], dΓnν(t)+[(vnν(t)·)Γnν(t)+CνΓnν(t)νΔΓnν(t)+Cν[vnν(t)·,R]τnν(t)+(Cν2aCν)R(τnν(t))]dt=ZFn(vnν(t),z)N˜1(dt,dz)CνBRGn(l,τnν(t))N˜2(dt,dl)CνRbn(τnν(t))dt. Next, exploiting already obtained good bounds on τnν and properties of R, we find an uniform energy bound in L2 for Γnν, E[supt[0,T]|Γnν(t)|L22]C, which further yields E[supt[0,T]|wnν(t)|L22]C. Further, with the help of an associated elliptic problem, we prove E[supt[0,T]|vnν(t)|V2]C. As a next step, due to the above uniform bounds (which are independent of ν), tightness of the sequence of measures {L(vnν,τnν)}nN is tight on the space (Z,T) (see Equation(3.20)). Next by choosing ν=1n and employing Skorokhod-Jakubowski theorem, we have a new probability space L̂ and random variables (v̂,τ̂,N̂1,N̂2),(v̂nν,τ̂nν,N̂1n,N̂2n),nN on this new probability space, such that P̂-a.s. (v̂nν,τ̂nν)(v̂,τ̂)  in  Z. As a last step, we show that on the new probability space L̂, the limiting sequence (v̂,τ̂,N̂1,N̂2), i.e. the system (L̂,v̂,τ̂,N̂1,N̂2) is a weak martingale solution of the original problem Equation(3.1).

3.2. Proof of Theorem 3.1

We present the proof of Theorem 3.1 below.

Proof.

We split the proof in few steps.

Step I: (Faedo-Galerkin Approximation and Energy Estimate). By Remark 1, there exist orthonormal bases {ei},{e˜i} of H,H˜ respectively consist of eigenvectors of operators A,A˜. Let {λi},{λ˜i} be the corresponding eigenvalues of A and A˜ respectively. Therefore we have Aei=λiei and A˜e˜i=λ˜ie˜i,  iN. Let us denote Hn=span{e1,,en},  H˜n=span{e˜1,,e˜n}. Let Pn,P˜n be defined as Pn:UU, P˜n:U˜U˜ by Pnvν:=i=1nUvν,eiUei,P˜nτν=i=1nU˜τν,e˜iU˜e˜i,  vνU, τνU˜. For vνHn,τνH˜n and zZ, let us consider the following maps:

  1. Bn:HnHnbyBn(vν):=PnB(vν),

  2. B˜n:Hn×H˜nH˜nbyB˜n(vν,τν):=P˜nB˜(vν,τν),

  3. Fn:Hn×ZHnbyFn(vν,z):=PnF(vν,z),

  4. gin:H˜nH˜nbygin(τν):=P˜n(hiτν).

Let us denote vnν(0)=Pnv0ν and τnν(0)=P˜nτ0ν.

Let Φn(t,l,vnν(0)) be a flow on H˜n assosciated to the vector field i=1kligin, i.e., dΦndt(t,l,τnν(0))=i=1kliginΦn(t,l,τnν(0)),  t0,Φn(t,l,τnν(0))=τnν(0)H˜n. For τnνH˜n, we define Gn(l,τnν):=Φn(l,τnν)τnν,Kn(l,τnν):=Φn(l,τnν)τnνi=1kligin(τnν)bn(τnν):=B[Φn(l,τnν)τnνi=1kligin(τnν)]λ2(dl)=BKn(l,τnν)λ2(dl). Consider the following mappings Bn(vν):=PnB(χn(vν),vν),vνHn,B˜n(vν,τν):=PnB˜(χn(vν),τν),vνHn,τνH˜n, where χn:HH is defined by χn(vν)=θn(|vν|V)vν, where θn:R[0,1] of class C such that θn(r)=1 if rnandθn(r)=0 if rn+1. Since HnH and H˜nH˜, the mappings Bn and B˜n are well defined. Moreover, Bn:HnHn and B˜n:Hn×H˜nH˜n are globally Lipschitz continuous.

We now consider the following Hn×H˜n-valued modified approximated system: (3.4a) dvnν(t)+[νPnAvnν(t)+Bn(vnν(t))Div τnν(t)]dt=ZFn(vnν(t),z)N˜1(dt,dz),t0(3.4a) (3.4b) dτnν(t)+[P˜nA˜τnν(t)+B˜n(vnν(t),τnν(t))+aτnν(t)ν2D(vnν(t))]dt=BGn(l,τnν(t))N˜2(dt,dl)+bn(τnν(t))dt.(3.4b) Since all relevant maps are globally Lipschitz (see Lemma A.1A.3 in Appendix A), we have the following standard result, see e.g. [Citation34] for a reference.

Lemma 3.2.

For each nN, there exists a unique global, F-progressively measurable, Hn×H˜n-valued càdlàg process (vnν,τnν) satisfying the modified Galerkin approximated system Equation(3.4).

It follows easily, see for instance [Citation34], that for each nN, the approximated system Equation(3.4) has a unique local maximal solution. In two dimensional setting, by a combination of the proof of [Citation34, Theorem 3.1] with Lemma 3.3 below, we infer that the approximated system Equation(3.4) has a unique global solution.

Step II: (Estimates for the velocity field and stress tensor)

Lemma 3.3.

Let p2. Let v0Lp(Ω;H) and τ0Lp(Ω;H˜). Then for every T > 0, there exists C>0, depending on v0,τ0 (but is independent of n and ν), such that supnN E[sup0sT(|vnν(s)|Hp+|τnν(s)|H˜p)]+νp supnNE[0T|vnν(s)|H2|vnν(s)|Hp2ds]  (3.5) +p supnN E[0T|τnν(s)|H˜2|τnν(s)|H˜p2ds]C.(3.5)

Proof.

Estimate Equation(3.5) can be proved by applying Itô’s Lemma to the function f(x)=1p|x|Hp (for the process vnν(t)) and 1p|x|H˜p (for the process τnν(t)) to the system Equation(3.4). □

Step III: (Estimate of the vorticity)

Lemma 3.4.

Let w0L2(Ω;H) and τ0L2(Ω;V˜). Then for every T > 0 there exists a constant C>0, depending on v0,τ0 (but is independent of n and ν), such that (3.6) E[sup0sT|wnν(s)|H2]C,(3.6) where wν:=Curl vν and wnν is the finite dimensional approximation of wν.

