Elsevier

Journal of Differential Equations

Volume 297, 5 October 2021, Pages 226-245
Journal of Differential Equations

Asymptotic stability of homogeneous solutions of incompressible stationary Navier-Stokes equations

https://doi.org/10.1016/j.jde.2021.06.033Get rights and content

Abstract

It was proved by Karch and Pilarczyk that Landau solutions are asymptotically stable under any L2-perturbation. In our earlier work with L. Li, we have classified all (1)-homogeneous axisymmetric no-swirl solutions of incompressible stationary Navier-Stokes equations in three dimension which are smooth on the unit sphere minus the south and north poles. In this paper, we study the asymptotic stability of the least singular solutions among these solutions other than Landau solutions, and prove that such solutions are asymptotically stable under any L2-perturbation.

Introduction

Consider the incompressible stationary Navier-Stokes Equations in R3,{Δu+(u)u+p=0,divu=0. These equations are invariant under the scaling u(x)λu(λx) and p(x)λ2p(λx), λ>0 and it is natural to study solutions which are invariant under this scaling. These solutions are referred to as (1)-homogeneous solutions (although p is (2)-homogeneous).

Let x=(x1,x2,x3) be Euclidean coordinates and e1=(1,0,0),e2=(0,1,0),e3=(0,0,1) be the corresponding unit normal vectors. Denote x=(x1,x2). Let (r,θ,ϕ) be the spherical coordinates, where r is the radial distance from the origin, θ is the angle between the radius vector and the positive x3-axis, and ϕ is the meridian angle about the x3-axis. A vector field u can be written asu=urer+uθeθ+uϕeϕ, whereer=(sinθcosϕsinθsinϕcosθ),eθ=(cosθcosϕcosθsinϕsinθ),eϕ=(sinϕcosϕ0). A vector field u is called axisymmetric if ur, uθ and uϕ are independent of ϕ, and is called no-swirl if uϕ=0.

In 1944, L.D. Landau [7] discovered a 3-parameter family of explicit (1)-homogeneous solutions of the stationary NSE in C(R3{0}). These solutions, now called Landau solutions, are axisymmetric with no-swirl and have exactly one singularity at the origin. Tian and Xin proved in [23] that all (1)-homogeneous, axisymmetric nonzero solutions of (1) in C(R3{0}) are Landau solutions. Šverák proved in [21] that all (-1)-homogeneous nonzero solutions of (1) in C(R3{0}) are Landau solutions. There have also been works on (1)-homogeneous solutions of (1), see [2], [14], [15], [16], [17], [19], [20], [24], [25]. In [10], [11], [12], the (1)-homogeneous axisymmetric solutions of (1) in C(R3{(x1,x2)=0}) with a possible singular ray {(x1,x2)=0} was studied, where such solutions with no-swirl were classified in [10] and [11], and existence of such solutions with nonzero swirl was proved in [10] and [12].

There has been much work in literature on the existence of weak solutions and L2-decay of weak solutions of the evolutionary Navier-Stokes equations, see e.g. [1], [3], [6], [8], [9], [13], [18], [22] and the references therein. Such L2-decay of weak solutions can be viewed as the asymptotically stability of the zero stationary solution of (1). The asymptotic stability problem has been studied for other nonzero stationary solutions of (1) with some possible singularities in R3. Karch and Pilarczyk proved in [4] that small Landau solutions are asymptotically stable under L2-perturbations. The L2 asymptotic stability of other solutions with singularities is also studied in [5]. With special (1)-homogeneous solutions which are different from Landau solutions obtained in [10], [11], [12], it is worth to explore the asymptotic stability or instability of these solutions. In this paper, we start this study for a family of solutions which are the simplest and least singular solutions among the solutions found in [10], [11], [12].

Denote U=ursinθ and y=cosθ. By the divergence free property of u we have ur=1rUθ. For (-1)-homogeneous axisymmetric no-swirl solutions, (1) can be reduced to(1y2)Uθ+2yUθ+12Uθ2=c1(1y)+c2(1+y)+c3(1y2). For c11 and c21, letc¯3(c1,c2):=12(1+c1+1+c2)(1+c1+1+c2+2), where c1,c2,c3 are real numbers. Denote c=(c1,c2,c3) andJ:={cR3|c11,c21,c3c¯3(c1,c2)}. In [11], it was proved that there exist γ,γ+C0(J,R), satisfying γ(c)<γ+(c) if c3>c¯3(c1,c2), and γ(c)=γ+(c) if c3=c¯3(c1,c2), such that equation (2) has a unique solution Uθc,γ in C(1,1)C0[1,1] satisfying Uθc,γ(0)=γ for every c in J and γ(c)γγ+(c). In particular, γ+(0)>0 and γ(0)<0. Moreover, letuc,γurc,γer+uθc,γeθ=(Uθc,γ)er+Uθc,γsinθeθ,pc,γ=1r2(urc,γ12(uθc,γ)2)=1r2((Uθc,γ)(Uθc,γ)22sin2θ). {(uc,γ,pc,γ)|cJ,γ(c)γγ+(c)} are all (1)-homogeneous axisymmetric no-swirl solutions of (1) in C(R3{(x1,x2)=0}). It was also obtained in [11] thatUθc,γ(1)={2+21+c1,when γ=γ+(c),221+c1,otherwise,Uθc,γ(1)={221+c2,when γ=γ(c),2+21+c2,otherwise.

