Home Stability of stationary solutions to the three-dimensional Navier-Stokes equations with surface tension
Article Open Access

Stability of stationary solutions to the three-dimensional Navier-Stokes equations with surface tension

  • Keiichi Watanabe EMAIL logo
Published/Copyright: January 3, 2023

Abstract

This article studies the stability of a stationary solution to the three-dimensional Navier-Stokes equations in a bounded domain, where surface tension effects are taken into account. More precisely, this article considers the stability of equilibrium figure of uniformly rotating viscous incompressible fluid in R 3 , which are rotationally symmetric about a certain axis. It is proved that this stability result can be obtained by the positivity of the second variation of the energy functional associated with the equation that determines an equilibrium figure, provided that initial data are close to an equilibrium state. The unique global solution is constructed in the L p -in-time and L q -in-space setting with ( p , q ) ( 2 , ) × ( 3 , ) satisfying 2 / p + 3 / q < 1 , where the solution becomes real analytic, jointly in time and space. It is also proved that the solution converges exponentially to the equilibrium.

1 Introduction

This article is concerned with the stability of equilibrium figure of uniformly rotating viscous incompressible fluid in R 3 with surface tension, where the equilibrium figure is rotationally symmetric about a certain axis. The fluid occupies a region Ω ( t ) at time t 0 , which is surrounded by a free interface Γ ( t ) . We denote the initial position of Γ ( t ) by Γ 0 . Besides, we denote the velocity and pressure of the fluid by v ( x , t ) and π ( x , t ) , respectively, and the unit outward normal on Γ ( t ) by ν Γ . The normal velocity and the doubled mean curvature of Γ ( t ) with respect to ν Γ are denoted by V Γ and H Γ , respectively. Then, the motion of the fluid is governed by the following system:

(1.1) t v + ( v ) v μ Δ v + π = 0 , in  Ω ( t ) , div v = 0 , in  Ω ( t ) , S ( v , π ) ν Γ = σ H Γ ν Γ , on  Γ ( t ) , V Γ = v ν Γ , on  Γ ( t ) , v ( 0 ) = v 0 , in  Ω ( 0 ) , Γ ( 0 ) = Γ 0 .

In this article, Ω ( 0 ) is assumed to be bounded. Here, S ( v , π ) is the stress tensor defined by

S ( v , π ) μ ( v + [ v ] ) π I in  Ω ( t ) ,

where μ is a positive constant and M means the transpose of M . Without loss of generality, we may assume that external (atmospheric) pressure is zero in this article.

It is well known that the Navier-Stokes equations admits the following stationary solution that corresponds to a rigid rotation:

v ( x ) = ω e 3 × x , π ( x ) = ω 2 2 x 2 + p 0 ,

where ω R describes a constant angular velocity and p 0 is some constant, and we have set x = ( x 1 , x 2 , 0 ) R 3 . Substituting ( v , π ) into the boundary condition (1.1) 3 , we obtain the equation for the doubled mean curvature H Γ of Γ :

(1.2) σ H Γ + ω 2 2 x 2 + p 0 = 0 on Γ Ω .

This defines the equilibrium figure Ω of the rotating liquid. Notice that, if Ω is rotationally symmetric about the x 3 axis, the boundary condition (1.1) 4 is automatically satisfied. In the sequel, Ω is assumed to be rotationally symmetric with respect to the axis defined by e 3 . Obviously, in the case ω = 0 , we observe the classical Young-Laplace law σ H Γ + p 0 = 0 . However, our interest here is to investigate the case ω 0 and, especially, characterize the stability of the equilibrium figure by means of the second variation of the energy functional instead of a restriction on the value of ω .

We note that a solution to (1.1) satisfies the following equalities:

(1.3) Ω ( t ) = Ω ( 0 ) , Ω ( t ) v ( x , t ) d x = Ω ( 0 ) v 0 ( x ) d x , Ω ( t ) ( v × x ) d x = Ω ( 0 ) ( v 0 × x ) d x .

By passing a uniformly moving coordinate system x ^ = x V ^ 0 t and v ^ = v V ^ 0 , where V ^ 0 = Ω ( 0 ) 1 Ω ( 0 ) v 0 d x , and by rotating coordinate axes, we suppose that

(1.4) Ω ( t ) v ( x , t ) d x = Ω ( 0 ) v 0 ( x ) d x = 0 , Ω ( t ) ( v × x ) d x = Ω ( 0 ) ( v 0 × x ) d x = γ e 3 γ ( 0 , 0 , 1 ) , Ω ( 0 ) x d x = 0 .

It follows form the Reynolds transport theorem that ( 1.4 ) 3 gives

(1.5) Ω ( t ) x d x = 0 t d d s Ω ( s ) x d x d s = 0 t Ω ( s ) v d x d s = 0 ,

which means that the barycenter point of the domain Ω ( t ) is always suited at the origin. Finally, to guarantee that ( 1.4 ) 2 holds with v = v and Ω ( t ) = Ω , the angular velocity ω and the value γ should satisfy the relation

(1.6) ω Ω x 2 d x = γ .

Notice that the value ω should be determined from a given quantity γ . Namely, Γ is determined from γ , where Γ is a smooth solution to (1.2) subject to (1.6). If γ 1 , there exists a unique Γ satisfying (1.2) and (1.6), see Solonnikov [30, Thm. 5.1] and Watanabe [40, Prop. A.1]. Finally, the multiplication of (1.2) by x ν Γ and integration over Γ leads to the expression for p 0 , where ν Γ is the unit outward normal on Γ . In fact, it follows from (1.2) that

Γ σ H Γ x ν Γ d Γ + Γ ω 2 2 x 2 x ν Γ d Γ + Γ p 0 x ν Γ d Γ = 0 .

The relation H Γ ν Γ = Δ Γ x , x Γ , and the divergence theorem imply

σ Γ Γ x 2 d Γ + 5 ω 2 2 Ω x 2 d x + 3 Ω p 0 = 0 .

Hence, we deduce that

p 0 = 2 σ Γ 3 Ω + 5 γ ω 6 Ω .

Notice that, by (1.6), we see that p 0 is strictly positive.

The free boundary problem of (1.1) is said to be finding a family of hypersurfaces { Γ ( t ) } t 0 and appropriately smooth solutions v and π . Notice that finding a family of hypersurfaces { Γ ( t ) } t 0 is equivalent to finding a family of { Ω ( t ) } t 0 . Since many authors have considered problems similar to (1.1) in various settings, we only mention the details of those papers that dealt with the effect of surface tension included on the free interface (the case σ > 0 ) and with assuming that Ω ( 0 ) is bounded. For the case of surface tension on the free boundary, we refer the reader to papers [1,3,4, 5,11,17, 18,35,36, 37] that deal with the case where the initial domain Ω ( 0 ) is an infinite layer of finite depth with a rigid bottom, see also a recent article by Saito and Shibata [23] that dealt with the case of the bottomless ocean. When Ω ( 0 ) is bounded, the first contribution to the solvability of (1.1) traces back to a long series of papers by Solonnikov [29,30,31]. Specifically, Solonnikov investigated the problem in L 2 regularity framework, i.e., he showed the local existence and uniqueness of solutions for (1.1) in Sobolev-Slobodetskiĭ spaces W 2 2 + α , 1 + α 2 with 1 / 2 < α < 1 . To obtain the local-in-time solutions in Hölder and anisotropic Sobolev regularity frameworks, we refer to the works by Moglilevskiĭ and Solonnikov [16] and Shibata [24], respectively.[1] The unique global existence theorem in the L 2 regularity frameworks was proved by Solonnikov [30], where the solution converges to a uniform rigid rotation of the liquid about a certain axis, provided that the initial velocity and the initial angular momentum (i.e., γ ) are sufficiently small and Γ 0 is sufficiently close to a sphere. The similar result was also established in Hölder regularity framework, see Padula and Solonnikov [19]. More recently, the author [40] extends the result obtained in [19] the class of anisotropic Sobolev spaces W p , q 2 , 1 with 2 < p < and 3 < q < satisfying 2 / p + 3 / q < 1 , which can also be regarded as an extension of Shibata [26] dealing with the case ω = 0 . Notice that, in the previous studies [16,30,40], the equilibrium figure is uniquely determined by the constant γ . However, it was necessary to assume that γ is small since this assumption yields the smallness of ω , so that we could find a unique smooth solution to equation (1.2) subject to (1.6) based on a standard contraction mapping theorem; see Solonnikov [30, Thm. 5.1] and Watanabe [40, Prop. A.1]. On the other hand, Solonnikov [34] showed that the smallness condition on γ can be replaced by the condition of the positivity of the second variation of the functional E ω given by

(1.7) E ω ( h ) = Γ ( t ) σ d Γ Ω ( t ) ω 2 2 x 2 d x Ω ( t ) p 0 d x ,

provided that there exists a smooth surface Γ such that Γ satisfies (1.2) and (1.6) and is rotationally symmetric about the x 3 axis. In particular, he showed that the stability result can be obtained by the positivity of the second variation of the energy functional E ω ( h ) within the Hölder regularity framework.

The aim of this article is to extend the aforementioned Hölder regularity result obtained by Solonnikov [34, Thm. 2.1] in the L p -in-time and L q -in-space ( L p L q ) setting with 2 < p < and 3 < q < satisfying 2 / p + 3 / q < 1 , which provides an optimal regularity on the initial data. In particular, in contrast to Shibata [26], we investigate the stability of nontrivial stationary solutions (i.e., ( v , π , Γ ) with ω 0 ) to the three-dimensional Navier-Stokes equations with surface tension. To prove our main result, we transform the system (1.1) into a problem on a domain F surrounded by a fixed surface G . To observe smoothing of the unknown interface, we rely on the direct mapping method via the Hanzawa transform, i.e., we approximate the free surface Γ ( t ) by a real analytic hypersurface G , in terms of the Hausdorff distance of the second-order normal bundles being as small as we wish. As the transformed problem becomes a quasilinear parabolic type PDE with inhomogeneous boundary data, it is widely known that the key idea to show the local-in-time existence of regular solution is to use the maximal L p L q regularity result for an associated linearized problem. In this article, however, we will address the global existence issue, and hence, it is also required to derive some decay property of an associated linearized system – combining the local existence result and the decay property yields the global existence result due to a standard argument. To this end, we use the maximal L p L q regularity result for the shifted linearized system to rewrite the linearized system as an abstract evolution equation with homogeneous boundary data. Since the shifted linearized system admits a unique solution in the maximal L p L q regularity class that decays exponentially, it suffices to study the decay property for a solution to the abstract evolution equation. Although Solonnikov [34] established the decay estimate by energy estimates, his approach does not seem to be available in a general L p L q setting. To overcome this difficulty, we study the decay property of an analytic C 0 -semigroup associated with the linearized system. On the basis of a spectral analysis of the corresponding linear operator A q defined in X 0 = J q ( F ) × B q , q 2 1 / q ( G ) , we will show that A ˜ q A q X ˜ 0 generates an analytic C 0 -semigroup { e A ˜ q t } t 0 in X ˜ 0 , which is exponentially stable on some subspace X ˜ 0 of X 0 , see Section 5. Here, J q ( F ) and B q , q 2 1 / q ( G ) stand for the solenoidal space of ( L q ( F ) ) 3 and the inhomogeneous Besov spaces, respectively, and the space X ˜ 0 is defined as the set of all ( f , g ) X 0 that satisfies the following orthogonal conditions:

F f d y = F f ( e 3 × y ) d y = 0 , F f ( e α × y ) d y = ω G g y α y 3 d G , ( α = 1 , 2 ) , G g d G = G g y d G = 0 , ( = 1 , 2 , 3 ) ,

which are similar to but essentially different from the orthogonal condition considered in Shibata [26]. Thanks to these orthogonal conditions, we will observe that the resolvent set of A ˜ q contains the right half-plane C + { λ C : Re λ 0 } including λ = 0 , which implies the exponential stability of the semigroup. Hence, our discussion generalizes the argument employed in Shibata [26] who deals with the case ω = 0 , and see also Shibata and Shimizu [28] for the case σ = ω = 0 . Notice that, if γ = 0 (i.e., ω = 0 ), then the equilibrium surface becomes a sphere so that a natural choice of G is a sphere as well. If we choose G as a sphere, we obtain a nice spectral property of the Laplace-Beltrami operator defined on G , which arises from the surface tension, see [26, Lemm. 4.5] (cf. [22, Prop. 10.2.1]). In our case, however, it follows from ω 0 and (1.2) that the equilibrium surface is not sphere. In addition, since we do not impose the smallness condition on γ (as well as ω ), the equilibrium surface cannot be understood as a normal perturbation of a sphere in general, which means that Shibata’s approach fails. If we assume that γ is sufficiently small, we can recover Shibata’s argument because the equilibrium surface is given by a normal perturbation of a sphere, and see a recent contribution by the author [40]. Thus, we have to introduce another technique to handle the term arising from the surface tension. To this end, we introduce the quadratic form that determines from the functional E ω given in (1.7).

It should be noted that our result even in the case γ = ω = 0 refines Shibata’s result [26]. In fact, in view of the trace theory (cf. Denk et al. [9]), if we study the linearized problem, the boundary data have to be in the intersection space:

(1.8) F p , q s ( 0 , T ; L q ( G ) ) L p ( 0 , T ; B q , q s / 2 ( G ) ) , 0 < s < 1 ,

where F p , q s denotes the vector-valued inhomogeneous Triebel-Lizorkin spaces. Hence, to find the strong solution to (1.1) via a contraction mapping principle, we have to estimate the nonlinear terms in this intersection space. However, in the argument in Shibata [26], the boundary data were not lying in (1.8), and hence, his result is not optimal in view of trace theory. Besides, Shibata [26] did not investigate that the height function h admits the higher regularity F p , q 2 1 / q ( J ; L q ( G ) ) , which implies that the free interface can be understood in the classical sense due to the embedding

F p , q 2 1 / q ( 0 , T ; L q ( G ) ) H 1 , p ( 0 , T ; B q , q 2 1 / q ( G ) ) L p ( 0 , T ; B q , q 3 1 / q ( G ) ) C ( [ 0 , T ] ; C 2 ( G ) ) C 1 ( [ 0 , T ] ; C 1 ( G ) ) .

To overcome these fallacious, we rely on the recent contributions established by the author [39,40], which were based on the studies of Lindemulder [13] and Meyries and Veraar [14,15]. Furthermore, it was not showed in the study by Shibata [25,26] that the solution can be understood in the classical sense, but the well-known parameter trick (cf. Prüss-Simonett [22, Ch. 9]) implies that the solution is indeed real analytic, jointly in time and space. This shows that the solutions to (1.1) are indeed classical.