Proof.

Let us recall that R denote the singular integral operator that operates on matrices in the following way: R(·):=(Δ)1Curl Div. We note that R is a degree 0 singular integral operator having the following properties (see [Citation4]):

  1. R is a bounded linear transformation in L2(R2×2).

  2. For divergence free vν, we have (3.7) |[vν·,R]τν|L2C|vν|H|τν|V˜.(3.7)

Now applying Curl(:=) and R to Equation(3.4), we get for t[0,T], dwnν(t)+[(vnν(t)·)wnν(t)νΔwnν(t)]dt=Div τnν(t)dt+ZFn(vnν(t),z)N˜1(dt,dz)dR(τnν)(t)+[RΔτnν(t)+R(vnν(t)·)τnν(t)+aR(τnν)(t)]dt=R(D(vnν)(t))dt+BRGn(l,τnν(t))N˜2(dt,dl)+Rbn(τnν(t))dt. Define Cν:=11ν>0,  for  ν(0,ν0],  1>ν0 being fixed.

Let Γnν:=wnνCνR(τnν). Then using properties of R, we see that Γnν satisfies (3.8) dΓnν(t)+[(vnν(t)·)Γnν(t)+CνΓnν(t)νΔΓnν(t)+Cν[vnν(t)·,R]τnν(t)+(Cν2aCν)R(τnν(t))]dt=ZFn(vnν(t),z)N˜1(dt,dz)CνBRGn(l,τnν(t))N˜2(dt,dl)CνRbn(τnν(t))dt(3.8) Exploiting the properties of R and Young’s inequality, we now estimate the following two terms which lead to (3.9) |(Cν2aCν)R(τnν(t)),Γnν(t)L2|Cν4|Γnν(t)|L22+Cν|Cνa|2|R(τnν(t))|L22|Cν[vnν(t)·,R]τnν(t),Γnν(t)L2|Cν4|Γnν(t)|L22+Cν|[vnν(t)·,R]τnν(t)|L22 (3.9) (3.10) Cν4|Γnν(t)|L22+C Cν|vnν(t)|H2|τnν(t)|V˜2.(3.10) For NN fixed, let us define a stopping time by SNn=inft0{t:supt[0,T]|Γnν(t)|L2>N}inft0{t:supt[0,T]|wnν(t)|H>N}. Therefore applying Itô’s formula to the function 12|Γnν(t)|L22 for the EquationEquation (3.8), using Equation(3.9), Equation(3.10) then integrating in [0,T] and then taking supremum and expectation, we get (3.11) 12E[supt[0,TSNn]|Γnν(t)|L22]+νE[0TSNn|Γnν(t)|L22dt]+Cν2E[0TSNn|Γnν(t)|L22dt]12E[|Γν(0)|L22]+|Cν||Cνa|2E[0TSNn|R(τnν(t))|L22dt]+CνC E[0TSNn|vnν(t)|H2|τnν(t)|V˜2dt]+i=12E[supt[0,TSNn]|0tMi(s)ds|]+i=12E[0TSNn|Li(t)|dt],(3.11) where (3.12a) M1(t):=Z{|Γnν(t)+Fn(vnν(t),z)|L22|Γnν(t)|L22}N˜1(dt,dz),(3.12a) (3.12b) L1(t):=Z{|Γnν(t)+Fn(vnν(t),z)|L22|Γnν(t)|L222Fn(vnν(t),z),Γnν(t)L2}λ1(dz)dt,(3.12b) (3.12c) M2(t):=B{|CνRΦn(l,τnν(t))|L22|Γnν(t)|L22}N˜2(dt,dl),(3.12c) (3.12d) L2(t):=B{|CνRΦn(l,τnν(t))|L22|Γnν(t)|L222i=1kliCνRgi(τnν(t)),Γnν(t)L2}λ2(dl)dt.(3.12d) Using Burkholder-Davis-Gundy inequality, (A.8) in Assumption 2.3, and Young’s inequality, we get (3.13) E[supt[0,TSNn]|0tM1(s)ds|]CT+18E[supt[0,TSNn]|Γnν(t)|L22]+C(T) E[0TSNn|Γnν(t)|L22dt].(3.13) In order to estimate M2 term, we apply the boundedness property of R (see Equation(3.7)) and part 1 of Lemma 2.4, and obtain (3.14) |RΦn(τnν(t))|L24C|Φn(l,τnν(t)|L24C(|τnν(t)|H˜4+|l|Rk2|τnν(t)|H˜4).(3.14) Thus using Equation(3.14), (A.4) in Assumption 2.3 and bounds on τn we have, (3.15) E[supt[0,TSNn]|0tM2(s)ds|]CE[(0TSNnB{|CνRΦn(l,τnν(t))|L22|Γnν(t)|L22}2λ2(dl)dt)1/2]CνCE[(0TSNn|τnν(t)|H˜4dt)1/2]+CE[(0TSNn|Γnν(t)|L24dt)1/2]CCνT+18E[supt[0,TSNn]|Γnν(t)|L22]+C(T) E[0TSNn|Γnν(t)|L22dt].(3.15) Likewise, using (A.8) in Assumption 2.3, Lemma 3.3 and TSNnT as N we have (3.16) E[0TSNn|L1(t)|dt]E[0TSNnZ|Fn(vnν(t),z)|L22λ1(dz)dt]CE[0TSNn(1+|vnν(t)|H2)dt]C(T),(3.16) and employing Lemma 2.4, (A.4) in Assumption 2.3, boundedness property of R and bounds on τnν, we achieve (3.17) E[0TSNn|L2(t)|dt]E[0TSNnB|CνRΦn(l,τnν(t))|L22+2Cνi=1k|liRgi(τnν(t)),Γnν(t)L2|λ2(dl)dt]C(T)+C (Cν+1)E[0TSNn|Γnν(t)|L22dt].(3.17) Finally we compute the third term on the left hand side of the inequality Equation(3.11) to obtain, (3.18) E[0TSNn|vnν(t)|L22|τnν(t)|V˜2dt]T E[supt[0,TSNn]|vnν(t)|H˜4]+E[0TSNn|τnν(t)|V˜4dt](3.18) which is finite because of Lemma 3.3 and Equation(3.5).