As mentioned earlier, we would like to study the asymptotic stability or instability of the (1)-homogeneous axisymmetric stationary solutions found in [10], [11], [12]. Different from Landau solutions, these solutions are singular at the north pole N and/or south pole S. These solutions u satisfy either 0<limsup|x|=1,x0|x||x||u(x)|< or limsup|x|=1,x0|x|2|u(x)|>0, while Landau solutions satisfy sup|x|=1|x|2|u|<. In this paper, we study the stability of (1)-homogeneous axisymmetric no-swirl solutions satisfying 0<limsupx0|x||x||u(x)|<. These solutions are the family {(uc,γ,pc,γ)|(c,γ)M}, whereM:={(c,γ)|c1=c2=0,c3>4,γ(c)<γ<γ+(c)}.

For any (c,γ)M, Uθc,γ satisfies{(1y2)(Uθc,γ)+2yUθc,γ+12(Uθc,γ)2=c3(1y2),1<y<1,Uθ(0)=γ.

Proposition 1.1

Let (c,γ)M, then (uc,γ(x),pc,γ(x)) satisfies{Δuc,γ+uc,γuc,γ+pc,γ=(4πc3ln|x3|x3δ(0,0,x3)bc,γδ0)e3,xR3,divuc,γ=0,xR3, wherebc,γ=11(y|Uθ|22y21y2Uθy1y2Uθ2)dy. Equations (6) and (7) are understood in the following distribution sense: for any φCc(R3), j=1,2,3,R3(ujφuiujxiφpxjφ)=[4πc3ln|x3|x3φ(0,0,x3)dx3bc,γφ(0)]δj3e3, andR3uc,γφ=0.

We now study the stability of the family of solutions {uc,γ|(c,γ)M}. Let H˙1(R3) denote the closure of Cc(R3,R3) under the norm uL2(R3), and for 1p<,Lσp(R3)={uLp(R3)|divu=0},H˙σ1(R3)={uH˙1(R3)|divu=0}, anduLσp(R3):=uLp(R3),uH˙σ1(R3)=uL2(R3).

For a given solution (uc,γ,pc,γ) of (1), let u=u(x,t) denote a solution of{utΔu+(u)u+p=(4πc3ln|x3|x3δ(0,0,x3)bc,γδ0)e3,(x,t)R3×(0,),divu=0,(x,t)R3×(0,),u(x,0)=uc,γ+w0, where w0Lσ2(R3) and bc,γ is given by (7). Then w(x,t)=u(x,t)uc,γ and π(x)=p(x)pc,γ(x) satisfy the initial value problem{wtΔw+(w)w+(w)uc,γ+(uc,γ)w+π=0,(x,t)R3×(0,),divw=0,(x,t)R3×(0,),w(x,0)=w0(x). We study the existence and asymptotic behavior of global-in-time weak solutions of (11). Let the energy spaceX:=L([0,),Lσ2(R3))L2([0,),H˙σ1), and for w in XwX:=wL([0,),Lσ2(R3))+wL2([0,),H˙σ1). Let (,) denote the L2-inner product, i.e. (f,g)=R3fgdx. A vector wX is a weak solution of (11) if for any 0st< and φC([0,),Hσ1(R3)C1([0,),Lσ2(R3)),(w(t),φ(t))+st[(w,φ)+(ww,φ)+(wuc,γ,φ)+(uc,γw,φ)]dτ=(w(s),φ(s))+st(w,φτ)dτ.

Theorem 1.1

There exists some μ0>0, such that for any c=(0,0,c3), |(c,γ)|<μ0, w0Lσ2(R3), there exists a weak solution w of (11) in the energy space X. Moreover, w is weakly continuous from [0,) to Lσ2(R3), and satisfies thatw(t)22+stw(τ)22dτw(s)22 for almost all s0, including s=0 and all ts.

Recall that γ+(0)>0 and γ(0)<0. So there is some μ0, such that {(c,γ)|c1=c2=0,|(c3,γ)|μ0}M. We also have

Theorem 1.2

There exists some μ0>0, such that for any c=(0,0,c3), |(c,γ)|<μ0 and weak solution wX of (11) satisfying (12),limtw(t)2=0. Moreover, if w0Lp(R3)Lσ2(R3) for some 65<p<2, then there exists some constant C>0, depending only on (c,γ), n,p and w0p, such that w(t)2Ct32(1p12), for all t>0.

