Skip to main content
Springer logoLink to Springer
. 2020 Jul 11;84(2):1971–2035. doi: 10.1007/s00245-020-09702-2

Stochastic Navier–Stokes Equations on a Thin Spherical Domain

Zdzisław Brzeźniak 1, Gaurav Dhariwal 2,, Quoc Thong Le Gia 3
PMCID: PMC8550046  PMID: 34720249

Abstract

Incompressible Navier–Stokes equations on a thin spherical domain Qε along with free boundary conditions under a random forcing are considered. The convergence of the martingale solution of these equations to the martingale solution of the stochastic Navier–Stokes equations on a sphere S2 as the thickness converges to zero is established.

Keywords: Stochastic Navier–Stokes equations, Navier–Stokes equations on a sphere, Singular limit

Introduction

For various motivations, partial differential equations in thin domains have been studied extensively in the last few decades; e.g. Babin and Vishik [4], Ciarlet [16], Ghidaglia and Temam [18], Marsden et al. [37] and references there in. The study of the Navier–Stokes equations (NSE) on thin domains originates in a series of papers by Hale and Raugel [2022] concerning the reaction-diffusion and damped wave equations on thin domains. Raugel and Sell [44, 45] proved the global existence of strong solutions to NSE on thin domains for large initial data and forcing terms, in the case of purely periodic and periodic-Dirichlet boundary conditions. Later, by applying a contraction principle argument and carefully analysing the dependence of the solution on the first eigenvalue of the corresponding Laplace operator, Arvin [2] showed global existence of strong solutions of the Navier–Stokes equations on thin three-dimensional domains for large data. Temam and Ziane [52] generalised the results of [44, 45] to other boundary conditions. Moise et al. [41] proved global existence of strong solutions for initial data larger than in [45]. Iftimie [26] showed the existence and uniqueness of solutions for less regular initial data which was further improved by Iftimie and Raugel [27] by reducing the regularity and increasing the size of initial data and forcing.

In the context of thin spherical shells, large-scale atmospheric dynamics that play an important role in global climate models and weather prediction can be described by the 3-dimensional Navier–Stokes equations in a thin rotating spherical shell [34, 35]. Temam and Ziane in [53] gave the mathematical justification for the primitive equations of the atmosphere and the oceans which are known to be the fundamental equations of meteorology and oceanography [36, 43]. The atmosphere is a compressible fluid occupying a thin layer around the Earth and whose dynamics can be described by the 3D compressible Navier–Stokes equations in thin layers. In [53] it was assumed that the atmosphere is incompressible and hence a 3D incompressible NSE on thin spherical shells could be used as a mathematical model. They proved that the averages in the radial direction of the strong solutions (whose existence for physically relevant initial data was established in the same article) to the NSE on the thin spherical shells converge to the solution of the NSE on the sphere as the thickness converges to zero. In a recent paper Saito [46] studied the 3D Boussinesq equations in thin spherical domains and proved the convergence of the average of weak solutions of the 3D Boussinesq equations to a 2D problem. More recent work on incompressible viscous fluid flows in a thin spherical shell was carried out in [2325].

For the deterministic NSE on the sphere, Il’in and Filatov [2830] considered the existence and uniqueness of solutions while Temam and Wang [51] considered inertial forms of NSE on spheres. Brzeźniak et al. proved the existence and uniqueness of the solutions to the stochastic NSE on the rotating two dimensional sphere and also proved the existence of an asymptotically compact random dynamical system [9]. Recently, Brzeźniak et al. established [10] the existence of random attractors for the NSE on two dimensional sphere under random forcing irregular in space and time deducing the existence of an invariant measure.

The main objective of this article is to establish the convergence of the martingale solution of the stochastic Navier–Stokes equations (SNSE) on a thin spherical domain Qε, whose existence can be established as in the forthcoming paper [7] to the martingale solution of the stochastic Navier–Stokes equations on a two dimensional sphere S2 [9] as thickness ε of the spherical domain converges to zero. In this way we also give another proof for the existence of a martingale solution for stochastic NSE on the unit sphere S2.

We study the stochastic Navier–Stokes equations (SNSE) for incompressible fluid

du~ε-[νΔu~ε-(u~ε·)u~ε-p~ε]dt=f~εdt+G~εdW~ε(t)inQε×(0,T), 1
divu~ε=0inQε×(0,T), 2

in thin spherical shells

Qε:=yR3:1|y|1+ε,where0<ε<1/2, 3

along with free boundary conditions

u~ε·n=0,curlu~ε×n=0onQε×(0,T), 4
u~ε(0,·)=u~0εinQε. 5

In the above, u~ε=(u~εr,u~ελ,u~εφ) is the fluid velocity field, p is the pressure, ν>0 is a (fixed) kinematic viscosity, u~0ε is a divergence free vector field on Qε and n is the unit outer normal vector to the boundary Qε and W~ε(t), t0 is an RN-valued Wiener process in some probability space Ω,F,F,P to be defined precisely later.

The main result of this article is Theorem 3, which establishes the convergence of the radial averages of the martingale solution (see Definition 1) of the 3D stochastic equations (1)–(5), as the thickness of the shell ε0, to a martingale solution u (see Definition 2) of the following stochastic Navier–Stokes equations on the unit sphere S2R3:

du-νΔu-(u·)u-pdt=fdt+GdWinS2×(0,T), 6
divu=0inS2×(0,T), 7
u(0,·)=u0inS2, 8

where u=(uλ,uφ) and Δ, are the Laplace–de Rham operator and the surface gradient on S2 respectively. Assumptions on initial data and external forcing will be specified later.

The paper is organised as follows. We introduce necessary functional spaces in Sect. 2. In Sect. 3, we define some averaging operators and give their properties. Stochastic Navier–Stokes equations on thin spherical domains are introduced in Sect. 4 and a priori estimates for the radially averaged velocity are obtained which are later used to prove the convergence of the radial average of a martingale solution of stochastic NSE on thin spherical shell (see (1)–(5)) to a martingale solution of the stochastic NSE on the sphere (see (6)–(8)) with vanishing thickness.

Preliminaries

A point yQε could be represented by the Cartesian coordinates y=x,y,z or y=r,λ,φ in spherical coordinates, where

x=rsinλcosφy=rsinλsinφz=rcosλ,

for r(1,1+ε), λ[0,π] and φ[0,2π).

For p[1,), by Lp(Qε), we denote the Banach space of (equivalence-classes of) Lebesgue measurable R-valued pth power integrable functions on Qε. The R3-valued pth power integrable vector fields will be denoted by Lp(Qε). The norm in Lp(Qε) is given by

uLp(Qε):=Qε|u(y)|pdy1/p,uLp(Qε).

If p=2, then L2(Qε) is a Hilbert space with the inner product given by

u,vL2(Qε):=Qεu(y)·v(y)dy,u,vL2(Qε).

By H1(Qε)=W1,2(Qε), we will denote the Sobolev space consisting of all uL2(Qε) for which there exist weak derivatives DiuL2(Qε), i=1,2,3. It is a Hilbert space with the inner product given by

u,vH1(Qε):=u,vL2(Qε)+u,vL2(Qε),u,vH1(Qε),

where

u,vL2(Qε)=i=13QεDiu(y)·Div(y)dy.

The Lebesgue and Sobolev spaces on the sphere S2 will be denoted by Lp(S2) and Ws,q(S2) respectively for p,q1 and s0. In particular, we will write H1(S2) for W1,2(S2).

Functional Setting on the Shell Qε

We will use the following classical spaces on Qε:

Hε=uL2(Qε):divu=0inQε,u·n=0onQε,Vε=uH1(Qε):divu=0inQε,u·n=0onQε,=H1(Qε)Hε.

On Hε, we consider the inner product and the norm inherited from L2(Qε) and denote them by ·,·Hε and ·Hε respectively, that is

u,vHε:=u,vL2(Qε),uHε:=uL2(Qε),u,vHε.

Let us define a bilinear map aε:Vε×VεR by

aε(u,v):=curlu,curlvL2(Qε),u,vVε, 9

where

curlu=×u,

and for uVε, we define

uVε2:=aε(u,u)=curluL2(Qε)2. 10

Note that for uVε, uVε=0 implies that u is a constant vector and u·n=0 on Qε i.e., u is tangent to Qε for every yQε, and thus must be 0. Hence ·Vε is a norm on Vε (other properties can be verified easily). Under this norm Vε is a Hilbert space with the inner product given by

u,vVε:=curlu,curlvL2(Qε),u,vVε.

We denote the dual pairing between Vε and Vε by ·,·ε, that is ·,·ε:=Vε·,·Vε. By the Lax–Milgram theorem, there exists a unique bounded linear operator Aε:VεVε such that we have the following equality:

Aεu,vε=u,vVε,u,vVε. 11

The operator Aε is closely related to the Stokes operator Aε defined by

D(Aε)=uVε:AεuHε,curlu×n=0onQε,Aεu=Aεu,uD(Aε). 12

The Stokes operator Aε is a non-negative self-adjoint operator in Hε (see Appendix B). Also note that

D(Aε)=uH2(Qε):divu=0inQε,u·n=0andcurlu×n=0onQε.

We recall the Leray–Helmholtz projection operator Pε, which is the orthogonal projector of L2(Qε) onto Hε. Using this, the Stokes operator Aε can be characterised as follows:

Aεu=Pε(-Δu),uD(Aε). 13

We also have the following characterisation of the Stokes operator Aε [53, Lemma 1.1]:

Aεu=curlcurlu,uD(Aε). 14

For uVε, vD(Aε), we have the following identity (see Lemma 28)

curlu,curlvL2(Qε)=u,AεvL2(Qε). 15

Let bε be the continuous trilinear from on Vε defined by:

bε(u,v,w)=Qεu·v·wdy,u,v,wVε. 16

We denote by Bε the bilinear mapping from Vε×Vε to Vε by

Bε(u,v),wε=bε(u,v,w),u,v,wVε,

and we set

Bε(u)=Bε(u,u).

Let us also recall the following properties of the form bε, which directly follows from the definition of bε :

bε(u,v,w)=-bε(u,w,v),u,v,wVε. 17

In particular,

Bε(u,v),vε=bε(u,v,v)=0,u,vVε. 18

Functional Setting on the Sphere S2

Let s0. The Sobolev space Hs(S2) is the space of all scalar functions ψL2(S2) such that (-Δ)s/2ψL2(S2), where Δ is the Laplace–Beltrami operator on the sphere (see (177)). We similarly define Hs(S2) as the space of all vector fields uL2(S2) such that (-Δ)s/2uL2(S2), where Δ is the Laplace–de Rham operator on the sphere (see (180)).

For s0, Hs(S2),·Hs(S2) and Hs(S2),·Hs(S2) are Hilbert spaces under the respective norms, where

ψHs(S2)2=ψL2(S2)2+(-Δ)s/2ψL2(S2)2,ψHs(S2) 19

and

uHs(S2)2=uL2(S2)2+(-Δ)s/2uL2(S2)2,uHs(S2). 20

By the Hodge decomposition theorem [3, Theorem 1.72] the space of C smooth vector fields on S2 can be decomposed into three components:

C(TS2)=GVH, 21

where

G={ψ:ψC(S2)},V={curlψ:ψC(S2)}, 22

and H is the finite-dimensional space of harmonic vector fields. Since the sphere is simply connected, H={0}. We introduce the following spaces

H=closureofVinL2(S2),V=closureofVinH1(S2).

Note that it is known (see [50])

H={uL2(S2):divu=0},V=HH1(S2).

Given a tangential vector field u on S2, we can find vector field u~ defined on some neighbourhood of S2 such that their restriction to S2 is equal to u, that is u~|S2=uTS2. Then we define

curlu(x):=(x·(×u~))|S2=(x·curlu~)|S2. 23

Since x is orthogonal to the tangent plane TxS2, curlu is the normal component of ×u~. It could be identified with a normal vector field when needed.

We define the bilinear form a:V×VR by

a(u,v):=(curlu,curlv)L2(S2),u,vV.

The bilinear from a satisfies a(u,v)uH1(S2)vH1(S2) and hence is continuous on V. So by the Riesz representation theorem, there exists a unique operator A:VV such that a(u,v)=VAu,vV for u,vV. Using the Poincaré inequality, we also have a(u,u)αuV2, for some positive constant α, which means a is coercive in V. Hence, by the Lax–Milgram theorem, the operator A:VV is an isomorphism.

Next we define an operator A in H as follows:

D(A):={uV:AuH},Au:=Au,uD(A). 24

By Cattabriga [15], see also Temam [49, p. 56], one can show that A is a non-negative self-adjoint operator in H. Moreover, V=D(A1/2), see [49, p. 57].

Let P be the orthogonal projection from L2(S2) to H, called the Leray–Helmholtz projection. It can be shown, see [19, p. 104], that

D(A)=H2(S2)H,Au=P-Δu,uD(A). 25

D(A) along with the graph norm

uD(A)2:=uL2(S2)2+AuL2(S2)2,uD(A),

forms a Hilbert space with the inner product

u,vD(A):=u,vL2(S2)+Au,AvL2(S2),u,vD(A).

Note that D(A)-norm is equivalent to H2(S2)-norm. For more details about the Stokes operator on the sphere and fractional power As for s0, see [9, Sec. 2.2].