To explain our main result, we shall introduce the quadratic form that characterize the stability of the stationary solution to (1.1). As we will explain in the next section, the free surface Γ ( t ) can be approximated by a real analytic surface G in the sense that the Hausdorff distance of the second-order bundles of Γ ( t ) and G is as small as we wish. In this case, we can write

(1.9) Γ ( t ) = { p + h ( p , t ) ν G ( p ) : p G } , Γ 0 = { p + h 0 ( p ) ν G ( p ) : p G } ,

where h is an unknown function but h 0 B q , p 2 + δ 1 / p 1 / q ( G ) is a given function. Here, we will prove that h is indeed smooth function defined on G for each t > 0 . Since we seek global solutions that converges to an equilibrium, in the following, we may set G = Γ . First, we observe that (1.2) is the Euler-Lagrange equation associated with E ω . In fact, the first variation of E ω at h = 0 is given by

(1.10) δ 0 E ω = G σ H G h d Γ G ω 2 2 y 2 h d G G p 0 h d G .

In addition, the second variation of E ω at h = 0 is given by

(1.11) δ 0 2 E ω = G σ ( h Δ G h 2 K G h 2 ) d G G ω 2 2 ν G y 2 y 2 H G h 2 d G + G p 0 H G h 2 d G ,

where K G is the Gaussian curvature. We refer to [32, Sec. 2] for the derivations of (1.10) and (1.11), and see also [22, Ch. 2] for further geometric background. Recalling (1.2), we have

σ H G 2 + ω 2 2 y 2 H G + p 0 H G = 0 on G ,

and hence, the second variation δ 0 2 E ω of E ω can be rewritten as follows:

δ 0 2 E ω = G σ h H G ( 0 ) h d G G ω 2 2 ν G y 2 h 2 d G G h G h d G ,

where H G ( 0 ) stands for the first variation of H G at h = 0 , since it holds

Δ G h 2 K G h = H G ( 0 ) h H G 2 h .

Since δ 0 E ω = 0 , we observe that G is a minimal surface. We now normalize G h by

^ G h = G h 1 G G G h d G

and define the quadratic form Ψ G by

(1.12) Ψ G ( g , h ) G g ^ G h d G

for g , h H 2 , 2 ( G ) . Then, we have δ 0 2 E ω = Ψ G ( h , h ) and G ^ G h d G = 0 .

The aim of this article is to show the stability of the stationary solution ( v , π , G ) to (1.1), provided that there exists a solution G to (1.2), and the quadratic form Ψ G satisfies the following assumption.

Assumption 1.1

The quadratic form Ψ G ( h , h ) is positive definite on

L ( 0 ) 2 ( G ) h L 2 ( G ) : G h d G = G h y j d G = 0 , j = 1 , 2 , 3 . ,

that is, there exists a constant c such that

Ψ G ( h , h ) L 2 ( G ) c h L 2 ( G ) 2

holds for every h L ( 0 ) 2 ( G ) H 2 , 2 ( G ) .

Our main result reads as follows.

Theorem 1.2

Suppose that Ω ( 0 ) satisfies ( 1.4 ) 3 . Assume that there exists a smooth solution Γ to (1.2), which is rotationally symmetric about the x 3 axis and that the quadratic form Ψ G satisfies Assumption 1.1. Let ( v , π ) be given by (1.3). If p , q , and δ satisfy

(1.13) 2 < p < , 3 < q < , 1 p + 3 2 q < δ 1 2 1 2 ,

then an equilibrium ( v , π , Γ ) is stable in the following sense: Let Γ 0 be given by ( 1.9 ) 2 with a given function h 0 B q , p 2 + δ 1 / p 1 / q ( G ) . There exists a positive constant ε > 0 such that for all v 0 v B q , p 2 ( δ 1 / p ) ( Ω ( 0 ) ) and h 0 B q , p 2 + δ 1 / p 1 / q ( G ) satisfying the smallness condition:

v 0 v B q , p 2 ( δ 1 / p ) ( Ω ( 0 ) ) + h 0 B q , p 2 + δ 1 / p 1 / q ( G ) ε ,

the compatibility conditions:

(1.14) div v 0 = 0 in Ω , P Γ 0 [ μ ( v 0 + [ v 0 ] ) ] = 0 on Γ 0

and the conditions Ω ( 0 ) v 0 ( x ) d x = 0 and Ω ( 0 ) ( v 0 × x ) d x = γ e 3 with a constant γ R , there exists a unique global classical solution ( v ( t ) , π ( t ) , Γ ( t ) ) of problem (1.1). In particular, the set t ( 0 , ) ( Γ ( t ) × { t } ) is real analytic manifold and the function ( v , π ) : { ( x , t ) Ω ( t ) × ( 0 , ) } R 4 is real analytic. In addition, if Γ ( t ) is parameterized over G by means of a height function h ( t ) , that is, if Γ ( t ) is given by ( 1.9 ) 1 with an unknown function h, it holds

v ( t ) v B q , p 2 ( δ 1 / p ) ( Ω ( t ) ) = O ( e c t ) and h ( t ) B q , p 2 + δ 1 / p 1 / q ( G ) = O ( e c t ) as t

with some positive constant c. Here, we have set P Γ 0 I ν Γ 0 ν Γ 0 .

Remark 1.3

The restriction (1.13) implies the embeddings

B q , p 2 ( δ 1 / p ) ( Ω ( t ) ) BUC 1 ( Ω ( t ) ) and B q , p 2 + δ 1 / p 1 / q ( G ) BUC 2 ( G ) .

Hence, we clearly observe that v ( t ) v C 1 ( Ω ( t ) ) = O ( e c t ) and h ( t ) C 2 ( G ) = O ( e c t ) as t . We also notice that the well known norm equivalence of the Hölder spaces C s B , s ( s > 0 , s N ) implies the embedding properties

C 2 + α ( Ω ( 0 ) ) B , 2 + α ( Ω ( 0 ) ) B q , p 2 ( δ 1 / p ) ( Ω ( 0 ) ) , α ( 0 , 1 ) , C 3 + α ( G ) B , 3 + α ( G ) B q , p 2 + δ 1 / p 1 / q ( G ) , α ( 0 , 1 ) ,

which shows that the regularity of v 0 v and h 0 in Theorem 1.2 are less than the assumption imposed in [34, Thm. 2.1]. Hence, Theorem 1.2 indeed improves the result due to Solonnikov [34, Thm. 2.1]. Furthermore, in contrast to [34, Thm. 2.1], solutions regularize and immediately become real analytic in space and time, which encodes typical parabolic behavior.

Remark 1.4

If the initial velocity v 0 satisfies the orthogonal condition (1.4) with γ = 0 , Theorem 1.2 can be regarded as a generalization of Shibata [26].

Remark 1.5

Using elliptic integrals, the existence of the equilibrium surface Γ that satisfies (1.6) may be shown and, especially, it is simply connected if the value of the nondimensional parameter

h = ω 2 a 3 8 σ

is strictly less than some value h max 2.32911 , and see Chandrasekhar [7] (cf. Appell [2, Ch. IX]). Here, a is the equatorial radius of the liquid drop and Γ is symmetric with respect to an x 1 - x 2 plane. In the following, let us briefly explain the result presented by Chandrasekhar [7]. If h = 0 (i.e., ω = 0 ), it is clear that Γ is the sphere with the radius R > 0 ; if 0 < h 1 , then Γ becomes an ellipsoid; if 1 < h < h max , then a dimple appears. See also [7, Fig. 1] for a precise information. On the other hand, by [2, Ch. IX], if h h max , the equilibrium surface Γ becomes a toroid (i.e., not simply connected), which replicates the classical experiment due to Plateau in the middle of the 19th century. In this article, it is not necessary to assume that the free boundary Γ ( t ) is simply connected, which is the same as Solonnikov [32], since we do not need to use polar coordinates to express the free boundary like Padula and Solonnikov [19]. Notice that Theorem 1.2 justifies the stability result, obtained by Brown and Scriven [6], for an axisymmetric equilibrium figure in the sense that we also consider the motion of the incompressible viscous fluid occupied inside the free boundary.

In contrast to the previous article [40], Theorem 1.2 replaces the smallness condition on γ by the condition of the positivity of δ 0 2 E ω = Ψ G ( h , h ) , where it is assumed the existence of the equilibrium surface Γ satisfying (1.6). However, Theorem 1.2 does not conflict with the result obtained in [40]. In fact, if γ is suitably small, it follows from [40, Prop A.1] that there exists a unique Γ satisfying (1.2) and (1.6) with

Γ = { p + h ( p ) ν G ( p ) : p S R } ,

where ( h , h ) L ( S R ) are small. Here, S R is a sphere centered at the origin with a radius R > 0 , and ( h , h ) L ( S R ) can be dominated by γ , see [40, Prop. A.1]. In this case, setting p 0 = 2 σ / R , the functional δ 0 2 E ω can be regarded as a perturbation from

S R σ ( h ¯ Δ S R h ¯ 2 K S R h ¯ 2 ) d S R ( h ¯ h h ) ,

which is positive definite on L ( 0 ) 2 ( S R ) H 2 , 2 ( S R ) (see [22, Prop. 10.2.1]). Hence, taking γ as small as possible, we observe that the smallness of γ implies the positivity of δ 0 2 E ω . Thus, we can conclude that Theorem 1.2 is an extension of the result obtained in the previous article. Accordingly, Theorem 1.2 includes all results obtained in [26,34,40].

The rest of the article is folded as follows. In Section 2, we recall the notation of functional spaces and preliminary propositions used throughout this article. In Section 3, we transform the system (1.1) to a problem on a domain F surrounded by a fixed interface G in terms of the Hanzawa transform. Section 4 is devoted to showing that the principal part of the linearization has the property of maximal L p L q regularity. Then, some exponential decay property of the linearized system is proved in Section 5. Finally, the Section 6 presents the proof of the main result, Theorem 1.2.

2 Preliminaries

2.1 Notations

Let us fix the notations in this article. Let N be the set of all natural numbers and N 0 N { 0 } , and let R and C be, respectively, the set of all real numbers and the set of all complex numbers. Set R + { a R : a > 0 } . By C > 0 , we will often denote a generic constant that does not depend on the quantities at stake.

2.2 Functional spaces

In this subsection, we introduce functional spaces used throughout this article. Let p , q [ 1 , ] . For any D domain of R 3 , we denote the standard K -valued Lebesgue spaces and Sobolev spaces on D by L q ( D ) and H m , q ( D ) , m N , respectively, where K { R , C } . The standard K -valued inhomogeneous Besov spaces on D are denoted by B p , q s ( D ) , s R . The homogeneous spaces H ˙ 1 , q ( D ) are given by H ˙ 1 , q ( D ) { w L loc 1 ( D ) : w L q ( D ) } for q ( 1 , ) and a domain D R 3 , while the dual space of H ˙ 1 , q ( D ) is written by H ˙ 1 , q ( D ) with the Hölder conjugate q q / ( q 1 ) of q .

Let X be a Banach space. Then the m -product space of X is denoted by X m , while its norm is usually denoted by X instead of X m when no confusion can arise.

Let Banach spaces X 1 and X 2 the norm of X 1 X 2 is denoted by X 1 X 2 X 1 + X 2 , and ( X 1 , X 2 ) stands for the Banach space of all bounded linear operators from X 1 to X 2 . We may write ( X 1 ) instead of ( X 1 , X 1 ) to shorten the notation. The symbol Hol ( Λ , ( X 1 , X 2 ) ) represents the set of all ( X 1 , X 2 ) -valued holomorphic functions defined on Λ C . We set Σ θ , z 0 { z C { 0 } : arg z π θ , z z 0 } with θ ( 0 , π / 2 ) and z 0 > 0 .

For I R and p ( 1 , ] , let L p ( I ; X ) and H 1 , p ( I ; X ) be the X -valued Lebesgue spaces on I and the X -valued Sobolev spaces on I , respectively. For p ( 1 , ) and δ ( 1 / p , 1 ] , we set

L δ p ( I ; X ) { f : I X : t 1 δ f L p ( I ; X ) } , H δ 1 , p ( I ; X ) { f L δ p ( I ; X ) H 1 , 1 ( I ; X ) : t f L δ p ( I ; X ) } .

For p ( 1 , ) , q [ 1 , ] , and s R , the symbol F p , q , δ s ( I ; X ) stands for the X -valued Triebel-Lizorkin spaces with the power weight t p ( 1 δ ) . In addition, the Banach space of all X -valued bounded uniformly continuous functions on I is denoted by BUC ( I ; X ) . Finally, BUC m ( I ; X ) is the subset of BUC ( I ; X ) that has bounded partial derivatives up to order m N . Here, BUC ( D ) and BUC m ( D ) are defined similarly as mentioned earlier. For further information on function spaces, we refer the reader to [22,38].

2.3 The space of data for divergence equation

As we will see in Section 2.4, in a transformed system, we have the divergence equation div u = g d in F . Hence, to deal with this equation, it is required to introduce a space DI q ( F ) that is the set of all g d H 1 , q ( F ) such that there exists a solution g ˜ d L q ( F ) 3 to ( g d , φ ) F = ( g ˜ d , φ ) F for every φ H 0 1 , q ( F ) { φ H 1 , q ( F ) : φ G = 0 } . Here and in the following, ( , ) F denotes the L 2 inner product in F . We now define G ( g d ) { g L q ( F ) 3 : div g ˜ d = div g } and denote the representative elements of G ( g d ) by [ G ( g d ) ] . For brevity, and when there is no danger of confusion, we write G ( g d ) instead of [ G ( g d ) ] . Here, it holds div G ( g d ) = g d in F for every g d DI q ( F ) . Setting

g d DI q ( F ) g d H 1 , q ( F ) + inf g G ( g d ) g L q ( F ) for g d DI q ( F ) ,

we observe that DI q ( F ) is a Banach space endowed with the norm DI q ( F ) .

2.4 -bounded families of operators

We next recall the basic theory of the -boundedness of a family of operators, which will be used in Section 4. Here, we refer to [8] for the fundamental concept of -boundedness. In the following, the -bound of a family of operators T ( X , Y ) is denoted by X Y { T } , where X and Y are Banach spaces. If X = Y , we ofren write X { T } for short. The following result is a direct consequence of [8, Rem. 3.2 (4)].

