Skip to content
BY 4.0 license Open Access Published by De Gruyter February 28, 2021

Non-stationary Navier–Stokes equations in 2D power cusp domain

II. Existence of the solution

  • Konstantin Pileckas and Alicija Raciene

Abstract

The initial boundary value problem for the non-stationary Navier-Stokes equations is studied in 2D bounded domain with a power cusp singular point O on the boundary. We consider the case where the boundary value has a nonzero flux over the boundary. In this case there is a source/sink in O and the solution necessary has infinite energy integral. In the first part of the paper the formal asymptotic expansion of the solution near the singular point was constructed. In this, second part, the constructed asymptotic decomposition is justified, i.e., existence of the solution which is represented as the sum of the constructed asymptotic expansion and a term with finite energy norm is proved. Moreover, it is proved that the solution represented in this form is unique.

MSC 2010: 35Q30; 35A20; 76M45; 76D03; 76D10

1 Introduction

In this paper we continue to study the boundary value problem for the non-stationary Navier–Stokes system

(1.1) utνΔu+(u)u+p=f,divu=0,u|ΩO=a,u(x,0)=b(x)

in a two-dimensional bounded domain Ω with the cusp point O = (0, 0) at the boundary: Ω = GHΩ0, where GH=xR2:|x1|<φ(x2),x2(0,H],φ(x2)=γ0x2λ,γ0=const,λ>1 (see Figure 1). For simplicity we assume that the boundary ∂Ω∂Ω0 is infinitely smooth. Here u = (u1, u2) stands for the velocity field, p stands for the pressure, ν > 0 is the constant kinematic viscosity.

Fig. 1 Domain Ω
Fig. 1

Domain Ω

It is supposed that supp a∂Ω0∂Ω, i.e., the support of the boundary value aL2(0,T;W3/2,2(Ω)) is separated from the cusp point O. We also assume that the flux F (t) of a is nonzero:

(1.2) ΩandS=F(t),F(t)/0,F(0)=0.

The initial velocity b ∈ W1,2(Ω) and the boundary value a have to satisfy the necessary compatibility conditions

(1.3) divb(x)=0,b(x)|Ω=a(x,0).

The solution u of (1.1) has to satisfy the condition

σ(h)undS+ΩΩ0andS=0

where σ(h) = {x ∈ GH : x2 = h = const}, which means that the total flux of the fluid is equal to zero. Thus,

(1.4) σ(h)undx1=F(t)/0,

and we can regard the cusp point O as a source (or a sink) of intensity F (t).

More information and references concerning the Navier–Stokes equations in domains with singular boundaries are given in the introduction to the first part of the paper, see [1]. In [1] the formal asymptotic decomposition of the solution (u, p) near the cusp point was constructed. This asymptotic expansion has the form

(1.5) U[J]x1x2λ,x2,t,τ=UO,[J]x1x2λ,x2,t+UB,[J]x1x2λ,x2,τ,P[J]x1x2λ,x2,t,τ=PO,[J]x1x2λ,x2,t+PB,[J]x1x2λ,x2,τ,

where the pair (UO,[J], PO,[J]) is an approximate solution (outer asymptotic expansion) of the Navier–Stokes problem in variables y1=x1x2λ,y2=x2,t=t; the "slow" time variable t plays the role of a parameter and, in general, the initial condition is not satisfied. The pair (UB,[J], PB,[J]) is the boundary layer corrector (the inner part of the asymptotic expansion) which compensate the discrepancy in the initial condition. Notice that (UB,[J], PB,[J]) exponentially vanishes as the fast time τ=tx22λ tends to infinity. The number J is taken so large that the discrepancy H[J]of(U[J],P[J]) in the Navier–Stokes equations belongs to the space L2(0, T; L2(Ω)), while the discrepancy in the initial condition is zero (see [1] for details). Moreover, in order to ensure the existence of all terms of asymptotic decomposition up to the order J, we have to assume that

latlL2(0,T;W1/2,2(Ω)),l=1,2,...,J+1.

Here we justify this asymptotics. We prove that there exists a solution of problem (1.1) which is represented as a sum of the singular part (the constructed asymptotic decomposition) and the function having finite energy. To be more precise, we construct a solenoidal extension V of the boundary value a which coincides with U[J] near the cusp point and we look for the solution (u, p) of (1.1) in the form u = V + v, p = ζP[J] + q, where ζ is a smooth cut-off function localising the asymptotical part of the pressure near the cusp point O. Then for (v, q) we obtain the problem

(1.6) vtνΔv+(v)v+(V)v+(v)V+q:=fˆ,divv=0,v|Ω=0,u(x,0)=b(x)V(x,0):=uˆ0(x)

with fˆL2(0,T;L2(Ω)),uˆ0W 1,2(Ω). In the paper we prove the existence of a unique regular solution v to (1.6).

The existence of singular solutions to the time-periodic and the non-stationary Stokes problem in the domain with a cusp point was studied in [2, 3]. We can also mention the recent paper [4] where the Dirichlet problem for the non-stationary Stokes system is studied in a three-dimensional cone. The non-stationary Navier–Stokes equations in tube-structures were studied in [5, 6]. The solvability of the stationary Navier–Stokes system in the cusp domain with source or sink in the cusp point was proved in [7]. The steady Navier-Stokes equations are also studied in a punctured domain Ω = Ω0 \ {O} with O ∈ Ω0 assuming that the point O is a sink or source of the fluid, see [8, 9, 10] and [11] for the review of these results. We also mention the papers [12, 13, 14, 15] where the stationary Navier–Stokes equations were studied in domains with paraboloidal outlets to infinity. Such geometry has similarities with the cusp domains, the difference is that in the case of a domain with outlet to infinity x2 → ∞, while in the cusp domain x2 → 0.

The paper is organised as follows. In Section 2 we introduce the main notation, function spaces and prove certain inequalities needed in subsequent sections. In Section 3 we study the Stokes problem and the Stokes operator in the cusp domain. Finally, the main result of the paper, the unique solvability of problem (1.6), is proved in Section 4.

2 Notation, function spaces and auxiliary results

Let G be a domain in Rn. We use usual notation of functional spaces (e.g., [16]). By Lp(G) and Wm,p(G), 1 ≤ p < ∞, we denote the Lebesgue and Sobolev spaces, respectively. The norm in a Banach space X is denoted by X.C(G) is the set of all infinitely differentiable functions defined on G and C0(G) - the subset of all functions from C(G) having compact supports in G. By W k,q(G) we denote the completion of the C0(G) in the Wm,p(G) -norm and by Wm−1/p,p(∂G) the space of traces on ∂G of functions from Wm,p(G). The space Lp(0, T; X) consists of all measurable functions u : [0, T] → X with

uLp(0,T;X)=0Tu(t)Xpdt1/p<,1p<.

We do not distinguish in notation the spaces of vector and scalar functions; from the context it will be clear which space we have in mind.

Denote J0(G)={vC0(G):divv=0} the set of all divergence free vector fields from C0(G) and by J0(G) be the closure of J0 (G) in L2(G)-norm. Let H(G)={vW1,2(G): div v = 0}. If G is a bounded domain with Lipschitz boundary, then H (G) coincides with the closure of J0(G) in the norm W1,2(G) (see [17]).

Let us consider the cusp domain Ω. Let h0 = H, hk=hk1φ(hk1)2L, where L is a Lipschitz constant for the function φ, k = 1,2, . . . . The sequence {hk} is decreasing and bounded from below. Assume that the limit of this sequence is a0 ≠ 0. From the definition of the sequence it follows that a0=a0φ(a0)2L. Then φ(a0) = 0. However, φ(a0) ≠ 0 for a0 ≠ 0 and, hence the limit a0 = 0. Since the sequence is decreasing and the limit is equal to 0, all its elements are positive.

Denote ωl={xR2:|x1|<φ(x2),x2(hl,hl1)},l=1,. Note that

(2.1) 12φ(hl)φ(t)32φ(hl),t[hl+1,hl].

Define the transformation y = Plx by the formulas

(2.2) y1=2Lx1φ(x2),y2=2L(x2hl1)φ(hl1)

and introduce the domains

G0={yR2:|y1|<2L,1<y2<0},G1={yR2:|y1|<2L,1g(y1)<y2<1},G2={yR2:|y1|<2L,2<y2<1}.

In the definition of G1 the function g ∈ C satisfies the conditions g(±2L) = 0,0 < g(y1) < 1 for |y1| < 2L and it is such that the curve {y:y1=2L}{y:y1=2L}{y:|y1|<2L,y2=1g(y1)} is infinitely smooth.

Obviously the transformation Pl1 maps G0 onto ωl. Consider the domain ωl=Pl1G2. Then ωlωl= {xR2:|x1|<φ(x2),hl1φ(hl1)L<x2<hl1+φ(hl1)2L}ωl+1ωlωl1,l=2,3, (see Figure 2). It is easy to see that

Ω=Ω0(l=1ωl)=Ω0ω1(l=2ωl).
Fig. 2 Domains 
ωj,ωj∗andωˆj+1.
$ \omega_j, \omega_j^* \;{\text {and}}\; \hat\omega_{j+1}.$
in different coordinate systems
Fig. 2

Domains ωj,ωjandωˆj+1. in different coordinate systems

Let us fix K ≥ 2 (sufficiently large) and define ωˆK=PK11G1,ωˆK=PK11(G0G1). Obviously, ωˆKωK and ωˆKωˆK. Let

ΩK=Ω0(l=1K1ωl)ωˆK.