Given two tangential vector fields u and v on S2, we can find vector fields u~ and v~ defined on some neighbourhood of S2 such that their restrictions to S2 are equal to, respectively, u and v. Then we define the covariant derivative

[vu](x)=πxi=13v~i(x)iu~(x)=πx((v~(x)·)u~(x)),xS2,

where πx is the orthogonal projection from R3 onto the tangent space TxS2 to S2 at x. By decomposing u~ and v~ into tangential and normal components and using orthogonality, one can show that

πx(u~×v~)=u×((x·v)x)+(x·u)x×v=u×((x·v)x),xS2, 26

where in the last equality, we use the fact that x·v=0 for any tangential vector v.

We set v=u and use the formula

(u~·)u~=|u~|22-u~×(×u~)

to obtain

[uu](x)=|u|22-πxu~×(×u~).

Using (26) for the vector fields u~ and v~=×u~=curlu~, we have

πx(u~×(×u~))=u×((x·curlu)x)=u×curlu.

Thus

uu=|u|22-u×curlu.

We consider the trilinear form b on V×V×V, defined by

b(v,w,z)=(vw,z)=S2vw·zdσ(x),v,w,zV, 27

where dσ(x) is the surface measure on S2.

Averaging Operators and Their Properties

In this section we recall the averaging operators which were first introduced by Raugel and Sell [44, 45] for thin domains. Later, Temam and Ziane [53] adapted those averaging operators to thin spherical domains, introduced some additional operators and proved their properties using the spherical coordinate system. Recently, Saito [46] used these averaging operators to study Boussinesq equations in thin spherical domains. We closely follow [46, 53] to describe our averaging operators and provide proofs for some of the properties mentioned below.

Let Mε:C(Qε,R)C(S2,R) be a map that projects functions defined on Qε to functions defined on S2 and is defined by

Mεψ(x):=1ε11+εrψ(rx)dr,xS2. 28

Remark 1

We will use the Cartesian and spherical coordinates interchangeably in this paper. For example, if xS2 then we will identify it by x=(λ,φ) where λ[0,π] and φ[0,2π).

Lemma 1

The map Mε as defined in (28) is continuous (and linear) w.r.t norms L2(Qε) and L2(S2). Moreover,

MεψL2(S2)21εψL2(Qε)2,ψL2(Qε). 29

Proof

Take ψC(Qε) then by the definition of Mε we have

Mεψ(x)=1ε11+εrψ(rx)dr.

Thus, using the Cauchy–Schwarz inequality we have

MεψL2(S2)2=S2|Mεψ(x)|2dσ(x)=S21ε11+εrψ(rx)dr2dσ(x)1ε2S211+εr2|ψ(rx)|2dr11+εdrdσ(x)=1ε2·ε11+εS2r2|ψ(rx)|2drdσ(x)=1εψL2(Qε)2,

where the last equality follows from the fact that

Qεdy=11+εS2r2drdσ(x)

is the volume integral over the spherical shell Qε in spherical coordinates, with

dσ(x)=sinλdλdφ,

being the Lebesgue measure over a unit sphere. Therefore, we obtain

MεψL2(S2)21εψL2(Qε)2, 30

and hence the map is bounded and we can infer (29).

Corollary 1

The map Mε as defined in (28) has a unique extension, which without the abuse of notation will be denoted by the same symbol Mε:L2(Qε)L2(S2).

Proof

Since C(Qε) is dense in L2(Qε) and Mε:L2(Qε)L2(S2) is a bounded map thus by the Riesz representation theorem there exists a unique extension.

Lemma 2

The following map

Rε:L2(S2)ψ1|·|ψ·|·|L2(Qε) 31

is bounded and

RεL(L2(S2),L2(Qε))2=ε.

Proof

It is sufficient to consider ψC(S2). For ψC(S2), we have

RεψL2(Qε)2=Qε|v(y)|2dy=11+εr2S2|v(rx)|2dσ(x)dr,

where v(y)=1|y|ψy|y|. But, for xS2

v(rx)=1|rx|ψrx|rx|=1rψ(x).

So

RεψL2(Qε)2=11+εr21r2S2|ψ(x)|2dσ(x)dr=εψL2(S2)2,

thus, showing that the map Rε is bounded w.r.t. L2(S2) and L2(Qε) norms.

Lemma 3

Let ψW1,p(S2) for p2. Then for ε(0,1) there exists a constant C>0 independent of ε such that

RεψLp(Qε)pCεψW1,p(S2)p.

Proof

By the definition of the map Rε (see (31)), and identities (172), (176) for the scalar function ψW1,p(S2), we have for Qεy=rx, r(1,1+ε) and xS2,

(Rε[ψ](y))=rRε[ψ](y)er^+1rλRε[ψ](y)eλ^+1rsinλφRε[ψ](y)eφ^=rψ(x)rer^+1rλψ(x)reλ^+1rsinλφψ(x)reφ^=-1r2ψ(x)er^+1r2λψ(x)eλ^+1r2sinλφψ(x)eφ^.

Hence,

RεψLp(Qε)p=Qε|(Rε[ψ](y))|pdy=11+εS21r2p|ψ(x)|p+|ψ(x)|pr2dσ(x)dr=-r3-2p2p-3|11+εψLp(S2)p+ψLp(S2)pC(p)εψW1,p(S2)p.

Lemma 4

Let ψH2(S2). Then for ε(0,1)

ΔRεψL2(Qε)2εΔψL2(S2)2. 32

Proof

Let ψH2(S2), then

[Rεψ](y)=1|y|ψy|y|,yQε.

Therefore, for every Qεy=rx, r(1,1+ε) and xS2, we have (see (171) and (177) for the definition of Laplace–Beltrami operator)

Δ([Rεψ](y))=2r2Rεψ(y)+2rrRεψ(y)+1r2ΔRεψ(y)=2r2ψ(x)r+2rrψ(x)r+1r2Δψ(x)r=2r3ψ(x)-2r3ψ(x)+1r3Δψ(x)=1r3Δψ(x).

Hence

ΔRεψL2(Qε)2=Qε|Δ[Rεψ](y)|2dy=11+εS21r6|Δψ(x)|2r2dσ(x)dr=-13r3|11+εΔψL2(S2)2=(1+ε)3-13(1+ε)3ΔψL2(S2)2=ε3+3ε2+3ε(1+ε)3ΔψL2(S2)2.

Since ε(0,1), the inequality (32) holds.

Remark 2

It is easy to check that the dual operator Mε:L2(S2)L2(Qε) is given by

Mεψ(y)=1ε(Rεψ)(y),yQε. 33

Next we define another map

M^ε=RεMε:L2(Qε)L2(Qε). 34

Courtesy of Corollary 1 and Lemma 2, M^ε is well-defined and bounded. Using definitions of maps Rε and Mε , we have

M^ε:ψy1|y|Mεψy|y|.

Lemma 5

Let ψL2(Qε), then we have the following scaling property

M^εψL2(Qε)2=εMεψL2(S2)2,ψL2(Qε). 35

Proof

Let ψL2(Qε). Then by the defintion of the map M^ε, we have

M^εψL2(Qε)2=Qε|M^εψ(y)|2dy=11+εS21r2|Mεψ(x)|2r2dσ(x)dr=εMεψL2(S2)2.

The normal component of a function ψ defined on Qε when projected to S2 is given by the map N^ε which is defined by

N^ε=Id-M^ε, 36

i.e.

N^ε:L2(Qε)ψψ-M^εψL2(Qε).

The following result establishes an important property of the map N^ε.

Lemma 6

Let ψL2(Qε), then

11+εrN^εψdr=0,a.e.onS2. 37

Proof

Let us choose and fix ψL2(Qε). Then by the definitions of the operators involved we have the following equality in L2(S2):

11+εrN^εψdr=εMεN^εψ.

Therefore ,we deduce that in order to prove equality (37), it is sufficient to show that

MεN^ε=0.

Hence, by taking into account definitions (36) of N^ε and (34) of M^ε , we infer that it is sufficient to prove that

Mε=MεRεMε.

Let us choose ψC(Qε) and put ϕ=Mεψ, i.e.

ϕ(x)=1ε11+εrψ(rx)dr.

Note that

Rεϕ(ρx)=1ρϕ(x),ρ(1,1+ε),xS2.

Thus, we infer that

MεRεϕ(x)=1ε11+ερRεϕ(ρx)dρ=1ε11+ερ1ρϕ(x)dr=ϕ(x)=Mεψ(x),xS2.

Thus, we proved MεRεMεψ=Mεψ for every ψC(Qε). Since C(Qε) is dense in L2(Qε) and the maps Mε and MεRεMε are bounded in L2(Qε), we conclude that we have proved (37).

Lemma 7

For all ψ,ξL2(Qε), we have

M^εψ,N^εξL2(Qε)=0. 38

Proof

Let ψ,ξL2(Qε), then

M^εψ,N^εξL2(Qε)=QεM^εψ(y)·N^εξ(y)dy=S211+εM^εψ(rx)·N^εξ(rx)r2drdσ(x).

By the definition (34) of the map M^ε and by Lemma 6, we infer that

M^εψ,N^εξL2(Qε)=S2Mεψ(x)11+εrN^εξ(rx)dr=0a.e.onS2byLemma6dσ(x)=0.

Next we define projection operators for R3-valued vector fields using the above maps (for scalar functions), as follows :

M~ε:L2(Qε)u=ur,uλ,uφ0,M^εuλ,M^εuφL2(Qε), 39
N~ε=Id-M~εL(L2(Qε)). 40

Lemma 8

Let uL2(Qε). Then

M~εu·n=0onQε.

Moreover, if u satisfies the boundary condition u·n=0, then

N~εu·n=0onQε.

Proof

The normal vector n to Qε is given by n=1,0,0. Thus by the definition of M~ε we have

M~εu·n=0onQε.

Now for the second part, from the definition of N~ε we have

N~εu·n=u·n=0fromb.c.-M~εu·n=0fromfirstpart=0.

We also have the following generalisation of Lemma 6.

Lemma 9

Let uL2(Qε), then

11+εrN~εudr=0,a.e.onS2. 41

The following Lemma makes sense only for vector fields.

Lemma 10

Let uHε, then

divM~εu=0anddivN~εu=0inQε.

Proof

Let uUε:=vC(Qε):divv=0inQεandv·n=0onQε, then using the definition of divergence for a vector field v=vr,vλ,vφ in spherical coordinates (see (174)), we get for Qεy=rx, xS2, r(1,1+ε),

div(M~εu)(y)=0+1rsinλλ(M^εuλ)(y)sinλ+1rsinλφ(M^εuφ)(y)=1rsinλλ1|y|(Mεuλ)y|y|sinλ+φ1|y|(Mεuφ)y|y|=1rsinλλsinλrε11+ερuλ(ρ,λ,φ)dρ+φ1rε11+ερuφ(ρ,λ,φ)dρ=:1r2sinλI+II. 42

Now considering each of the terms individually, we have

II=1εφ11+ερuφdρ=1ε11+ερuφφdρ. 43
I=1ελsinλ11+ερuλdρ=1ελ11+ερuλsinλdρ=1ε11+ερλuλsinλdρ. 44

Using (43) and (44) in the equality (42), we obtain

divM~εu=1r2sinλ1ε11+ερλuλsinλ+uφφdρ. 45

Since uUε, divu=0 in Qε, which implies

1ρ2ρρ2uρ+1ρsinλλuλsinλ+1ρsinλuφφ=0.

Using this in (45), we get

divM~εu=-1r2sinλ1ε11+ερsinλρρρ2uρdρ=-1εr211+ερρ2uρds=-1εr2ρ2uρ|11+ε=-1εr2(1+ε)2uρ(1+ε,·,·)-uρ(1,·,·)=0(sinceu·n=0).

Thus, we have proved that divM~εu=0, for every uUε. Since, Uε is dense in Hε, it holds true for every uHε too. The second part follows from the definition of N~ε and Hε.

From Lemmas 8 and 10, we infer the following corollary:

Corollary 2

If uHε then M~εu and N~εu belong to Hε.

Using the definition of maps M~ε and N~ε and Lemma 7, we conclude:

Proposition 1

For all u,vL2(Qε), we have

M~εu,N~εvL2(Qε)=0. 46

Moreover,

uL2(Qε)2=M~εuL2(Qε)2+N~εuL2(Qε)2,uHε. 47

Finally we define a projection operator that projects R3-valued vector fields defined on Qε to the “tangent” vector fields on sphere S2.

graphic file with name 245_2020_9702_Equ48_HTML.gif 48

Lemma 11

Let uL2(Qε), then

graphic file with name 245_2020_9702_Equ346_HTML.gif

Proof

Let uL2(Qε), then

graphic file with name 245_2020_9702_Equ347_HTML.gif

Remark 3

Similar to the scalar case, one can prove that the dual operator Inline graphic is given by

graphic file with name 245_2020_9702_Equ49_HTML.gif 49

Indeed, for uL2(Qε),vL2(S2)

graphic file with name 245_2020_9702_Equ348_HTML.gif

Using the identities (168)–(170), we can show that for a divergence free smooth vector field u

-Δu,uL2(Qε)=curlu,curluL2(Qε)=curluL2(Qε)2. 50

We define a weighted L2-product on Hε by

u,vr=Qεr2u·vdy,u,vHε, 51

and the corresponding norm will be denoted by ·r which is equivalent to ·L2(Qε), uniformly for ε(0,12) :

uL2(Qε)2ur294uL2(Qε)2,uL2(Qε). 52

We end this section by recalling a lemma and some Poincaré type inequalities from [53].

Lemma 12

[53, Lemma 1.2] For u,vVε, we have

curlM~εu,curlN~εvr=0,u,vVε.