Proposition 2.1

Let 1 q < and G R 3 be a domain. Let m ( λ ) be a bounded function defined on a subset Λ of C , and let M m ( λ ) be a multiplication operator given by M m ( λ ) f m ( λ ) f for every f L q ( G ) . Then it holds

L q ( G ) ( { M m ( λ ) : λ Λ } ) K q 2 ( sup λ Λ m ( λ ) ) ,

where K q > 0 is a constant appearing in the Khintchine inequality.

3 Reduction to a fixed reference surface

In this section, we transform problem (1.1) to one on the fixed spatial domain t 0 . Putting V v v and P = π π , the problem of a stability of the equilibrium ( v , π , Γ ) reduces to a free boundary problem for the perturbation ( V , P ) :

(3.1) t V + ( v ) V + ( V ) v + ( V ) V μ Δ V + P = 0 , in Ω ( t ) , div V = 0 , in Ω ( t ) , S ( V , P ) ν Γ = σ H Γ + ω 2 2 x 2 + p 0 ν Γ , on Γ ( t ) , V Γ = ( v + V ) ν Γ , on Γ ( t ) , V ( 0 ) = v 0 v , in Ω ( 0 ) , Γ ( 0 ) = Γ 0 .

In our analysis, it will be useful to eliminate the terms ( v ) V and ( V ) v . To this end, we introduce the coordinate system rotating about the x 3 axis with a constant angular velocity ω R , i.e., x = O ( ω t ) z . We now set

V ˜ ( z , t ) O 1 ( ω t ) V ( O ( ω t ) z , t ) and P ˜ ( z , t ) P ( O ( ω t ) z , t )

with

O ( θ ) cos θ sin θ 0 sin θ cos θ 0 0 0 1 .

Then problem (3.1) can be rewritten as follows:

(3.2) t V ˜ + ( V ˜ ) V ˜ μ Δ V ˜ + 2 ω ( e 3 × V ˜ ) + P ˜ = 0 , in  Ω ˜ ( t ) , div V ˜ = 0 , in  Ω ˜ ( t ) , S ( V ˜ , P ˜ ) ν Γ ˜ = σ H Γ ˜ + ω 2 2 z 2 + p 0 ν Γ ˜ , on  Γ ˜ ( t ) , V Γ ˜ = V ˜ ν Γ ˜ , on  Γ ˜ ( t ) , V ˜ ( 0 ) = v 0 v V ˜ 0 , in  Ω ˜ ( 0 ) , Γ ( 0 ) = Γ 0 ,

where we have set Ω ˜ ( t ) = O 1 ( ω t ) Ω ( t ) and Γ ˜ ( t ) = Ω ˜ ( t ) . Here, V Γ ˜ ( , t ) and H Γ ˜ ( , t ) stand for the normal velocity and the doubled mean curvature of Γ ˜ ( t ) with respect to the norm ν Γ ˜ ( , t ) = O 1 ( ω t ) ν Γ ( , t ) to Γ ˜ ( t ) , respectively. For further details of the derivation of (3.2), the readers may consult the discussion in [34]. It follows from conditions ( 1.4 ) 3 and (1.5) that

Ω ˜ ( t ) z d z = Ω ˜ ( 0 ) d z = 0 ,

i.e., the barycenter of the fluid is still located at the origin. In addition, condition ( 1.4 ) 1 becomes

Ω ˜ ( t ) V ˜ ( z , t ) d z = 0

due to Ω ˜ ( t ) z d z = 0 . Notice that ( 3.2 ) 4 is equivalent to

V Γ ˜ ( z , t ) = V ˜ ν Γ ˜ 1 F ν Γ ˜ Ω ˜ ( t ) V ˜ ( z , t ) d z .

We will see in Section 5 that this modification will be convenient.

Next, we transform (3.2) to a system on a domain F surrounded by a fixed surface G via the Hanzawa transform, where the strategy is due to Köhne et al. [12, Sec. 2] (cf. Prüss and Siminett [22]). For the necessary geometric background, we refer to Chapter 2 in [22]. Recall that the second-order bundle of Γ ˜ is given by

N 2 Γ ˜ { ( p , ν Γ ˜ ( p ) , Γ ˜ ν Γ ˜ ( p ) ) : p Γ ˜ } ,

where ν Γ ˜ is the surface gradient on Γ ˜ . In addition, let us write d H to denote the Hausdorff distance between the closed subsets K 1 , K 2 R 3 , which is defined by

d H ( K 1 , K 2 ) max ( sup a K 1 dist ( a , K 2 ) , sup b K 2 dist ( b , K 1 ) ) .

Then, the unknown surface Γ ˜ can be approximated by a real analytic hypersurface G in the sense that d H ( N 2 Γ ˜ , N 2 G ) is as small as we wish, i.e., for each η > 0 , we can find a real analytic closed hypersurface G such that d H ( N 2 Γ ˜ , N 2 G ) η . Since it is well known that the hypersurface G admits a tubular neighborhood, there exists some positive constant d 0 such that the mapping Λ : G × ( d 0 , d 0 ) R 3 defined by

Λ ( p , r ) p + r ν G ( p ) , p G , r < d 0

is a diffeomorphism from G × ( d 0 , d 0 ) onto R ( Λ ) . Here, R ( Λ ) denotes the range of Λ . Then, the inverse Λ 1 : R ( Λ ) G × ( d 0 , d 0 ) is conveniently decomposed as Λ 1 ( y ) = ( Π G ( y ) , d G ( y ) ) for y R ( Λ ) , where Π G ( y ) and d G ( y ) stand for the orthogonal projection of y onto G and the signed distance from y onto G ; so d G ( y ) = dist ( y , G ) and d G ( y ) < 0 if and only if y F . Noting the compactness of G , there exists a radius r G > 0 such that for each point p G there are balls B ( y , r G ) F satisfying G B ( y , r G ) ¯ = { p } . Choosing r G maximal, it holds r G > d 0 . In the following, we fix d 0 = r G / 2 and d = d 0 / 3 .

We write the derivatives of d G ( y ) and Π G ( y ) by

d G ( y ) = ν G ( Π G ( y ) ) , D Π G ( y ) = M 0 ( d G ( y ) ) P G ( Π G ( y ) ) ,

respectively. Here, M 0 ( r ) ( I r L G ) 1 is the Weingarten tensor L G G ν G and P G ( p ) = I ν G ( p ) ν G ( p ) represents the orthogonal projection onto the tangent space of G at p G . Here, it holds

M 0 ( r ) ( 1 r L G ) 3

for all r 2 r G / 3 . Using the mapping Λ , we may approximate the unknown free surface Γ ˜ ( t ) over G by means of a height function h via

Γ ˜ ( t ) { p + h ( p , t ) ν G ( p ) : p G }

for small t 0 , at least. We extend this diffeomorphism to all of F ¯ and define the Hanzawa transform by

Ξ h ( y , t ) y + χ d G ( y ) d h ( Π G ( y ) , t ) ν G ( Π G ( y ) ) y + ξ h ( y , t ) ,

where χ C ( R ) is a cutoff function such that 0 χ 1 , 1 < χ < 3 , χ ( r ) = 1 for r 1 , and χ ( r ) = 0 for r > 2 . According to the definition of Ξ h , we obtain

Ξ h ( y , t ) = z if d G ( y ) > 2 d , Π G ( Ξ h ( y , t ) ) = Π G ( y ) if d G ( y ) < d , d G ( Ξ h ( y , t ) ) = d G ( y ) + χ d G d h ( Π G ( y ) , t ) if d G ( y ) < 2 d ,

as well as

Ξ h 1 ( y , t ) = y h ( Π G ( y ) , t ) ν G ( Π G ( y ) ) if d G ( y ) < d .

By using the transform Ξ h , we define the pull backs of V ˜ and P ˜ by

u ( y , t ) V ˜ ( Ξ h ( y , t ) , t ) , q ( y , t ) P ˜ ( Ξ h ( y , t ) , t ) , y F , t > 0 ,

respectively. Then, we observe that ( u , q , h ) satisfies the following system:

(3.3) t u μ ( h ) u + 2 ω ( e 3 × u ) + G ( h ) q = R ( u , h ) , in F , G ( h ) u = 0 , in F , ( μ ( G ( h ) u + [ G ( h ) u ] ) q I ) ν Γ ˜ ( h ) = σ H Γ ˜ ( h ) + ω 2 2 z 2 + p 0 ν Γ ˜ ( h ) , on G , t h u ν G = u a ( h ) 1 F ( ν G a ( h ) ) Ω ˜ ( t ) V ˜ d y , on G , u ( 0 ) = u 0 , in F , h ( 0 ) = h 0 , on G .

Here, ( h ) , G ( h ) , and H Γ ˜ ( h ) are the transformed Laplacian, gradient, and doubled mean curvature, respectively, while the functions R ( u , h ) and a ( h ) are defined below. Furthermore, it holds

D Ξ h = I + D ξ h , ( D Ξ h ) 1 = I ( I + D ξ h ) 1 ξ h I [ M 1 ( h ) ] ,

where

D ξ h ( y , t ) = 1 a χ d G ( y ) d h ( Π G ( y ) , t ) ν G ( Π G ( y ) ) ν G ( Π G ( y ) ) + χ d G ( y ) d ν G ( y ) M 0 ( Π G ( y ) ) G h ( Π G ( y ) , t ) χ d G ( y ) d h ( Π G ( y ) , t ) L G ( Π G ( y ) ) M 0 ( Π G ( y ) ) P G ( Π G ( y ) )

if d G ( y ) < 2 d , while D ξ h ( y , t ) = 0 if d G ( y ) > 2 d . In particular, if d G ( y ) < d , we have

D ξ h ( y , t ) = ν G ( y ) M 0 ( Π G ( y ) ) G h ( Π G ( y ) , t ) h ( Π G ( y ) , t ) L G ( Π G ( y ) ) M 0 ( Π G ( y ) ) P G ( Π G ( y ) ) .

On the basis of these representations, we find that ( I + D ξ h ) is boundedly invertible if

h L ( G ) < 1 3 min d , 1 L G L ( G ) and G h L ( G ) < 1 3 .

Then we can write

[ P ˜ ] Ξ h = G ( h ) q = [ D Ξ h 1 Ξ h ] q = [ ( D Ξ h ) 1 ] q = q M 1 ( h ) q , [ div V ˜ ] Ξ h = G ( h ) u = ( ( I M 1 ( h ) ) ) u ,

and

[ Δ V ˜ ] Ξ h = ( h ) u = [ ( I M 1 ( h ) ) ] [ ( I M 1 ( h ) ) h ] = Δ u [ M 1 ( h ) + [ M 1 ( h ) ] M 1 ( h ) [ M 1 ( h ) ] ] : 2 u + [ [ ( I M 1 ( h ) ) : M 1 ( h ) ] ] u Δ u M 2 ( h ) : 2 u M 3 ( h ) u .

Here, we have used the notation A : B = i , j = 1 3 a i j b i j = tr ( A B ) . In our analysis, it is required to obtain another representation formula for [ div V ˜ ] Ξ h . To this end, we use the L 2 inner product in Ω ˜ ( t ) , which is denoted by ( , ) Ω ˜ ( t ) . For any test function φ ˜ C c ( Ω ˜ ( t ) ) , we write φ ( y ) = φ ˜ ( z ) . Here, C c ( D ) is the set of all C -functions on R 3 , which have compact supports contained in D R 3 . In addition, the Jacobian of the transform Ξ h ( y , t ) is denoted by J = J ( h ) . Recalling the expression of D ξ h , we shall write J ( h ) = 1 + J 0 ( h ) with some function J 0 ( h ) that vanishes at h = 0 . Then, we see that

( div V ˜ , φ ˜ ) Ω ˜ ( t ) = ( V ˜ , φ ˜ ) Ω ˜ ( t ) = ( J u , ( I M 1 ( h ) ) φ ) F = ( div ( ( I [ M 1 ( h ) ] ) J u ) , φ ) F = ( J 1 div ( ( I [ M 1 ( h ) ] ) J u ) , φ ˜ ) Ω ˜ ( t ) ,

where ( , ) F is the L 2 inner product in F . This gives

[ div V ˜ ] Ξ h = div u M 1 ( h ) : u = J 1 ( div ( ( I [ M 1 ( h ) ] ) J u ) ) ,

i.e., the divergence free condition ( 3.2 ) 2 is rewritten as follows:

div u = J 0 ( h ) div u + ( 1 + J 0 ( h ) ) M 1 ( h ) : u = div ( ( 1 + J 0 ( h ) ) [ M 1 ( h ) ] u ) .

Next, we note that

[ t V ˜ ] Ξ h = t u [ ( V ˜ ) Ξ h ] ( t Ξ h ) = t u [ [ D Ξ h 1 Ξ h ] u ] ( t Ξ h ) = t u u [ ( I + D ξ h ) 1 t ξ h ] t u M 4 ( h ) u ,

which implies

R ( u , h ) = ( u G ( h ) u ) + M 4 ( h ) u .

Furthermore, it holds

ν Γ ˜ = b ( h ) ( ν G a ( h ) ) , a ( h ) = M 0 ( h ) G h , b ( h ) = 1 1 + a ( h ) 2 , M 0 ( h ) = ( I h L G ) 1 ,

and

V Γ ˜ = ( t Ξ h ) ν Γ ˜ = t h ( ν Γ ˜ ν G ) = b ( h ) t h .

Here, ν G and a ( h ) are linearly independent. The term Ω ^ ( t ) V ˜ ( y , t ) d y can be read as follows:

Ω ˜ ( t ) V ˜ d y = F u J ( h ) d z = F u d z + F u J 0 ( h ) d z .

The doubled mean curvature H Γ ˜ is given by

(3.4) H Γ ˜ ( h ) = b ( h ) ( tr [ M 0 ( h ) ( L G + G a ( h ) ) ] b 2 ( h ) ( M 0 ( h ) a ( h ) ) ( [ G a ( h ) ] a ( h ) ) ) .

Its linearization at h = 0 is given by

(3.5) H Γ ˜ ( 0 ) = tr L G 2 + Δ G = H G 2 2 K G + Δ G ,

where Δ G is the Laplace-Beltrami operator on G . Here and in the following, for sufficiently smooth functions a , b C ( G ) and F ( a ) : G R k , we use the notation

F ( a ) b d d s F ( a + s b ) s = 0 ,

which denotes the first variation of F ( a ) . We refer to [22, Ch. 2] for the derivations of (3.4) and (3.5).