The boundary of ΩK consist of ΩΩˉK and the curve ΓK which is defined as ΓK=PK11E0, where E0 = {y : |y1| < 2L, y2 = −1 − g(y1)}. By construction, the boundary ∂ΩK is smooth (see Figure 3).

We can take also the other covering of the domain ΩK. Namely,

ΩK=Ω0ω1(l=1K2ωl)ωˆK1ωˆK,

where ωˆK1=PK11({yR2:|y1|<2L,1g(y1)<y2<1})=PK11(G1).

We also introduce domains Ωl=Ω0(j=1lωj),l=1,2, (see Figure 4).

In ΩK we define the function space L12(ΩK) with the weighted norm

fL12(ΩK)2=Ω0|f(x)|2dx+ΩKΩ0φ2(x2)|f(x)|2dx.

Let us prove some auxiliary inequalities for functions defined in Ω.

Fig. 3 Domains ΩK.
Fig. 3

Domains ΩK.

Fig. 4 Domains 
Ωj♯
$ {\it\Omega}_{j}^{\sharp}$
Fig. 4

Domains Ωj

Lemma 2.1

(Poincaré inequalities). Let uWloc1,2(Ωˉ),u|Ω=0. Then the following inequalities

(2.3) ωl|u(x)|2dx9π2φ2(hl1)ωl|u(x)|2dx,
(2.4) hHφ(x2)φ(x2)|φ(x2)|κ2|u(x)|2dx4π2hHφ(x2)φ(x2)|φ(x2)|κ|u(x)|2dx

hold for any κ ∈ R and any h ∈ (0, H).

Proof

By the classical Poincaré inequality on the interval ( − φ(x2), φ(x2)), we have

φ(x2)φ(x2)|u(x1,x2)|2dx14π2φ2(x2)φ(x2)φ(x2)|u(x1,x2)x1|2dx1.

Integrating this inequality over (hl , hl−1) and applying (2.1) we derive (2.3). To prove (2.4) it is enough to multiply the above inequality by |φ(x2)|κ−2 and integrate over the interval (h, H).

Lemma 2.2

Let uWloc1,2(Ωˉ). Then

(2.5) uL4(ωl)4cφ2(hl1)uL2(ωl)2(uL2(ωl)2+φ2(hl1)uL2(ωl)2)

holds with a constant c independent of l. In particular, if uWloc2,2(Ωˉ), then the following estimate

holds.

uL4(ωl)4cφ2(hl1)uL2(ωl)2(uL2(ωl)2+φ2(hl1)2uL2(ωl)2)

Proof

After the transformation Pl, the domain ωl is transformed into the domain G0={y:|y1|<1,1< y2<0} which is independent of l. In G0 holds the inequality (see [18])

uL4(G0)4cuL2(G0)2(uL2(G0)2+uL2(G0)2).

Passing in the last inequality to variables x we obtain

ωl|u|44L2φ(x2)φ(hl1)dxc(ωl|u|24L2φ(x2)φ(hl1)dx)(ωl|u|24L2φ(x2)φ(hl1)dx+ωl|yu|24L2φ(x2)φ(hl1)dx)
cωl|u|24L2φ(x2)φ(hl1)dx(ωl|u|24L2φ(x2)φ(hl1)dx+ωl[|ux1|2(φ2(x2)+|x1|2φ2(hl1)φ2(x2)|φ(x2)|2)+|ux2|2φ2(hl1)]4L2φ(x2)φ(hl1)dx).

Since |x1| < φ(x2) and |φ′(x2)| ≤ const, from the last inequality using (2.1) we derive (2.5). From (2.3) and (2.5) we obtain

Lemma 2.3

Let uWloc1,2(Ωˉ),u|Ω=0. Then

(2.6) uL4(ωl)cuL2(ωl)1/2uL2(ωl)1/2cφ1/2(hl1)uL2(ωl)

with a constant c is independent of l.

Let us consider in ωl the divergence problem

(2.7) divv=gin ωl,v=0on ωl.

Lemma 2.4

Let g ∈ L2(ωl) and

(2.8) ωlg(x)dx=0.

Then there exists a solution v ∈ W˚1,2(ωl) of (2.7) satisfying the estimate

(2.9) vL2(ωl)cgL2(ωl)

with a constant c independent of l.

Proof

The transformation Pl (see (2.2)) maps the domain ωl onto G0 = {y : |y1| < 1,−1 < y2 < 0}. Because of (2.8),

G0g(Pl1(y))Jl1(Pl1(y))dy=ωlg(x)dx=0,

where Jl(x)=4L2φ(hl1)φ(x2) is the Jacobian. Therefore (see [17]), there exists a function vˆW 1,2(G0) such that

divyvˆ(y)=Jl1(Pl1(y))g(Pl1(y))

and

(2.10) yvˆL2(G0)cJˆl1gˆL2(G0),

where gˆ(y)=g(Pl1(y)), etc. Let us define the vector field v(x) with the components

v1(x)=vˆ1(y)|y=Pl(x)2Lφ(hl1)+vˆ2(y)|y=Pl(x)2Lx1φ(x2)φ2(x2),v2(x)=vˆ2(y)|y=Pl(x)2Lφ(x2).

Then it is straightforward to verify that

divxv=4L2φ(hl1)φ(x2)divyvˆ(y)|y=Pl(x)=g(x).

Thus, we have only to show estimate (2.9). Let us estimate the norm xvL2(ωl). Using inequality (2.1) for φ, Poincaré inequality (2.3) and the relations |φ(x2)|const,|φ′′(x2)φ2(x2)|+|φ2(x2)φ(x2)|const (recall that φ(x2)=γ0x2λ,λ>1) we obtain

ωl|xv1(x)|2dxcφ2(hl1)ωl(|xvˆ1(Pl(x))|2+|xvˆ2(Pl(x))|2+|vˆ2(Pl(x))|2φ2(x2))dxcφ2(hl1)G0|yvˆ(y)|2dy,ωl|xv2(x)|2dxcφ2(hl1)ωl(|xvˆ2(Pl(x))|2+|vˆ2(Pl(x))|2φ2(x2))dxcφ2(hl1)G0|yvˆ(y)|2dy.

Estimating now yvˆL2(G0)2 by inequality (2.10), using the expression for the Jacobean and returning to coordinates x, we derive estimate (2.9) with a constant independent of l.

Remark 2.1

It is easy to see that Lemmas 2.1–2.4 remain valid if we take the domains ωl,l=2,,k,orωˆk, ωˆk instead of ωl.

3 Stokes problem and Stokes operator

3.1 Estimates of solutions to the Stokes problem

In ΩK consider the Dirichlet boundary value problem for the Stokes system

(3.1) νΔv+p=f,divv=0,v|ΩK=0.

The weak solution v ∈ H (ΩK) to (3.1) satisfies the integral identity

νΩKvηdx=ΩKfηdxηH(ΩK).

Lemma 3.1

Let fL12(ΩK). Then problem (3.1) has a unique solution v ∈ H (ΩK) and there holds the estimate

(3.2) vL2(ΩK)cfL12(ΩK)

with a constant c independent of K.

Proof

By Poincaré’s inequality (2.4) with κ = 0,

|ΩKfηdx|cfL12(ΩK)2ηL2(ΩK).

Hence, the statement of the lemma follows from Lax–Milgram’s theorem.

Lemma 3.2

Let fL2(ΩK)L12(ΩK). Then the weak solution v of (3.1) satisfies the estimate

(3.3) Ω0|v|2dx+ΩKΩ0φ2(x2)|v|2dxcΩK|f|2dx

with a constant c independent of K.

Proof

Let

Φ(x2)=1φ2(hl),xΩ0(j=1lωj)=Ωl,1φ2(x2),x(j=l+1K1ωj)ωˆK=ΩKΩl.

Here and below the number l is fixed; we specify it during the proof.

Consider the function u = Φ(x2)v. Then div u = Φ′(x2)v2. Since v ∈ H (ΩK), the flux of v over any section x2 = const of GH is equal to zero, i.e.,

φ(x2)φ(x2)Φ(x2)v2(x1,x2)dx1=0x2(hK,H].

Integrating over x2 we conclude that

ωjΦ(x2)v2(x1,x2)dx=0,j=l+1,,K1;ωˆKΦ(x2)v2(x1,x2)dx=0.

Then by Lemma 2.4, there exist functions wjW 1,2(ωj),j=l+1,,K1,wKW 1,2(ωˆK) such that

divwj=Φ(x2)v2inωj,j=l+1,,K1,divwK=Φ(x2)v2inωˆK.

Moreover, the following estimates

(3.4) wjL2(ωj)cΦv2L2(ωj)cmaxxωj|φ(x2)|φ3v2L2(ωj),j=l+1,,K1,wKL2(ωˆK)cmaxxωˆK|φ(x2)|φ3v2L2(ωˆK)

hold with a constant c independent of j and K. Taking into account inequalities (2.1), from (3.4) we obtain

(3.5) φ(x2)wjL2(ωj)cmaxxωj|φ(x2)|φ2(x2)v2L2(ωj),j=l+1,,K1,φ(x2)wKL2(ωˆK)cmaxxωˆK|φ(x2)|φ2(x2)v2L2(ωˆK).

Define the function

w(x)=0,xΩl,wj(x),xωj,j=l+1,,K1,wK(x),xωˆK

(recall that Φ(x2)=0inΩl). Take in the integral identity η = u + w. By construction, div (u + w) = 0 and hence, u + w ∈ H (ΩK). This yields

νΩKΦ|v|2dx=νΩKvΦvdxνΩKvwdx+ΩKf(Φv+w)dx=J1+J2+J3.