Now substituting Equation(3.13), Equation(3.15)Equation(3.18) in Equation(3.11), we have 12E[supt[0,TSNn]|Γnν(t)|L22]+νE[0TSNn|Γnν(t)|L22dt]+Cν2E[0TSNn|Γnν(t)|L22dt]12E[|Γν(0)|L22]+C(T)+C(T)(Cν+Cν|Cνa|2)+C(T)(1+Cν)E[0TSNn|Γnν(t)|L22dt]. Using Gronwall’s inequality and noting that as N, SNnTT, we finally achieve E[supt[0,T]|Γnν(t)|L22]C(T,Cν). Finally, using the definition of wnν:=Γnν+CνRτn, we have E[supt[0,T]|wnν(t)|L22]2{E[supt[0,T](|Γnν(t)|L22+Cν2|R(τnν(t))|L22)]}C(T,Cν). Note that, as 1>ν0ν>0, for a fixed ν0, we have by the definition of Cν that 1<CνCν0. Hence the right hand side of the above inequality can be further made less than some constant, which is independent of n and ν. □

Corollary 3.5.

Under the above mathematical settings, there exists a constant C > 0, which depends on the parameters of the problem but is independent of n and ν, such that E[supt[0,T]|vnν(t)|V2]C.

Proof.

Following Lemma 3.2 of [Citation35], let us consider the following elliptic problem: (3.19) Δvnν=wnνinR2,vnν(x,·)0as|x|(3.19) where =(x2,x1).

Multiplying Equation(3.19) by vnν and using integration by parts on R2, we have R2|vnν(x)|R22dx=R2wnν(x),vnν(x)R2dx=R2|wnν(x)|R22dx. This directly implies sup0st|vnν(s)|H2=sup0st|wnν(s)|H2. Then taking expectation we have by Equation(3.6) E[0t|vnν(s)|H2ds]CTE(sup0st|vnν(s)|H2)=CTE(sup0st|wnν(s)|H2)C. This completes the proof. □

Step IV: (Tightness)

Let the spaces Zi;  i=1,2 defined by (3.20) {Z1:=D([0,T];U)D([0,T];Hw)Lw2(0,T;V)L2(0,T;Hloc),Z2:=D([0,T];U˜)D([0,T];H˜w)Lw2(0,T;V˜)L2(0,T;H˜loc).(3.20) and let Ti be the supremum of the corresponding topologies in the spaces Zi. Let Z:=Z1×Z2 and T be the supremum of T1 and T2. We note that the solutions (vnν,τnν) obtained from Galerkin approximation induces sequence of measures L(vnν,τnν) on the space (Z,T). Then we have the following result:

Lemma 3.6.

(Tightness of law) The sequence {L(vnν,τnν)}nN of measures is tight on the space (Z,T).

Proof.

We observe from Step II that there exists a constant CT=CT(ν1,ν2,κ,a)>0 such that for all t[0,T],   nN (3.21a) supnNE[supt[0,T]|vnν(t)|H2],supnNE[0T|vnν(t)|V2dt]CT,(3.21a) (3.21b) supnNE[supt[0,T]|τnν(t)|H˜2],supnNE[0T|τnν(t)|V˜2dt]CT.(3.21b) Therefore from Equation(3.21) we conclude that first two conditions of Theorem A.9 for (vnν,τnν) are satisfied. Now it is sufficient to prove that (vnν)nN and (τnν)nN satisfy Aldous condition in the space U and U˜ respectively. We start with the former sequence. We will use Lemma A.7. Let (τn)nN be a sequence of stopping times such that 0τnT.

Let (ρn)nN be a sequence of stopping times such that 0ρnT. Then for t(0,ρnT) we get vnν(t)=v0nνν0tAvnν(s)ds0tBn(vnν(s))ds+0tDiv τnν(s)ds+0tZFn(vnν(s),z)N˜1(ds,dz):=i=15Jin(t). We will show that each of {Jin(t)}i=15 satisfies condition Equation(A.27) in Lemma A.7. Let us fix θ>0. Jn1 being independent of time clearly satisfies Equation(A.27) in Lemma A.7. Now exploiting the fact A:VV given by |Av|V|v|V, the embedding VU is continuous, then by Hölder’s inequality and Equation(3.21), J2n can be estimated as E[|J2n(ρn+θ)J2n(ρn)|U]C1E[θ12(0T|vnν(s)|V2ds)12]c2θ12. Thus J2n satisfies Equation(A.27) in Lemma A.7 with α = 1 and ζ=12.

Using B:V×VV is continuous and the continuous embedding VU, and Equation(3.21), we estimate J3n as E[|J3n(ρn+θ)J3n(ρn)|U]CE[0T|vnν(s)|V2ds]·θ=:c3θ. Thus J3n satisfies Equation(A.27) in Lemma A.7 with α = 1 and ζ=1.

Using Equation(3.21) and the embedding HVU we have E[|J4n(ρn+θ)J4n(ρn)|U]E[|ν1ρnρn+θ|Div τnν(s)|Vds]c4θ12. Therefore, J4n satisfies Equation(A.27) in Lemma A.7 with α = 1 and ζ=12.

Using Equation(3.21), the continuous embedding HU, Itô-Lévy isometry and Assumption 2.3, we have E[|J5n(ρn+θ)J5n(ρn)|U2]CE[ρnρn+θZ|Fn(vnν(s),z)|H2λ1(dz)ds]c5θ. Thus J5n satisfies Equation(A.27) in Lemma A.7 with α = 2 and ζ = 1.

Let us consider the equation for τnν for t(0,ρnT], τnν(t)=τ0n0tA˜τnν(s)ds0tB˜n(vnν(s),τnν(s))dsa0tτnν(s)ds+0tD(vnν(s))ds+0tBGn(l,τnν(s))N˜2(dl,ds)+0tbn(τnν(s))ds:=i=17Kin(t). Kn1 being independent of time, Equation(A.27) in Lemma A.7 is automatically satisfied for any α,β>0. Now exploiting the fact A˜:V˜V˜ given by |A˜τ|V˜|τ|V˜, the continuous embedding V˜U˜, Hölder’s inequality and Equation(3.21), K2n can be estimated as E[|K2n(ρn+θ)K2n(ρn)|U˜]C1E[θ12(0T|τnν(s)|V˜2ds)12]c˜2θ12. Thus K2n satisfies Equation(A.27) in Lemma A.7 with α = 1 and ζ=12.