Moreover,

curlur2=curlM~εur2+curlN~εur2,uVε. 53

Corollary 3

Let ε(0,12) and uVε. Then

curlM~εuL2(Qε)94curluL2(Qε)2,curlN~εuL2(Qε)94curluL2(Qε)2.

Proof

Let ε(0,12) and uVε. Then, by relation (50), equivalence of norms (52) and Eq. (53), we have

curluL2(Qε)249curlur249curlM~εur249curlM~εuL2(Qε)2.

The second inequality can be proved similarly.

The following two lemmas are taken from [53]. For the sake of completeness and convenience of the reader we have provided the proof in Appendix C.

Lemma 13

(Poincaré inequality in thin spherical shells) [53, Lemma 2.1] For 0<ε12, we have

N~εuL2(Qε)2εcurlN~εuL2(Qε),uVε. 54

Lemma 14

(Ladyzhenskaya’s inequality) [53, Lemma 2.3] There exists a constant c1, independent of ε, such that

N~εuL6(Qε)c1N~εuVε,uVε. 55

Corollary 4

For ε(0,12), there exists a constant c2>0 such that

N~εuL3(Qε)2c2εN~εuVε2,uVε. 56

Proof

Let uVε, then by the Hölder inequality, we have

N~εuL3(Qε)3=Qε|N~εu(y)|3dy=Qε|N~εu(y)|3/2|N~εu(y)|3/2dyQε|N~εu(y)|6dy1/4Qε|N~εu(y)|2dy3/4=N~εuL6(Qε)3/2N~εuL2(Qε)3/2.

Thus, by Lemmas 13 and 14, we get

N~εuL3(Qε)2c1N~εuVε2εN~εuVε=c2εN~εuVε2.

In the following lemma we enlist some properties of operators M^ε, N^ε, M~ε and N~ε.

Lemma 15

Let ε>0. Then

  • (i)
    for ψL2(Qε)
    M^εM^εψ=M^εψ, 57
    N^εN^εψ=N^εψ, 58
    M^εN^εψ=0,andN^εM^εψ=0, 59
  • (ii)
    and for uL2(Qε)
    M~εM~εu=M~εu, 60
    N~εN~εu=N~εu, 61
    M~εN~εu=0,andN~εM~εu=0. 62

Proof

Let ψC(Qε). Put

ϕ=M^εψ,

i.e. for yQε

ϕ(y)=1|y|Mεψy|y|=1|y|1ε11+εrψry|y|dr.

Next for zQε

|z|M^εϕ(z)=Mεϕz|z|=1ε11+ερϕρz|z|dρ=1ε211+ερ1ρz|z|11+εrψrρz|z|ρz|z|drdρ=1ε211+ε11+εrψrz|z|drdρ=1ε11+εrψrz|z|dr=Mεψz|z|=|z|M^εψ(z).

Hence, we proved (57) for every ψC(Qε). Since C(Qε) is dense in L2(Qε), it holds true for every ψL2(Qε).

Proof of first part of (59). Let ψC(Qε). Put ϕ=N^εψC(Qε). By Lemma 6

11+εrϕ(y)dr=0,yQε.

Therefore, for yQε,

Mεϕy|y|=1ε11+εrϕ(y)dr=0.

Therefore, we infer that

M^εϕ(y)=1|y|Mεϕy|y|=0,

for all yQε. Thus, we have established first part of (59) for all ψC(Qε). Using the density argument, we can prove it for all ψL2(Qε).

Now for (58), by the definition of N^ε and (59), we obtain

N^εN^εψ=N^εψ-M^εN^εψ=N^εψ.

Again using the definition of N^ε and Eq.(57), we have

N^εM^εψ=M^εψ-M^εM^εψ=M^εψ-M^εψ=0,

concluding the proof of second part of (59).

Proof of (60). Let uC(Qε,R3). Write u=ur,uλ,uφ. Put v=M~εu, i.e.

v=0,vλ,vφ,

where

vλ=M^εuλ,andvφ=M^εuφ.

Thus, by the definition of M~ε and identity (57)

M~εM~εu=M~εv=0,M^εvλ,M^εvφ=0,M^εM^εuλ,M^εM^εuφ=0,M^εuλ,M^εuφ=v=M~εu.

We can extend this to uL2(Qε) by the density argument. The remaining identities can be also established similarly as in the case of scalar functions.

Later in the proof of Theorem 3, in order to pass to the limit we will use an operator Inline graphic defined by

graphic file with name 245_2020_9702_Equ63_HTML.gif 63

where

L2(S2)u=0,uλ(x),uφ(x),xS2.

Using the definition of map Rε from Lemma 2, we can rewrite Inline graphic as

graphic file with name 245_2020_9702_Equ64_HTML.gif 64

Note that Inline graphic is a bounded linear map from L2(S2) to L2(Qε).

This operator Inline graphic is retract of Inline graphic, i.e. a map Inline graphic such that

graphic file with name 245_2020_9702_Equ65_HTML.gif 65

One can easily show that if uD(A) then Inline graphic. In particular, for uH, Inline graphic. Next we establish certain scaling properties for the map Inline graphic.

Lemma 16

Let ε>0, then

graphic file with name 245_2020_9702_Equ66_HTML.gif 66

Proof

Let ε>0 and consider L2(S2)u=0,uλ,uφ. Then, by the definition of the retract operator Inline graphic and L2(Qε)-norm we have

graphic file with name 245_2020_9702_Equ349_HTML.gif

Using the definition of the map Inline graphic and Lemmas 3, 4 we can deduce the following two lemmas (we provide the detailed proof of the latter in Appendix C):

Lemma 17

Let uW1,p(S2) for p2. Then for ε(0,1) there exists a constant C>0 independent of ε such that

graphic file with name 245_2020_9702_Equ67_HTML.gif 67

Lemma 18

Let uH2(S2)H and ε(0,1). Then

graphic file with name 245_2020_9702_Equ68_HTML.gif 68

where Δ is defined in (180).

Stochastic NSE on Thin Spherical Domains

This section deals with the proof of our main result, Theorem 3. First we introduce our two systems; stochastic NSE in thin spherical domain and stochastic NSE on the sphere, then we present the definition of martingale solutions for both systems. We also state the assumptions under which we prove our result. In Sect. 4.1, we obtain a priori estimates (formally) which we further use to establish some tightness criterion (see Sect. 4.2) which along with Jakubowski’s generalisation of Skorokhod Theorem gives us a converging (in ε) subsequence. At the end of this section we show that the limiting object of the previously obtained converging subsequence is a martingale solution of stochastic NSE on the sphere (see Sect. 4.3).

In thin spherical domain Qε, which was introduced in (3), we consider the following stochastic Navier–Stokes equations (SNSE)

du~ε-[νΔu~ε-(u~ε·)u~ε-p~ε]dt=f~εdt+G~εdW~ε(t)inQε×(0,T), 69
divu~ε=0inQε×(0,T), 70
u~ε·n=0,curlu~ε×n=0onQε×(0,T), 71
u~ε(0,·)=u~0εinQε. 72

Recall that, u~ε=(u~εr,u~ελ,u~εφ) is the fluid velocity field, p is the pressure, ν>0 is a (fixed) kinematic viscosity, u~0ε is a divergence free vector field on Qε and n is the unit normal outer vector to the boundary Qε. We assume that1NN. We consider a family of maps

G~ε:R+T2(RN;Hε)

such that

G~ε(t)k:=j=1Ng~εj(t)kj,k=(kj)j=1NRN, 73

for some g~εj:R+Hε, j=1,,N. The Hilbert–Schmidt norm of G~ε is given by

G~ε(s)T2(RN;Hε)2=j=1Ng~εj(s)L2(Qε)2. 74

Finally we assume that W~ε(t), t0 is an RN-valued Wiener process defined on the probability space Ω,F,F,P. We assume that βjj=1N are i.i.d real valued Brownian motions such that W(t)=(βj(t))j=1N, t0.

In this section, we shall establish convergence of the radial averages of the martingale solution of the 3D stochastic equations (69)–(72), as the thickness of the shell ε0, to a martingale solution u of the following stochastic Navier–Stokes equations on the sphere S2:

du-νΔu-(u·)u-pdt=fdt+GdWinS2×(0,T), 75
divu=0inS2×(0,T), 76
u(0,·)=u0inS2, 77

where u=(0,uλ,uφ) and Δ, are as defined in (176)–(180). Assumptions on initial data and external forcing will be specified later (see Assumptions 12). Here, G:R+T2(RN;H) and W(t), t0 is an RN-valued Wiener process such that

G(t)dW(t):=j=1Ngj(t)dβj(t), 78

where NN, βjj=1N are i.i.d real valued Brownian motions as before and gjj=1N are elements of H, with certain relation to g~εj, which is specified later in Assumption 2.

Remark 4

We are aware of other formulations of the Laplacian in (75) such as the one with an additional Ricci tensor term [47, 48]. However, as it was written in [47, p. 144], “Deriving appropriate equations of motion involves dynamical considerations which do not seem adapted to Riemannian space; in particular it is not evident how to formulate the principle of conservation of momentum.” Therefore, in this paper, we follow the approach presented in [53], that the Navier–Stokes equations on the sphere is the thin shell limit of the 3-dimensional Navier–Stokes equations defined on a thin spherical shell.

Now, we specify assumptions on the initial data u~0ε and external forcing f~ε, g~εj.

Assumption 1

Let Ω,F,F,P be the given filtered probability space. Let us assume that p2 and that u~0εHε, for ε(0,1], such that for some C1>0

u~0εL2(Qε)=C1ε1/2,ε(0,1]. 79

We also assume that f~εLp([0,T];Vε), for ε(0,1], such that for some C2>0,

0Tf~ε(s)VεpdsC2εp/2,ε(0,1]. 80

Let W~ε be an RN-valued Wiener process as before and assume that

G~εLp(0,T;T2(RN;Hε)),forε(0,1],

such that, using convention (73), for each j=1,,N,

0Tg~εj(t)L2(Qε)pdt=O(εp/2),ε(0,1]. 81

Projecting the stochastic NSE (on thin spherical shell) (69)–(72) onto Hε using the Leray–Helmholtz projection operator and using the definitions of operators from Sect. 2.1, we obtain the following abstract Itô equation in Hε, t[0,T]

du~ε(t)+νAεu~ε(t)+Bε(u~ε(t),u~ε(t))dt=f~ε(t)dt+G~ε(t)dW~ε(t),u~ε(0)=u~0ε. 82

Definition 1

Let ε(0,1]. A martingale solution to (82) is a system

Ω,F,F,P,W~ε,u~ε

where Ω,F,P is a probability space and F=Ftt0 is a filtration on it, such that

  • W~ε is a RN-valued Wiener process on Ω,F,F,P,

  • u~ε is Vε-valued progressively measurable process, Hε-valued weakly continuous F-adapted process such that2P-a.s.
    u~ε(·,ω)C([0,T],Hεw)L2(0,T;Vε),E12sup0sTu~ε(s)L2(Qε)2+ν0Tcurlu~ε(s)L2(Qε)2ds<
    and, for all t[0,T] and vVε, P-a.s.,
    (u~ε(t),v)L2(Qε)+ν0t(curlu~ε(s),curlv)L2(Qε)ds+0tBε(u~ε(s),u~ε(s)),vεds=(u~0ε,v)L2(Qε)+0tf~ε(s),vεds+0tG~ε(s)dW~ε(s),vL2(Qε). 83

In the following remark we show that a martingale solution u~ε of (82), as defined above, satisfies an equivalent equation in the weak form.

Remark 5

Let u~ε=u~ε(t), t0 be a martingale solution of (82). We will use the following notations

graphic file with name 245_2020_9702_Equ84_HTML.gif 84

and also from Lemma 15 we have

u~ε(t)=α~ε(t)+β~ε(t),t[0,T].

Then, for ϕD(A), we have

graphic file with name 245_2020_9702_Equ350_HTML.gif

and using Lemma 6, Proposition 1 and Lemma 12, we can rewrite the weak formulation identity (83) as follows.

graphic file with name 245_2020_9702_Equ85_HTML.gif 85

where ·,· denotes the duality between V and V.

Next, we present the definition of a martingale solution for stochastic NSE on S2.

Definition 2

A martingale solution to equation (75)–(77) is a system

Ω^,F^,F^,P^,W^,u^

where Ω^,F^,P^ is a probability space and F^=F^tt0 is a filtration on it, such that

  • W^ is an RN-valued Wiener process on Ω^,F^,F^,P^,

  • u^ is V-valued progressively measurable process, H-valued continuous F^-adapted process such that
    u^(·,ω)C([0,T],H)L2(0,T;V),E^sup0sTu^(s)L2(S2)2+ν0Tcurlu^(s)L2(S2)2ds<
    and
    u^(t),ϕL2(S2)+ν0tcurlu^(s),curlϕL2(S2)ds+0tu^(s)·u^(s),ϕL2(S2)ds=u0,ϕL2(S2)+0tf(s),ϕL2(S2)ds+0tG(s)dW^(s),ϕL2(S2), 86
    for all t[0,T] and ϕV.

Assumption 2

Let p2. Let Ω^,F^,F^,P^ be the given probability space, u0H such that

graphic file with name 245_2020_9702_Equ87_HTML.gif 87

Let fLp([0,T];V), such that for every s[0,T],

graphic file with name 245_2020_9702_Equ88_HTML.gif 88

And finally, we assume that GLp(0,T;T2(RN;H)), such that for each j=1,,N and s[0,T], Inline graphic converges weakly to gj(s) in L2(S2) as ε0 and

0Tgj(t)L2(S2)2dtM, 89

for some M>0.