Let us estimate the integrals Ji in the right-hand side of the last relation. Using (2.4) and (3.5) we get

|J1|=|νΩKΩlvΦvdx|2νΩKΩl|v||v||φ(x2)|φ3(x2)dxcsupxΩKΩl|φ(x2)|(ΩKΩlφ2(x2)|v|2dx)1/2(ΩKΩlφ4(x2)|v|2dx)1/2csupxΩKΩl|φ(x2)|ΩKΩlφ2(x2)|v|2dx;|J2|νj=l+1K1|ωjvwjdx|+|ωˆKvwKdx|εj=l+1K1ωjφ2(x2)|v|2dx+εωˆKφ2(x2)|v|2dx+cεj=l+1K1ωjφ2(x2)|wj|2dx+cεωˆKφ2(x2)|wK|2dxεΩKΩlφ2(x2)|v|2dx+cεsupxΩKΩl|φ(x2)|2(j=l+1K1ωjφ2(x2)|v|2dx+ωˆKφ2(x2)|v|2dx)(ε+cεsupxΩKΩl|φ(x2)|2)ΩKΩlφ2(x2)|v|2dx;
|ΩKfΦvdx|1φ2(hl)|Ωlfvdx|+j=l+1K1|ωjf1φ2(x2)vdx|+|ωˆKf1φ2(x2)vdx|cε(1φ2(hl)Ωl|f|2dx+j=l+1K1ωj|f|2dx+ωˆK|f|2dx)+ε(1φ2(hl)Ωl|v|2dx+j=l+1K1ωj1φ4(x2)|v|2dx+ωˆK1φ4(x2)|v|2dx)cΩK|f|2dx+ε(cφ2(hl)Ωl|v|2dx+j=l+1K1ωj1φ2(x2)|v|2dx+ωˆK1φ2(x2)|v|2dx)cΩK|f|2dx+cεΩKΦ|v|2dx;|ΩKfwdx|=|ΩKΩlfwdx|j=l+1K1|ωjfwjdx|+|ωˆKfwKdx|cΩK|f|2dx+j=l+1K1ωj|wj|2dx+ωˆK|wK|2dxcΩK|f|2dx+j=l+1K1ωjφ2(x2)|wj|2dx+ωˆKφ2(x2)|wK|2dxcΩK|f|2dx+cj=l+1K1supxωj|φ(x2)|2ωjφ4(x2)|v2|2dx+csupxωˆK|φ(x2)|2ωˆKφ4(x2)|v2|2dxcΩK|f|2dx+csupxΩKΩl|φ(x2)|2ΩKΩlφ2(x2)|v|2dx.

Collecting the obtained estimates yields

(3.6) νΩKΦ|v|2dxc1ΩK|f|2dx+c2(ε+supxΩKΩl|φ(x2)|)ΩKΦ|v|2dx,

where the constant c1 is independent of K (but c1 depends on l) and c2 is independent of K and l. The function φ′(x2) is monotonically decreasing and tends to zero as x2 → 0. Hence supxΩKΩl|φ(x2)|=|φ(hl)| and limlφ(hl)=0. We choose and fix l such that |φ(hl)|ν/(4c2) and take ε=ν/(4c2). Then from (3.6) follows the inequality

ΩKΦ|v|2dxc1ΩK|f|2dx

which is equivalent to (3.3).

Lemma 3.3

For sufficiently large K the weak solution v of problem (3.1) satisfies the estimate

(3.7) 2vL2(ΩK)cfL2(ΩK)

with a constant c independent of K.

Proof

Consider the solution *(v, p) of problem (3.1) in the domain ωl,l=2,,K2. Changing the variables y=Pl(x) (see (2.2)) we rewrite problem (3.1) in coordinate y in the domain G2:

(3.8) νΔyvˆ(y)+yqˆ(y))=φ2(hl1)4L2fˆ(y)+Hˆ(y),divyv˜(y)=gˆ(y),v˜|y1=±2L=0,

where vˆ(y)=v(Pl1y),etc.,qˆ(y)=φ(hl1)2Lpˆ(y),

(3.9) Hˆ(y)=α11(y)2vˆy12+α12(y)2vˆy1y2+β1(y)vˆy1+γ(y)qˆy1,α11(y)=ν((φ2(hl1)φ2(x2))φ2(x2)+x12φ(x2)φ2(hl1)φ4(x2))|x=Pl1(y),α12(y)=ν2x1φ(x2)φ(hl1)φ2(x2)|x=Pl1(y),β1(y)=νx1(φ′′(x2)φ(x2)2φ2(x2))φ2(hl1)2Lφ3(x2)|x=Pl1(y),γ(y)=(γ1(y),γ2(y))T,γ1(y)=φ(hl1)φ(x2)φ(x2)|x=Pl1(y),γ2(y)=x1φ(x2)φ(hl1)φ2(x2)|x=Pl1(y),gˆ(y)=μ1(y)vˆ1y1+μ2(y)vˆ2y1,μ1(y)=φ(x2)φ(hl1)φ(x2)|x=Pl1(y),μ2(y)=x1φ(x2)φ(hl1)φ2(x2)|x=Pl1(y).

Applying the usual local ADN-estimates for elliptic problems (see [19, 20]) in the pair of domains G0 ⊂ G2, we obtain the estimate

(3.10) vˆL2(G0)2+yvˆL2(G0)2+y2vˆL2(G0)2+yqˆL2(G0)2c(φ4(hl1)fˆL2(G2)2+HˆL2(G2)2+gˆW1,2(G2)2+vˆL2(G2)2+qˆqˉL2(G2)2),

where qˉ=1|G2|G2qˆ(y)dy.SinceG2(qˆ(y)qˉ)dy=0, there exists wW 1,2(G2) such that div yw=qˆ(y)qˉ in G2 and

ywL2(G2)cqˆqˉL2(G2)

(see [17]). Multiplying (3.8) by w and integrating by parts yields

qˆqˉL2(G2)2=G2qˆ(y)(qˆ(y)qˉ)dy=G2qˆ(y)divywdy=νG2yvˆywdyφ2(hl1)4L2G2fˆwdyG2HˆwdyνyvˆL2(G2)ywL2(G2)+φ2(hl1)4L2fˆL2(G2)wL2(G2)+HˆL2(G2)wL2(G2)cyvˆL2(G2)qˆqˉL2(G2)+cφ2(hl1)fˆL2G2)qˆqˉL2(G2)+cHˆL2(G2)qˆqˉL2(G2).

Therefore,

(3.11) qˆqˉL2(G2)c(yvˆL2(G2)+φ2(hl1)fˆL2(G2)+HˆL2(G2)).

From (3.10) using (3.11) and Poincaré’s inequality (2.3) we derive

(3.12) vˆL2(G0)2+yvˆL2(G0)2+y2vˆL2(G0)2+yqˆL2(G0)2c(φ4(hl1)fˆL2(G2)2+HˆL2(G2)2+gˆW1,2(G2)2+yvˆL2(G2)2).

By definition ωl={x:|x1|φ(x2),hl1φ(hl1)L<x2<hl1+φ(hl1)2L} and the following inequality

|φ(hl1)φ(x2)|maxx2ωl|φ(x2)||hl1x2|cmaxx2ωl|φ(x2)|φ(hl1)

holds. For y=Pl(x),xωl, using this inequality, (2.1) and the definition φ(x2)=γ0x2λ,λ>1, we obtain

|α11(y)|+|α12(y)|+|β1(y)|+|γ1(y)|+|γ1(y)|+|μ1(y)|+|μ2(y)|ε(l),

where limlε(l)=0 see (3.9)). Therefore, from (3.12) follows the estimate

vˆL2(G0)2+yvˆL2(G0)2+y2vˆL2(G0)2+yqˆL2(G0)2c(φ4(hl1)fˆL2(G2)2+yvˆL2(G2)2)+cε(l)(y2vˆL2(G2)2+yqˆL2(G2)2).

Passing to coordinates x and using the same arguments we derive

(3.13) 1φ2(hl1)vL2(ωl)2+xvL2(ωl)2+φ2(hl1)x2vL2(ωl)2+xqL2(ωl)2c(φ2(hl1)fL2(ωl)2+xvL2(ωl)2)+cε(l)(φ2(hl1)x2vL2(ωl)2+xqL2(ωl)2),l=1,,K2.

The constant c in (3.13) is independent of l. Multiplying (3.13) by 1φ2(hl1) yields

(3.14) 1φ4(hl1)vL2(ωl)2+1φ2(hl1)xvL2(ωl)2+x2vL2(ωl)2+xpL2(ωl)2c(fL2(ωl)2+1φ2(hl1)xvL2(ωl)2)+cε(l)(x2vL2(ωl)2+xpL2(ωl)2),l=1,,K2.

By the same ADN-estimate together with the properties of the domain ωˆK , we get the inequalities

(3.15) 1φ4(hK2)vL2(ωK1)2+1φ2(hK2)xvL2(ωK1)2+x2vL2(ωK1)2+xpL2(ωK1)2c(fL2(ωK1)2+1φ2(hK2)xvL2(ωK1)2)+cε(K)(x2vL2(ωK1)2+xpL2(ωK1)2)

and

(3.16) 1φ4(hK1)vL2(ωK)2+1φ2(hK1)xvL2(ωk)2+x2vL2(ωK)2+xpL2(ωK)2c(fL2(ωK)2+1φ2(hK1)xvL2(ωK)2)+cε(K)(x2vL2(ωK)2+xpL2(ωK)2),

with constants independent of K.

Let l0 < K−2 be a positive natural number (l0 be fixed later). Arguing as above we can prove the following local estimate for the pair of domains Ωl0+1Ωl0+2:

(3.17) vW2,2(Ωl0+1)2+pL2(Ωl0+1)2c(fL2(Ωl0+2)2+xvL2(Ωl0+2)2).