Theorem 1.1 and Theorem 1.2 can be established using the same arguments as [4], as long as the special stationary solutions uc,γ satisfy the following condition|R3(vuc,γ)wdx|KwL2vL2, for some constant K small enough, for any divergence free v,wCc(R3). In [4], (13) is proved by Hardy's inequality when uc,γ is replaced by small Landau solutions. In this paper, we analyze the solutions uc,γ where (c,γ)M, and obtain |uc,γ|C(|c|+|γ|)/(|x||x|). So (13) is true if we haveR3|v|2|x||x|dxKvL22, for any vCc(R3). Notice (14) cannot be proved by the classical Hardy's inequality. In Section 4, we prove the following extended Hardy-type inequality, which includes (14).

Theorem 1.3

Let n2, 1p<n, uCc1(Rn), αp>1n, (α+β)p>n, then there exists some constant C, depending on p, α and β, such that|x|β|x|αuLp(Rn)C|x|β+αα|x|α+1uLp(Rn), for all αα. Moreover, for any α>α and any C>0, (15) fails in general.

Estimate (14) is the special case of (15) with p=2, α=α=β=12. Then we also have (13). Given (13), Theorem 1.1 and Theorem 1.2 can be proved by the same arguments used in [4], see also [5]. So in this paper we will only prove Theorem 1.3 and (13).

Remark 1.1

In [5], Karch, Pilarczyk and Schonbek proved the asymptotic stability of a class of general time-dependent solutions u of (10) using Fourier analysis, where (13) with uc,γ replaced by u is an essential assumption. A list of spaces were given in [5] where (13) is true if uc,γ is in one of those spaces. But the solutions uc,γ we discuss here are not in those spaces.

We will analyze in Section 2 the singular behaviors of uc,γ, (c,γ)M. In Section 3 we study the force of uc,γ, (c,γ)M. Theorem 1.3 will be proved in Section 4. Then as stated above, Theorem 1.1 and Theorem 1.2 follow with the same arguments as in [4].

Acknowledgment. We thank Vladimír Šverák for bringing to our attention the work [4] of Karch and Pilarczyk.

Section snippets

Estimate of the special solutions uc,γ

Lemma 2.1

Let K be a compact subset of M. Then there exists some positive constant C, depending only on K, such that for any (c,γ) in K and 1y1,Uθc,γ(y)=c32sgn(y)(1y2)ln(1y2)+O(1)(|c|+|γ|)(1y2),(Uθc,γ)(y)=c3ln(1y2)+O(1)(|c|+|γ|), and(Uθc,γ)(y)=2c3y1y2+O(1)(|c|+|γ|)(|ln(1y2)|2), where O(1) denotes some quantity satisfying |O(1)|C for some positive constant C depending only on K.

Proof

For convenience, let C be a constant depending only on K, O(1) be a function satisfying |O(1)|C for all 1y1

Force of uc,γ, (c,γ)M

In this section, we study the force of the special solutions uc,γ and prove Proposition 1.1, where (c,γ) in M and M is the set defined by (4). Recall that (r,θ,ϕ) are the polar coordinates, let ρ=rsinθ, (ρ,ϕ,z) be the cylindrical coordinates, y=cosθ. Recall (uc,γ,pc,γ) are given by (3), where Uθc,γ(y) is a solution of (5). For convenience, denote u=uc,γ, p=pc,γ and Uθ=Uθc,γ. In Euclidean coordinates, x=(x1,x2,x3) and u=(u1,u2,u3).

Proof of Proposition 1.1

Let (c,γ)M. For any R>0, letΩ:={xR3||x|R,R<x3<R}. We prove

Proof of Theorem 1.3

Proof of Theorem 1.3

For convenience, let C denote a constant depending only on p,α,β,α and n, which may vary from line to line. We first prove that if (15) holds for some C, then αα.

Let 0<δ<1, fδ(x) be a smooth function of x, such thatfδ(x):={1,2δ|x|3δ,0,|x|δor|x|4δ, and |xf|C/δ. Let g(xn) be a smooth function such thatg(xn):={1,2|xn|3,0,|xn|1or|xn|4, and |g(xn)|C. Define uδ(x):=fδ(x)g(xn), then uδ is in Cc1(Rn). By computation,|x|β|x|αuδLp(Rn)p232δ|x|3δ|x|βp|x|αpdxdxnδαp+n1/C.

References (25)

  • L. Li et al.

    Homogeneous solutions of stationary Navier-Stokes equations with isolated singularities on the unit sphere. I. One singularity

    Arch. Ration. Mech. Anal.

    (2018)
  • L. Li et al.

    Homogeneous solutions of stationary Navier-Stokes equations with isolated singularities on the unit sphere. III. Existence of axisymmetric solutions with nonzero swirl

    Discrete Contin. Dyn. Syst., Ser. A

    (2019)
  • Cited by (0)

    1

    Partially supported by NSF grant DMS-1501004.

    2

    Partially supported by AMS-Simons Travel Grant and AWM-NSF Travel Grant 1642548.

    View full text