Remark 6

(Existence of martingale solutions) In a companion paper [7] we will address an easier question about the existence of a martingale solution for (1)–(5) in a more general setting with multiplicative noise. The key idea of the proof is taken from [11], where authors prove existence of a martingale solution for stochastic NSE in unbounded 3D domains.

The existence of a pathwise unique strong solution (hence a martingale solution) for the stochastic NSE on a sphere S2 is already established by two of the authors and Goldys in [9]. Through this article we give another proof of the existence of a martingale solution for such a system.

We end this subsection by stating the main theorem of this article.

Theorem 3

Let the given data u~0ε, u0, f~ε, f, g~εj, gj, j{1,,N} satisfy Assumptions 1 and 2. Let Ω,F,F,P,W~ε,u~ε be a martingale solution of (69)–(72) as defined in Definition 1. Then, the averages in the radial direction of this martingale solution i.e. Inline graphic converge to a martingale solution, Ω^,F^,F^,P^,W^,u^, of (75)–(77) in L2(Ω^×[0,T]×S2).

Remark 7

According to Remark 6, for every ε[0,1] there exists a martingale solution of (69)–(72) as defined in Definition 1, i.e. we will obtain a tuple Ωε,Fε,Fε,Pε,W~ε,u~ε as a martingale solution. It was shown in [31] that is enough to consider only one probability space, namely,

Ωε,Fε,Pε=[0,1],B([0,1]),Lε(0,1],

where L denotes the Lebesgue measure on [0, 1]. Thus, it is justified to consider the probability space Ω,F,P independent of ε in Theorem 3.

Estimates

From this point onward we will assume that for every ε(0,1] there exists a martingale solution Ω,F,F,P,W~ε,u~ε of (82). Please note that we do not claim neither we use the uniqueness of this solution.

The main aim of this subsection is to obtain estimates for αε and β~ε uniform in ε using the estimates for the process u~ε.

The energy inequality (90) and the higher-order estimates (105)–(106), satisfied by the process u~ε, as obtained in Lemmas 19 and 22 is actually a consequence (essential by-product) of the existence proof. In principle, one obtains these estimates (uniform in the approximation parameter N) for the finite-dimensional process u~ε(N) (using Galerkin approximation) with the help of the Itô lemma. Then, using the lower semi-continuity of norms, convergence result (u~ε(N)u~ε in some sense), one can establish the estimates for the limiting process. Such a methodology was employed in a proof of Theorem 4.8 in the recent paper [13] by the first named author, Motyl and Ondreját.

In Lemmas 19 and 22 we present a formal proof where we assume that one can apply (ignoring the existence of Lebesgue and stochastic integrals) the Itô lemma to the infinite dimensional process u~ε. The idea is to showcase (though standard) the techniques involved in establishing such estimates.

Lemma 19

Let u~0εHε, f~εL2(Ω×[0,T];Vε) and G~εL2(Ω×[0,T];T2(RN;Hε)). Then, the martingale solution u~ε of (82) satisfies the following energy inequality

E12sup0sTu~ε(s)L2(Qε)2+ν0Tcurlu~ε(s)L2(Qε)2dsu~0εL2(Qε)2+1ν0Tf~ε(s)Vε2ds+K0TG~ε(s)T2(RN;Hε)2ds, 90

where K is some positive constant independent of ε.

Proof

Using the Itô formula for the function ξL2(Qε)2 with the process u~ε, for a fixed t[0,T] we have

u~ε(t)L2(Qε)2+2ν0tcurlu~ε(s)L2(Qε)2dsu~0εL2(Qε)2+20tf~ε(s),u~ε(s)εds+20tG~ε(s)dW~ε(s),u~ε(s)L2(Qε)+0tG~ε(s)T2(RN;Hε)2ds. 91

Using the Cauchy–Schwarz inequality and the Young inequality, we get the following estimate

f~ε,u~εεu~εVεf~εVεν2u~εVε2+12νf~εVε2,

which we use in (91), to obtain

u~ε(t)L2(Qε)2+ν0tcurlu~ε(s)L2(Qε)2dsu~0εL2(Qε)2+1ν0tf~ε(s)Vε2ds+20tG~ε(s)dW~ε,u~ε(s)L2(Qε)+0tG~ε(s)T2(RN;Hε)2ds. 92

Using the Burkholder–Davis–Gundy inequality (see [32, Prop. 2.12]), we have

Esup0tT0t(G~ε(s)dW~ε(s),u~ε(s))L2(Qε)CE0Tu~ε(s)L2(Qε)2G~ε(s)T2(RN;Hε)2ds1/2CEsup0tTu~ε(t)L2(Qε)21/20TG~ε(s)T2(RN;Hε)2ds1/214Esup0tTu~ε(t)L2(Qε)2+C0TG~ε(s)T2(RN;Hε)2ds. 93

Taking the supremum of (92) over the interval [0, T], then taking expectation and using inequality (93) we infer the energy inequality (90).

Let us recall the following notations, which we introduced earlier, for t[0,T]

graphic file with name 245_2020_9702_Equ94_HTML.gif 94

Lemma 20

Let u~ε be a martingale solution of (82) and Assumption 1 hold, in particular, for p=2. Then

E12supt[0,T]αε(t)L2(S2)2+ν0Tcurlαε(s)L2(S2)2dsC12+C2ν+C3, 95

where C1,C2 are positive constants from (79) and (80) and C3>0 (determined within the proof) is another constant independent of ε.

Proof

Let u~ε be a martingale solution of (82), then it satisfies the energy inequality (90). From Eq. (47), we have

α~ε(t)L2(Qε)2u~ε(t)L2(Qε)2,t[0,T]. 96

Moreover, by Corollary 3

49curlα~ε(t)L2(Qε)2curlu~ε(t)L2(Qε)2,t[0,T]. 97

Therefore, using (96) and (97) in the energy inequality (90), we get

E12supt[0,T]α~ε(t)L2(Qε)2+4ν90Tcurlα~ε(s)L2(Qε)2dsu~0εL2(Qε)2+1ν0Tfε(s)Vε2ds+K0TG~ε(s)T22ds,

and hence from the scaling property, Lemma 11, we have

E12εsupt[0,T]αε(t)L2(S2)2+4ν9ε0Tcurlαε(s)L2(S2)2dsu~0εL2(Qε)2+1ν0Tfε(s)Vε2ds+K0TG~ε(s)T22ds. 98

By the assumptions on g~εj (81), there exists a positive constant c such that for every j{1,,N}

0Tg~εj(t)L2(Qε)2dtcε. 99

Therefore, using Assumption 1 and (99) in (98), cancelling ε on both sides and defining C3=NKc, we infer inequality (95).

From the results of Lemma 20, we deduce that

αεε>0isboundedinL2(Ω;L(0,T;H)L2(0,T;V)). 100

Since V can be embedded into L6(S2), by using interpolation between L(0,T;H) and L2(0,T;L6(S2)) we obtain

E0Tαε(s)L3(S2)2dsC. 101

Lemma 21

Let u~ε be a martingale solution of (82) and Assumption 1 hold, in particular, for p=2. Then

E12supt[0,T]β~ε(t)L2(Qε)2+ν0Tcurlβ~ε(s)L2(Qε)2dsC12ε+C2εν+C3ε. 102

Proof

Let u~ε be a martingale solution of (82), then it satisfies the energy inequality (90). From (47), we have

β~ε(t)L2(Qε)2u~ε(t)L2(Qε)2,t[0,T]. 103

Thus, by Corollary 3

49curlβ~ε(t)L2(Qε)2curlu~ε(t)L2(Qε)2,t[0,T]. 104

Therefore, using Assumption 1, (99), inequalities (103)–(104), in the energy inequality (90), we infer (102).

In the following lemma we obtain some higher order estimates (on a formal level) for the martingale solution u~ε, which will be used to obtain the higher order estimates for the processes αε and β~ε.

Lemma 22

Let Assumption 1 hold true and u~ε be a martingale solution of (82). Then, for p>2 we have following estimates

Esup0sTu~ε(s)L2(Qε)pC2p,u~0ε,f~ε,G~εexpKpT 105

and

E0Tu~ε(t)L2(Qε)p-2u~ε(t)Vε2dtC2p,u~0ε,f~ε,G~ε1+KpTexpKpT, 106

where

C2p,u~0ε,f~ε,G~ε:=u~0εL2(Qε)p+ν-p/2f~εLp(0,T;Vε)p+14p2(p-1)+K12pG~εLp(0,T;T2)p,Kp:=K12p+p(p-2)2,

and K1 is a constant from the Burkholder–Davis–Gundy inequality.

Proof

Let F(x)=xL2(Qε)p then

Fx=F=pxL2(Qε)p-2x,

and

2Fx2p(p-1)xL2(Qε)p-2. 107

Applying the Itô lemma with F(x) and process u~ε for t[0,T], we have

u~ε(t)L2(Qε)p=u~ε(0)L2(Qε)p+p0tu~ε(s)L2(Qε)p-2-νAεu~ε(s)-Bε(u~ε(s),u~ε(s))+f~ε(s),u~ε(s)εds+p0tu~ε(s)L2(Qε)p-2u~ε(s),G~ε(s)dW~ε(s)L2(Qε)+120tTr2Fx2(G~ε(s),G~ε(s))ds.

Using the fact that Bε(u~ε,u~ε),u~εε=0 and Aεu~ε,u~εε=u~εVε2 we arrive at

u~ε(t)L2(Qε)p=u~ε(0)L2(Qε)p-pν0tu~ε(s)L2(Qε)p-2u~ε(s)Vε2+p0tu~ε(s)L2(Qε)p-2f~ε(s),u~ε(s)εds+p0tu~ε(s)Hεp-2u~ε(s),G~ε(s)dW~ε(s)L2(Qε)+120tTr2Fx2(G~ε(s),G~ε(s))ds.

Using (107) and the Cauchy–Schwarz inequality, we get

u~ε(t)L2(Qε)p+pν0tu~ε(s)L2(Qε)p-2u~ε(s)Vε2dsu~ε(0)L2(Qε)p+p0tu~ε(s)L2(Qε)p-2f~ε(s)Vεu~ε(s)Vεds+p0tu~ε(s)L2(Qε)p-2u~ε(s),G~ε(s)dW~ε(s)L2(Qε)+p(p-1)20tu~ε(s)L2(Qε)p-2G~ε(s)T2(RN;Hε)2ds,

where we recall

G~ε(s)T2(RN;Hε)2=j=1Ng~εj(s)L2(Qε)2.

Using the generalised Young inequality abcaq/q+br/r+cs/s (where 1/q+1/r+1/s=1) with a=νu~εL2(Qε)p/2-1u~εVε, b=u~εL2(Qε)p/2-1, c=1νfεVε and exponents q=2,r=p,s=2p/(p-2) we get

νu~εL2(Qε)p-2fεVεu~εVεν2u~εL2(Qε)p-2u~εVε2+1pνp/2f~εVεp+p-22pu~εL2(Qε)p. 108

Again using the Young inequality with exponents p/(p-2), p/2 we get

u~εL2(Qε)p-2G~εT2(RN;Hε)2p-2pu~εL2(Qε)p+p2G~εT2(RN;Hε)p. 109

Using (108) and (109) we obtain

u~ε(t)L2(Qε)p+pν20tu~ε(s)L2(Qε)p-2u~ε(s)Vε2dsu~ε(0)L2(Qε)p+p(p-2)20tu~ε(s)L2(Qε)pds+ν-p/20tf~ε(s)Vεpds+14p2(p-1)0tG~ε(s)T2(RN;Hε)pds+p0tu~ε(s)L2(Qε)p-2u~ε(s),G~ε(s)dW~ε(s)L2(Qε). 110

Since u~ε is a martingale solution of (82) it satisfies the energy inequality (90), hence the real-valued random variable

με(t)=0tu~ε(s)L2(Qε)p-2u~ε(s),G~ε(s)dW~ε(s)L2(Qε)

is a Ft-martingale. Taking expectation both sides of (110) we obtain

Eu~ε(t)L2(Qε)p+pν2E0tu~ε(s)L2(Qε)p-2u~ε(s)Vε2dsu~ε(0)L2(Qε)p+p(p-2)2E0tu~ε(s)L2(Qε)pds+ν-p/20tf~ε(s)Vεpds+14p2(p-1)0tG~ε(s)T2(RN;Hε)pds. 111

Therefore, by the Gronwall lemma we obtain

Eu~ε(t)L2(Qε)pCu~0ε,f~ε,G~εexpp(p-2)t2,

where

Cu~0ε,f~ε,G~ε:=u~0εL2(Qε)p+ν-p/2fLp(0,T;Vε)p+14p2(p-1)G~εLp(0,T;T2(RN;Hε))p.

By Burkholder–Davis–Gundy inequality, we have

Esup0st0su~ε(σ)L2(Qε)p-2u~ε(σ),G~ε(σ)dW~ε(σ)L2(Qε)K1E0tu~ε(s)L2(Qε)2p-4|u~ε(s)|L2(Qε)2G~ε(s)T2(RN;Hε)2ds1/2K1Esup0stu~ε(s)L2(Qε)p/20tu~ε(s)L2(Qε)p-2G~ε(s)T2(RN;Hε)2ds1/212Esup0stu~ε(s)L2(Qε)p+K122E0tu~ε(s)L2(Qε)p-2G~ε(s)T2(RN;Hε)2ds12Esup0stu~ε(s)L2(Qε)p+K122(p-2)pE0tu~ε(s)L2(Qε)pds+K12p0tG~ε(s)T2(RN;Hε)pds 112

where in the last step we have used the Young inequality with exponents p/(p-2) and p/2.