We now decompose the stress boundary condition into tangential and normal parts. Multiplying ( 3.3 ) 3 with ν G / b , we obtain

q + σ H Γ ˜ ( h ) + ω 2 2 z 2 + p 0 = ( μ ( G ( h ) u + [ G ( h ) u ] ) ( ν G a ( h ) ) ) ν G

for the normal part of ( 3.3 ) 3 , while

(3.6) P G ( μ ( G ( h ) u + [ G ( h ) u ] ) ( ν G a ( h ) ) ) = 0

for the tangential part of ( 3.3 ) 3 . It should be emphasized that (3.6) neither contains the pressure nor the curvature. Finally, we have

z 2 y 2 ν G y 2 h = h 2 ( ( ν G ( 1 ) ) 2 + ( ν G ( 2 ) ) 2 ) .

Consequently, from the aforementioned discussion, problem (3.2) can be rewritten as follows:

(3.7) t u μ Δ u + 2 ω ( e 3 × u ) + q = F u ( u , q , h ) , in  F , div u = G d ( u , h ) = div G div ( u , h ) , in  F , P G ( μ ( u + [ u ] ) ν G ) = G u τ ( u , h ) , on  G , μ ( u + [ u ] ) ν G ν G q + G h = G u v ( u , h ) + G 0 ( h ) , on  G , t h u ν G + 1 F ν G F u d z = F h ( u , h ) + F ( u , h ) , on  G , u ( 0 ) = u 0 , in  F , h ( 0 ) = h 0 , on  G .

Here, the right-hand members of (3.7) can be written as follows:

F u ( u , q , h ) = M 4 ( h ) u u ( I M 1 ( h ) ) u μ ( M 2 ( h ) : 2 u + M 3 ( h ) u ) + M 1 ( h ) q , G d ( u , h ) = J 0 ( h ) div u + ( 1 + J 0 ( h ) ) M 1 ( h ) : u , G div ( u , h ) = ( 1 + J 0 ( h ) ) M 1 ( h ) u , G u τ ( u , h ) = P G ( μ ( M 1 ( h ) u + [ M 1 ( h ) u ] ) ( ν G M 0 ( h ) G h ) ) + P G ( ( u + [ u ] ) M 0 ( h ) G h ) , G u v ( u , h ) = ( u + [ u ] ) M 0 ( h ) G h ν G + ( M 1 ( h ) u + [ M 1 ( h ) u ] ) ( ν G M 0 ( h ) G h ) ν G G 0 ( h ) = σ ( H Γ ˜ ( h ) H Γ ˜ ( 0 ) h ) + ω 2 2 { h 2 ( ( ν G ( 1 ) ) 2 + ( ν G ( 2 ) ) 2 ) } , F h ( u , h ) = ( M 0 ( h ) G h ) u , F ( u , h ) = 1 F a ( h ) F u ( 1 + J 0 ) d z 1 F ν G F u J 0 ( h ) d z .

Notice that the right-hand members of (3.7) vanish at ( u , q , h ) = ( 0 , 0 , 0 ) .

4 Maximal regularity

To simplify the notation, in the following, we write D ( u ) 2 1 ( u + [ u ] ) for every vector field u . The principal part of the linearized problem reads as follows:

(4.1) t u μ Δ u + 2 ω ( e 3 × u ) + q = f u , in  F , div u = g d , in  F , P G ( 2 μ D ( u ) ν G ) = g u τ , on  G , 2 μ D ( u ) ν G ν G q + G h = g u v , on  G , t h ( P 0 G u ) ν G = f h , on  G , u ( 0 ) = u 0 , in  F , h ( 0 ) = h 0 , on  G ,

where we have set

P 0 G u = u 1 F F u ( y , t ) d y .

We define the function spaces

E δ ( J ; F ) E 1 , δ ( J ; F ) × E 2 , δ ( J ; F ) × E 3 , δ ( J ; G ) × E 4 , δ ( J ; G ) , E 1 , δ ( J ; F ) H δ 1 , p ( J ; L q ( F ) 3 ) L δ p ( J ; H 2 , q ( F ) 3 ) , E 2 , δ ( J ; F ) L δ p ( J ; H ˙ 1 , q ( F ) ) , E 3 , δ ( J ; G ) F p , q , δ 1 / 2 1 / ( 2 q ) ( J ; L q ( G ) ) L δ p ( J ; B q , q 1 1 / q ( G ) ) , E 4 , δ ( J ; G ) F p , q , δ 2 1 / q ( J ; L q ( G ) ) H δ 1 , p ( J ; B q , q 2 1 / q ( G ) ) L δ p ( J ; B q , q 3 1 / q ( G ) ) , F δ ( J ; F ) F 1 , δ ( J ; F ) × F 2 , δ ( J ; F ) × F 3 , δ ( J ; F ) 2 × F 4 , δ ( J ; G ) , F 1 , δ ( J ; F ) L δ p ( J ; L q ( F ) 3 ) , F 2 , δ ( J ; F ) H δ 1 , p ( J ; H ˙ 1 , q ( F ) ) L δ p ( J ; DI q ( F ) ) , F 3 , δ ( J ; G ) F p , q , δ 1 / 2 1 / ( 2 q ) ( J ; L q ( G ) ) L δ p ( J ; B q , q 1 1 / q ( G ) ) , F 4 , δ ( J ; G ) F p , q , δ 1 1 / ( 2 q ) ( J ; L q ( G ) ) L δ p ( J ; B q , q 2 1 / q ( G ) )

for J R + . The main theorem of this section states that problem (4.1) has maximal regularity.

Theorem 4.1

Let 1 < p , q < , 1 / p < δ 1 , and 1 / p + 1 / ( 2 q ) δ 1 / 2 . There exists a constant β 0 > 0 such that for all β β 0 , problem (4.1) has a unique solution ( u , q , Tr G [ q ] , h ) e β t E δ ( R + ; F ) if and only if

  1. ( u 0 , h 0 ) B q , p 2 ( δ 1 / p ) ( F ) 3 × B q , p 2 + δ 1 / p 1 / q ( G ) ;

  2. ( f u , g d , g u τ , g u v , f h ) e β t F δ ( R + ; F ) ;

  3. g d t = 0 = div u 0 ;

  4. g u τ t = 0 = P G ( 2 μ D ( u ) ν G ) if 1 / p + 1 / ( 2 q ) < δ 1 / 2 .

Furthermore, the solution ( u , q , Tr G [ q ] , h ) enjoys the estimate

e β t ( u , q , Tr G [ q ] , h ) E δ ( R + ; F ) C ( u 0 B q , p 2 ( δ 1 / p ) ( F ) + η 0 B q , p 2 + δ 1 / p 1 / q ( G ) + e β t ( f u , g d , g u τ , g u v , f h ) F δ ( R + ; F ) )

with some constant C independent of ω , β 0 , β , and T .

To prove Theorem 4.1, we consider the corresponding resolvent problem:

(4.2) λ u ^ μ Δ u ^ + 2 ω ( e 3 × u ^ ) + q ^ = f ^ u , in  F , div u ^ = g ^ d , in  F , P G ( 2 μ D ( u ^ ) ν G ) = g ^ u τ , on  G , 2 μ D ( u ^ ) ν G ν G q ^ + G h ^ = g ^ u v , on  G , λ h ^ ( P 0 G u ^ ) ν G = f ^ h , on  G ,

which can be obtained by the Laplace transform, with respect to time t , applied to the system (4.1). Here, λ is the resolvent parameter varying in

Σ ε , λ 0 { λ C : arg λ π ε , λ λ 0 }

for ε ( 0 , π / 2 ) and λ 0 > 0 . By using the result obtained by Shibata [27, Thm. 4.8], we show the existence of -bounded solution operator families for (4.2). This section is mainly devoted to the proof of the following lemma.

Lemma 4.2

Assume 1 < q < and 0 < ε < π / 2 . Let

X q ( f ^ u , g ^ d , g ^ u τ , g ^ u v , f ^ h ) : f ^ u L q ( F ) 3 , g ^ d DI q ( F ) , g ^ u τ H 1 , q ( F ) 2 , g ^ u v H 1 , q ( F ) , f ^ h B q , q 2 1 / q ( G ) , X q F = ( F 1 , , F 9 ) : F 1 , F 4 L q ( F ) 3 , F 2 L q ( F ) , F 3 H 1 , q ( F ) , F 5 , F 7 L q ( F ) 2 , F 6 , F 8 H 1 , q ( F ) 2 , F 9 B q , q 2 1 / q ( G ) . .

Then there exists a constant λ 1 C 0 max ( ω 2 , 1 ) with some constant C 0 and families of operators U ω ( λ ) , P ω ( λ ) , and ω ( λ ) with

U ω ( λ ) Hol ( Σ ε , λ 1 ; ( X q , H 2 , q ( F ) 3 ) ) , Q ω ( λ ) Hol ( Σ ε , λ 1 ; ( X q , H 1 , q ( F ) ) ) , ω ( λ ) Hol ( Σ ε , λ 1 ; ( X q , B q , q 3 1 / q ( G ) ) ) ,

such that ( u ^ , q ^ , η ^ ) ( U ω ( λ ) , P ω ( λ ) , ω ( λ ) ) F λ G is the unique solution to (4.2), where G and F λ are given by

G ( f ^ u , g ^ d , G ( g ^ d ) , g ^ u τ , g ^ u v , f ^ h ) , F λ G ( f ^ u , λ 1 / 2 g ^ d , g ^ d , λ G ( g ^ d ) , λ 1 / 2 g ^ u τ , g ^ u τ , λ 1 / 2 g ^ u v , g ^ u v , f ^ h ) .

Besides, it holds

(4.3) X q H 2 j , q ( F ) 3 ( { ( τ τ ) ( λ j / 2 U ω ( λ ) ) : λ Σ ε , λ 1 } ) c , X q L q ( F ) 3 ( { ( τ τ ) Q ω ( λ ) : λ Σ ε , λ 1 } ) c , X q B q , q 3 1 / q k ( G ) ( { ( τ τ ) ( λ k ω ( λ ) ) : λ Σ ε , λ 1 } ) c

for = 0 , 1 , j = 0 , 1 , 2 , k = 0 , 1 , and τ Im λ , where c is independent of ω .

Proof

According to Shibata [27, Thm. 4.8], we know the existence of -bounded solution operator for the following problem:

(4.4) λ u ^ μ Δ u ^ + q ^ = f ^ u , in  F , div u ^ = g ^ d , in  F , P G ( 2 μ D ( u ^ ) ν G ) = g ^ u τ , on  G , 2 μ D ( u ^ ) ν G ν G q ^ σ H ( 0 ) h ^ = g ^ u v , on  G , λ h ^ ( P 0 G u ^ ) ν G = f ^ h , on  G .

More precisely, he prove that there exist a constant λ 0 and families of operators U 0 ( λ ) , P 0 ( λ ) , and 0 ( λ ) with

U 0 ( λ ) Hol ( Σ ε , λ 0 ; ( X q , H 2 , q ( F ) 3 ) ) , Q 0 ( λ ) Hol ( Σ ε , λ 0 ; ( X q , H 1 , q ( F ) ) ) , 0 ( λ ) Hol ( Σ ε , λ 0 ; ( X q , B q , q 3 1 / q ( G ) ) )

such that for every λ Σ ε , λ 0 and G X q , the triplet ( u ^ , q ^ , h ^ ) = ( U 0 ( λ ) , Q 0 ( λ ) , 0 ( λ ) ) F λ G is a unique solution to problem (4.4) satisfying

(4.5) X q H 2 j , q ( F ) 3 ( { ( τ τ ) ( λ j / 2 U 0 ( λ ) ) : λ Σ ε , λ 0 } ) c 0 , X q L q ( F ) 3 ( { ( τ τ ) Q 0 ( λ ) : λ Σ ε , λ 0 } ) c 0 , X q B q , q 3 1 / q k ( G ) ( { ( τ τ ) ( λ k 0 ( λ ) ) : λ Σ ε , λ 0 } ) c 0

for = 0 , 1 , j = 0 , 1 , 2 , k = 0 , 1 , and τ Im λ . Here, λ 0 and c 0 are independent of ω . Then we see that the solution ( u ^ , q ^ , h ^ ) of (4.4) satisfies

λ u ^ μ Δ u ^ + 2 ω ( e 3 × u ^ ) + q ^ = f ^ u + 2 ω ( e 3 × u ^ ) , in  F , div u ^ = g ^ d , in  F , P G ( 2 μ D ( u ^ ) ν G ) = g ^ u τ , on  G , 2 μ D ( u ^ ) ν G ν G q ^ + G h ^ = g ^ u v + ω 2 ( y ν G ) h ^ , on  G , λ h ^ ( P 0 G u ^ ) ν G = f ^ h , on  G .

Now we define

R 1 ( λ ) G 2 ω ( e 3 × U 0 ( λ ) F λ G ) , R 1 ( λ ) F 2 ω ( e 3 × U 0 ( λ ) F ) , R 2 ( λ ) G ω 2 ( y ν G ) 0 ( λ ) F λ G , R 2 ( λ ) F ω 2 ( y ν G ) 0 ( λ ) F

for G X q and F X q . Setting R ( λ ) ( R 1 ( λ ) , 0 , 0 , R 2 ( λ ) , 0 ) and R ( λ ) ( R 1 ( λ ) , 0 , 0 , R 2 ( λ ) , 0 ) , we have the relation

(4.6) R ( λ ) = R ( λ ) F λ ,

which maps from X q to X q . Since it holds

2 ω ( e 3 × u ^ ) L q ( F ) 2 ω u ^ L q ( F ) , ω 2 ( y ν G ) h ^ H 1 , q ( F ) C ω 2 y ν G H 1 , ( F ) h ^ H 1 , q ( F ) C q , G ω 2 h ^ B q , q 2 1 / q ( G )

for ( u ^ , h ^ ) H 2 , q ( F ) 3 × B q , q 3 1 / q ( G ) , q ( 1 , ) , it follows from Proposition 2.1 and (4.5) that

X q ( { ( τ τ ) F λ R ( λ ) : λ Σ ε , λ 1 } ) c 0 ( 2 ω λ 1 1 + C q , G ω 2 λ 1 1 )

for any λ 1 λ 0 . We shall choose λ 1 large enough such that

λ 1 4 ( C q , G + 1 ) c 0 max ( 1 , ω 2 ) C 0 max ( 1 , ω 2 ) .

Then, we have

(4.7) X q ( { ( τ τ ) F λ R ( λ ) : λ Σ ε , λ 1 } ) 1 2 .

Hence, from (4.6) and (4.7), we obtain

F λ R ( λ ) G X q 1 2 F λ G X q .