Summing inequalities (3.14) from l0 to K −2, adding (3.15)-(3.17) and taking into account that φ(hl1)φ(x2) in ωl and ωl (see (2.1)), we get

(3.18) vW2,2(Ωl0+1)2+pL2(Ωl0+1)2+φ2vL2(ΩKΩl0+1)2+φ1vL2(ΩKΩl0+1)2+2vL2(ΩKΩl0+1)2+pL2(ΩKΩl0+1)2c1(fL2(ΩK)2+vL2(Ωl0+2)2+φ1vL2(ΩKΩl0)2)+c2ε(l0)(2vL2(ΩKΩl0)2+pL2(ΩKΩl0)2).

Since limlε(l)=0, we can chose l0 to satisfy c2ε(l0)≤ κ, where κ is such that

κ(2vL2(ΩKΩl0)2+pL2(ΩKΩl0)2)12(2vL2(Ωl0+1)2+pL2(ΩK)2+2vL2(ΩKΩl0)2).

Then estimate (3.18) takes the form

vW2,2(Ωl0+1)2+pL2(Ωl0+1)2+φ2vL2(ΩKΩl0+1)2+φ1vL2(ΩKΩl0+1)2+2vL2(ΩKΩl0+1)2+pL2(ΩKΩl0+1)2c1(fL2(ΩK)2+vL2(Ωl0+2)2+φ1vL2(ΩKΩl0)2).

In particular,

(3.19) 2vL2(ΩK)2c1(fL2(ΩK)2+vL2(Ωl0+2)2+φ1vL2(ΩKΩl0)2).

Estimating the last two term in the right-hand side of (3.19) by (3.2) and (3.3) we obtain (3.7).

3.2 Stokes operator

The most results we present in this subsection are standard (e.g., [21]). Problem (3.1) can be rewritten in the operator form (without loss of generality we suppose that f ∈ J0(ΩK)[1], adding the gradient part to the pressure)

Δ˜v=f,

where Δ˜=PΔ:H(ΩK)W2,2(ΩK)J0(ΩK) is an unbounded operator with the domain H(ΩK)W2,2(ΩK), where P is the projector from L2ΩK) onto J0(ΩK) (Leray’s projector). For given wH(ΩK)W2,2(ΩK) the operator Δ˜w is defined by

ΩKΔ˜wvdx=ΩK(νΔw+p)vdx=νΩKΔwvdx=νΩKwvdxvJ0(ΩK)

(for vJ0(ΩK) holds div v = 0). By density argument,

(3.20) ΩKΔ˜wvdx=νΩKΔwvdxvJ0(ΩK).

Hence,

(3.21) ΩK|Δ˜w|2dx=νΩKΔwΔ˜wdx.

From (3.20) also follows the estimate

Δ˜w(H(ΩK))wL2(ΩK)=wH(ΩK).

Since H(ΩK)W2,2(ΩK) is dense in H (ΩK), there exists a unique extension of the operator Δ˜ (denoted again by Δ˜ from H(ΩK)W2,2(ΩK) to the whole space H (ΩK). Moreover, the extension Δ˜:H(ΩK)(H(ΩK)) is a bijection. Δ˜ is called the Stokes operator.

It is known (see, e.g., [21, 22]) that

(i) The Stokes operator has a discrete spectrum:

Δ˜w=λw,wH(ΩK),w0;
λi>0,limiλi+.

(ii) The set {wl} of eigenfunctions of Δ˜ is an orthogonal basis in J0(ΩK) and H(ΩK),wlL2(ΩK)=λk, wlL2(ΩK)=1. Since ∂ΩK is smooth, we have wlH(ΩK)W2,2(ΩK) [2]

Relation (3.21) yields

(3.22) Δ˜wL2(ΩK)cΔwL2(ΩK)c2wL2(ΩK).

From (3.3) follows the estimate

(3.23) Ω0|v|2dx+ΩKΩ0φ2(x2)|v|2dxcΩK|Δ˜v|2dx,

and from (3.7) we get the inequality

2wL2(ΩK)cΔ˜wL2(ΩK),

which together with (3.22) implies

(3.24) c12wL2(ΩK)Δ˜wL2(ΩK)2c22wL2(ΩK).

Note that constants in (3.22),(3.24) are independent of K.

4 Solvability of Navier–Stokes problem

4.1 Construction of the extension of boundary data

Consider problem (1.1)–(1.4). Suppose that aL2(0,T;Wloc3/2,2(Ω{O})),atL2(0,T;Wloc1/2,2(Ω{O})), supp suppaΩ0ΩΩ1.

First we consider the linear extension operator E in the domain Ω3,E:W3/2,2(Ω3)W2,2(Ω3) given by Ea=w(1), where w(1)|Ω3=a. Since the boundary ΩΩ3 is smooth and supp aΩ0ΩΩ1, the linear operator E is bounded:

(4.1) EaW2,2(Ω3)2=w(1)W2,2(Ω3)2caW3/2,2(Ω)2.

Moreover, w(1) can be constructed so that supp w(1)Ω2 (see, e.g., [16]).

If a=a(x,t)andat(,t)W1/2,2(Ω), then, due to the fact that the operator E is linear, we have Eat= wt(1) and

(4.2) EatW1,2(Ω3)2catW1/2,2(Ω)2.

Moreover, if aL2(0,T;W3/2,2(Ω)),atL2(0,T;W1/2,2(Ω)), then integrating (4.1), (4.2) by t yields

(4.3) w(1)L2(0,T;W2,2(Ω3))2caL2(0,T;W3/2,2(Ω))2,wt(1)L2(0,T;W1,2(Ω3))2catL2(0,T;W1/2,2(Ω))2.

Let U[J](x1x2λ,x2,t,τ) be the formal asymptotic decomposition of the velocity component near the cusp point O constructed in [1] . Recall that

U[J](x1x2λ,x2,t,τ)=UO,[J](x1x2λ,x2,t)+UB,[J](x1x2λ,x2,tx22λ),

where UO,[J] is the outer asymptotic expansion and UB,[J] is the boundary layer expansion (see also formulas (1.5) in Introduction). In order to insure the existence of U[J], the following regularity requirements for the boundary value a are needed:

latlL2(0,T;W1/2,2(Ω)),l=1,2,...,J+1.

It is proved (see inequality (4.15) in [1]) that the vector field U[J] satisfies the following estimates

(4.4) supt[0,T]U[J](,y2,t)W1,2(Υ)2+U[J]L2(0,T;W2,2(Υ)2+Ut[J]L2(0,T;L2(Υ))2cφ2(y2)0T|||F|||J+12dt,supt[0,T]U[J](,y2,t)W2,2(Υ)2+supt[0,T]Ut[J]L2(Υ)2+Ut[J]L2(0,;L2(Υ))2cφ2(y2)0T|||F|||J+22dt,supt[0,T]U[J](,y2,t)y2W1,2(Υ)2cφ4(y2)0T|||F|||J+12dt,

where y1=x1x2λ,y2=x2,φ(y2)=γ0y2λ,λ>1,Υ=(γ0,γ0),|||F|||J2=k=0J|kF(t)tk|2.

Since W1,2(Υ)C(Υ), we have

(4.5) supt(0,T)y1Υ(|U[J](y1,y2,t)|2+|U[J](y1,y2,t)y1|2)csupt(0,T)U[J](,y2,t)W2,2(Υ)2cφ2(y2)0T|||F|||J+22dt,supt(0,T)y1Υ|U[J](y1,y2,t)y2|2dtcsupt(0,T)U[J](,y2,t)y2W1,2(Υ)2cφ4(y2)0T|||F|||J+12dt.

Passing to coordinates x we obtain

(4.6) supt(0,T)x1(φ(x2),φ(x2))|U[J](x1,x2,t)|2dtcφ2(x2)0T|||F|||J+12dt,supt(0,T)x1(φ(x2),φ(x2))|xU[J](x1,x2,t)|2cφ4(x2)0T|||F|||J+12dt.

Notice that F(t)=ΩandS, and hence,

(4.7) 0T|||F|||J2dtc0T|||a|||J2dt,supt(0,T)|||F|||J2dtc0T|||a|||J+12dt,

where |||a|||J2=k=0Jka(,t)tkW1/2,2(Ω)2.

Consider the function B=w(1)+ζU[J], where ζ=ζ(x2) is a smooth cut-off function equal to one in ΩΩ2 and equal to zero in Ω1. Obviously, B|∂Ω = a, however, B is not solenoidal, divB=divw(1)+ζU[J]:=h. Notice that

Ω2hdx=Ω2(w(1)+ζU[J])ndS=Ω0ΩandS+Ω2ΩU[J]ndS=F(t)F(t)=0.

Since supp hΩ2 and the boundary Ω3Ω is smooth, there exist a function w(2)W2,2(Ω3) such that supp suppw(2)Ω3,w(2)=0 in the neighbourhood of Ω3Ω and

(4.8) divw(2)=hinΩ3,w|Ω3(2)=0.

Moreover,

(4.9) w(2)(,t)W2,2(Ω3)2ch(,t)W1,2(Ω3)2c(w(1)(,t)W2,2(Ω3)2+U[J](,t)W1,2(Ω3)2)

(see [23]). By construction in [23], it follows that the operator 𝓓 of problem (4.8), D:w(2)(,t)W2,2(Ω3) div w(2)(,t)=h(,t)W1,2(Ω3) is linear and the inverse operator 𝓓−1 defined on functions hW1,2(Ω3), satisfying the condition Ω3hdx=0, is bounded. Moreover, equations (4.8) can be differentiated with respect

to t:

divwt(2)=htinΩ3,wt|Ω3(2)=0,

and

(4.10) wt(2)(,t)W1,2(Ω3)2cht(,t)L2(Ω3)2c(wt(1)(,t)W1,2(Ω3)2+Ut[J](,t)L2(Ω3)2).