Taking supremum over 0st in (110) and using (112) we get

12Esup0stu~ε(s)L2(Qε)p+νpE0tu~ε(s)L2(Qε)p-2u~ε(s)Vε2dsu~ε(0)L2(Qε)p+K12p+p(p-2)20tEsup0sσu~ε(s)L2(Qε)pdσ+ν-p/20tf~ε(s)Vεpds+14p2(p-1)+K12p0tG~ε(s)T2(RN;Hε)pds. 113

Thus using the Gronwall lemma, we obtain

Esup0stu~ε(s)L2(Qε)pC2p,u~0ε,f~ε,G~εexpKpt,

where C2p,u~0ε,f~ε,G~ε and Kp are the constants as defined in the statement of lemma. We deduce (106) from (113) and (105).

In the following lemma we will use the estimates from previous lemma to obtain higher order estimates for αε and β~ε.

Lemma 23

Let p>2. Let u~ε be a martingale solution of (82) and Assumption 1 hold with the chosen p. Then, the processes αε and β~ε (as defined in (94)) satisfy the following estimates

Esupt[0,T]αε(t)L2(S2)pK(ν,p)expKpT, 114

and

Esupt[0,T]β~ε(t)L2(Qε)pεp/2K(ν,p)expKpT, 115

where K(ν,p) is a positive constant independent of ε and Kp is defined in Lemma 22.

Proof

The lemma can be proved following the steps of Lemmas 20 and 21 with the use of Proposition 1, scaling property from Lemma 11, the Cauchy–Schwarz inequality, Assumptions 1, 2 and the estimates obtained in Lemma 22.

Tightness

In this subsection we will prove that the family of laws induced by the processes αε is tight on an appropriately chosen topological space ZT. In order to do so we will consider the following functional spaces for fixed T>0:

C([0,T];D(A-1)):= the space of continuous functions u:[0,T]D(A-1) with the topology T1 induced by the norm uC([0,T];D(A-1)):=supt[0,T]u(t)D(A-1),

Lw2(0,T;V):= the space L2(0,T;V) with the weak topology T2,

L2(0,T;H):= the space of measurable functions u:[0,T]H such that

uL2(0,T;H)=0Tu(t)L2(S2)2dt12<,

with the topology T3 induced by the norm uL2(0,T;H).

Let Hw denote the Hilbert space H endowed with the weak topology.

C([0,T];Hw):= the space of weakly continuous functions u:[0,T]H endowed with the weakest topology T4 such that for all hH the mappings

C([0,T];Hw)uu(·),hL2(S2)C([0,T];R)

are continuous. In particular unu in C([0,T];Hw) iff for all hH:

limnsupt[0,T]un(t)-u(t),hL2(S2)=0.

Let

ZT=C([0,T];D(A-1))Lw2(0,T;V)L2(0,T;H)C([0,T];Hw), 116

and let T be the supremum3 of the corresponding topologies.

Lemma 24

The set of measures L(αε),ε(0,1] is tight on ZT,T.

Proof

Let u~ε, for some fixed ε>0, be a martingale solution of problem (82). Let us choose and fix ϕD(A). Then, recalling the definition (40) of the operator N~ε, by Lemma 9, we infer that for t[0,T] we have

graphic file with name 245_2020_9702_Equ117_HTML.gif 117

Similarly we have, also for t[0,T],

graphic file with name 245_2020_9702_Equ118_HTML.gif 118

Thus, by Proposition 1, identity (83), equalities (117), (118), and the notations from (94), we infer that martingale solution u~ε satisfies the following equality, for t[0,T], P-a.s.

graphic file with name 245_2020_9702_Equ119_HTML.gif 119

The proof of lemma turns out to be a direct application of Corollary 6. Indeed, by Lemma 20, assumptions (a) and (b) of Corollary 6 are satisfied and therefore, it is sufficient to show that the sequence αεε>0 satisfies the Aldous condition [A], see Definition 6, in space D(A-1).

Let θ(0,T) and τεε>0 be a sequence of stopping times such that 0τετε+θT. We start by estimating each term in the R.H.S. of (119). We will use the Hölder inequality, the scaling property from Lemma 11, the Poincaré type inequality (54), the Ladyzhenskaya inequality (55), inequality (56), the a priori estimates from Lemmas 20, 21, results from Lemmas 17 and 18.

In what follows, we will prove that each of the eight process from equality (119) satisfies the Aldous condition [A]. In order to help the reader, we will divide the following part of the proof into eight parts.

  • For the first term, we obtain
    graphic file with name 245_2020_9702_Equ120_HTML.gif 120
  • Similarly for the second term we have
    graphic file with name 245_2020_9702_Equ121_HTML.gif 121
  • Now we consider the first non-linear term.
    graphic file with name 245_2020_9702_Equ122_HTML.gif 122
  • Similarly for the second non-linear term, we have
    graphic file with name 245_2020_9702_Equ123_HTML.gif 123
  • Now as in the previous case, for the next mixed non-linear term, we obtain
    graphic file with name 245_2020_9702_Equ124_HTML.gif 124
  • Finally, for the last non-linear term, we get
    graphic file with name 245_2020_9702_Equ125_HTML.gif 125
  • Now for the term corresponding to the external forcing f~ε, we have using the radial invariance of Mεf~ε and assumption (80)
    graphic file with name 245_2020_9702_Equ126_HTML.gif 126
  • At the very end we are left to deal with the last term corresponding to the stochastic forcing. Using the radial invariance of Mεg~εj, Itô isometry, scaling (see Lemma 11) and assumption (81), we get
    graphic file with name 245_2020_9702_Equ127_HTML.gif 127
    Eτετε+θj=1Ng~εj(s)L2(Qε)2dsϕL2(S2)2εNcϕD(A)2θ:=c8ε·θϕD(A)2. 128

After having proved what we had promised, we are ready to conclude the proof of Lemma 24. Since for every t>0

graphic file with name 245_2020_9702_Equ351_HTML.gif

one has for ϕD(A),

graphic file with name 245_2020_9702_Equ129_HTML.gif 129

Let us fix κ>0 and γ>0. By equality (119), the sigma additivity property of probability measure and (129), we have

Pαε(τε+θ)-αε(τε)D(A-1)κ1εi=18PsupϕD(A)=1Jiε(τε+θ)-Jiε(τε)κ.

Using the Chebyshev’s inequality, we get

Pαε(τε+θ)-αε(τε)D(A-1)κ1κεi=17EsupϕD(A)=1Jiε(τε+θ)-Jiε(τε)+1κ2εEsupϕD(A)=1J8ε(τε+θ)-J8ε(τε)2. 130

Thus, using estimates (120)–(128) in (130), we get

Pαε(τε+θ)-αε(τε)D(A-1)κ1κεεθ1/2c1θ1/2+c2ε+c3θ1/4+c4ε2/3+c5ε2/3+c6ε2/3+c7+1κ2εc8εθ. 131

Let δi=κ8ciγ2, for i=1,,7 and δ8=κ28c8γ. Choose δ=maxi{1,,8}δi. Hence,

supε>0sup0θδPαε(τε+θ)-αε(τε)D(A-1)κγ.

Since αε satisfies the Aldous condition [A] in D(A-1), we conclude the proof of Lemma 24 by invoking Corollary 6.

Proof of Theorem 3

For every ε>0, let us define the following intersection of spaces

YTε=Lw2(0,T;Vε)C([0,T];Hεw). 132

Now, choose a countable subsequence εkkN converging to 0. For this subsequence define a product space YT given by

YT=ΠkNYTεk,

and η:ΩYT by

η(ω)=β~ε1(ω),β~ε2(ω),,YT.

Now with this YT-valued function we define a constant YT-sequence

ηkη,kN.

Then by Lemma 24 and the definition of sequence ηk, the set of measures Lαεk,ηk,kN is tight on ZT×YT.

Thus, by the Jakubowski–Skorohod theorem4 there exists a subsequence knnN, a probability space (Ω^,F^,P^) and, on this probability space, ZT×YT×C([0,T];RN)-valued random variables (u^,η^,W^), α^εkn,η^kn,W^εkn,nN such that

α^εkn,η^kn,W^εknhasthesamelawasαεkn,ηkn,W~onBZT×YT×C([0,T];RN) 133

and

α^εkn,η^kn,W^εknu^,η^,W^inZT×YT×C([0,T];RN),P^-a.s. 134

In particular, using marginal laws, and definition of the process ηk, we have

Lα^εkn,β^εkn=Lαεkn,β~εknonBZT×YTεkn 135

where β^εkn is the knth component of YT-valued random variable η^kn. We are not interested in the limiting process η^ and hence will not discuss it further.

Using the equivalence of law of W^εkn and W~ on C([0,T];RN) for nN one can show that W^ and W^εkn are RN-valued Wiener processes (see [8, Lemma 5.2 and Proof] for details).

α^εknu^ in ZT, P^-a.s. precisely means that

α^εknu^inC([0,T];D(A-1)),α^εknu^inL2(0,T;V),α^εknu^inL2(0,T;H),α^εknu^inC([0,T];Hw),

and

W^εknW^inC([0,T];RN).

Let us denote the subsequence (α^εkn,β^εkn,W^εkn) again by (α^ε,β^ε,W^ε)ε(0,1].

Note that since BZT×YT×C([0,T];RN)B(ZT)×B(YT)×BC([0,T];RN), the functions u^, η^ are ZT, YT Borel random variables respectively.

Using the retract operator Inline graphic as defined in (63)–(65), we define new processes α^ε_ corresponding to old processes α~ε on the new probability space as follows

graphic file with name 245_2020_9702_Equ136_HTML.gif 136

Moreover, by Lemma 16 we have the following scaling property for these new processes, i.e.

α^ε_L2(Qε)=εα^εL2(S2). 137

The following auxiliary result which is needed in the proof of Theorem 3, cannot be deduced directly from the Kuratowski Theorem (see Theorem 7).

Lemma 25

Let T>0 and ZT be as defined in (116). Then the following sets C([0,T];H)ZT, L2(0,T;V)ZT are Borel subsets of ZT.

Proof

See Appendix E.2.

By Lemma 25, C([0,T];H) is a Borel subset of ZT. Since αεC([0,T];H), P^-a.s. and α^ε, αε have the same laws on ZT, thus

L(α^ε)C([0,T];H)=1,ε>0, 138

and from estimates (95) and (114), for p[2,)

supε>0E^sup0sTα^ε(s)L2(S2)pK1(p). 139

Since L2(0,T;V)ZT is a Borel subset of ZT (Lemma 25), αε and α^ε have same laws on ZT; from (95), we have

supε>0E^0Tcurlα^ε(s)L2(S2)2dsK2. 140

Since the laws of ηkn and η^kn are equal on YT, we infer that the corresponding marginal laws are also equal. In other words, the laws on BYTεkn of L(β^εkn) and L(β~εkn) are equal for every kn.

Therefore, from the estimates (102) and (115) we infer for p[2,)

E^sup0sTβ^ε(s)L2(Qε)pK3(p)εp/2,ε(0,1] 141

and

E^0Tcurlβ^ε(s)L2(Qε)2dsK4ε,ε(0,1]. 142

By inequality (140) we infer that the sequence (α^ε)ε>0 contains a subsequence, still denoted by (α^ε)ε>0 convergent weakly (along the sequence εkn) in the space L2([0,T]×Ω^;V). Since α^εu^ in ZT P^-a.s., we conclude that u^L2([0,T]×Ω^;V), i.e.

E^0Tcurlu^(s)L2(S2)2ds<. 143

Similarly by inequality (139), for every p[2,) we can choose a subsequence of (α^ε)ε>0 convergent weak star (along the sequence εkn) in the space Lp(Ω^;L(0,T;H)) and, using (134), we infer that

E^sup0sTu^(s)L2(S2)p<. 144

Using the convergence from (134) and estimates (139)–(144) we will prove certain term-by-term convergences which will be used later to prove Theorem 3. In order to simplify the notation, in the result below we write limε0 but we mean limkn.

Before stating the next lemma, we introduce a new functional space U as the space of compactly supported, smooth divergence free vector fields on S2:

U:={v:=(0,vλ,vφ)Cc(S2;R3):divv=0inS2}. 145

Lemma 26

For all t[0,T], and ϕU, we have (along the sequence εkn)

  1. limε0E^0Tα^ε(t)-u^(t),ϕL2(S2)dt=0,

  2. limε0α^ε(0)-u0,ϕL2(S2)=0,

  3. limε0E^0T0tν1+εcurlα^ε(s)-curlu^(s),curlϕL2(S2)dsdt=0,

  4. limε0E^0T0t11+εα^ε(s)·α^ε(s)-u^(s)·u^(s),ϕL2(S2)dsdt=0,

  5. Inline graphic

  6. Inline graphic

Proof

Let us fix ϕU.

(a) We know that α^εu^ in ZT. In particular,

α^εu^inC([0,T];Hw)P^-a.s.