Namely, we have F λ R ( λ ) F λ 1 ( X q ) with F λ R ( λ ) F λ 1 ( X q ) 1 / 2 . Then, the Neumann series argument implies the existence of the inverse ( I F λ R ( λ ) F λ 1 ) 1 of F λ R ( λ ) F λ 1 in ( X q ) . Hence, the operator F λ 1 ( I F λ R ( λ ) F λ 1 ) 1 F λ = ( I R ( λ ) ) 1 exists in ( X q ) for each λ Σ ε , λ 1 . Setting

U ω ( λ ) U 0 ( λ ) ( I R ( λ ) ) 1 , Q ω ( λ ) Q 0 ( λ ) ( I R ( λ ) ) 1 , ω ( λ ) 0 ( λ ) ( I R ( λ ) ) 1 ,

we see that ( U ω ( λ ) , Q ω ( λ ) , ω ( λ ) ) F λ G solves (4.2) such that the estimates (4.3) are valid with c = 4 c 0 . Finally, the uniqueness of solutions to (4.2) follows from the duality argument. Suppose that u ^ H 2 , q ( F ) 3 , q ^ H 1 , q ( F ) , and h ^ B q , q 3 1 / q ( G ) satisfy problem (4.2) with ( f ^ u , g ^ d , g ^ u τ , g ^ u v , f ^ h ) vanishing. For any λ Σ ε , λ 1 , Φ C c ( F ) 3 , we consider that

(4.8) λ ¯ v ^ μ Δ v ^ + 2 ( ω ) ( e 3 × v ^ ) + p ^ = Φ , in  F , div v ^ = 0 , in  F , P G ( 2 μ D ( v ^ ) ν G ) = 0 , on  G , 2 μ D ( v ^ ) ν G ν G p ^ + G θ ^ = 0 , on  G , λ ¯ θ ^ ( P 0 G v ^ ) ν G = 0 , on  G ,

where λ ¯ stands for the complex conjugate of λ . Then, according to the aforementioned discussion, there exists a solution ( v ^ , p ^ , θ ^ ) H 2 , q ( F ) 3 × H 1 , q ( F ) × B q , q 2 1 / q ( G ) of (4.8) with 1 < q < . Hence, the divergence theorem gives

( u ^ , Φ ) F = ( u ^ , λ ¯ v ^ ) F ( u ^ , μ Δ v ^ + p ^ ) F + ( u ^ , 2 ( ω ) ( e 3 × v ^ ) ) F = ( u ^ , λ ¯ v ^ ) F ( u ^ , ( 2 μ D ( v ^ ) p ^ I ) ν G ) G + 2 μ ( D ( u ^ ) , D ( v ^ ) ) F + ( 2 ω ( e 3 × u ^ ) , v ^ ) F = ( u ^ , λ ¯ v ^ ) F ( u ^ ν G , G θ ^ ) G + 2 μ ( D ( u ^ ) , D ( v ^ ) ) F + ( 2 ω ( e 3 × u ^ ) , v ^ ) F .

Here, we have used the identity ( e 3 × u ^ , v ^ ) F = ( u ^ , e 3 × v ^ ) F . By ( 4.8 ) 5 , it holds

( u ^ ν G , G θ ^ ) G = ( λ h ^ , G θ ^ ) G + 1 F F v ^ d y , G θ ^ G = ( λ h ^ , G θ ^ ) G ,

since integrating ( 4.8 ) 5 over G gives ( 1 , θ ^ ) G = 0 due to λ 0 . Noting the self-adjointness of G , we deduce that

( u ^ , Φ ) F = ( u ^ , λ ¯ v ^ ) F + ( λ G h ^ , θ ^ ) G + 2 μ ( D ( u ^ ) , D ( v ^ ) ) F + ( 2 ω ( e 3 × u ^ ) , v ^ ) F .

Analogously, we obtain

0 = ( λ u ^ μ Δ u ^ + 2 ω ( e 3 × u ^ ) + q ^ , v ^ ) F = ( u ^ , λ ¯ v ) F ( ( 2 μ D ( u ^ ) p ^ I ) ν G , v ^ ) G + 2 ( D ( u ^ ) , D ( v ^ ) ) F + ( 2 ω ( e 3 × u ^ ) , v ^ ) F = ( u ^ , λ ¯ v ) F ( G h ^ , v ^ ν G ) G + 2 ( D ( u ^ ) , D ( v ^ ) ) F + ( 2 ω ( e 3 × u ^ ) , v ^ ) F = ( u ^ , λ ¯ v ) F ( G h ^ , λ ¯ θ ^ ) G + 2 ( D ( u ^ ) , D ( v ^ ) ) F + ( 2 ω ( e 3 × u ^ ) , v ^ ) F = ( u ^ , λ ¯ v ) F ( λ h ^ , G θ ^ ) G + 2 ( D ( u ^ ) , D ( v ^ ) ) F + ( 2 ω ( e 3 × u ^ ) , v ^ ) F = ( u ^ , λ ¯ v ) F ( u ^ ν G , G θ ^ ) G + 2 ( D ( u ^ ) , D ( v ^ ) ) F + ( 2 ω ( e 3 × u ^ ) , v ^ ) F .

Hence, we arrive at ( u ^ , Φ ) F = 0 for any Φ C c ( F ) 3 . Since Φ is arbitrary, we obtain u ^ = 0 in F . Then, it follows from the equation ( 4.2 ) 5 that h ^ = 0 on G , as λ 0 . This completes the proof.□

Using the operator valued Fourier multiplier theorem due to Prüss [21], we find that Theorem 4.1 immediately follows from Lemma 4.2, see also the discussion in the proof of [39, Thm. 3.5].

5 Decay estimate for the linearized equations

In this section, we shall address some exponential decay property of the linearized system (4.1). To this end, let { φ m } m = 1 4 be an orthogonal basis of N ( G ) C with respect to L 2 inner-product ( , ) G , where N ( G ) stands for the null space of G . Here, φ m , m = 1 , 2 , 3 , 4 , are given by φ 1 = G 1 / 2 , and φ + 1 = C y , = 1 , 2 , 3 , respectively, normalized by ( φ j , φ k ) G = δ j k , where C are constants. Assume that λ 1 is the same constant as in Lemma 4.2 in what follows. This section is dedicated to show the following theorem.

Theorem 5.1

Let 1 < p , q < , 1 / p < δ 1 , and 1 / p + 1 / ( 2 q ) δ 1 / 2 . Set J = ( 0 , T ) with T > 0 . Then, there exists a constant ε 0 > 0 such that the following assertion is valid: Let u 0 B q , p 2 ( δ 1 / p ) ( F ) 3 , h 0 B q , p 2 + δ 1 / p 1 / q ( G ) , and ( f u , g d , g u τ , g u v , f h ) F δ ( J ; F ) that satisfy the compatibility conditions given in Theorem 4.1. The problem (4.1) admits a unique solution ( u , q , Tr G [ q ] , h ) E δ ( J ; F ) possessing the estimate

e ε 0 t ( u , q , Tr G [ q ] , h ) E δ ( J ; F ) C u 0 B q , p 2 ( δ 1 / p ) ( F ) + h 0 B q , p 2 + δ 1 / p 1 / q ( G ) + e ε 0 t ( f u , g d , g u τ , g u v , f h ) F δ ( J ; F ) + 0 T ( e ε 0 s ( u ( , s ) , 1 ) F ) p d s 1 / p + α = 1 , 2 0 T e ε 0 s ( u ( , s ) , e α × y ) F ω G h ( , s ) y α y 3 d G p d s 1 / p + 0 T ( e ε 0 s ( u ( , s ) , e 3 × y ) F ) p d s 1 / p + m = 1 4 0 T ( e ε 0 s ( h ( , s ) , φ m ) G ) p d s 1 / p

with some constant C independent of T.

We first decompose ( u , q , h ) of (4.1). To this end, we recall a unique solvability result of the weak Dirichlet problem: For any f L q ( F ) 3 , 1 < q < , there exists a unique θ H ˙ 0 1 , q ( F ) satisfying

(5.1) ( θ , φ ) F = ( f , φ ) F for any  φ H ˙ 0 1 , q ( F )

and possessing the estimate θ L q ( F ) C f L q ( F ) with a constant C independent of the choices of θ , φ , and f . Then, for any f L q ( F ) 3 , we define the operators P F and Q F by Q F f θ and P F f ( I Q F ) f J q ( F ) with θ H ˙ 0 1 , q ( F ) satisfying (5.1). Here, J q ( F ) is the solenoidal space given by

J q ( F ) { f L q ( F ) 3 : ( f , φ ) F = 0 for all  φ H ˙ 0 1 , q ( F ) } .

Next, we consider the following systems:

(5.2) t v + 2 λ 2 v μ Δ v + 2 ω ( e 3 × v ) + π = f u , in  F , div v = g d , in  F , P G ( 2 μ D ( v ) ν G ) = g u τ , on  G , 2 μ D ( v ) ν G ν G π + G η = g u v , on  G , t η + 2 λ 2 η ( P 0 G u ) ν G = f h , on  G , v ( 0 ) = u 0 , in  F , η ( 0 ) = h 0 , on  G .

(5.3) t u μ Δ u + 2 ω ( e 3 × u ) + q = 2 λ 2 P F v ˜ , in  F , div u = 0 , in  F , P G ( 2 μ D ( u ) ν G ) = 0 , on  G , 2 μ D ( u ) ν G ν G q + G h = 0 , on  G , t h ( P 0 G u ) ν G = 2 λ 2 η ˜ , on  G , u ( 0 ) = 0 , in  F , h ( 0 ) = 0 , on  G .

Here, we have set

v ˜ v ( v , 1 ) F α = 1 , 2 ( v , e α × y ) F ω G η ˜ y α y 3 d y ( e α × y ) ( v , e 3 × y ) F ( e 3 × y )

and η ˜ η m = 1 4 ( η , φ m ) G φ m , respectively. Now, setting

(5.4) u v + u + 2 λ 2 0 t ( v ( s ) , 1 ) F + α = 1 , 2 ( v ( s ) , e α × y ) F ω G η ˜ ( s ) y α y 3 d y ( e α × y ) + ( v ( s ) , e 3 × y ) F ( e 3 × y ) d s ,

and

(5.5) q π + q 2 λ 2 Q F v ˜ , h η + h + 2 λ 2 m = 1 4 0 t ( η ( s ) , φ m ) G φ m d s ,

we see that ( u , q , h ) is a solution to (4.1). In the following, we construct the solutions of (5.2) and (5.3), respectively.

Step 1: Study of (5.2). For the shifted system (5.2), we have the next theorem.

Theorem 5.2

Assume that 1 < p , q < , 1 / p < δ 1 , and 1 / p + 1 / ( 2 q ) δ 1 / 2 . Set J = ( 0 , T ) with T > 0 . Let u 0 B q , p 2 ( δ 1 / p ) ( F ) 3 , h 0 B q , p 2 + δ 1 / p 1 / q ( G ) , and ( f u , g d , g u τ , g u v , f h ) F δ ( J ; F ) that satisfy the compatibility conditions given in Theorem 4.1. Then, for any λ 2 > λ 1 / 2 , system (5.2) has a unique solution ( v , π , Tr G [ π ] , η ) E δ ( J ; F ) possessing the estimate

( v , π , Tr F [ π ] , η ) E δ ( J ; F ) C ( u 0 B q , p 2 ( δ 1 / p ) ( F ) + h 0 B q , p 2 + δ 1 / p 1 / q ( G ) + ( f u , g d , g u τ , g u v , f h ) F δ ( J ; F ) )

for some constant C independent of T , ω , λ 1 , and λ 2 .

Proof

For λ Σ ε , λ 1 , let U ω ( λ ) , Q ω ( λ ) , and ω ( λ ) be operators given in Lemma 4.2. Let λ 2 and ε 1 be numbers for which 2 λ 2 λ 1 > ε 1 > 0 . Then, for every λ ε 1 + Σ ε , 0 , it holds λ + 2 λ 2 λ 1 + Σ ε , 0 . Hence, by Lemma 4.2, we obtain

X q H 2 j , q ( F ) 3 ( { ( τ τ ) ( λ j / 2 U ω ( λ + 2 λ 2 ) ) : ε 1 + λ Σ ε , 0 } ) c , X q L q ( F ) 3 ( { ( τ τ ) Q ω ( λ + 2 λ 2 ) : ε 1 + λ Σ ε , 0 } ) c , X q B q , q 3 1 / q k ( G ) ( { ( τ τ ) ( λ k ω ( λ + 2 λ 2 ) ) : ε 1 + λ Σ ε , 0 } ) c

for = 0 , 1 , j = 0 , 1 , 2 , and k = 0 , 1 . Employing the argument in the proof of [39, Thm. 3.5] readily implies the required assertion. This completes the proof.□

For every ε 0 > 0 , we see that e ε 0 t ( v , π , η ) satisfies

t ( e ε 0 t v ) + ( 2 λ 2 ε 0 ) ( e ε 0 v ) μ Δ ( e ε 0 t v ) + 2 ω ( e 3 × ( e ε 0 t v ) ) + ( e ε 0 t π ) = e ε 0 t f u , in  F , div ( e ε 0 t v ) = e ε 0 t g d , in  F , P G ( 2 μ D ( e ε 0 t v ) ν G ) = e ε 0 t g u τ , on  G , 2 μ D ( e ε 0 t v ) ν G ν G ( e ε 0 t π ) + G ( e ε 0 t η ) = e ε 0 t g u v , on  G , t ( e ε 0 t η ) + ( 2 λ 2 ε 0 ) ( e ε 0 t η ) ( P 0 G ( e ε 0 t v ) ) ν G = e ε 0 t f h , on  G , v ( 0 ) = u 0 , in  F , η ( 0 ) = h 0 , on  G .

Given ε 0 > 0 , we choose λ 2 > 0 suitably large such that 2 λ 2 λ 1 > ε 0 > 0 . Then, from Theorem 5.2, we obtain the following corollary, which gives the decay property of solutions to (5.2).