Using |B˜(v,τ)|Vs˜|v|H|τ|H˜ and the embedding Vs˜U˜ we have E[|K3n(ρn+θ)K3n(ρn)|U˜]CE[(ρnρn+θ|vnν(s)|H2ds)12(ρnρn+θ|τnν(s)|H˜2ds)12]c˜3θ. Thus K3n satisfies Equation(A.27) in Lemma A.7 with α = 1 and ζ=1.

Using the embedding H˜U˜ and Equation(3.21) we have E[|Kin(ρn+θ)Kin(ρn)|U˜]c˜iθ12, for i=4,5. Thus Kin:i=4,5 satisfies Equation(A.27) in Lemma A.7 with α = 1 and ζ=12.

Using Itô-Lévy isometry, embedding H˜U˜, Lemma A.3 and Equation(3.21) we have E[|K6n(ρn+θ)K6n(ρn)|U˜2]CE[ρnρn+θB|Gn(l,τnν(s))N˜2(dl,ds)|H˜2]c˜6θ. Thus K6n satisfies Equation(A.27) in Lemma A.7 with α = 2 and ζ = 1.

Since H˜U˜ is continuous, using Schwarz’s inequality, from Lemma A.3 and Equation(3.21) we get, E[|K7n(ρn+θ)K7n(ρn)|U˜]C E[ρnρn+θ|bn(τnν(s))|H˜ ds]c˜7θ. Thus K7n(t) satisfies the condition Equation(A.27) in Lemma A.7 with α = 1 and ζ=1. This proves that the sequences (vnν)nN and (τnν)nN satisfy the Aldous condition in the space U and U˜ respectively. Thus the proof is complete. □

Step V: (Use of Skorokhod Representation Theorem)

In separable metric spaces, one can apply Prokhorov Theorem (see [Citation36], Theorem II.6.7) and classical Skorokhod Representation Theorem (see [Citation37], Theorem 6.7) to obtain a.s. convergence from tightness. However, for our purpose, we need a modified version of Skorokhod Representation Theorem due to Jakubowski [Citation29, Citation30], which can be used for non-metrizable topological space. The version used in this paper is stated below (see also [Citation31], Proposition 9).

Proposition 3.7.

Let X1 be a complete separable metric space and X2 a topological space such that there is a sequence of continuous functions fm:X2R that separates points of X2. Denote X:=X1×X2 and equip X with the topology induced by the canonical projections πj:X1×X2Xj. Let (Ω,F,P) be a probability space and (χn)nN be a tight sequence of random variables in (X,B(X1)A), where A is the σalgebra generated by fm,  mN. Assume that there is a random variable η in X1 such that Pπ1°χη=Pη. Then there are a subsequence χnk  kN and random variables χ̂k,  χ̂ in X for kN on a common probability space (Ω̂,F̂,P̂) with

  • P̂χ̂k=Pχnk for kN,

  • χ̂kχ̂ in X almost surely for k,

  • π2°χ̂k=π1°χ̂ almost surely.

Note that, by MN̂([0,T]×Z), we denote the set of all N̂-valued Borel measures ξ on [0,T]×Z with ξ(Sn)< for all nN, for some sequence Sn[0,T]×Z of Borel sets with Sn[0,T]×Z and Lebλ1(Sn)<, for all nN. It is well known that MN̂([0,T]×Z) is a complete separable metric space (see [Citation31]). Similarly, MN̂([0,T]×B) is a complete separable metric spaces.

In the line of notations used in Proposition 3.7, we denote X1=MN̂([0,T]×Z)×MN̂([0,T]×B),X2=Z1×Z2=Z. Clearly, X1 is a complete separable metric space. Since any infinite dimensional normed space with respect to weak topology in that space is not metrizable, both Z1 and Z2 are not metrizable, and so is X2 or Z.

Lemma 3.6 ensures that the sequence of measures {L(vnν,τnν)}nN is tight on the space (Z,T). Let Nin:=Ni for i=1,2. Then the set of measures {L(vnν,τnν,N1n,N2n), nN} is tight on Z×MN̂([0,T]×Z)×MN̂([0,T]×B).

We choose ν=1n. By the Skorokhod-Jakubowski theorem (see Proposition 3.7 and also Lemma A.10), there exists a subsequence (nk)kN, a filtered probability space L̂=(Ω̂,F̂,F̂,P̂) with a filtration F̂=(F̂t)t0, and on this space, Z×MN̂([0,T]×Z)×MN̂([0,T]×B)-valued random variables (v̂,τ̂,N̂1,N̂2),(v̂nν,τ̂nν,N̂1n,N̂2n),nN such that

  • L((v̂nν,τ̂nν,N̂1n,N̂2n))=L((vmnν,τmnν,N1mn,N2mn)) for all nN;

  • (v̂nν,τ̂nν,N̂1n,N̂2n)(v̂,τ̂,N̂1,N̂2) in Z×MN̂([0,T]×Z)×MN̂([0,T]×B) with probability 1 on L̂ as n;

  • (N̂1n(ω̂),N̂2n(ω̂))=(N1(ω̂),N2(ω̂)) for all ω̂Ω̂.

We will denote these sequences again by ((vnν,τnν,N1n,N2n))nN and ((v̂nν,τ̂nν,N̂1n,N̂2n))nN. Using the definiton of Z, we have P̂-a.s. (3.22a) v̂nνv̂ in Lw2(0,T;V)L2(0,T;Hloc)D([0,T];U)D([0,T];Hw),(3.22a) (3.22b) τ̂nντ̂ in Lw2(0,T;V˜)L2(0,T;H˜loc)D([0,T];U˜)D([0,T];H˜w).(3.22b) It is easy to verify that the spaces Zi,i=1,2 are not a Polish spaces. The space D([0,T];U)L2(0,T;H) is a Polish space. Then by Kuratowski theorem, D([0,T];Hn) is a Borel subset of D([0,T];U)L2(0,T;H). Hence D([0,T];Hn)Z1 is a Borel subset of D([0,T];U)L2(0,T;H)Z1, which is equal to Z1 (see Lemma 4.2 in [Citation38]). Thus we have the following proposition.