Hence, for t[0,T]

limε0α^ε(t),ϕL2(S2)=u^(t),ϕL2(S2),P^-a.s. 146

Since by (139), for every ε>0, supt[0,T]α^ε(s)L2(S2)2K1(2), P^-a.s., using the dominated convergence theorem we infer that

limε00Tα^ε(t)-u^(t),ϕL2(S2)dt=0,P^-a.s. 147

By the Hölder inequality, (139) and (144) for every ε>0 and every r>1

E^0Tα^ε(t)-u^(t)L2(S2)dtrcE^0Tα^ε(t)L2(S2)r+u^(t)L2(S2)rdtcTK1(r), 148

where c is some positive constant. To conclude the proof of assertion (a) it is sufficient to use (147), (148) and the Vitali’s convergence theorem.

(b) Since α^εu^ in C([0,T];Hw) P^-a.s. we infer that

α^ε(0),ϕL2(S2)u^(0),ϕL2(S2),P^-a.s. 149

Also, note that by condition (87) in Assumption 2, Inline graphic converges weakly to u0 in L2(S2).

On the other hand, by (133) we infer that the laws of α^ε(0) and αε(0) on H are equal. Since αε(0) is a constant random variable on the old probability space, we infer that α^ε(0) is also a constant random variable (on the new probability space) and hence, by (72) and (94), we infer that Inline graphic almost surely (on the new probability space). Therefore we infer that

u^(0),ϕL2(S2)=u0,ϕL2(S2),

concluding the proof of assertion (b).

(c) Since α^εu^ in C([0,T];Hw) P^-a.s.,

limε0ν1+ε0tcurlα^ε(s),curlϕL2(S2)ds=-limε0ν1+ε0tα^ε(s),ΔϕL2(S2)ds=-ν0tu^,ΔϕL2(S2)ds=ν0tcurlu^,curlϕL2(S2)ds. 150

The Cauchy–Schwarz inequality and estimate (140) infer that for all t[0,T] and ε(0,1]

E^0tν1+εcurlα^ε(s),curlϕL2(S2)ds2ν2curlϕL2(S2)2E^0tcurlα^ε(s)L2(S2)2dscK2 151

for some constant c>0. By (150), (151) and the Vitali’s convergence theorem we conclude that for all t[0,T]

limε0E^0tν11+εcurlα^ε(s)-curlu^(s),curlϕL2(S2)ds=0.

Assertion (c) follows now from (140), (143) and the dominated convergence theorem.

(d) For the non-linear term, using the Sobolev embedding H2(S2)L(S2), we have

0tα^ε(s)·α^ε(s),ϕL2(S2)ds-0tu^(s)·u^(s),ϕL2(S2)ds0tS2α^ε(s,x)-u^(s,x)·u^(s,x)·ϕ(x)dxds+0tS2α^ε(s,x)·α^ε(s,x)-u^(s,x)·ϕ(x)dxdsα^ε-u^L2(0,T;H)u^L2(0,T;V)ϕH2(S2)+0tα^ε(s,x)-u^(s,x),α^ε(s)ϕL2(S2)ds. 152

The first term converges to zero as ε0, since α^εu^ strongly in L2(0,T;H) P^-a.s., u^L2(0,T;V) and the second term converges to zero too as ε0 because α^εu^ weakly in L2(0,T;V). Using the Hölder inequality, estimates (139) and the embedding H2(S2)L(S2) we infer that for all t[0,T], ε(0,1], the following inequalities hold

E^0t11+εα^ε(s)·α^ε(s),ϕL2(S2)ds2E^0tα^ε(s)L2(S2)2ϕL(S2)ds2cϕH2(S2)tE^0tα^ε(s)L2(S2)4dscϕH3(S2)tE^sups[0,t]α^ε(s)L2(S2)4c~K1(4). 153

By (152), (153) and the Vitali’s convergence theorem we obtain for all t[0,T],

limε0E^0t11+εα^ε(s)·α^ε(s)-u^(s)·u^(s),ϕL2(S2)ds=0. 154

Using the Hölder inequality and estimates (139), (144), we obtain for all t[0,T], ε(0,1]

E^0t11+εα^ε(s)·α^ε(s)-u^(s)·u^(s),ϕL2(S2)dscϕH3(S2)E^supt[0,T]α^ε(s)L2(S2)2+supt[0,T]u^(s)L2(S2)22c~K1(2)

where c,c~>0 are constants. Hence by (154) and the dominated convergence theorem, we infer assertion (d).

(e) Assertion (e) follows because by Assumption 2 the sequence Inline graphic converges weakly in L2(0,T;L2(S2)) to f.

(f) By the definition of maps G~ε and G, we have

graphic file with name 245_2020_9702_Equ352_HTML.gif

Since, by Assumption 2, for every j{1,,N}, and s[0,t], Inline graphic converges weakly to gj(s) in L2(S2) as ε0, we get

graphic file with name 245_2020_9702_Equ155_HTML.gif 155

By assumptions on g~εj, we obtain the following inequalities for every t[0,T] and ε(0,1]

graphic file with name 245_2020_9702_Equ156_HTML.gif 156

where c,c~>0 are some constants. Using the Vitali’s convergence theorem, by (155) and (156) we infer

graphic file with name 245_2020_9702_Equ157_HTML.gif 157

Hence, by the properties of the Itô integral we deduce that for all t[0,T],

graphic file with name 245_2020_9702_Equ158_HTML.gif 158

By the Itô isometry and assumptions on g~εj and gj we have for all t[0,T] and ε(0,1]

graphic file with name 245_2020_9702_Equ159_HTML.gif 159

where c>0 is a constant. Thus, by (158), (159) and the dominated convergence theorem assertion (f) holds.

Lemma 27

For all t[0,T] and ϕU, we have (along the sequence εkn)

  1. Inline graphic

  2. Inline graphic

  3. Inline graphic

  4. Inline graphic

where the process α^ε_ is defined in (136).

Proof

Let us fix ϕU.

(a) Let t[0,T], then by the Hölder inequality, Lemma 18, Poincaré inequality and estimate (142), we have the following inequalities

graphic file with name 245_2020_9702_Equ353_HTML.gif

Thus

graphic file with name 245_2020_9702_Equ160_HTML.gif 160

We infer assertion (a) by dominated convergence theorem, estimate (142) and convergence (160).

(b) Using the Hölder inequality, scaling property (Lemma 11), Corollary 4, relations (67), (68), and estimates (139), (142), we get for t[0,T]

graphic file with name 245_2020_9702_Equ354_HTML.gif

Thus

graphic file with name 245_2020_9702_Equ161_HTML.gif 161

We infer assertion (b) by dominated convergence theorem, estimates (139), (142) and convergence (161). Assertion (c) can be proved similarly.

(d) Now for the last one, using the Hölder inequality, Corollary 4, relations (67), (68), and estimates (141), (142), we get for t[0,T]

graphic file with name 245_2020_9702_Equ355_HTML.gif

Thus

graphic file with name 245_2020_9702_Equ162_HTML.gif 162

We infer assertion (d) by dominated convergence theorem, estimates (141), (142) and convergence (162).

Finally, to finish the proof of Theorem 3, we will follow the methodology as in [42] and introduce some auxiliary notations (along sequence εkn)

graphic file with name 245_2020_9702_Equ163_HTML.gif 163
Λ(u^,W^,ϕ):=u0,ϕL2(S2)-ν0tcurlu^(s),curlϕL2(S2)ds-0tu^(s)·u^(s),ϕL2(S2)ds+0tf(s),ϕL2(S2)ds+0tG(s)dW^(s),ϕL2(S2). 164

Corollary 5

Let ϕU. Then (along the sequence εkn)

limε0α^ε(·),ϕL2(S2)-u^(·),ϕL2(S2)L1(Ω^×[0,T])=0 165

and

limε0Λε(α^ε,β^ε,W^ε,ϕ)-Λ(u^,W^,ϕ)L1(Ω^×[0,T])=0. 166

Proof

Assertion (165) follows from the equality

α^ε(·),ϕL2(S2)-u^(·),ϕL2(S2)L1(Ω^×[0,T])=E^0Tα^ε(t)-u^(t),ϕL2(S2)dt

and Lemma 26 (a). To prove assertion (166), note that by the Fubini Theorem, we have

Λε(α^ε,β^ε,W^ε,ϕ)-Λ(u^,W^,ϕ)L1(Ω^×[0,T])=0TE^Λε(α^ε,β^ε,W^ε,ϕ)(t)-Λ(u^,W^,ϕ)(t)dt.

To conclude the proof of the corollary, it is sufficient to note that by Lemma 26(b)-(f) and Lemma 27, each term on the right hand side of (163) tends at least in L1(Ω^×[0,T]) to the corresponding term (to zero in certain cases) in (164).

Conclusion of proof of Theorem 3

Let us fix ϕU. Since αε is a solution of (85), for all t[0,T],

αε(t),ϕL2(S2)=Λε(αε,β~ε,W~ε,ϕ)(t),P-a.s.

In particular,

0TEαε(t),ϕL2(S2)-Λε(αε,β~ε,W~ε,ϕ)(t)dt=0.

Since L(αε,ηkn,W~ε)=L(α^ε,η^kn,W^ε), on BZT×YT×C([0,T];RN) (along the sequence εkn),

0TE^α^ε(t),ϕL2(S2)-Λε(α^ε,β^ε,W^ε,ϕ)(t)dt=0.

Therefore by Corollary 5 and the definition of Λ, for almost all t[0,T] and P^-almost all ωΩ^

u^(t),ϕL2(S2)-Λ(u^,W^,ϕ)(t)=0,

i.e. for almost all t[0,T] and P^-almost all ωΩ^

u^(t),ϕL2(S2)+ν0tcurlu^(s),curlϕL2(S2)ds+0tu^(s)·u^(s),ϕL2(S2)ds=u0,ϕL2(S2)+0tf(s),ϕL2(S2)ds+0tG(s)dW^(s),ϕL2(S2). 167

Hence (167) holds for every ϕU. Since u^ is a.s. H-valued continuous process, by a standard density argument, we infer that (167) holds for every ϕV (U is dense in V).

Putting U^:=Ω^,F^,F^,P^, we infer that the system U^,W^,u^ is a martingale solution to (75)–(77).

Acknowledgements

Open access funding provided by Austrian Science Fund (FWF).

Vector Analysis in Spherical Coordinates

In this appendix we collect some basic results from vector algebra and formulas for Laplace and gradient of scalar function and vector fields in spherical coordinates.

The following identities are very well known [53, Appendix] in vector algebra: Let u,v and w be R3-valued smooth vector fields then

curlcurlu=-Δu+divu, 168
u·v×w=u×v·w, 169
Qεv·curludy=Qεu·curlvdy+Qεu×v·ndσ. 170

The Laplace–Beltrami operator of a scalar function ψ, in spherical coordinates (r,λ,φ) is given by

2ψ=2ψr2+2rψr+1r2sinλλsinλψλ+1r2sin2λ2ψφ2, 171

and its gradient is given by

ψ=ψrer^+1rψλeλ^+1rsinλψφeφ^. 172

For a vector field u written in the spherical coordinates, u=urer^+uλeλ^+uφeφ^, the curl and the divergence are given as follows

curlu=1rsinλλ(sinλuφ)-uλφer^+1rsinλurφ-1rr(ruφ)eλ^+1rr(ruλ)-1rurλeφ^, 173
divu=1r2r(r2ur)+1rsinλλ(uλsinλ)+1rsinλuφφ. 174

The Laplacian of a vector field in spherical coordinates is

Δur=2ur-2urr2-2r2uλλ-2cotλr2uλ-2r2sinλuφφ,Δuλ=2uλ+2r2urλ-uλr2sin2λ-2cosλr2sin2λuφφ,Δuφ=2uφ-uφr2sin2λ+1r2sinλurφ+2cosλr2sin2λuλφ, 175

where 2ur, 2uλ and 2uφ are as in (171).

We recall some standard differential operators on the unit sphere S2. For ψ a scalar function defined on S2, the tangential gradient is given by

ψ=ψλeλ^+1sinλψφeφ^. 176

The Laplace–Beltrami of a scalar function ψ is

Δψ=1sinλλsinλψλ+1sinλ2ψφ2. 177

For a tangential vector field v defined on S2, v=vλe^λ+vφe^φ, the tangential divergence is expressed by

divv=1sinλλvλsinλ+1sinλvφφ, 178

and the curlv is the scalar function defined by

curlv=1sinλλvφsinλ-1sinλvλφ. 179

The Laplace–de Rham operator applied to a vector field v is given by

Δv=Δvλ-2cosλsin2λvφφ-vλsin2λeλ^+Δvφ+2cosλsin2λvλφ-vφsin2λeφ^, 180

where Δvφ and Δvλ are as in (177).

The curl and the Stokes operator

In this section we present a integration by parts formula corresponding to curl operator and later we use it to give a relation between the Stokes operator Aε and curl.

Let OR3 be a bounded domain with a regular boundary O. Define

H(curl):={uL2(O):curluL2(O)}. 181

H(curl) with the graph norm

vH(curl)2:=vL2(O)2+curlvL2(O)2,vH(curl),

is a Hilbert space.

The following theorem is a reformulation of Lemma 4.2 from [17, p. 341].

Theorem 4

Assume that OR3 be a bounded domain with a regular boundary and n be unit normal vector field on Γ=O (directed towards exterior of O). Then there exists a unique bounded linear map

n×·|Γ:H(curl)H-1/2(Γ), 182

such that

n×·(u)=n×u|Γ, 183

if uC01(O¯) and

Ocurlu·vdy-Ou·curlvdy=H-1/2(Γ)(n×·)(u)|Γ,v|ΓH1/2(Γ) 184

for every uH(curl) and vH1(O).