Corollary 5.3

Let 1 < p , q < , 1 / p < δ 1 , and 1 / p + 1 / ( 2 q ) δ 1 / 2 . Set J = ( 0 , T ) with T > 0 . Let u 0 B q , p 2 ( δ 1 / p ) ( F ) 3 , h 0 B q , p 2 + δ 1 / p 1 / q ( G ) , and ( f u , g d , g u τ , g u v , f h ) F δ ( J ; F ) that satisfy the compatibility conditions given in Theorem 4.1. There exists a constant λ 2 > 0 such that the system (5.2) admits a unique solution ( v , π , Tr G [ π ] , η ) E δ ( J ; F ) possessing the estimate

e ε 0 t ( v , π , Tr G [ π ] , η ) E δ ( J ; F ) C ( u 0 B q , p 2 ( δ 1 / p ) ( F ) + h 0 B q , p 2 + δ 1 / p 1 / q ( G ) + e ε 0 t ( f u , g d , g u τ , g u v , f h ) F δ ( J ; F ) )

with a constant C independent of T , ω , λ 1 , and λ 2 , where ε 0 is a number such that 0 < ε 0 < 2 λ 2 λ 1 .

Step 2: Study of (5.3). To study (5.3), we rely on the following decomposition:

(5.6) u ( y , t ) = U ( y , t ) + d 3 [ h ] ( e 3 × y ) , q ( y , t ) = Q ( y , t ) + ω d 3 [ h ] y 2 + 1 G G C G h d G ,

where we have set

d [ h ] ω S G h ( y , t ) ( e 3 × y ) ( e × y ) d G , S e × y L 2 ( F ) 2 = F ( y 2 y 2 ) d y , C G h G h ω d 3 [ h ] y 2

with = 1 , 2 , 3 . We further set

C ^ G h C G h 1 G G C G h d G .

Noting that

2 e 3 × ( e 3 × y ) = ( 2 y 1 , 2 y 2 , 0 ) = y 2 , t d 3 [ h ] = ω S 3 G t h y 2 d G = ω S 3 G ( ( P 0 G U ) ν G + 2 λ 2 η ˜ ) y 2 d G ,

the system (5.3) can be rewritten as follows:

(5.7) t U L ω , y U + Q = 2 λ 2 f ˜ w , in  F , div U = 0 , in  F , P G ( 2 μ D ( U ) ν G ) = 0 , on  G , 2 μ D ( U ) ν G ν G Q + C ^ G h = 0 , on  G , t h ( P 0 G U ) ν G = 2 λ 2 η ˜ , on  G , U ( 0 ) = 0 , in  F , h ( 0 ) = 0 , on  G ,

where we have set

L ω , y w 2 ω S 3 F ( P 0 G w ) y d y ( e 3 × y ) + μ Δ w 2 ω ( e 3 × w ) , f ˜ w P F v ˜ + ω S 3 G η ˜ y 2 d G ( e 3 × y ) .

Notice that it holds ( e 3 × y ) ν G = 0 on G due to the axisymmetry of G .

As a base space for our analysis, we use

X 0 = J q ( F ) × B q , q 2 1 / q ( G ) ,

and we set

X ¯ 1 H 2 , q ( F ) 3 × B q , q 3 1 / q ( G ) .

Define a closed linear operator in X 0 by means of

A q ( U , h ) ( L ω , y U + K ( U , h ) , ( P 0 G U ) ν G )

with domain X 1 D ( A q ) X ¯ 1 defined by

D ( A q ) { ( U , h ) X 0 X ¯ 1 : P G ( 2 μ D ( U ) ν G ) = 0 on  G } .

Here, K ( U , h ) H ˙ 0 1 , q ( F ) is a functional that is a unique solution to the weak Dirichlet problem

(5.8) ( K ( U , h ) , φ ) F = ( L ω , y U , φ ) F for any  φ H ˙ 0 1 , q ( F ) , K ( U , h ) = 2 μ ( D ( U ) ν G ) ν G + C ^ G h on  G

for ( U , h ) X 0 X ¯ 1 . Since ( U , h ) belongs to X 0 X ¯ 1 , we have

L ω , y U L q ( F ) C q , F ( μ U H 2 , q ( F ) + ω U L q ( F ) ) , 2 μ ( D ( U ) ν G ) ν G + C ^ G h B q , q 1 1 / q ( G ) C q , G ( μ U H 2 , q ( F ) + σ h B q , q 3 1 / q ( G ) + ω 2 h B q , q 2 1 / q ( G ) ) ,

which yields the estimate for K ( U , h ) :

K ( U , h ) L q ( F ) C q , G ( μ U H 2 , q ( F ) + ω U L q ( F ) + σ h B q , q 3 1 / q ( G ) + ω 2 h B q , q 2 1 / q ( G ) ) .

As F is a bounded smooth domain, the weak Dirichlet problem (5.8) admits a unique solution K ( U , h ) H ˙ 0 1 , q ( F ) , i.e., the functional K ( U , h ) is well defined. Notice that the solution K ( U , h ) to (5.8) depends on ω . Applying the similar argument in [25, Sec. 9.2.1], we see that the system (5.7) is equivalent to the abstract evolution equation

z ˙ + A q z = ( 2 λ 2 f ˜ w , 2 λ 2 η ˜ ) , t > 0 , z ( 0 ) = 0

with z = ( U , h ) . Employing the standard Neumann series argument, we can prove that the operator A q generates an analytic C 0 -semigroup in X 0 .

Lemma 5.4

Let 1 < q < . Then A q with domain D ( A q ) generates an analytic C 0 -semigroup in X 0 .

Proof

Let us consider the following resolvent problem:

(5.9) λ U ^ L ω , y U ^ + Q ^ = f ^ u , in  F , div U ^ = 0 , in  F , P G ( 2 μ D ( U ^ ) ν G ) = 0 , on  G , 2 μ D ( U ^ ) ν G ν G Q ^ + C ^ G h ^ = g ^ u v , on  G , λ h ^ ( P 0 G U ^ ) ν G = f ^ h , on  G

for any λ Σ ε , λ 4 and ( f ^ u , 0 , 0 , g ^ u v , f ^ h ) X q , σ with some positive constant λ 4 , where we have set

X q , σ X q ( J q ( F ) × DI q ( F ) × H 1 , q ( F ) 2 × H 1 , q ( F ) × B q , q 2 1 / q ( G ) ) .

To find solutions of the aforementioned problem, let us consider

(5.10) λ u ^ μ Δ u ^ + 2 ω ( e 3 × u ^ ) + q ^ = f ^ u , in  F , div u ^ = 0 , in  F , P G ( 2 μ D ( u ^ ) ν G ) = 0 , on  G , 2 μ D ( u ^ ) ν G ν G q ^ + G h ^ = g ^ u v , on  G , λ h ^ ( P 0 G u ^ ) ν G = f ^ h , on  G

for ( f ^ u , 0 , 0 , g ^ u v , f ^ h ) X q , σ and λ Σ ε , λ 2 . According to Lemma 4.2, there exists a unique solution

( u ^ , q ^ , h ^ ) ( H 2 , q ( F ) 3 J q ( F ) ) × H 1 , q ( F ) × B q , q 3 1 / q ( G )

of (5.10). Let U ω and ω be the operators given in Lemma 4.2. If we set q ˜ q ^ G 1 G C G h ^ d G , for λ Σ ε , λ 2 , we see that ( u ^ , q ˜ , h ^ ) satisfies

λ u ^ L ω , y u ^ + q ˜ = f ^ u + 2 ω S 3 F ( P 0 G U ω ( λ ) F 0 ) y d y ( e 3 × y ) , in  F , div u ^ = 0 , in  F , P G ( 2 μ D ( u ^ ) ν G ) = 0 , on  G , 2 μ D ( u ^ ) ν G ν G q ˜ + C ^ G h ^ = f ^ u v + ω d 3 [ ω ( λ ) F 0 ] y 2 , on  G , λ h ^ ( P 0 G u ^ ) ν G = f ^ h , on  G ,

where we have set F 0 = ( f ^ u , 0 , , 0 , λ 1 / 2 g ^ u v , g ^ u v , f ^ h ) X q . Since it holds

2 ω S 3 F ( P 0 G U ω ( λ ) F 0 ) y d y ( e 3 × y ) L q ( F ) C 1 ω λ 3 1 F 0 X q , ω d 3 [ ω ( λ ) F 0 ] y 2 B q , q 2 1 / q ( G ) C 1 ω 2 λ 3 1 F 0 X q

for any λ 3 λ 2 , we shall take λ 3 sufficiently large such that λ 3 4 C 1 max ( ω 2 , 1 ) so that it follows from the Neumann series argument that (5.9) admits a unique solution

( U ^ , Q ^ , h ^ ) ( H 2 , q ( F ) 3 J q ( F ) ) × H 1 , q ( F ) × B q , q 3 1 / q ( G )

with λ 4 λ 3 , where the resolvent estimate

( λ U ^ , λ 1 / 2 U ^ , 2 U ^ ) L q ( F ) + Q ^ L q ( F ) + ( λ h ^ , h ^ ) B q , q 2 1 / q ( G ) C ( f ^ u L q ( F ) + λ 1 / 2 f ^ u v L q ( F ) + f ^ u v H 1 , q ( F ) + f ^ h B q , q 2 1 / q ( G ) )

is valid with a constant C depending only on q , G , μ , and σ . If we set f ^ u v 0 , the resolvent problem

λ z ^ + A q z ^ = f ^ with  z ^ = ( U ^ , h ^ )   and  f ^ = ( f ^ u , f ^ h )

is equivalent to (5.9) with Q ^ = K ( U ^ , h ^ ) and admits a unique solution z ^ X 1 satisfying the resolvent estimate

λ z ^ X 0 + z ^ X ¯ 1 C f ^ X 0 , λ Σ ε , λ 4 .

Finally, the closedness of A q follows from the fact that the resolvent set is not empty. This completes the proof.□

To show the exponential decay property of the system (5.7), we introduce the following functional spaces: Let B ˜ q , q 2 1 / q ( G ) be the set of all g B q , q 2 1 / q ( G ) satisfying ( g , φ m ) G = 0 for m = 1 , 2 , 3 , 4 , i.e.,

G g d G = G g y d G = 0 , ( = 1 , 2 , 3 ) .

The subspace J ˜ q ( F ) of J q ( F ) stands for the set of all f J q ( F ) satisfying the orthogonal conditions

F f d y = F f ( e 3 × y ) d y = 0 , F f ( e α × y ) d y = ω G g ˜ y α y 3 d G , ( α = 1 , 2 ) ,

where g ˜ B ˜ q , q 2 1 / q ( G ) . We then define

X ˜ 0 J ˜ q ( F ) B ˜ q , q 2 1 / q ( G ) .

We set the restriction operator A ˜ q A q X ˜ 0 with its domain given by D ( A ˜ q ) D ( A q ) X ˜ 0 . Then the operator A ˜ q generates an analytic C 0 -semigroup on X ˜ 0 .

Lemma 5.5

Let 1 < q < . The induced operator A ˜ q A q X ˜ 0 with domain D ( A ˜ q ) D ( A q ) X ˜ 0 is the generator of an analytic C 0 -semigroup in X ˜ 0 .

To prove this lemma, we need the following proposition given in [33, Prop. 2.3].

Proposition 5.6

Let r be a function defined on G and G r be a normal perturbation of G given by

G r { s = y + r ν G ( y ) : y G } ,

where r L ( G ) and G r L ( G ) are assumed to be suitably small such that G r is contained in the tubular neighborhood of G . For arbitrary function ζ ( y ) = a + b × y (with constants a and b ) defined on G , the equality

G C ^ G r ν G ζ d G = ω 2 G r ζ y d G

holds.

Proof

There are only a few words for the proof in [33], but we record the proof here for the reader’s convenience. We denote by F r the closed domain surrounded by G r . We consider the integral

(5.11) I [ r ] G r σ H G r + ω 2 2 s 2 + p 0 ν G r ζ d G r ,

where s ( s 1 , s 2 , 0 ) . Since H G r ν G r = Δ G r s for s G r and G r p 0 ν G r ζ d G r = F r p 0 div ζ d s = 0 , integration by parts and the divergence theorem imply

(5.12) I [ r ] = G r ω 2 2 s 2 ν G r ζ d G r = ω 2 2 F r div ( s 2 ζ ) d G r = ω 2 F r ζ s d s .

We next calculate the first variation of I ( r ) . By using the well known formulas

δ 0 H G r = Δ G r + ( H G 2 2 K G ) r , δ 0 s 2 = 2 ν G y r ,

it follows from (5.11) that

δ 0 I [ r ] = G G r ν G ζ d G = G C ^ G r ν G ζ d G .

Conversely, the first variation of (5.12) gives

δ 0 I [ r ] = ω 2 G r ζ y d G

due to a formula δ 0 F r f d s = G f r d G . Combining the aforementioned equalities, we obtain the required equality.□

We next give the proof of Lemma 5.5.

Proof of Lemma 5.5

First, we show that the closed subspace X ˜ 0 of X 0 is e A q t -invariant, i.e., e A q t X ˜ 0 X ˜ 0 for any t 0 . According to the classical semigroup theory (cf. [20, Thm. 4.5.1]), it suffices to show that there is a real number c such that for every λ > c , the space X ˜ 0 is an invariant subspace of R ( λ ; A q ) , the resolvent of A q . To this end, we shall consider the following resolvent problem

(5.13) λ U ^ L ω , y U ^ + Q ^ = F ˜ , in  F , div U ^ = 0 , in  F , P G ( 2 μ D ( U ^ ) ν G ) = 0 , on  G , 2 μ D ( U ^ ) ν G ν G Q ^ + C ^ G h ^ = 0 , on  G , λ h ^ ( P 0 G U ^ ) ν G = G ˜ , on  G

for given ( F ˜ , G ˜ ) X ˜ 0 . Here, λ Σ ε , λ is a resolvent parameter, where ε ( 0 , π / 2 ) and λ max ( λ 4 , 3 ω ) are constants. From Lemma 5.4, the resolvent set ρ ( A q ) of A q contains Σ ε , λ .

Integrating (5.13) 5 , for m = 1 , 2 , 3 , 4 , it holds

0 = ( G ˜ , φ m ) G = ( λ h ^ ( P 0 G U ^ ) ν G , φ m ) G = λ ( h ^ , φ m ) G F div ( ( P 0 G U ^ ) φ m ) d y .

In the following, we write U ^ ( U ^ ( 1 ) , U ^ ( 2 ) , U ^ ( 3 ) ) . As div ( P 0 G U ^ ) = div U ^ = 0 in F and φ m , = 1 , 2 , 3 , are constants, we observe

F div ( ( P 0 G U ^ ) φ m ) d y = F ( div ( P 0 G U ^ ) ) φ m d y + = 1 3 F U ^ ( ) 1 F F U ^ ( ) d y ( φ m ) d y = 0 .

Since λ 0 , this gives ( h ^ , φ m ) G = 0 for m = 1 , 2 , 3 , 4 .