Proposition 3.8.

The set D([0,T];Hn)Z1 is a Borel subset of Z1 and the corresponding embedding transforms Borel sets into Borel subsets. A similar result is true for Z2.

The above result leads us to the following conclusion:

Corollary 3.9.

v̂nν and τ̂nν take values in Hn and H˜n respectively. The laws of vnν and v̂nν are equal on D([0,T];Hn) and the laws of τnν and τ̂nν are equal on D([0,T];H˜n).

Thus one may conclude that the limiting processes v̂ and τ̂ satisfy the following estimates for r2,   Ê[sups[0,T]|v̂(s)|Hr]Cr,Ê[sups[0,T]|τ̂(s)|H˜r]Cr, for some constant Cr (depending on r).

Step VI: (Convergence)

Let us define the following functionals for t[0,T], and for all ϕU,  ψU˜, (3.23a) Sn(v̂nν,τ̂nν,N̂1n,ϕ)(t):=(v̂nν(0),ϕ)Hν0tAv̂nν(s),ϕ ds0tBn(v̂nν(s)),ϕ ds+0tDiv τ̂nν(s),ϕ ds+0tZ(Fn(v̂nν(s),z),ϕ)H N̂˜1n(ds,dz),(3.23a) (3.23b) Λn(v̂nν,τ̂nν,N̂2n,ψ)(t):=(τ̂nν(0),ψ)H˜0tA˜τ̂nν(s),ψ ds0tB˜n(v̂nν(s),τ̂nν(s)),ψ dsa0tτ̂nν(s)),ψ ds+0tD(v̂nν(s)),ψ ds+0tB(Gn(l,τ̂nν(s)),ψ)H˜ N̂˜2n(ds,dl)+0tbn(τ̂nν(s)),ψ ds,(3.23b) (3.23c) S(v̂,τ̂,N̂1,ϕ)(t):=(v̂(0),ϕ)H0tB(v̂(s)),ϕ ds+0tDiv τ̂(s),ϕ ds+0tZ(F(v̂(s),z),ϕ)H N̂˜1(ds,dz),(3.23c) (3.23d) Λ(v̂,τ̂,N̂2,ψ)(t):=(τ̂(0),ψ)H˜0tA˜τ̂(s),ψ ds0tB˜(v̂(s),τ̂(s)),ψ dsa0tτ̂(s)),ψ ds+0tD(v̂(s)),ψ ds+0tB(Gn(l,τ̂(s)),ψ)H˜ N̂˜2(ds,dl)+0tb(τ̂(s)),ψ ds.(3.23d) Next, we claim that (3.24a) 1.  limn(v̂nν(·),ϕ)H(v̂(·),ϕ)HL2([0,T]×Ω̂)=0,(3.24a) (3.24b) 2.  limnSn(v̂nν,τ̂nν,N̂1n,ϕ)S(v̂,τ̂,N̂1,ϕ)L2([0,T]×Ω̂)=0,(3.24b) (3.24c) 3.  limn(τ̂nν(·),ψ)H˜(τ̂(·),ψ)H˜L2([0,T]×Ω̂)=0,(3.24c) (3.24d) 4.  limnΛn(v̂nν,τ̂nν,N̂2n,ψ)Λ(v̂,τ̂,N̂2,ψ)L2([0,T]×Ω̂)=0,(3.24d) which is a direct consequence of the following Proposition.

Proposition 3.10.

If Sn, Λn, S, Λ are defined by Equation(3.23), then we have the following convergences: (3.25) 1.  limnSn(v̂nν,τ̂nν,N̂1n,ϕ)S(v,τ,N1,ϕ)L2([0,T]×Ω̂)=0,(3.25) (3.26) 2.  limnΛn(v̂nν,τ̂nν,N̂2n,ϕ)Λ(v,τ,N2,ϕ)L2([0,T]×Ω̂)=0.(3.26)

We postpone the proof of the Proposition in Appendix A.1. Hence from Equation(3.24), we infer that for Leb-almost all t[0,T], all ϕU and ψU˜, and P̂-almost all ωΩ̂, we infer that (v̂,τ̂) satisfies (v̂(t),ϕ)H+0tB(v̂(s)),ϕ ds+0tDiv τ̂(s),ϕ ds=(v̂0,ϕ)H+0tZ(F(v̂(s),z),ϕ)HN̂˜1(ds,dz),(τ̂(t),ψ)H˜+0tA˜τ̂(s),ψ ds+0tB˜(v̂(s),τ̂(s)),ψ ds+a0tτ̂(s),ψ ds=(τ̂(0),ψ)H˜+0tD(v̂(s)),ψ ds+0tb(τ̂(s)),ψ ds+0tB(G(l,τ̂(s)),ψ)H˜ N̂˜2(ds,dl). Finally, since U is dense in V and U˜ is dense in V˜, we infer that the above two equalities hold for all t[0,T] and all ϕV and ψV˜ respectively. This concludes that the system (L̂,v̂,τ̂,N̂1,N̂2) is a weak martingale solution of Equation(3.1). Thus the proof of Theorem 3.1 is complete. □

Acknowledgements

The first author would like to thank Professor Zdzisław Brzeźniak for many useful discussions on Marcus canonical SDEs. The second author would like to thank Professor Erika Hausenblas for useful discussions and partial support from the FWF-Project P32295.

Disclosure statement

Both the authors declare that they have no conflict of interest.

Additional information

Funding

The research of the first author is partially supported by his Royal Society International Exchange Grant entitled “Stochastic Landau-Lifshitz-Gilbert equation with Lévy noise and ferromagnetism” (ref: IE140328).

Notes

1 Supremum topology is the topology generated by the collection S=i=14Ti, i.e., the topology generated from the collection S as a subbasis. Let Ki be the supremum of the corresponding topologies in the spaces Zi,i=1,2. This supremum topology is different from the subspace topology.

References

Appendix A.

Technical lemmas

We now state some useful results obtained by the first author and his collaborator [Citation32].

Lemma A.1.

  • There exists M1>0 such that for any lB and τH˜n, |Gn(l,τ)|H˜nM1|l|Rk(1+|τ|H˜n).

  • There exists M2>0 such that for any lB and τ1,τ2H˜n, |Gn(l,τ2)Gn(l,τ1)|H˜nM2|l|Rk|τ2τ1|H˜n.