Integrating inequalities (4.9), (4.10) with respect to t and using estimates (4.3), (4.4) and (4.7) we obtain

(4.11) w(2)L2(0,T;W2,2(Ω3))2c0T(a(,t)W3/2,2(Ω)2+|||a(,t)|||J+12+|||F(t)|||J+12)dtc0T(a(,t)W3/2,2(Ω)2+|||a(,t)|||J+12)dt=c0TaJ+12dt,wt(2)L2(0,T;W1,2(Ω3))2c0T(at(,t)W1/2,2(Ω)2+|||F(t)|||J+12)dt0T|||a(,t)|||J+12dt,

where aJ+12=a(,t)W3/2,2(Ω)2+|||a(,t)|||J+1.

Define

W=w(1)+w(2),V=W+ζU[J],

where ζ is a smooth cut-off function defined above. By construction div V = 0, V|∂Ω = a and V = U[J] for xΩΩ3. Therefore, for xΩΩ3 the function V satisfies estimates (4.4), (4.6), while for xΩ3 from (4.11) and (4.4) it follows that

(4.12) 0TV(,t)W2,2(Ω3)2dt+0TVt(,t)L2(Ω3)2dtc0T(aW3/2,2(Ω)2+|||a|||J+12+|||F|||J+12)dtc0TaJ+12dt.

We look for the solution (u, p) of problem (1.1) in the form

u=v+V,p=q+ζP[J].

Then for (v, q) we obtain the following problem

(4.13) vtνΔv+(v)v+(V)v+(v)V+q=fˆ,divv=0,v|ΩO=0,v(x,0)=uˆ0(x),

where fˆ=ff1,f1=VtνΔV+(V)(V)+(ζP[J]),uˆ0=bW|t=0 (since U[J](x,0)=0). Recall that the number J was chosen such that

Ut[J]νΔU[J]+(U[J])U[J]+P[J]=H(J)L2(0,T;L2(GH)).

(see [1]). Therefore, taking into account that W has compact support Ω3, we conclude

fˆL2(0,T;L2(Ω)),uˆ0W 1,2(Ω).

Moreover, using results obtained in [1] we get (see estimate (4.19) in [1])

(4.14) fˆL2(0,T;L2(Ω))2c(fL2(0,T;L2(Ω))2+0TaJ+12dt.),uˆ0W1,2(Ω)2c(bW1,2(Ω)2+0TaJ+12dt).

In the next subsections we construct the sequence of weak solutions vK to the Navier–Stokes equations in regular domains ΩK and prove the uniform (with respect to K) estimates for them. The solution of problem (4.13) is then found as a limit of {vK}.

4.2 Existence of the solution in ΩK

Consider in ΩK the following boundary value problem

(4.15) vKtνΔvK+(vK)vK+(V)vK+(vK)V+qK=fˆ,divvK=0,vK|ΩK=0,vK(x,0)=uˆK0(x),

where uˆK0W 1,2(ΩK),divuˆK0=0anduˆK0uˆ0W1,2(Ω)0 as K → +∞ (here we suppose that uˆK0 is extended by zero into Ω \ ΩK).

In this subsection we omit the subscript K in notation of the solution vK.

By a weak solution of problem (4.15) we mean the function v ∈ L2(0, T; H(ΩK)) with vt,2v L2(0,T;L2(ΩK)) satisfying the initial condition v(x,0)=uˆK0(x), and for all t ∈ [0, T] satisfying the integral identity

(4.16) 0TΩK(vtη+νvη((v+V))ηv(v)ηV)dxdt=0TΩKfˆηdxdt

for any test function ηL2(0,T;H(ΩK)) with ηtL2(0,T;L2(ΩK)).

Lemma 4.1

Let uˆ0W 1,2(Ω). There exists a sequence uˆK0W 1,2(ΩK) such that div uˆK0=0andlimKuˆK0 uˆ0W1,2(Ω)=0. Moreover there hods the estimate

(4.17) uˆK0W1,2(Ω)2c(bW1,2(Ω)2+0TaJ+12dt)

with the constant c independent of K.

Proof

Let χK(x2) be a smooth cut-off function such that χK(x2)=0forx2hK1,χK(x2)=1forx2hK2 and |χ(x2)|cφ(hK1). Consider the sequence of functions uˆK0=χK(x2)uˆ0+wˆK, where wˆK is a solution of the problem

divwˆK=χKuˆ0,xωK1,wˆK=0,xωK1.

Obviously, ωK1χKuˆ0dx=0. Therefore, by Lemma 2.4, there exists wˆKW 1,2(ωK1) satisfying the estimate

(4.18) wˆKL2(ωK1)cχKuˆ0L2(ωK1)cφ1uˆ0L2(ωK1)cuˆ0L2(ωK1)

with the constant c independent of K. By construction div uˆK0=0 and

uˆK0uˆ0W1,2(Ω)(1χK)uˆ0W1,2(ωK1)+wˆKW1,2(ωK1)cuˆ0W1,2(ωK1).

Since the right-hand side of this inequality tend to zero as K → ∞, we get limKuˆK0uˆ0W1,2(Ω)=0. Estimate (4.17) follows from (4.14).

Theorem 4.1

Suppose that f ∈ L2(0, T; Ω), b ∈ W1,2(Ω), the boundary value a has the finite norm 0TaJ+22dt and let there hold the compatibility conditions div b = 0, b|∂Ω = a(x, 0). There exists a positive constant κ0 such that if the flux F (t) satisfies the inequality 0T|||F|||J+22dtκ0, then problem (4.15) admits at least one weak solution v. The following estimates

(4.19) supt[0,T]v(,t)L2(ΩK)2+0Tv(,t)L2(ΩK)2dtcB1
(4.20) supt[0,T]v(,t)L2(ΩK)2ceB2(A0+A1+B1):=cB3;
(4.21) 0T2v(,t)L2(ΩK)2dtcB4
(4.22) 0Tvt(,t)L2(ΩK)2dtcB5

hold. The constants in estimates (4.19)(4.22) are independent of K and

A0=bW1,2(Ω)2+fL2(0,T;L2(Ω))2,A1=0TaJ+12dt,B1=(1+ec4A1A1)(A0+A1),B2=c9A1+c10B12,B3=eB2(A0+A1+B1),B4=A0+A1+A1B3+B3B12+B1,B5=A0+A1+(1+A1)B4+(1+B1B3)B1+B3A1.

The constants c4, c9 and c10 are defined in the proof of the theorem.

Proof

We follow the scheme of O.A. Ladyzhenskaya book [22] (see also [21, 24]) where the solvability of problem (4.15) is proved by the Galerkin method. Let {ψl}l=1 be an orthogonal basis in the space H (ΩK). Consider Galerkin approximations v(N)(x,t)=l=1Nγl(N)(t)ψl(x) of the solution v of problem (4.15) which are defined by the following system of ordinary differential equations (with respect to functions γl(N)(t),l=1,,N):

(4.23) ΩK(vt(N)ψl+νv(N)ψl((v(N)+V))ψlv(N)(v(N))ψlV)dx=ΩKfˆψldx,l=1,,N,γl(N)(0)=αl,l=1,,N,

where αl are the coefficients of the initial function uˆK0 in the basis {ψk}k=1,i.e.,uˆK0=l=1αlψl(x).

Multiplying (4.23) by γl(N)(t) and summing by l from 1 to N we obtain

12ddtΩK|v(N)|2dx+νΩK|v(N)|2dx=ΩK(v(N))v(N)Vdx+ΩKfˆv(N)dx.

Consider the integral

|ΩK(v(N))v(N)Vdx||Ω3(v(N))Vv(N)dx|+|ΩKΩ3(v(N))Vv(N)dx|=I1+I2.

For I1 we have the estimate

|I1|VL2(Ω3)v(N)L4(Ω3)2cVL2(Ω3)v(N)L2(Ω3)v(N)L2(Ω3)ν4ΩK|v(N)|2dx+cVL2(Ω3)2ΩK|v(N)|2dx.

Consider I2. By Poincaré inequality (2.4) and (4.6),

|ΩKΩ3(v(N))Vv(N)dx|ΩKΩ3|v(N)|2|U[J]|dxc(0T|||F|||J+22dt)1/2ΩKΩ3|v(N)(x)|2φ2(x2)dxc1(0T|||F|||J+22dt)1/2ΩK|v(N)|2dx.

Further,

|ΩKfˆv(N)dx|c(ε)(Ω3|fˆ|2dx+ΩKΩ3φ2(x2)|fˆ|2dx)+ε(ΩKΩ3|v(N)|2φ2(x2)dx+Ω3|v(N)|2dx)c(ε)ΩK|fˆ|2dx+c2εΩK|v(N)|2dx.

Thus, for F(t) such that (0T|||F|||J+22dt)1/2ν4c1andε=ν4c2 we obtain

(4.24) 12ddtΩK|v(N)|2dx+ν4ΩK|v(N)|2dxc3VL2(Ω3)2ΩK|v(N)|2dx+cΩK|fˆ|2dx.

By Gronwall’s inequality, (4.24) yields

(4.25) ΩK|v(N)(x,t)|2dxec30tV(,s)L2(Ω3)2dsΩK|v(N)(x,0)|2dx+cec30tV(,s)L2(Ω3)2ds0tec30τV(,s)L2(Ω3)2dsfˆ(,τ)L2(ΩK)2dτec30TV(,t)L2(Ω3)2dt(ΩK|uˆK0(x)|2dx+c0Tfˆ(,τ)L2(Ω)2dτ).