Employing the similar argument as mentioned earlier, we can show U ^ satisfies the orthogonal conditions. In fact, integrating (5.13) 1 , for = 1 , 2 , 3 , we have

0 = ( F ˜ , e ) F = ( λ U ^ μ Δ U ^ + 2 ω ( e 3 × U ^ ) + Q ^ , e ) F = λ ( U ^ , e ) F + 2 ω ( e 3 × U ^ , e ) F + ( C ^ G h ^ , ν G e ) G .

By Proposition 5.6, it holds

( C ^ G h ^ , ν G e α ) G = ω 2 G h ^ y α d G , ( α = 1 , 2 ) , ( C ^ G h ^ , ν G e 3 ) G = 0 ,

and thus, we obtain

λ F U ^ ( 1 ) d y 2 ω F U ^ ( 2 ) d y ω 2 G h ^ y 1 d G = 0 , λ F U ^ ( 2 ) d y + 2 ω F U ^ ( 1 ) d y ω 2 G h ^ y 2 d G = 0 , λ F U ^ ( 3 ) d y = 0 .

Recalling that h ^ satisfies ( h ^ , φ m ) G = 0 for m = 1 , 2 , 3 , 4 , we obtain ( U ^ , e ) F = 0 , = 1 , 2 , 3 , due to λ 0 , ± 2 i ω .

We next multiply the equation (5.13) 1 by e × y , = 1 , 2 , 3 , and integrate over F , which gives

( F ˜ , e × y ) F = λ ( U ^ , e × y ) F 2 ω δ , 3 F U ^ y d y + 2 ω ( e 3 × U ^ , e × y ) F + ( C ^ G h ^ , ν G ( e × y ) ) G .

Here, δ , 3 stands for the Kronecker delta. Since Proposition 5.6 gives

( C ^ G h ^ , ν G ( e 1 × y ) ) G = ω 2 G h ^ y 2 y 3 d G , ( C ^ G h ^ , ν G ( e 2 × y ) ) G = ω 2 G h ^ y 1 y 3 d G , ( C ^ G h ^ , ν G ( e 3 × y ) ) G = 0 ,

it holds

( F ˜ , e 1 × y ) F = λ ( U ^ , e 1 × y ) F 2 ω F U ^ ( 1 ) y 3 d y + ω 2 G h ^ y 2 y 3 d G , ( F ˜ , e 2 × y ) F = λ ( U ^ , e 2 × y ) F 2 ω F U ^ ( 2 ) y 3 d y ω 2 G h ^ y 1 y 3 d G , ( F ˜ , e 3 × y ) F = λ ( U ^ , e 3 × y ) F .

Conversely, we have

ω G G ˜ y α y 3 d G = ω G ( λ h ^ ( P 0 G U ^ ) ν G ) y α y 3 d G = λ ω G h ^ y α y 3 d G ω F ( U ^ ( α ) y 3 + U ^ ( 3 ) y α ) d y

with α = 1 , 2 . Since F ˜ and G ˜ satisfy the conditions:

F F ˜ ( e α × y ) d y = ω G G ˜ y α y 3 d G , ( α = 1 , 2 ) , F F ˜ ( e 3 × y ) d y = 0 ,

we observe

(5.14) λ ( U ^ , e 1 × y ) F ω G h ^ y 1 y 3 d G ω F ( U ^ ( 1 ) y 3 U ^ ( 3 ) y 1 ) d y ω G h ^ y 2 y 3 d G = 0 , λ ( U ^ , e 2 × y ) F ω G h ^ y 2 y 3 d G + ω F ( U ^ ( 3 ) y 2 U ^ ( 2 ) y 3 ) d y ω G h ^ y 1 y 3 d G = 0 , λ ( U ^ , e 3 × y ) F = 0 .

Noting that

U ^ ( 1 ) y 3 U ^ ( 3 ) y 1 = U ^ ( e 2 × y ) , U ^ ( 2 ) y 3 U ^ ( 3 ) y 2 = U ^ ( e 1 × y ) ,

it follows from ( 5.14 ) 1 , 2 that

λ ( U ^ , e 1 × y ) F ω G h ^ y 1 y 3 d G ω ( U ^ , e 2 × y ) F ω G h ^ y 2 y 3 d G = 0 , λ ( U ^ , e 2 × y ) F ω G h ^ y 2 y 3 d G + ω ( U ^ , e 1 × y ) F ω G h ^ y 1 y 3 d G = 0 .

Hence, by λ 0 , ± i ω , we arrive at

( U ^ , e α × y ) F ω G h ^ y α y 3 d G = 0 , ( α = 1 , 2 ) , ( U ^ , e 3 × y ) F = 0 ,

which implies ( U ^ , h ^ ) D ( A ˜ q ) . Accordingly, we have the R ( λ ; A q ) -invariance of X ˜ 0 for any λ Σ ε , λ , which implies Σ ε , λ ρ ( A ˜ q ) (cf. [10, Prop. A.2.8]). Hence, the C 0 -semigroup { e A q t X ˜ 0 } t 0 is indeed analytic.□

In the following, we denote the induced C 0 -semigroup by

{ e A ˜ q t } t 0 { e A q t X ˜ 0 } t 0 .

As F and G are compact, we can prove that 0 ρ ( A ˜ q ) , which immediately implies the exponential stability of the analytic C 0 -semigroup generated by A ˜ q .

Theorem 5.7

Let 1 < q < . For any ( F ˜ , G ˜ ) X ˜ 0 and t > 0 , there exist positive constants C and β such that

e A ˜ q t ( F ˜ , G ˜ ) X 0 C e β t ( F ˜ , G ˜ ) X 0

is valid, i.e., { e A ˜ q t } t 0 is exponentially stable on X ˜ 0 .

Proof

From Lemma 5.5, there exists λ > 0 such that Σ ε , λ ρ ( A ˜ q ) for ε ( 0 , π / 2 ) . It remains to prove out result for λ Q λ { λ C : Re λ 0 , λ λ } . We define

R R ( 2 λ ; A ˜ q ) : X ˜ 0 X 1 X ˜ 0 X ˜ 0 .

Since F and G are compact, it follows from the Rellich theorem that R is a compact operator from X ˜ 0 into itself. For any λ Q λ , rewriting I + A ˜ q by

( λ I + A ˜ q ) ( F ˜ , G ˜ ) = ( I + ( λ 2 λ ) R ) ( 2 λ I + A q ) ( F ˜ , G ˜ )

for ( F ˜ , G ˜ ) X ˜ 0 , we observe that Q λ ρ ( A ˜ q ) follows from the Fredholm alternative theorem and the injection of I + ( λ 2 λ ) R . To see this, for any λ Q λ , take ( F ˜ , G ˜ ) Ker ( I + ( λ 2 λ ) R ) X ˜ 0 , i.e.,

( I + ( λ 2 λ ) R ) ( F ˜ , G ˜ ) = 0 for any ( F ˜ , G ˜ ) X ˜ 0 .

By the definition of R , we see that ( F ˜ , G ˜ ) belongs to D ( A ˜ q ) and satisfies

(5.15) ( λ + A ˜ q ) ( F ˜ , G ˜ ) = 0 for any λ Q λ .

Notice that the equation (5.15) is equivalent to

(5.16) λ F ˜ L ω , y F ˜ + K ( F ˜ , G ˜ ) = 0 , in  F , P G ( 2 μ D ( F ˜ ) ν G ) = 0 , on  G , 2 μ D ( F ˜ ) ν G ν G K ( F ˜ , G ˜ ) + C ^ G G ˜ = 0 , on  G , λ G ˜ ( P 0 G F ˜ ) ν G = 0 , on  G

with the orthogonal conditions

(5.17) F F ˜ d y = F F ˜ ( e 3 × y ) d y = 0 , F F ˜ ( e α × y ) d y = ω G G ˜ y α y 3 d G , ( α = 1 , 2 ) , G G ˜ d G = G G ˜ y d G = 0 , ( = 1 , 2 , 3 ) .

In the following, we may assume ( F ˜ , G ˜ ) D ( A ˜ 2 ) . In fact, the boundedness of F implies D ( A ˜ q ) D ( A ˜ 2 ) for 2 q < . Besides, when 1 < q 2 , by the bootstrap argument and Sobolev embedding theorem, we see ( F ˜ , G ˜ ) D ( A ˜ 2 ) as well. Using the divergence theorem and (5.17), it follows from (5.16) that

(5.18) 0 = λ F ˜ L 2 ( F ) 2 + 2 μ D ( F ˜ ) L 2 ( F ) 2 + ( C ^ G G ˜ , F ˜ ν G ) G = λ F ˜ L 2 ( F ) 2 + 2 μ D ( F ˜ ) L 2 ( F ) 2 + λ ¯ ( C ^ G G ˜ , G ˜ ) G = λ F ˜ L 2 ( F ) 2 + 2 μ D ( F ˜ ) L 2 ( F ) 2 + λ ¯ Ψ G ( G ˜ , G ˜ ) + ω 2 S 3 G G ˜ y 2 d G 2 ,

where Ψ G is the quadratic form given in (1.12). Now, we decompose F ˜ = F ˜ + α = 1 , 2 d α [ G ˜ ] ( e α × y ) with F ˜ satisfying

F F ˜ d y = F F ˜ ( e × y ) d y = 0 , ( = 1 , 2 , 3 ) ,

and

F ˜ L 2 ( F ) 2 = F ˜ L 2 ( F ) 2 + α = 1 , 2 d α [ G ˜ ] 2 S α .

Then, (5.18) becomes

0 = λ F ˜ L 2 ( F ) 2 + α = 1 , 2 d α [ G ˜ ] 2 S α + 2 μ D ( F ˜ ) L 2 ( F ) 2 + λ ¯ Ψ G ( G ˜ , G ˜ ) + ω 2 S 3 G G ˜ y 2 d G 2 .

Taking the real part yields

0 = 2 μ D ( F ˜ ) L 2 ( F ) 2 + ( Re λ ) F ˜ L 2 ( F ) 2 + α = 1 , 2 d α [ G ˜ ] 2 S α + Ψ G ( G ˜ , G ˜ ) + ω 2 S 3 G G ˜ y 2 d G 2 .

According to Assumption 1.1, we see that Re λ 0 implies D ( F ˜ ) = 0 in F . Hence, it follows from the second Korn inequality that F ˜ = 0 in F . Then, the first equation of (5.16) takes the form

λ α = 1 , 2 d α [ G ˜ ] ( e α × y ) L ω , y α = 1 , 2 d α [ G ˜ ] ( e α × y ) + K ( F ˜ , G ˜ ) = 0 in  F .

Taking the curl in this equation leads to

λ α = 1 , 2 d α [ G ˜ ] e α ω ( d 2 [ G ˜ ] e 1 d 1 [ G ˜ ] e 2 ) = 0 .

Hence, we obtain λ d 1 [ G ˜ ] = ω d 2 [ G ˜ ] and λ d 2 [ G ˜ ] = ω d 1 [ G ˜ ] . Besides, it follows from (5.17) 2 that

λ d α [ G ˜ ] = λ ω S α G G ˜ y α y 3 d G = λ S α F β = 1 , 2 d β [ G ˜ ] ( e β × y ) ( e α × y ) d y , ( α = 1 , 2 ) .

Setting

S ˜ α = F ( y α 2 y 3 2 ) d y , ( α = 1 , 2 ) ,

we easily observe

λ d 1 [ G ˜ ] = ω S ˜ 1 S 1 d 2 [ G ˜ ] = λ S ˜ 1 S 1 d 1 [ G ˜ ] , λ d 2 [ G ˜ ] = ω S ˜ 2 S 2 d 1 [ G ˜ ] = λ S ˜ 2 S 2 d 2 [ G ˜ ] ,

i.e., d 1 [ G ˜ ] = d 2 [ G ˜ ] = 0 , as follows:

S ˜ α + S α = F ( ( y α 2 y 3 2 ) + ( y 2 y α 2 ) ) d y = F y 2 d y 0 .

Thus, we arrive at F ˜ = 0 in F , which, combined with (5.16) 1 , implies that the pressure term K ( F ˜ , G ˜ ) is equal to some constant p 0 . From (5.16) 2 , we have C ^ G G ˜ = p 0 on G , but it follows from G C ^ G G ˜ d G = 0 that p 0 = 0 , i.e., C ^ G G ˜ = 0 on G . Hence, we observe that

0 = ( C ^ G G ˜ , G ˜ ) G = Ψ G ( G ˜ , G ˜ ) + ω 2 S 3 G G ˜ y 2 d G 2 c G ˜ L 2 ( G ) 2 + ω 2 S 3 G G ˜ y 2 d G 2 ,

which implies G ˜ = 0 on G . Therefore, there are no eigenvalues λ Q λ of A ˜ 2 . This completes the proof.□

By using Theorem 5.7, we complete the decay estimate of solutions to (5.3).

Theorem 5.8

Assume that 1 < p , q < , 1 / p < δ 1 , and 1 / p + 1 / ( 2 q ) δ 1 / 2 . Let v and η be functions obtained in Theorem 5.2. Then, the solution ( u , h ) to (5.3) enjoys the estimate

e ε 0 t ( u , h ) E 1 , δ ( J ; F ) × E 4 , δ ( J ; G ) C ( u 0 B q , p 2 ( δ 1 / p ) ( F ) + h 0 B q , p 2 + δ 1 / p 1 / q ( G ) + e ε 0 t ( f u , g d , g u τ , g u v , f h ) F δ ( J ; F ) ) .

with a constant C independent of T.

Proof

By the variation of constants formula, we see that

( U , h ) ( , t ) = 0 t e A ˜ q ( t s ) ( 2 λ 2 f ˜ w ( , s ) , 2 λ 2 η ˜ ( , s ) ) d s

are solutions to (5.7) with Q = K ( U , h ) . Then, Theorem 5.7 yields

( U , h ) ( , t ) X 0 C λ 2 0 t e β ( t s ) ( f ˜ w , η ˜ ) ( , s ) X 0 d s C λ 2 0 t e β ( t s ) d s 1 / p 0 t e β ( t s ) ( f ˜ w , η ˜ ) ( , s ) X 0 p d s 1 / p C λ 2 β 1 / p 0 t e β ( t s ) ( f ˜ w , η ˜ ) ( , s ) X 0 p d s 1 / p .