  • There exists M3>0 such that for any lB and τH˜n, |Kn(l,τ)|H˜nM3|l|Rk2(1+|τ|H˜n).

  • There exists M4>0 such that for any lB and τ1,τ2H˜n, |Kn(l,τ2)Kn(l,τ1)|H˜nM4|l|Rk2|τ2τ1|H˜n.

Lemma A.2.

There exists a constant C1>0 such that for any τ1,τ2H˜n, |bn(τ2)bn(τ1)|H˜n2+B|Gn(l,τ2)Gn(l,τ2)|H˜n2 λ2(dl)C1|τ2τ1|H˜n2.

Lemma A.3.

There exists a constant C2>0 such that for any τH˜n, (A.1) |bn(τ)|H˜n2+B|Gn(l,τ)|H˜n2 λ2(dl)C2|τ|H˜n2.(A.1)

Below we provide Proof of Lemma 2.4.

Proof of Lemma 2.4.

Proof. Since ψ(z)=|z|H˜p, for all g,kH˜, ψ(z)g=p|z|H˜p2 z,gH˜, andψ(z)(g,k)=[p(p2)|z|H˜p4 (zz)+p|z|H˜p2]g,kH˜.

  1. Note that ψ(Φ(l,τ))ψ(τ)=ψ(y(1))ψ(y(0))=01dds[ψ°y](s)ds=01ψ(y(s))(Ry(s))ds. Hence |ψ(Φ(l,τ))ψ(τ)|p01|y(s)|H˜p1|Ry(s)|H˜dspRL(H˜)01|y(s)|H˜pdsNp|l|RkgL(H˜)|τ|H˜p. This proves first part of the Lemma.

  2. We begin with the observation that (A.2) ψ(Φ(l,τ))ψ(τ)ψ(τ)Rτ=ψ(y(1))ψ(y(0))ψ(y(0))R(y(0)):=01(ψ°y)(s)ds(ψ°y)(0)=010sψ(y(r))(Ry(r),Ry(r))drds+010sψ(y(r))(R2y(r))drds:=I1+I2.(A.2) Hence as 0s1, (A.3) |I1|p(p1)01|y(s)|H˜p2 |Ry(s)|H˜2dsp(p1)RL(H˜)201|y(s)|H˜pdsp(p1)RL(H˜)2|τ|H˜p01esRpdsNp(p1)RL(H˜)2|τ|H˜p.(A.3) Similarly, one obtains (A.4) |I2|pRL(H˜)201|y(s)|H˜pdsNpRL(H˜)2|τ|H˜p.(A.4) Therefore using the estimates Equation(A.3) and Equation(A.4) in Equation(A.2) we have |ψ(Φ(l,τ))ψ(τ)ψ(τ)Rτ|Np2RL(H˜)2|τ|H˜pNp2|l|Rk2gL(H˜)2|τ|H˜p.

A.1. Proof of Proposition 3.10

We now see using Fubini’s Theorem that (A.5) Sn(v̂nν,τ̂nν,N̂1n,ϕ)S(v,τ,N1,ϕ)L2([0,T]×Ω̂)2=0TÊ[|Sn(v̂nν,τ̂nν,N̂1n,ϕ)(t)S(v,τ,N1,ϕ)(t)|2] dt.(A.5) Lemma A.4 (the following Lemma) ensures that each term on the right hand side of Equation(3.23a) converges to the right hand side of corresponding term in Equation(3.23c) in L2([0,T]×Ω̂) which further assures that right hand side of Equation(A.5) goes to zero as n. This verifies Equation(3.25). Lemma A.5 (which is proven below) ensures that each term on the right hand side of Equation(3.23b) converges to the right hand side of corresponding term in Equation(3.23d) in L2([0,T]×Ω̂) which further verifies Equation(3.26). This completes the proof.

Lemma A.4.

For all ϕU,

  1. limnÊ[0T|(v̂nν(t)v(t),ϕ)H|2 dt]=0,

  2. limn0TÊ[|(v̂nν(0)v(0),ϕ)H|2]dt=0,

  3. limn0TÊ[|0tAv̂nν(s)Av(s),ϕ ds|2] dt=0,

  4. limn0TÊ[|0tBn(v̂nν(s))B(v(s)),ϕ ds|2] dt=0,

  5. limn0TÊ[|0tDiv τ̂nν(s)Div τ(s),ϕ ds|2] dt=0,

  6. limn0TÊ[|0tZFn(v̂nν(s),z)F(v(s),z),ϕ λ1(dz)ds|2]dt=0.

  7. limn0TÊ[|0tZFn(v̂nν(s),z)F(v(s),z),ϕ N˜1(ds,dz)|2]dt=0.

Proof.

Let ϕU be fixed.

  1. Owing to Equation(3.22), in particular for almost all t[0,T], limn(v̂nν(t),ϕ)H=(v(t),ϕ)H  P̂a.s. Also using Ê[0T|(v̂nν(t)v(t),ϕ)H|2dt]C, and by employing Vitali Theorem we have limn(v̂nν,ϕ)H(v,ϕ)HL2([0,T]×Ω̂)2=limnÊ[0T|(v̂nν(t)v(t),ϕ)H|2dt]=0.

  2. Equation Equation(3.22) ensures that (v̂nν(0),ϕ)H(v(0),ϕ)H  P̂a.s., which further with Vitali Theorem infer limn(v̂nν(0)v(0),ϕ)HL2([0,T]×Ω̂)2=0.

  3. Owing to Equation(3.22), we have for any ϕ˜L2(0,T;V) (A.6) limn0TAv̂nν(s)Av(s),ϕ˜(s)ds=0.(A.6) Let ϕU. Let t[0,T] be fixed. We choose ϕ˜(s)=χ(0,t)(s)ϕ and note that ϕ˜L2(0,T;V). Hence for r>2, using Equation(A.6) and Hölder’s inequality, we achieve for all t[0,T], and nN, (A.7) Ê[|0tAv̂nν(s),ϕ ds|r]Ê[0t|(v̂nν(s),ϕ)V|r ds]2rTrϕUr Ê[sup0st|v̂nν(s)|Vr]cr(A.7) for some positive constant cr (depending on r). Therefore using Equation(A.7) and by Vitali Theorem, we conclude for all t[0,T] (A.8) limnÊ[|0tAv̂nν(s)Av(s),ϕds|2]=0.(A.8) Hence Equation(A.7), Equation(A.8) and the Dominated Convergence Theorem proves (iii).