Remark 8

We will call formula (184) the generalised Stokes formula. From now on we will write n×u|Γ instead of (n×·)(u)|Γ for uH(curl).

Recall that

H={uL2(O):divu=0inOandu·n=0onΓ}, 185

and the Stokes operator is given by

D(A)={uH2(O):divu=0inO,u·n=0andn×curlu=0onΓ}, 186
Au=π(-Δu),uD(A). 187

Remark 9

To define n×curlu as an element of H-1/2(Γ), we need to know that curl(curlu)L2(O). But if uH2(O), then obviously this condition is satisfied.

Theorem 5

A is self-adjoint and non-negative on H.

Proof

Here we will only show that A is symmetric and non-negative. Let u,vD(A), then

Au,vL2(O)=π(-Δu),vL2(O)=-Δu,πvL2(O)=-Δu,vL2(O). 188

Recall that (from (168)) for smooth R3-valued vector fields,

curl(curlu)=-Δu+(divu).

Using the above identity in (188) along with the fact that uD(A), in particular, divu=0 and generalised Stokes formula (184), we have

Au,vL2(O)=Ocurl(curlu)·vdy=Ocurlu·curlvdy+H-1/2(Γ)n×(curlu)|Γ=0,sinceuD(A),v|ΓH1/2(Γ)=Ocurlu·curlvdy.

Similarly

u,AvL2(O)=Av,uL2(O)=Ocurlv·curludy=Ocurlu·curlvdy.

This establishes that A is symmetric on H. The non-negativity follows from the above identity by taking v=uD(A).

Using the definition of D(A), we can characterise D(A1/2) as

D(A1/2)={uL2(O):divu=0,curluL2(O)andu·n=0onΓ}.

By Theorem 6.1 [17, Pg 358] we have

D(A1/2){uH1(O):u·n=0onΓ}.

We use the following relation repeatedly in our calculations.

Lemma 28

Let uD(A1/2) and vD(A). Then

curlu,curlvL2(O)=u,AvL2(O).

Proof

Note that for uD(A1/2) and vD(A), the LHS makes sense. Using the generalised Stokes formula (184), we get

Ocurlu·curlvdy=Ou·curl(curlv)dy+H-1/2(Γ)n×u|Γ,curlv|ΓH1/2(Γ).

To finish the proof we need the following lemma:

Lemma 29

Let uH1(O) such that u·n=0 on Γ and ΦH1(O) with n×Φ=0 on Γ. Then

H-1/2(Γ)n×u|Γ,Φ|ΓH1/2(Γ)=0. 189

Proof

It is sufficient to prove (189) for uC01(O¯) with u·n=0 on Γ. In this case for all xΓ

n×u·Φ=-u×n·Φ=-u·n×Φ=0.

The proof of Lemma 28 is finished by observing that

u,AvL2(O)=Ou·curl(curlv)dy,

from the proof of Theorem 5.

Let us consider an abstract framework. Let H be a Hilbert space and A be a non-negative self-adjoint operator on H.

Lemma 30

Let uD(A1/2) and vD(A). Then

A1/2u,A1/2vH=(u,Av)H.

Proof

Take uD(A1/2), vD(A). Then A1/2vD(A1/2). So by self-adjointness of A1/2 and A1/2A1/2=A,

A1/2u,A1/2vH=u,A1/2(A1/2v)H=u,AvH.

Lemma 31

Let u,vD(A1/2), then by Lemmas 30 and 28,

curlu,curlvL2(O)=A1/2u,A1/2vL2(O).

Proof

Let u,vD(A), then

A1/2u,A1/2vL2(O)=u,AvL2(O)=curlu,curlvL2(O).

By density argument, this is true for all u,vD(A1/2).

As a consequence of the above lemma we have

uL2(O)2=curluL2(O)2,uD(A1/2).

Proof of the Poincaré and the Ladyzhenskaya Inequalities

Proof of Lemma 13

We will establish the Poincaré inequality (54) following the footsteps of Lemma 2.1 [53] with all the details. By density argument, it is enough to prove (54) for smooth functions. Let ψC(Qε¯) be a real continuous function. We write for any ξ,η[1,1+ε]:

ξ2ψ2(ξ,x)+η2ψ2(η,x)=2ξηψ(ξ,x)ψ(η,x)+ξψ(ξ,x)-ηψ(η,x)2=2ξηψ(ξ,x)ψ(η,x)+ηξrψr(r,x)dr2 190

with x=y|y|S2. We fix ξ and integrate w.r.t. η[1,1+ε] to obtain

εξ2ψ2(ξ,x)+11+εη2ψ2(η,x)dη=2ξψ(ξ,x)11+εηψ(η,x)dη+11+εηξrψr(r,x)dr2dη. 191

With ψ=ur and ξ=1, observing that ur(1,x)=0 (because of the boundary condition u·n=0 on Qε) from (191) we obtain

11+εη2ur2(η,x)dη=11+εηξrurr(r,x)dr2dη. 192

Applying (191) with ψ=N^εuλ, we get

εξ2N^εuλ(ξ,x)2+11+εη2N^εuλ(η,x)2dη=2ξN^εuλ(ξ,x)11+εηN^εuλ(η,x)dη+11+εηξrN^εuλr(r,x)dr2dη. 193

Observing from Lemma 6 for every ψL2(Qε)

11+εrN^εψ(r,x)dr=0xS2,

and since the first term on the LHS of (193) is positive we can simplify (193) as follows

11+εη2N^εuλ(η,x)2dη11+εηξrN^εuλr(r,x)dr2dη. 194

Similarly for ψ=N^εuφ, we have

11+εη2N^εuφ(η,x)2dη11+εηξrN^εuφr(r,x)dr2dη. 195

Thus, using (192), (194) and (195) for each of the cases ψ=ur, ψ=N^εuλ and ψ=uφ, we obtain

11+εη2ψ2(η,x)dη11+εηξrψr(r,x)dr2dη. 196

Using the Cauchy–Schwarz inequality, we find

11+εη2ψ2(η,x)dη11+ε|ξ-η|dη11+εrψr2drε211+εrψr2dr2ε211+εψr2r2dr+2ε211+ε|ψ|2dr2ε211+εψr2r2dr+2ε211+ε|rψ|2dr, 197

the last inequality follows since r1. On rearranging, we obtain

1-2ε211+εr2ψ2(r,λ,φ)dr2ε211+εψr2r2dr, 198

which implies for 0ε<12

11+εr2ψ2(r,λ,φ)dr4ε211+εψr2r2dr. 199

We then integrate w.r.t. λ and φ to obtain

Qεψ2(y)dy4ε2Qεψr2dy. 200

Adding (200) for ψ=ur, ψ=N^εuλ and ψ=N^εuφ; using finally

QεrN~εu2dyN~εuL2(Qε)2=curlN~εuL2(Qε)2uVε,

we conclude the proof of the inequality (54).

Proof of Lemma 14

We will prove the lemma for smooth vector fields uC(Qε). By Lemma 2.3 [53] there exists a constant c0>0 s.t.

N~εuL6(Qε)c0N~εuL2(Qε). 201

By Lemma 6.1 [17, Eq. 6.11, Pg 359] for vector fields vC1(Qε) with v·n=0 on Qε, we have a constant c2>0 s.t.

vL2(Qε)22divvL2(Qε)2+curlvL2(Qε)2+c2vL2(Qε)2. 202

Also by Poincaré inequality (54) for all uVε, we have

N~εuL2(Qε)24ε2curlN~εuL2(Qε)2. 203

Using (202) with v=N~εu along with the fact that divv=0 if divu=0 and v·n=0 on Qε if u·n=0 on Qε and combining it with the Poincaré inequality (203), we obtain

N~εuL6(Qε)22c02N~εuVε2+4c2ε2N~εuVε2.

Therefore, choosing c1=2c01+4c21/2 we establish (55) for every uC1(Qε) with divu=0 on Qε and u·n=0 on Qε. We finish the proof using density argument.

Proof of Lemma 18

Let u be a tangential vector field defined on S2, u=(0,uλ,uφ). Then using the definition of the map Inline graphic (see (63)), for Qεy=rx, r(1,1+ε) and xS2, we have

graphic file with name 245_2020_9702_Equ356_HTML.gif

Using the definitions of Laplace (Δ) for vector fields in spherical coordinates, Laplace–Beltrami (2) for scalars, tangential Laplace (Δ) for tangential vector fields and Laplace–Beltrami (Δ) for the scalar defined on S2, we have following relations:

graphic file with name 245_2020_9702_Equ357_HTML.gif

where

graphic file with name 245_2020_9702_Equ358_HTML.gif

If divu=0, then by the definition of div,

1sinλλ(uλsinλ)+1sinλuφφ=0

which is equivalent to

uλλ+uλcotλ+1sinλuφφ=0.

Hence using all the above relations, we obtain

graphic file with name 245_2020_9702_Equ359_HTML.gif

Since ε(0,1), the inequality (68) holds.

Compactness

Skorohod Theorem and Aldous Condition

Let E be a separable Banach space with the norm ·E and let B(E) be its Borel σ-field. The family of probability measures on (E,B(E)) will be denoted by P. The set of all bounded and continuous E-valued functions is denoted by Cb(E).

Definition 3

The family P of probability measures on E,B(E) is said to be tight if for arbitrary ε>0 there exists a compact set KεE such that

μ(Kε)1-ε,forallμP.

We used the following Jakubowski’s generalisation of the Skorokhod Theorem, in the form given by Brzeźniak and Ondreját [14, Theorem C.1], see also [31], as our topological space ZT is not a metric space.

Theorem 6

Let X be a topological space such that there exists a sequence {fm}mN of continuous functions fm:XR that separates points of X. Let us denote by S the σ-algebra generated by the maps {fm}. Then

  1. every compact subset of X is metrizable,

  2. if (μm)mN is a tight sequence of probability measures on (X,S), then there exists a subsequence (mk)kN, a probability space (Ω,F,P) with X-valued Borel measurable variables ξk,ξ such that μmk is the law of ξk and ξk converges to ξ almost surely on Ω.

Let (S,ϱ) be a separable and complete metric space.

Definition 4

Let uC([0,T];S). The modulus of continuity of u on [0, T] is defined by

m(u,δ):=sups,t[0,T],|t-s|δϱ(u(t),u(s)),δ>0.

Let (Ω,F,P) be a probability space with filtration F:=(Ft)t[0,T] satisfying the usual conditions, see [38], and let (Xn)nN be a sequence of continuous F-adapted S-valued processes.

Definition 5

We say that the sequence (Xn)nN of S-valued random variables satisfies condition [T] iff ε>0,η>0,δ>0:

supnNPm(Xn,δ)>ηε. 204
Lemma 32

[12, Lemma 2.4] Assume that (Xn)nN satisfies condition [T]. Let Pn be the law of Xn on C([0,T];S), nN. Then for every ε>0 there exists a subset AεC([0,T];S) such that

supnNPn(Aε)1-ε

and

limδ0supuAεm(u,δ)=0. 205

Now we recall the Aldous condition [A], which is connected with condition [T] (see [39] and [1]). This condition allows to investigate the modulus of continuity for the sequence of stochastic processes by means of stopped processes.

Definition 6

(Aldous condition) A sequence (Xn)nN satisfies condition [A] iff ε>0, η>0, δ>0 such that for every sequence (τn)nN of F-stopping times with τnT one has

supnNsup0θδPϱ(Xn(τn+θ),Xn(τn))ηε.
Lemma 33

[39, Theorem 3.2] Conditions [A] and [T] are equivalent.

Tightness Criterion

Now we formulate the compactness criterion analogous to the result due to Mikulevicus and Rozowskii [40], Brzeźniak and Motyl [12] for the space ZT, see also [5, Lemma 4.2].

Lemma 34

Let ZT, T be as defined in (116). Then a set KZT is T-relatively compact if the following three conditions hold

  1. supuKsups[0,T]u(s)L2(S2)<,

  2. supuK0Tu(s)V2ds<, i.e. K is bounded in L2(0,T;V),

  3. limδ0supuKsups,t[0,T]|t-s|δu(t)-u(s)D(A-1)=0.

Using Sect. D.1 and the compactness criterion from Lemma 34 we obtain the following corollary.

Corollary 6

(Tightness criterion) Let (αε)ε>0 be a sequence of continuous F-adapted H-valued processes such that

  1. there exists a constant C1>0 such that
    supε>0Esups[0,T]αε(s)H2C1,
  2. there exists a constant C2>0 such that
    supε>0E0Tcurlαε(s)L2(S2)2dsC2,
  3. (αε)ε>0 satisfies the Aldous condition [A] in D(A-1).

Let P~ε be the law of αε on ZT. Then for every δ>0 there exists a compact subset Kδ of ZT such that

supε>0P~ε(Kδ)1-δ.

Kuratowski Theorem and Proof of Lemma 25

This appendix is dedicated to the proof of Lemma 25. We will first recall the Kuratowski Theorem [33] in the next subsection and prove some related results which will be used later to prove Lemma 25 in Sect. E.2.

Kuratowski Theorem and Related Results

Theorem 7

Assume that X1,X2 are the Polish spaces with their Borel σ-fields denoted respectively by B(X1),B(X2). If φ:X1X2 is an injective Borel measurable map then for any E1B(X1), E2:=φ(E1)B(X2).