Inequalities (4.24), (4.25) imply

supt[0,T]v(N)(,t)L2(ΩK)2+ν0TΩK|v(N)(x,t)|2dxdtc(0Tfˆ(,t)L2(ΩK)2dt+Ω|uˆK0|2dx)+c0T(V(,t)L2(Ω3)2ΩKu(N)(x,t)|2dx)dtc(1+ec30TV(,t)L2(Ω3)2dt0TV(,t)L2(Ω3)2dt)Aˆ,

where

Aˆ=ΩK|uˆK0(x)|2dx+0Tfˆ(,t)L2(Ω)2dtc(bW1,2(Ω)2+fL2(0,T;L2(Ω))2)+c0TaJ+12dt:=c(A0+A1)

(see (4.14), (4.17)). By (4.12),

(4.26) 0TV(,t)W2,2(Ω3)2dtc0TaJ+12dt.

Therefore, using (4.17) we obtain

(4.27) supt[0,T]v(N)(,t)L2(ΩK)2+ν0TΩK|v(N)(x,t)|2dxdtc5(1+ec4A1A1)(A0+A1):=c5B1.

Estimate (4.27) guaranties that the Cauchy problem (4.23) admits a unique solution for each fixed N. Now we derive a number of a priori estimates for Galerkin approximations v(N). Estimate (4.27) is valid for Galerkin approximations constructed using an arbitrary orthogonal basis. In order to estimate the higher derivatives of v(N), as a basis we shall use the eigenfunctions of the Stokes operator.

Taking in (4.23) ψl = wl, where wl are eigenfunctions of the Stokes operator, i.e., Δ˜wl=λlwl, multiplying the obtained relations by λlγl(N)(t) and summing by l from 1 to N we obtain

ΩK(vt(N)Δ˜v(N)νΔv(N)Δ˜v(N)+((v(N)+V))v(N)Δ˜v(N)+(v(N))VΔ˜v(N))dx=ΩKfˆΔ˜vdx.

This is equivalent to (see the properties of the Stokes operator)

(4.28) 12ddtΩK|v(N)|2dx+ΩK|Δ˜v(N)|2dx=ΩK((v(N)+V))v(N)Δ˜v(N)dxΩK(v(N))VΔ˜v(N)dx+ΩKfˆΔ˜v(N)dx=i=13Ji.

Let us estimate the right-hand side of (4.28). By Young’s inequality,

(4.29) |J3|εΩK|Δ˜v(N)|2dx+cεΩK|fˆ|2dx.

Further, by (4.6), (2.4) and (3.23),

(4.30) |J2|{VL4(Ω3)v(N)L4(Ω3)+(ΩKΩ3|v(N)|2|U[J]|2dx)1/2}×(ΩK|Δ˜v(N)|2dx)1/2εΩK|Δ˜v(N)|2dx+cεVW2,2(Ω3)2v(N)L2(ΩK)2+(ΩKΩ3|v(N)|2|U(J)|2dx)1/2(ΩK|Δ˜v(N)|2dx)1/2εΩK|Δ˜v(N)|2dx+cεVW2,2(Ω3)2v(N)L2(ΩK)2+(0T|||F|||J+22dt)1/2(ΩKΩ3|v(N)|2φ4(x2)dx)1/2(ΩK|Δ˜v(N)|2dx)1/2εΩK|Δ˜v(N)|2dx+cεVW2,2(Ω3)2v(N)L2(ΩK)2+(0T|||F|||J+22dt)1/2(ΩKΩ3|v(N)|2φ2(x2)dx)1/2(ΩK|Δ˜v(N)|2dx)1/2εΩK|Δ˜v(N)|2dx+cεVW2,2(Ω3)2v(N)L2(ΩK)2+(0T|||F|||J+22dt)1/2ΩK|Δ˜v(N)|2dx.

Similarly (J1 = J11 + J12) we obtain the estimates

(4.31) |J11|=|ΩK(V)v(N)Δ˜v(N)dx|VL4(Ω3)v(N)L4(Ω3)(ΩK|Δ˜v(N)|2dx)1/2+c(0T|||F|||J+22dt)1/2(ΩKΩ3|v(N)|2φ2(x2)dx)1/2(ΩK|Δ˜v(N)|2dx)1/2cVW1,2(Ω3)v(N)W2,2(Ω3)(ΩK|Δ˜v(N)|2dx)1/2+c(0T|||F|||J+22dt)1/2(ΩKΩ3|v(N)|2φ2(x2)dx)1/2(ΩK|Δ˜v(N)|2dx)1/2c(ε+(0T|||F|||J+22dt)1/2)|ΩK|Δ˜v(N)|2dx+cεVW2,2(Ω3)2v(N)L2(ΩK)2,

and

(4.32) |J12|=|ΩK(v(N))v(N)Δ˜v(N)dx|εΩK|Δ˜v(N)|2dx+cε(v(N)L4(Ω3)2v(N)L4(Ω3)2+l=1K1v(N)L4(ωl)2v(N)L4(ωl)2+v(N)L4(ωˆK)2v(N)L4(ωˆK)2).

Applying inequalities (2.6) and (2.5) we get

(4.33) v(N)L4(Ω3)2v(N)L4(Ω3)2cv(N)L2(Ω3)v(N)L2(Ω3)2(v(N)L2(Ω3)2+2v(N)L2(Ω3)2)1/2cε(v(N)L2(Ω3)2+v(N)L2(Ω3)2v(N)L2(Ω3)4)+ε2v(N)L2(Ω3)2;
(4.34) v(N)L4(ωl)2v(N)L4(ωl)2cφ1(hl1)v(N)L2(ωl)v(N)L2(ωl)2(v(N)L2(ωl)2+φ2(hl1)2v(N)L2(ωl)2)1/2cεv(N)L2(ωl)2v(N)L2(ωl)4+ε(2v(N)L2(ωl)2+φ2(hl1)v(N)2);
(4.35) v(N)L4(ωˆK)2v(N)L4(ωˆK)2cεv(N)L2(ωˆK)2v(N)L2(ωˆK)4+ε(2v(N)L2(ωˆK)2+φ2(hK1)v(N)L2(ωˆK)2).

Thus, (4.32)–(4.35) and (3.23), (3.24) imply

(4.36) |J12|cε(ΩK|Δ˜v(N)|2dx+ΩK|2v(N)|2dx+ΩKΩ3|v(N)|2φ2(x2)dx)+c(v(N)L2(Ω3)2+v(N)L2(Ω3)2v(N)L2(Ω3)4)+cl=1K1v(N)L2(ωl)2v(N)L2(ωl)4+cv(N)L2(ωˆK)2v(N)L2(ωˆK)4cεΩK|Δ˜v(N)|2dx+cv(N)L2(ΩK)2+v(N)L2(ΩK)2v(N)L2(ΩK)2×(Ω3|v(N)|2dx+l=1K1ωl|v(N)|2dx+ωˆK|v(N)|2dx)cεΩK|Δ˜v(N)|2dx+cv(N)L2(ΩK)2v(N)L2(ΩK)4+cv(N)L2(ΩK)2.

Substituting (4.29)–(4.31), (4.36) into (4.28) yields

(4.37) 12ddtΩK|v(N)|2dx+ΩK|Δ˜v(N)|2dxc6(ε+(0T|||F|||J+22dt)1/2)ΩK|Δ˜u(N)|2dx+cVW2,2(Ω3)2v(N)L2(ΩK)2+cv(N)L2(ΩK)2v(N)L2(ΩK)4+cv(N)L2(ΩK)2+cΩK|fˆ|2dx.

Taking in (4.37) ε=14c6 and assuming that (0T|||F|||J+22dt)1/214c6 we derive

(4.38) ddtΩK|v(N)|2dx+ΩK|Δ˜v(N)|2dxc7VW2,2(Ω3)2v(N)L2(ΩK)2+c8(v(N)L2(ΩK)2v(N)L2(ΩK)4+v(N)L2(ΩK)2)+c9ΩK|fˆ|2dx

Denoting

Y(t)=ΩK|v(N)(x,t)|2dx,A(t)=c9ΩK|fˆ|2dx+c8v(N)L2(ΩK)2,B(t)=c7VW2,2(Ω3)2+c8v(N)L2(ΩK)2v(N)L2(ΩK)2,

we rewrite (4.38) as

Y(t)B(t)Y(t)+A(t).

By Gronwall’s lemma,

(4.39) Y(t)e0tB(τ)dτ(Y(0)+0te0sB(τ)dτA(s)ds)
e0TB(τ)dτ(Y(0)+0TA(s)ds).

Estimates (4.26), (4.27) yield

0TB(t)dt=c70TVW2,2(Ω3)2dt+c80Tv(N)L2(ΩK)2v(N)L2(ΩK)2dtc0TVW2,2(Ω3)2dt+csupt[0,T]v(N)(,t)L2(ΩK)20TΩK|v(N)|2dxdtc9A1+c10B12:=B2.

Therefore, from (4.39) and (4.27) we have

(4.40) supt(0,T)]v(N)(,t)L2(ΩK)2ceB2(uˆk0L2(ΩK)2+v(N)L2(0,T;L2(ΩK))2+fˆL2(0,T;L2(Ω))2)c11eB2(A0+A1+B1)=c11B3.