  4. Let ϕU. We have, B(v̂nν)B(v):=B(v̂nν,v̂nν)B(v,v)=B(v̂nνv,v̂nν)+B(v,v̂nνv). Now, using Hölder’s inequality and v̂nνv in L2(0,T;H), we have limn0tB(v̂nν(s))B(v(s)),ϕds=0P̂a.s. For every ϕU, since Pnϕϕ in Vs and UV1, we conclude that for all ϕU and all t[0,T], (A.9) limn0tBn(v̂nν(s))B(v(s)),ϕds=0P̂a.s.(A.9) Using Hölder’s inequality we obtain for all t[0,T], r>2 and nN, (A.10) Ê[|0tBn(v̂nν(s)),ϕds|r]Ê[(0t|Bn(v̂nν(s))|Vs|ϕ|Vsds)r]CC1(r).(A.10) Considering Equation(A.9) and Equation(A.10) and by the Vitali Theorem we achieve for all t[0,T] (A.11) limnÊ[|0tBn(v̂nν(s))B(v(s)),ϕds|2]=0.(A.11) Hence, in view of Equation(A.10), Equation(A.11) and the Dominated Convergence Theorem, we infer that (A.12) limn0TÊ[|0tBn(v̂nν(s))B(v(s)),ϕds|2]dt=0.(A.12) Hence we have (iv).

  5. As ϕU˜, so ϕH˜. Now using Equation(3.22) we have (A.13) |0tDiv τ̂nν(s)Div τ(s),ϕ ds|2=|0tτ̂nν(s)τ(s),ϕds|20asn.(A.13) Let r2. Now using Equation(3.22) we have (A.14) 0TÊ[|0tDiv τ̂nν(s)Div τ(s),ϕds|r]=0TÊ[|0tτ̂nν(s)τ(s),ϕds|r]Cr,(A.14) for some positive constant Cr. Hence by Vitali Theorem, using Equation(A.13) and Equation(A.14) we have (v).

  6. Let us consider ϕU. Using Assumption 2.3 we have (A.15) 0tZ|(F(v̂nν(s),z)F(v(s),z),ϕ)H|2λ1(dz)ds0asn.(A.15) Furthermore, using Assumption 2.3, for every t[0,T], r>2 and nN, we have the following inequality (A.16) Ê[0tZ|(F(v̂nν(s),z)F(v(s),z),ϕ)H|2λ1(dz)ds|r]C˜r,(A.16) for some constant C˜r>0. Since the restriction of Pn to the space H is the (·,·)H-projection onto Hn, hence by Equation(A.15), Equation(A.16) and by the Vitali Theorem we infer that for all t[0,T],  ϕU (A.17) limnÊ[0tZ|(Fn(v̂nν(s),z)F(v(s),z),ϕ)H|2λ1(dz)ds]=0,ϕH.(A.17) Since UH, Equation(A.17) holds for all ϕU. Moreover, Ê[0tZ|(Fn(v̂nν(s),z)F(v(s),z),ϕ)H|2λ1(dz)ds]C˜2 and Equation(A.17) and the Dominated Convergence Theorem assures assertion (vi).

  7. Using Itô-Lévy isometry, and the fact that N̂in=Ni for i=1,2, and Equation(A.17) and then exploiting Dominated Convergence Theorem we ensure assertion (vii).

Lemma A.5.

For all ψU˜,

  1. limnÊ[0T|(τ̂nν(t)τ(t),ψ)H˜|2 dt]=0,

  2. limn0TÊ[|(τ̂nν(0)τ(0),ψ)H˜|2] dt=0,

  3. limn0TÊ[|0tA˜τ̂nν(s)A˜τ(s),ψ ds|2] dt=0,

  4. limn0TÊ[|0tB˜n(v̂nν(s),τ̂nν(s))B˜(v(s),τ(s)),ψ ds|2] dt=0,

  5. limn0TÊ[|0tD(v̂nν(s))D(v(s)),ψ ds|2] dt=0,

  6. limn0TÊ[|0tBGn(l,τ̂nν(s))G(l,τ(s)),ψ N˜2(dl,ds)|2 dt]=0,

  7. limn0TÊ[|0tbn(τ̂nν(s))b(τ(s)),ψ ds|2] dt=0.

Proof.

Proofs of (i)–(iv) are similar to the proofs of Lemma A.4 (i)–(iv), and hence omitted.

(v) Let ψU˜. Now using integration by parts and (3.22) we note that (A.18) |0tD(v̂nν(s))D(v(s)),ψds|20asn.(A.18) Let r2. Now using Equation(3.22) we have 0TÊ[|0tD(v̂nν(s))D(v(s)),ψ ds|r] dtCr, for some positive constant Cr depending on r. Hence by Vitali Theorem, using Equation(A.18) we ensure assertion (v).

(vi) Using Lipschitz property of G and Equation(3.22), we obtain for all ψH˜, for all t[0,T], P̂-a.s. (A.19) 0tB|(G(l,τ̂nν(s))G(l,τ(s)),ψ)H˜|2 λ2(dl)dsC0T|τ̂nν(s)τ(s)|H˜2 ds0asn.(A.19) Moreover, for every t[0,T], every r1 and every nN, (A.20) Ê[|0tB|(G(l,τ̂nν(s))G(l,τ(s)),ψ)H˜|2 λ2(dl)ds|r]C.(A.20) Since the restriction of P˜n to the space H˜ is the (·,·)H˜-projection onto H˜n, then by Equation(A.19), Equation(A.20) and by Vitali’s Theorem, we obtain (A.21) limnÊ[0tB|(Gn(l,τ̂nν(s))G(l,τ(s)),ψ)H˜|2 λ2(dl)ds]=0,ψH˜.(A.21) Since U˜H˜, Equation(A.21) holds for all ψU˜.