Next two lemmas are the main results of this appendix. For the proof of Lemma 35 please see [6, Appendix B].

Lemma 35

Let X1,X2 and Z be topological spaces such that X1 is a Borel subset of X2. Then X1Z is a Borel subset of X2Z, where X2Z is a topological space too, with the topology given by

τ(X2Z)=AB:Aτ(X2),Bτ(Z). 206

Proof of Lemma 25

In this subsection we recall Lemma 25 and prove it using the results from previous subsection.

Lemma 36

Let T>0 and ZT be as defined in (116). Then, the following sets C([0,T];H)ZT, L2(0,T;V)ZT are Borel subsets of ZT.

Proof

First of all C([0,T];H)C([0,T];D(A-1))L2(0,T;H). Secondly, C([0,T];H) and C([0,T];D(A-1))L2(0,T;H) are Polish spaces. And finally, since H is continuously embedded in D(A-1), the map

i:C([0,T];H)C([0,T];D(A-1))L2(0,T;H),

is continuous and hence Borel. Thus, by application of the Kuratowski Theorem (see Theorem 7), C([0,T];H) is a Borel subset of C([0,T];D(A-1))L2(0,T;H). Therefore, by Lemma 35, C([0,T];H)ZT is a Borel subset of C([0,T];D(A-1))L2(0,T;H)ZT which is equal to ZT.

Similarly we can show that L2(0,T;V)ZT is a Borel subset of ZT. L2(0,T;V)L2(0,T;H) and both are Polish spaces thus by application of the Kuratowski Theorem, L2(0,T;V) is a Borel subset of L2(0,T;H). Finally, we can conclude the proof of lemma by Lemma 35.

Compliance with Ethical Standards

Conflict of interest

The authors declare that they have no conflict of interest.

Footnotes

1

We could have considered the case N=. This case will be considered in the companion paper [7].

2

The space Xw denotes a topological space X with weak topology. In particular, C([0,T];Xw) is the space of weakly continuous functions v:[0,T]X.

3

T is the supremum of topologies T1, T2, T3 and T4, i.e. it is the coarsest topology on ZT that is finer than each of T1, T2, T3 and T4.

4

The space ZT×YT×C([0,T];RN) satisfies the assumption of Theorem 6. Indeed, since ZT and YTε, ε>0 satisfies the assumptions (see [5, Lemma 4.10]) and C([0,T];RN) is a Polish space and thus automatically satisfying the required assumptions.

The research of all three authors is partially supported by Australian Research Council Discover Project Grant DP180100506, “Uncertainty on Spheres and Shells: Mathematics and Methods for Applications”. Zdzisław Brzeźniak has been supported by the Leverhulme Project Grant Ref No RPG-2012-514 and by Australian Research Council Discover Project Grant DP160101755. The research of Gaurav Dhariwal was supported by Department of Mathematics, University of York and is partially supported by the Austrian Science Fund (FWF) Grants P30000, W1245, and F65.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Contributor Information

Zdzisław Brzeźniak, Email: zdzislaw.brzezniak@york.ac.uk.

Gaurav Dhariwal, Email: gaurav.dhariwal@tuwien.ac.at.

Quoc Thong Le Gia, Email: qlegia@unsw.edu.au.

References

  • 1.Aldous D. Stopping times and tightness. Ann. Probab. 1978;6(2):335–340. [Google Scholar]
  • 2.Avrin JD. Large-eigenvalue global existence and regularity results for the Navier-Stokes equation. J. Differ. Equ. 1996;127:365–390. [Google Scholar]
  • 3.Aubin T. Some Nonlinear Problems in Riemannian Geometry. New York: Springer; 1998. [Google Scholar]
  • 4.Babin AV, Vishik MI. Attractors of partial differential equations and estimate of their dimension. Russ. Math. Surv. 1983;38:151–213. [Google Scholar]
  • 5.Brzeźniak, Z., Dhariwal, G.: Stochastic constrained Navier-Stokes equations on T2. Submitted (2019). arXiv:1701.01385 [DOI] [PMC free article] [PubMed]
  • 6.Brzeźniak, Z., Dhariwal, G.: Stochastic tamed Navier-Stokes equations on R3: the existence and the uniqueness of solutions and the existence of an invariant measure. To appear in J. Math. Fluid Mech. (2020). arXiv:1904.13295
  • 7.Brzeźniak, Z., Dhariwal, G., Le Gia, Q.T.: Stochastic Navier-Stokes equations on a thin spherical domain: Existence of a martingale solution (In preparation) [DOI] [PMC free article] [PubMed]
  • 8.Brzeźniak Z, Goldys B, Jegaraj T. Weak solutions of a stochastic Landau-Lifshitz-Gilbert equation. Appl. Math. Res. eXpress. 2013;2013(1):1–33. [Google Scholar]
  • 9.Brzeźniak Z, Goldys B, Le Gia QT. Random dynamical systems generated by stochastic Navier-Stokes equations on a rotating sphere. J. Math. Anal. Appl. 2015;426:505–545. [Google Scholar]
  • 10.Brzeźniak Z, Goldys B, Le Gia QT. Random attractors for the stochastic Navier-Stokes equations on the 2D unit sphere. J. Math. Fluid. Mech. 2018;20:227–253. [Google Scholar]
  • 11.Brzeźniak Z, Motyl E. Existence of a martingale solution of the stochastic Navier-Stokes equations in unbounded 2D and 3D domains. J. Differ. Equ. 2013;254(4):1627–1685. [Google Scholar]
  • 12.Brzeźniak Z, Motyl E. The existence of martingale solutions to the stochastic Boussinesq equations. Glob. Stoch. Anal. 2014;1(2):175–216. [Google Scholar]
  • 13.Brzeźniak Z, Motyl E, Ondreját M. Invariant measure for the stochastic Navier-Stokes equations in unbounded 2D domains. Ann. Probab. 2017;45(5):3145–3201. [Google Scholar]
  • 14.Brzeźniak Z, Ondreját M. Stochastic wave equations with values in Riemanninan manifolds. Stochastic partial differential equations and applications. Quaderni di Matematica. 2011;25:65–97. [Google Scholar]
  • 15.Cattabriga L. Su un problema al contorno relativo al sistema di equazioni di Stokes. Rend. Semin. Mat. Univ. Padova. 1961;31:308–340. [Google Scholar]
  • 16.Ciarlet PG. Plates and Junctions in Elastic Multi-structures. An Asymptotic Analysis. Paris: Masson; 1990. [Google Scholar]
  • 17.Duvaut G, Lions JL. Inequalities in Mechanics and Physics. Berlin: Springer; 1976. [Google Scholar]
  • 18.Ghidaglia, J.M., Temam, R.: Lower bound on the dimension of the attractor for the Navier-Stokes equations in space dimension 3. In: Mechanics, Analysis and Geometry: 200 Years After Lagrange, pp. 33–60, North-Hollan Delta Ser., North-Holland, Amsterdam (1991)
  • 19.Grigoryan A. Heat Kernel and Analysis on Manifolds. Providence, RI: Amer. Math. Soc; 2009. [Google Scholar]
  • 20.Hale JK, Raugel G. A damped hyperbolic equation on thin domains. Trans. Am. Math. Soc. 1992;329:185–219. [Google Scholar]
  • 21.Hale, J.K., Raugel, G.: Partial differential equations on thin domains. In: Differential Equations and Mathematical Physics (Birmingham, AL, 1990), pp. 63–97, Math. Sci. Engrg., vol. 186. Academic Press, Boston (1992)
  • 22.Hale JK, Raugel G. Reaction-diffusion equation on thin domains. J. Math. Pures Appl. 1992;71:33–95. [Google Scholar]
  • 23.Ibragimov RN, Pelinovsky DE. Incompressible viscous fluid flows in a thin spherical shell. J. Math. Fluid Mech. 2009;11:60–90. [Google Scholar]
  • 24.Ibragimov RN. Nonlinear viscous fluid patterns in a thin rotating spherical domain and applications. Phys. Fluids. 2011;23:123102. [Google Scholar]
  • 25.Ibragimov NH, Ibragimov RN. Integration by quadratures of the nonlinear Euler equations modeling atmospheric flows in a thin rotating spherical shell. Phys. Lett. A. 2011;375:3858. [Google Scholar]
  • 26.Iftimie D. The 3D Navier-Stokes equations seen as a perturbation of the 2D Navier-Stokes equations. Bull. Soc. Math. France. 1999;127:473–517. [Google Scholar]
  • 27.Iftimie D, Raugel G. Some results on the Navier-Stokes equations in thin 3D domains. J. Differ. Equ. 2001;169:281–331. [Google Scholar]
  • 28.Il’in AA. The Navier-Stokes and Euler equations on two dimensional manifolds. Math. USSR Sb. 1991;69:559–579. [Google Scholar]
  • 29.Il’in AA. Partially dissipative semigroups generated by the Navier-Stokes system on two dimensional manifolds and their attractors. Russ. Acad. Sci. Sb. Math. 1994;78:47–76. [Google Scholar]
  • 30.Il’in AA, Filatov AN. On unique solvability of the Navier-Stokes equations on the two dimensional sphere. Sov. Math. Dokl. 1989;38:9–13. [Google Scholar]
  • 31.Jakubowski, A.: The almost sure Skorokhod representation for subsequences in nonmetric spaces. Teor. Veroyatn. Primen. 42(1), 209–216 (1998); translation in Theory Probab. Appl. 42(1), 167–174 (1998)
  • 32.Kruse R. Strong and Weak Approximation of Semilinear Stochastic Evolution Equations. Lecture Notes Math. Cham: Springer; 2014. [Google Scholar]
  • 33.Kuratowski, K.: Topologie, Vol. I (French)3’eme Ed. Monografie Matematyczne, Tom XX, Polskie Towarzystwo Matematyczne, Warszawa (1952)
  • 34.Lions JL, Temam R, Wang S. New formulations of the primitive equations of atmosphere and applications. Nonlinearity. 1992;5:237–288. [Google Scholar]
  • 35.Lions JL, Temam R, Wang S. On the equations of the large-scale ocean. Nonlinearity. 1992;5:1007–1053. [Google Scholar]
  • 36.Lions JL, Temam R, Wang SH. Mathematical theory for the coupled atmosphere-ocean models. J. Math. Pures Appl. (9) 1995;74(2):105–163. [Google Scholar]
  • 37.Marsden JE, Raitu TS, Raugel G. Les équation d’Euler dans des coques sphériques minces. C. R. Acad. Sci. Paris. 1995;321:1201–1206. [Google Scholar]
  • 38.Métivier M. Semimartingales: A Course on Stochastic Processes. Berlin: Walter de Gruyter & and Co.; 1982. [Google Scholar]
  • 39.Métivier M. Stochastic Partial Differential Equations in Infinite Dimensions. Pisa: Scuola Normale Superiore; 1988. p. 142. [Google Scholar]
  • 40.Mikulevicius R, Rozovskii BL. Global L2-solutions of stochastic Navier-Stokes equations. Ann. Prob. 2005;33(1):137–176. [Google Scholar]
  • 41.Moise I, Temam R, Ziane M. Asymptotic analysis of the Navier-Stokes equations in thin domains. Topol. Methods Nonlinear Anal. 1997;10:249–282. [Google Scholar]
  • 42.Motyl E. Stochastic hydrodynamic-type evolution equations driven by Lévy noise in 3D unbounded domains—abstract framework and applications. Stoch. Process. Appl. 2014;124:2052–2097. [Google Scholar]
  • 43.Pedlosky J. Geophysical Fluid Dynamics. 2. New York: Springer; 1987. p. xiv+710. [Google Scholar]
  • 44.Raugel G, Sell GR. Navier-Stokes equations on thin 3D domains I. Global attractors and global regularity of solutions. J. Am. Math. Soc. 1993;6:503–568. [Google Scholar]
  • 45.Raugel, G., Sell, G.R.: Navier-Stokes equations on thin 3D domains II. Global regularity of spatially periodic solutions. In: Nonlinear Partial Differential Equations and Their Applications, Collège de France Seminar, vol. XI, pp. 205–247, Longman, Harlow (1994)
  • 46.Saito J. Boussinesq equations in thin spherical domains. Kyushu J. Math. 2005;59:443–465. [Google Scholar]
  • 47.Serrin J. Mathematical Principles of Classical Fluid Mechanics, Encly. of Physics. New York: Springer; 1959. pp. 125–263. [Google Scholar]
  • 48.Taylor ME. Analysis on Morrey spaces and applications to Navier-Stokes and other evolution equations. Commun. Partial Differ. Equ. 1992;17:1407–1456. [Google Scholar]
  • 49.Temam R. Infinite-Dimensional Dynamical Systems in Mechanics and Physics. 2. New York: Springer; 1997. p. xxi+650. [Google Scholar]
  • 50.Temam R. Navier-Stokes Equations: Theory and Numerical Analysis. UK: American Mathematical Society; 2000. [Google Scholar]
  • 51.Temam R, Wang S. Inertial forms of Navier-Stokes equations on the sphere. J. Funct. Anal. 1993;117:215–242. [Google Scholar]
  • 52.Temam R, Ziane M. Navier-Stokes equations in three-dimensional thin domains with various boundary conditions. Adv. Differ. Equ. 1996;1:499–546. [Google Scholar]
  • 53.Temam R, Ziane M. Navier-Stokes equations in thin spherical domains. Contemp. Math. 1997;209:281–314. [Google Scholar]

Articles from Applied Mathematics and Optimization are provided here courtesy of Springer

RESOURCES