Substituting (4.40) into (4.38) and integrating over t yield

(4.41) 0TΩK|Δ˜v(N)|2dxdtΩK|uˆ0k|2dx+c0TΩ|fˆ|2dxdt+c0TVW2,2(Ω3)2v(N)L2(ΩK)2dt+0Tv(N)L2(ΩK)2dt+c0Tv(N)L2(ΩK)2v(N)L2(ΩK)4dtc(A0+A1)+csupt(0,T)v(N)L2(ΩK)20TVW2,2(Ω3)2dt+csupt(0,T)v(N)L2(ΩK)2supt(0,T)v(N)L2(ΩK)20Tv(N)L2(ΩK)2dt+c0Tv(N)L2(ΩK)dt2c12[A0+A1+B3A1+B3B12+B1]:=c12B4.

Let us estimate the norm of ut(N). Multiplying (4.23) by ddtγl(N)(t) and summing by l from l = 1 to l = N we obtain

(4.42) ΩK|vt(N)|2dx+ν2ddtΩK|v(N)|2dx=ΩK((v(N)+V))v(N)vt(N)dxΩK(v(N))Vvt(N)dx+ΩKfvt(N)dx12ΩK|vt(N)|2dx+cΩ|fˆ|2dx+cΩK(|v(N)|2+|V|2)|v(N)|2dx+cΩK|v(N)|2|V|2dx.

Let us estimate the last two terms in the right-hand side of (4.42). By the same argument as before,

ΩK|v(N)|2|V|2dxc((Ω3|v(N)|4dx)1/2(Ω3|V|4dx)1/2+0T|||F|||J+22dtΩKΩ3|v(N)|2φ4(x2)dx)c(VW2,2(Ω3)2ΩK|v(N)|2dx+0T|||F|||J+22dtΩK|Δ˜v(N)|2dx);
ΩK|V|2|v(N)|2dxc(supxΩ3|V(x,t)|2Ω3|v(N)|2dx+0T|||F|||J+22dtΩKΩ3|v(N)|2φ2(x2)dx)c(VW2,2(Ω3)2ΩK|v(N)|2dx+0T|||F|||J+22dtΩK|Δ˜v(N)|2dx).

To estimate the integral ΩK|v(N)|2|v(N)|2dx, we apply inequalities (4.33)-(4.34) and argue as in the proof of (4.36). In virtue of (2.5) applied to v(N), we obtain

ΩK|v(N)|2|v(N)|2dxcΩK|Δ˜v(N)|2dx+cv(N)L2(ΩK)2+cv(N)L2(ΩK)2v(N)L2(ΩK)4cΩK|Δ˜v(N)|2dx+cv(N)L2(ΩK)2+csupt[0,T]v(N)(,t)L2(ΩK)2supt[0,T]v(N)(,t)L2(ΩK)2)v(N)|L2(ΩK)2cΩK|Δ˜v(N)|2dx+c(1+B1B3)ΩK|v(N)|2dx.

Substituting the obtained inequalities into (4.42) yields

12ΩK|vt(N)|2dx+ν2ddtΩK|v(N)|2dxcΩ|fˆ|2dx+c(1+0T|||F|||J+22dt)ΩK|Δ˜v(N)|2dx+c(1+B1B3)ΩK|v(N)|2dx+cVW2,2(Ω3)2ΩK|v(N)|2dx.

Integrating this inequality over [0, T] and using estimates (4.27), (4.41) we derive

(4.43) 0TΩK|vt(N)|2dxdtc(ΩK|uˆ0k|2dx+cΩ|fˆ|2dx)+c(1+supt(0,T)0T|||F|||J+22dt)0TΩK|Δ˜v(N)|2dxdt+c(1+B1B3)0TΩK|v(N)|2dxdt+supt(0,T)v(N)(,t)L2(ΩK)20TVW2,2(Ω3)2dtc13[A0+A1+(1+A1)B4+(1+B1B3)B1+B3A1]:=c13B5.

Estimates (4.27), (4.41) and (4.43) ensure that there exist a subsequence {v(Nl)} such that {v(Nl)},{v(Nl)}, {2v(Nl)},{vt(Nl)} are weakly convergent in L2(0, T; L2(ΩK)). The limit v of this subsequence satisfies the integral identity (4.16) and thus, is a weak solution of (4.15). This part of the proof is standard (see [21, 22, 24]) and we omit the details. Remind that in this section we have omitted the subscript K for notation of the solution and Galerkin approximations, so v(N)=vK(N)andv=vK.

4.3 Existence and uniqueness of the solution to problem (4.13)

By a weak solution of problem (4.13) in the cusp domain Ω we mean the function v ∈ L2(0, T; H(Ω)) with vt ,2v ∈ L2(0, T; L2(Ω)) satisfying the initial condition v(x,0)=uˆ0(x), and for all t ∈ [0, T] satisfying the integral identity

(4.44) 0tΩ(vtη+νvη((v+V))ηv(v)ηV)dxdt=0tΩfˆηdxdt

for any test function η ∈ L2(0, T; H(Ω)), ηt ∈ L2(0, T; L2(Ω)) having compact support in Ωˉ{O} (i.e., the support of η is separated from the cusp point O)

Theorem 4.2

Assume that the conditions of Theorem 4.1 are valid. There exists a number κ1 > 0 such that if 0T|||F|||J+22dtκ1, then there exists a unique weak solution v of problem (4.13).

Proof

Let 0T|||F|||J+22dtκ0,

where κ0 is a number from Theorem 4.1. Then, due to estimates (4.19)–(4.22) for

vK,we can extract a subsequence {vKl} such that {vKl},{vKl},{2vKl}and{vKl,t} are weakly convergent in L2(0, T; L2(Ω)) as Kl Taking in (4.16) a test function η with the compact support and passing Kl to infinity we obtain for the limit v integral identity (4.44). Moreover, obviously v(x,0)=uˆ0(x) and thus, v is a weak solution of problem (4.13). For v remain valid estimates (4.19)–(4.22). Let us prove the uniqueness. Let u1 and u2 be two solutions of problem (1.1) having the same representations: u1=V+v1andu2=V+v2, where v1,v2 are solutions of problem (4.13). The difference u1u2=v1v2:= vW 1,2(Ω) satisfies zero initial condition v(x, 0) = 0 and the integral identity

(4.45) 0tΩ(vtη+νvη)dxdt0tΩ(V)ηvdxdt0tΩ(v)ηVdxdt0tΩ(v)ηv1dxdt=0.

Let us take in (4.45) η = χK(x2)v + wK, where χK is defined in the proof of Lemma 4.1 and wK is a solution of the problem

divwK=χKv,xωK1,wK=0,xωK1,

satisfying the estimate

(4.46) wKL2(ωK1)cχKvL2(ωK1)cφ1vL2(ωK1)cvL2(ωK1).

This gives

(4.47) 120tddtΩχK|v|2dxdt+ν0tΩχK|v|2dxdt=0tΩvtwKdxdtν0tΩvχKvdxdtν0tΩvwKdxdt+0tΩ(V)(χKv+wK)vdxdt+0tΩ(v)(χKv+wK)Vdxdt+0tΩ(v)(χKv+wK)v1dxdt:=i=16Ji.

Let us estimate the right-hand side of (4.47). Using the properties of χK and wK we obtain

(4.48) |J1|0tvtL2(ωK1)wKL2(ωK1)dtc0t(vtL2(ωK1)2+vL2(ωK1)2)dt;|J2|+|J3|0tvL2(ωK1)φ1vL2(ωK1)dt+0tvL2(ωK1)wKL2(ωK1)dtc0tvL2(ωK1)2dt;

Integrating by parts yields

0tΩ(V)(χKv)vdxdt=120tωK1(V)χK|v|2dxdt.

Therefore, by (4.6), (4.46),

(4.49) |J4|(0T|||F|||J+22dt)1/20t(φ1vL2(ωK1)2+φ1vL2(ωK1)wKL2(ωK1))dt(0T|||F|||J+22dt)1/20tvL2(ωK1)2dt.

Moreover, using (4.12), (4.6), (2.5) we get

(4.50) |J5||0tΩ(v)V(χKv+wK)dxdt||0tΩ3(v)Vvdxdt|+c(0T|||F|||J+22dt)1/20tΩΩ3χKφ2|v|2dxdt+|0tωK1(v)VwKdxdt|c(0T|||F|||J+22dt)1/20tΩΩ3χK|v|2dxdt+c0t((0T|||F|||J+22dt)1/2φ1vL2(ωK1)φ1wKL4(ωK1)+VL2(Ω3)vL4(Ω3)2)dtc(0T|||F|||J+22dt)J+11/20tΩΩ3χK|v|2dxdt+c0tVL2(Ω3)vL2(Ω3)vL2(Ω3)dt+c0tvL2(ωK1)2dtc((0T|||F|||J+22dt)1/2+ε)0tΩχK|v|2dxdt+cε0tvL2(ωK1)2dt+c0tVL2(Ω3)2χKvL2(Ω)2dt

and

(4.51) |J6||0tΩχK(v)vv1dxdt|+|0tωK1(v)χK(vv1)dxdt|+|0tωK1(v)wKv1dxdt|ε0tΩχK|v|2dxdt+cε0tχKvL4(Ω)2v1L4(Ω)2dt+c0tφ1vL2(ωK1)vL4(ωK1)v1L4(ωK1)dt+c0twKL2(ωK1)vL4(ωK1)v1L4(ωK1)dtε0tΩχK|v|2dxdt+c0tχKvL2(Ω)(χKv)L2(Ω)v1L2(Ω)2dt+c0tv1L2(ωK1)vL2(ωK1)2dtcε0tΩχK|v|2dxdt+cε0tv1L2(Ω)4χKvL2(Ω)2dt+c0tv1L2(ωK1)vL2(ωK1)2dt.