As N̂2n=N2, for all nN. From Equation(A.21) and the Itô isometry we have, (A.22) limnÊ[|0tBGn(l,τ̂nν(s))G(l,τ(s)),ψ N˜2(ds,dl)|2]=0.(A.22) Moreover, from Equation(A.20) and the Itô isometry, with r=1, we obtain, (A.23) Ê[|0tBGn(l,τ̂nν(s))G(l,τ(s)),ψ N˜2(ds,dl)|2]C.(A.23) Finally, from Equation(A.22), Equation(A.23) and using Dominated Convergence Theorem we obtain (vi).

(vii) Since, τ̂nντ in L2(0,T;V˜), exploiting the Lipschitz property of b, we obtain, (A.24) limn0tbn(τ̂nν(s))b(τ(s)),ψ ds=limn0tb(τ̂nν(s))b(τ(s)),P˜nψ ds=0(A.24) Now using Lemma A.1, for r1 we get, Ê[|0tbn(τ̂nν(s)),ψ ds|r]Cr, which further with Equation(A.24), and Vitali’s theorem we get, (A.25) limnÊ[|0tbn(τ̂nν(s))b(τ(s)),ψ ds|2]=0.(A.25) Finally from Equation(A.25) and Dominated convergence theorem we get limn0TÊ[|0tbn(τ̂nν(s))b(τ(s)),ψ ds|2] dt=0.

A.2. Compactness and tightness criteria

Let S be a complete separable metric space with a metric ρ. Let us denote by D([0,T];S), the set of all S-valued functions defined on [0,T], which are right continuous and have left limits (càdlàg functions) for every t[0,T]. The space D([0,T];S) is endowed with the Skorokhod J-topology. For more details see Métivier’s [Citation39, Chapter II] and Billingsley’s [Citation37, Chapter 3].

A.2.1. The Aldous condition

Let (S,ρ) be a complete, separable metric space. Let (Ω,F,F,P) be a probability space with usual hypotheses.

Definition A.1.

Let uD([0,T];S) and let δ>0 be given. A modulus of continuity is defined by (A.26) W[0,T],S(u;δ):=infΠδmaxtiω˜suptis<t<ti+1Tρ(u(t),u(s)),(A.26) where Πδ is the set of all increasing sequences ω˜={0=t0<t1<<tn=T} with the following property ti+1tiδ,i=0,1,,n1.

Definition A.2.

Let (Xn) be a sequence of S-valued random variables. The sequence of laws of these processes is Tight if and only if

[T]  ε>0 η>0 δ>0: supnNP{W[0,T],S(Xn,δ)>η}ε.

Definition A.3.

A sequence (Xn)nN of S-valued random variables satisfies the Aldous condtion if and only if

[A]  ϵ>0 η>0 δ>0 such that for every sequence (τn)nN of F-stopping times with τnT one has supnN sup0θδP{ρ(Xn(τn+θ),Xn(τn))η}ϵ.

Lemma A.6.

Condition [A] implies condition [T].

Proof.

See Theorem 2.2.2 of [Citation40]. □

Lemma A.7.

Let (E,|·|E) be a separable Banach space and let (Xn)nN be a sequence of E-valued random variables. Assume that for every sequence (τn)nN of F-stopping times with τnT and for every nN and θ0 the following condition holds (A.27) E[|Xn(τn+θ)Xn(τn)|Eα]Cθβ(A.27) for some α,β>0 and some constant C>0. Then the sequence (Xn)nN satisfies the Aldous condtion in the space E.

Proof.

See [Citation33] for the proof. □

Theorem A.8.

(Compactness criteria for (v,τ)) Let Z1:=D([0,T];U)D([0,T];Hw)Lw2(0,T;V)L2(0,T;Hloc) and T1 be the supremum of the corresponding topologies. Let Z2:=D([0,T];U˜)D([0,T];H˜w)Lw2(0,T;V˜)L2(0,T;H˜loc) and T2 be the supremum of the corresponding topologies. Define Z:=Z1×Z2 and T as the supremum of T1 and T2. Then (K1,K2)Z is Trelatively compact if the following three conditions are satisfied:

  1. (v,τ)K1×K2 and all t[0,T],(v(t),τ(t))H×H˜, and sup(v,τ)K1×K2(sups[0,T]|v(s)|H+sups[0,T]|τ(s)|H˜)<,

  2. sup(v,τ)K1×K20T(|v(s)|V2+|τ(s)|V˜2)ds<, i.e. (K1,K2) is bounded in L2(0,T;V)×L2(0,T;V˜),

  3. limδ0sup(v,τ)K1×K2(w[0,T],V(v,δ)+w[0,T],V˜(τ,δ))=0.

Proof of Theorem A.8 can be found in Lemma 3.3 in Brzeźniak and Motyl [Citation23], Lemma 4.1 in Motyl [Citation41], Theorem 2 of Motyl [Citation33], Lemma 2.7 in Mikulevicius and Rozovskii [Citation42].

Theorem A.9.

Let (vn,τn)nN be a sequence of càdlàg F-adapted V×V˜-valued processes such that

  1. there exists a positive constant C1 such that supnNE[sups[0,T](|vn(s)|H+|τn(s)|H˜)]C1,

  2. there exists a positive constant C2 such that supnNE[0T(|vn(s)|V2+|τn(s)|V˜2)ds]C2,

  3. (vn)nN (respectively (τn)nN) satisfies the Aldous condition in V (respectively V˜).

Let (Pn1,Pn2) be the law of (vn,τn)nN on Z:=Z1×Z2. Then, for every ϵ>0, there exist compact subset (Kϵ1,Kϵ2) of Z such that Pni(Kϵ1)1ϵ,  for i = 1, 2 and the sequence of measures {(Pn1,Pn2),nN} is said to be tight on (Z,T).

For a relevant proof see Corollary 1, Motyl [Citation33].

Lemma A.10.

Let (vn,τn)nN be a sequence of adapted U×U˜-valued processes satisfying the Aldous condition in U×U˜ and supnNE[||vn||L(0,T;H)2+||τn||L(0,T;H˜)2]<,supnNE[||vn||L2(0,T;V)2+||τn||L2(0,T;V˜)2]<. Then, there are a subsequence (vnk,τnk)kN and random variables (v̂k,τ̂k) and (v̂,τ̂) for kN on a common probability space L̂ with P̂v̂k=Pvnk and P̂τ̂k=Pτnk for kN, and (v̂k,τ̂k)(v̂,τ̂) P̂-almost surely in Z for k.