Substituting estimates (4.48)–(4.51) into (4.47) we obtain

(4.52) 120tddtΩχK|v|2dxdt+ν0tΩχK|v|2dxdtc14((0T|||F|||J+22dt)1/2+ε)0tΩχK|v|2dxdt+c0t(VL2(Ω3)2+v1L2(Ω)4)χKvL2(Ω)2dt+c0t[vtL2(ωK1)+v1L2(ωK1)vL2(ωK1)2]dt.

Taking ε=ν4c14 in (4.52) and assuming that c14(0T|||F|||J+22dt)1/2ν4 we get

(4.53) ΩχK|v(x,t)|2dxc0t(VL2(Ω3)2+v1L2(Ω)4)ΩχK|v(x,t)|2dxdt+c0t[vtL2(ωK1)+v1L2(ωK1)vL2(ωK1)2]dt.

Introduce the notation:

Z(t)=ΩχK|v(x,t)|2dx,β(t)=(V(,t)L2(Ω3)2+v1(,t)L2(Ω)4),α(t)=0t[vtL2(ωK1)+v1L2(ωK1)vL2(ωK1)2]dt.

Then (4.53) can be written as

Z(t)c15α(t)+c160tβ(τ)Z(τ)dτ.

By Gronwall lemma, the last inequality yields

(4.54) Z(t)c15α(t)exp(c160tβ(τ)dτ).

Estimates (4.19), (4.20) for the solution v1 and estimate (4.26) imply (see Theorem 4.1)

0Tv1L2(Ω)2dtcB1,supt(0,T)v1L2(Ω)2cB3,0TVL2(Ω3)2dtcA1.

This together with (4.54) yields

(4.55) 0TΩK2|v(x,t)|2dxdt0TZ(t)dtc150Tα(t)dtexp(c160Tβ(t)dt)cec(A1+B1B3)0TωK1(|vt|2+|v|2)dxdt.

Obviously, the right-hand side of (4.55) vanishes as K → ∞. Therefore, passing in (4.55) to the limit, we obtain 0TΩ|v(x,t)|2dxdt=0, and hence v(x,t)=v1(x,t)v2(x,t)=0.

Remark 4.1

The solution u of problem (1.1) considered in Theorem 4.2 has the representation u = V + v, where V is a singular part coinciding near the cusp point with the formal asymptotic decomposition of the solution, and v is a regular part having finite energy norm. Theorem 4.2 states only the uniqueness of regular part v. We do not prove the uniqueness of general singular solution of problem (1.1) having a source or sink in the cusp point.

Remark 4.2

The "smallness" assumption of Theorems 4.1 and 4.2 concerns only the smallness of fluxes F(t)=Ωa(x,t)n(x)dS, i.e. of the magnitude of 0T|||F|||J+2dt. We do not suppose that the norms of the boundary value a, initial condition b or the right-hand side f are small.

Acknowledgment

The research of authors was funded by the grant No. S-MIP-17-68 from the Research Council of Lithuania.

  1. Conflict of interest: The authors declare that they have no conflict of interest.

References

[1] K. Pileckas and A. Raciene, Non-stationary Navier–Stokes equations in 2D power cusp domain. I. Construction of the formal asymptotic decomposition, Advances in nonlinear analysis (2021), no 10, 982–1010.10.1515/anona-2020-0164Search in Google Scholar

[2] A. Eismontaite and K. Pileckas, On singular solutions of time-periodic and steady Stokes problems in a power cusp domain, Applicable Analysis, 97 (2018), no. 3, 415–437.10.1080/00036811.2016.1269321Search in Google Scholar

[3] A. Eismontaite and K. Pileckas, On singular solutions of the initial boundary value problem for the Stokes system in a power cusp domain, Applicable Analysis, 98 (2019), no. 13, 2400–2422.10.1080/00036811.2018.1460815Search in Google Scholar

[4] V. Kozlov and J. Rossmann, On the nonstationary Stokes system in a cone, J. Diff. Equations, 260 (2016), no. 12, 8277–8315.10.1016/j.jde.2016.02.024Search in Google Scholar

[5] G. Panasenko and K. Pileckas, Asymptotic analysis of the non-steady Navier-Stokes equations in a tube structure. I. The case without boundary-layer-in-time, Nonlinear Analysis: Theory, Methods & Applications, 122 (2015), 125–168.10.1016/j.na.2015.03.008Search in Google Scholar

[6] G. Panasenko and K. Pileckas, Asymptotic analysis of the non-steady Navier-Stokes equations in a tube structure. II. General case, Nonlinear Analysis: Theory, Methods & Applications, 125 (2015), 582–607.10.1016/j.na.2015.05.018Search in Google Scholar

[7] K. Kaulakyte and N. Kloviene and K. Pileckas, Nonhomogeneous boundary value problem for the stationary Navier-Stokes equations in a domain with a cusp, Z. Angew. Math. Phys., (2019), 70:36.10.1007/s00033-019-1075-5Search in Google Scholar

[8] H. Kim and H. Kozono, A removable isolated singularity theorem for the stationary Navier-Stokes equations, J. Diff. Equations, 220 (2006), 68–84.10.1016/j.jde.2005.02.002Search in Google Scholar

[9] V.V. Pukhnachev, Singular solutions of Navier-Stokes equations, Advances in Mathematical Analysis of PDEs, Proc. St. Petersburg Math. Soc. XV, AMS Transl. Series 2., 232 (2014), 193-218.10.1090/trans2/232/11Search in Google Scholar

[10] A. Russo and A. Tartaglione, On the existence of singular solutions of the stationary Navier-Stokes problem, Lithuanian Math. J., 53 (2013), no. 4, 423-437.10.1007/s10986-013-9219-3Search in Google Scholar

[11] M.B. Korobkov and K. Pileckas and V.V. Pukhnachev and R. Russo, The Flux Problem for the Navier-Stokes Equations, Us-pech Mat. Nauk, 69 (2014), no. 6, 115–176. Transl.: Russian Math. Surveys, 69 (2015), no. 6, 1065–1122.10.1070/RM2014v069n06ABEH004928Search in Google Scholar

[12] S.A. Nazarov and K. Pileckas, Asymptotics of Solutions to Stokes and Navier-Stokes equations in domains with paraboloidal outlets to infinity, Rend. Sem. Math. Univ. Padova, 99 (1998), 1–43.Search in Google Scholar

[13] K. Pileckas, Weighted Lq–solvability of the steady Stokes system in domains with noncompact boundaries, Math. Models and Methods in Applied Sciences, 6 (1996), no. 1, 97–136.10.1142/S0218202596000079Search in Google Scholar

[14] K. Pileckas, Classical solvability and uniform estimates for the steady Stokes system in domains with noncompact boundaries, Math. Models and Methods in Applied Sciences, 6 (1996), no. 2, 149–167.10.1142/S0218202596000523Search in Google Scholar

[15] K. Pileckas, Strong solutions of the steady nonlinear Navier–Stokes system in domains with exits to infinity, Rend. Semin. Mat. Univ. Padova, 97 (1996), 72–107.Search in Google Scholar

[16] R.A. Adams and J. Fournier, Sobolev Spaces, Second edition, Academic Press (Elsevier), (2003).Search in Google Scholar

[17] O.A. Ladyzhenskaya and V.A. Solonnikov, On some problems of vector analysis and generalized formulations of boundary value problems for the Navier–Stokes equations, Zap. Nauch. Sem. LOMI, 59 (1976), 8–116; Transl.: J. Sov. Math., 10 (1978), no. 2, 257–285.Search in Google Scholar

[18] O.A. Ladyzhenskaya and N.N. Ural’tseva, Linear and Quasilinear Elliptic Equations, Mathematics in Science and Engineering, 46, New York and London: Academic Press, (1968).Search in Google Scholar

[19] S. Agmon and A. Douglis and L. Nirenberg, Estimates near the boundary for solutions of elliptic partial differential equations satisfying general boundary conditions II, Commun. Pure Appl. Math., 17 (1964), 35–92.10.1002/cpa.3160170104Search in Google Scholar

[20] V.A. Solonnikov, On the boundary value problems for systems elliptic in the sense of A. Douglis, L. Nirenberg. I., Izv. Akad. Nauk SSSR, Ser. Mat., 28 (1964), 665–706; II. Trudy Mat. Inst. Steklov, 92 (1966), 233–297. Transl.: I. Amer. Math. Soc. Transl., 56 (1966), no. 2, 192–232; II, Proc. Steklov Inst. Math., 92 (1966), 269–333.Search in Google Scholar

[21] G. Seregin, Lecture notes on regularity theory for the Navier–Stokes equations, World Scientific, Singapore, (2015).10.1142/9314Search in Google Scholar

[22] O.A. Ladyzhenskaya, The mathematical theory of viscous incompressible flow, New York (NY), London, Paris: Gordon and Breach, (1969).Search in Google Scholar

[23] L.V. Kapitanskii and K. Pileckas, Certain problems of vector analysis, Zapiski Nauchn. Sem. LOMI, 138 (1984), 65–85; Transl.: J. Sov. Math., 32 (1986), no. 5, 469–483.Search in Google Scholar

[24] R. Temam, Navier–Stokes Equations, North-Holland, Amsterdam (1977).Search in Google Scholar

Received: 2020-06-04
Accepted: 2020-10-17
Published Online: 2021-02-28

© 2021 Konstantin Pileckas and Alicija Raciene, published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 25.4.2024 from https://www.degruyter.com/document/doi/10.1515/anona-2020-0165/html
Scroll to top button