Hence, for every 0 < ε 0 < β / p , it holds

(5.19) 0 T ( e ε 0 t ( U , h ) ( , t ) X 0 ) p d t C λ 2 β 1 / p 0 T 0 t e ε 0 t p e β ( t s ) ( f ˜ w , η ˜ ) ( , s ) X 0 p d s d t = C λ 2 β 1 / p 0 T 0 t e ( β ε 0 p ) ( t s ) ( e ε 0 s ( f ˜ w , η ˜ ) ( , s ) X 0 ) p d s d t = C λ 2 β 1 / p 0 T ( e ε 0 s ( f ˜ w , η ˜ ) ( , s ) X 0 ) p s T e ( β ε 0 p ) ( t s ) d t d s = C λ 2 β 1 / p ( β ε 0 p ) 1 0 T ( e ε 0 s ( f ˜ w , η ˜ ) ( , s ) X 0 ) p d s .

Besides, we have

( f ˜ w , η ˜ ) ( , s ) X 0 f ˜ w L q ( F ) + η ˜ B q , q 2 1 / q ( G ) C ( v ˜ L q ( F ) + ( 1 + ω ) η ˜ B q , q 2 1 / q ( G ) ) C ( v L q ( F ) + ( 1 + ω ) η B q , q 2 1 / q ( G ) )

for any s ( 0 , T ) . Hence, it follows from the estimate (5.19) and Corollary 5.3 that

(5.20) e ε 0 t ( U , h ) ( , t ) L p ( J ; X 0 ) C ( u 0 B q , p 2 ( δ 1 / p ) ( F ) + h 0 B q , p 2 + δ 1 / p 1 / q ( G ) + e ε 0 t ( f u , g d , g u τ , g u v , f h ) F δ ( J ; F ) )

for t ( 0 , T ) and ε 0 ( 0 , β / p ) , where C is independent of T and t . Besides, by ( 5.7 ) 5 , we also obtain the estimate

(5.21) e ε 0 t h ( , t ) F p , q , δ 1 1 / ( 2 q ) ( J ; L q ( G ) ) C e ε 0 t h ( , t ) H δ 1 , p ( J ; L q ( F ) ) L δ p ( J ; H 2 , q ( F ) ) C ( e ε 0 t s h ( , t ) L δ p ( J ; L q ( F ) ) + e ε 0 t h ( , t ) L δ p ( J ; H 2 , q ( F ) ) ) C ( e ε 0 t ( U , η ˜ ) ( , t ) L δ p ( J ; L q ( F ) ) + e ε 0 t h ( , t ) L δ p ( J ; H 2 , q ( F ) ) ) C ( u 0 B q , p 2 ( δ 1 / p ) ( F ) + h 0 B q , p 2 + δ 1 / p 1 / q ( G ) + e ε 0 t ( f u , g d , g u τ , g u v , f h ) F δ ( J ; F ) )

with a constant C independent of t and T . If ( U ˜ , h ˜ ) satisfies the shifted equations

t U ˜ + 2 λ 2 U ˜ L ω , y U ˜ + K ( U ˜ , h ˜ ) = 2 λ 2 ( f ˜ w + U ) , in  F , div U ˜ = 0 , in  F , P G ( 2 μ D ( U ˜ ) ν G ) = 0 , on  G , 2 μ D ( U ˜ ) ν G ν G K ( U ˜ , h ˜ ) + C ^ G h ˜ = 0 , on  G , t h ˜ + 2 λ 2 h ˜ ( P 0 G U ˜ ) ν G = 2 λ 2 ( η ˜ + h ) , on  G , U ˜ ( 0 ) = 0 , in  F , h ˜ ( 0 ) = 0 , on  G ,

we obtain from Corollary 5.3, (5.20), and (5.21) that

e ε 0 t ( U ˜ , h ˜ ) E 1 , δ ( J ; F ) × E 4 , δ ( J ; G ) C ( u 0 B q , p 2 ( δ 1 / p ) ( F ) + h 0 B q , p 2 + δ 1 / p 1 / q ( G ) + e ε 0 t ( f u , g d , g u τ , g u v , f h ) F δ ( J ; F ) ) .

Noting that U and h satisfy (5.7), the uniqueness of solutions implies U ˜ = U and h ˜ = h for any t ( 0 , T ) , and hence, we observe

e ε 0 t ( U , h ) E 1 , δ ( J ; F ) × E 4 , δ ( J ; G ) C ( u 0 B q , p 2 ( δ 1 / p ) ( F ) + h 0 B q , p 2 + δ 1 / p 1 / q ( G ) + e ε 0 t ( f u , g d , g u τ , g u v , f h ) F δ ( J ; F ) ) .

Recalling the decomposition (5.6), we arrive at

e ε 0 t ( u , h ) E 1 , δ ( J ; F ) × E 4 , δ ( J ; G ) C ( u 0 B q , p 2 ( δ 1 / p ) ( F ) + h 0 B q , p 2 + δ 1 / p 1 / q ( G ) + e ε 0 t ( f u , g d , g u τ , g u v , f h ) F δ ( J ; F ) ) .

This completes the proof of Theorem 5.8.□

Step 3: The completion of the proof of Theorem 5.1. Finally, let us derive the estimates of ( u , q , h ) . Recalling (5.4) and (5.5), it follows from Theorems 5.2 and 5.8 that

e ε 0 t t ( u , h ) L δ p ( J ; X 0 ) C ( u 0 B q , p 2 ( δ 1 / p ) ( F ) + h 0 B q , p 2 + δ 1 / p 1 / q ( G ) + e ε 0 t ( f u , g d , g u τ , g u v , f h ) F δ ( J ; F ) ) .

Besides, we also have

e ε 0 t ( u , h ) L δ p ( J ; X 1 ) e ε 0 t ( u , h ) L δ p ( J ; L q ( F ) × L q ( G ) ) + j = 1 , 2 e ε 0 t j u L δ p ( J ; L q ( F ) ) + k = 1 , 2 , 3 e ε 0 t k h L δ p ( J ; L q ( G ) ) C [ e ε 0 t ( v , η ) L δ p ( J ; H 2 , q ( F ) × B q , q 3 1 / q ( G ) ) + e ε 0 t ( u , h ) L δ p ( J ; H 2 , q ( F ) × B q , q 3 1 / q ( G ) ) + 0 T ( e ε 0 s ( u ( , s ) , 1 ) F ) p d s 1 / p + α = 1 , 2 0 T ( e ε 0 s ( u ( , s ) , e α × y ) F ω G h ( , s ) y α y 3 d G p d s 1 / p + 0 T ( e ε 0 s ( u ( , s ) , e 3 × y ) F ) p d s 1 / p + m = 1 4 0 T ( e ε 0 s ( h ( , s ) , φ m ) G ) p d s 1 / p C [ u 0 B q , p 2 ( δ 1 / p ) ( F ) + h 0 B q , p 2 + δ 1 / p 1 / q ( G ) + e ε 0 t ( f u , g d , g u τ , g u v , f h ) F δ ( J ; F ) + 0 T ( e ε 0 s ( u ( , s ) , 1 ) F ) p d s 1 / p + α = 1 , 2 0 T e ε 0 s ( u ( , s ) , e α × y ) F ω G h ( , s ) y α y 3 d G p d s 1 / p + 0 T ( e ε 0 s ( u ( , s ) , e 3 × y ) F ) p d s 1 / p + m = 1 4 0 T ( e ε 0 s ( h ( , s ) , φ m ) G ) p d s 1 / p ,

where C is independent of T . Finally, we estimate e ε 0 t h in the F p , q , δ 2 / q ( J ; L q ( G ) ) -norm. It holds

e ε 0 t h F p , q , δ 2 1 / q ( J ; L q ( G ) ) C e ε 0 t h H δ 2 , p ( J ; L q ( F ) ) H δ 1 , p ( J ; H 2 , q ( F ) ) C ( e ε 0 t t 2 h L δ p ( J ; L q ( F ) ) + e ε 0 t h H δ 1 , p ( J ; H 2 , q ( F ) ) ) .

From ( 4.1 ) 5 , we observe that

t 2 h ( P 0 G t u ) ν G = t f h on  G .

Hence, we obtain

e ε 0 t t 2 h L δ p ( J ; L q ( F ) ) C ( e ε 0 t ( P 0 G t u ) ν G L δ p ( J ; L q ( F ) ) + e ε 0 t t f h L δ p ( J ; L q ( F ) ) ) C ( e ε 0 t u E 1 , δ ( J ; F ) + e ε 0 t f h F 4 , δ ( J ; G ) ) ,

which implies

e ε 0 t h F p , q , δ 2 1 / q ( J ; L q ( G ) ) C ( e ε 0 t ( u , h ) E 1 , δ ( J ; F ) × E 4 , δ ( J ; G ) + e ε 0 t f h F 4 , δ ( J ; G ) ) .

It follows from (5.4) and (5.5) that

0 T ( e ε 0 s ( u ( , s ) , 1 ) F ) p d s 1 / p + 0 T ( e ε 0 s ( u ( , s ) , e 3 × y ) F ) p d s 1 / p + α = 1 , 2 0 T e ε 0 s ( u ( , s ) , e α × y ) F ω G h ( , s ) y α y 3 d G p d s 1 / p C 0 T ( e ε 0 s ( u ( , s ) , 1 ) F ) p d s 1 / p + 0 T ( e ε 0 s ( u ( , s ) , e 3 × y ) F ) p d s 1 / p + α = 1 , 2 0 T e ε 0 s ( u ( , s ) , e α × y ) F ω G h ( , s ) y α y 3 d G p d s 1 / p + u 0 B q , p 2 ( δ 1 / p ) ( F ) + h 0 B q , p 2 + δ 1 / p 1 / q ( G ) + e ε 0 t ( f u , g d , g u τ , g u v , f h ) F δ ( J ; F ) } , m = 1 4 0 T ( e ε 0 s ( h ( , s ) , φ m ) G ) p d s 1 / p C m = 1 4 0 T ( e ε 0 s ( h ( , s ) , φ m ) G ) p d s 1 / p + u 0 B q , p 2 ( δ 1 / p ) ( F ) + h 0 B q , p 2 + δ 1 / p 1 / q ( G ) + e ε 0 t ( f u , g d , g u τ , g u v , f h ) F δ ( J ; F ) ,

where C depends on ω but is independent of T . This completes the proof of Theorem 5.1.

6 The nonlinear problem

6.1 Local existence

We now show the local existence result for given initial data

( v 0 , Γ 0 ) B q , p 2 ( δ 1 / p ) ( Ω ( 0 ) ) 3 × B q , p 2 + δ 1 / p 1 / q ( G ) ,

which are subject to the compatibility conditions the compatibility conditions

div v 0 = 0 in  Ω ( 0 ) , P Γ 0 [ μ ( v 0 + [ v 0 ] ) ] = 0 on  Γ 0

and condition (1.4). We define the nonlinear mapping N = ( N 1 , N 2 , N 3 , N 4 , N 5 , N 6 ) by

N 1 F u ( u , q , h ) N 2 G d ( u , h ) = div G div ( u , h ) , N 3 G u τ ( u , h ) N 4 G u v ( u , h ) + G 0 ( h ) N 5 F h ( u , h ) + F ( u , h )

respectively. We set

U T { z = ( u , q , Tr G [ q ] , h ) E δ ( J ; F ) : h L ( G × J ) < ε }

with J = ( 0 , T ) , T > 0 , and 0 < ε < 1 . Then we have the following proposition.

Proposition 6.1

Let ( p , q , δ ) satisfy (1.13). Then it holds

  1. N is a real analytic mapping from U T to F δ ( J ; F ) and N ( 0 ) = D N ( 0 ) = 0 .

  2. D N ( z ) ( U T , F δ ( J ; F ) ) .

Here, D N represents the Fréchet derivative of N .

Proof

Since ( p , q , δ ) satisfies (1.13), we have the following assertions, see [39, Lem. 5.3]:

  1. E 1 , δ ( J ; F ) BUC 1 ( J ; BUC ( F ) ) .

  2. E 3 , δ ( J ; G ) BUC ( J ; BUC ( G ) ) .

  3. E 4 , δ ( J ; G ) BUC 1 ( J ; BUC 1 ( G ) ) BUC ( J ; BUC 2 ( G ) ) .

  4. E 3 , δ ( J ; G ) and F 4 , δ ( J ; G ) are multiplication algebras.

Here, in assertions (i)–(iii), the embedding constants are independent of T > 0 if the time traces vanish at t = 0 . The polynomial structure of the nonlinearity N with respect to u , q , and h gives mapping properties for N . This completes the proof.□

Using the aforementioned proposition, we can obtain the local existence of classical solution to (1.1). Since the proof is standard (cf. [39, Sec. 5]), we may omit the detail.

Theorem 6.2

Let T > 0 be a given constant. Assume conditions (1.13) hold. Then there exists a constant ε = ε ( T ) > 0 such that for arbitrary initial data ( u 0 , η 0 ) B q , p 2 ( δ 1 / p ) ( F ) 3 × B q , p 2 + δ 1 / p 1 / q ( G ) satisfying the compatibility conditions

(6.1) div u 0 = G d ( u 0 , h 0 ) = div G div ( u 0 , h 0 ) , in F , P F [ μ ( u 0 + [ u 0 ] ) ] = G u τ ( u 0 , h 0 ) , on F ,

and the smallness condition

u 0 B q , p 2 ( δ 1 / p ) ( F ) + h 0 B q , p 2 + δ 1 / p 1 / q ( F ) ε ,

the transformed problem (3.7) has a unique solution ( u , q , Tr G [ q ] , η ) E δ ( ( 0 , T ) ; F ) . Furthermore, the solution is indeed real analytic in F × ( 0 , T ) and, especially, t ( 0 , T ) ( Γ ( t ) × { t } ) is a real analytic manifold.

Remark 6.3

To remove the smallness condition on the initial velocity field v 0 , one has to consider the modified term F h ( u , h ) + b G h instead of F h ( u , h ) , where b F 4 , δ ( J ; G ) 3 is taken such that b ( 0 ) = Tr G [ P G u 0 ] . In fact, F h ( u , h ) cannot be small in the norm of F 5 , δ ( J ; G ) even if G h L ( G ) is small. However, to make our discussion simple, we keep the smallness assumption on v 0 . Notice that the local existence result for arbitrary large initial velocity already obtained by Shibata [24], and see also [27, Thm. 3.6.1].

6.2 Global existence and convergence

Finally, we prove Theorem 1.2. In the following, we suppose that the initial data ( u 0 , η 0 ) B q , p 2 ( δ 1 / p ) ( F ) 3 × B q , p 2 + δ 1 / p 1 / q ( G ) satisfies the smallness condition:

(6.2) u 0 B q , p 2 ( δ 1 / p ) ( F ) + h 0 B q , p 2 + δ 1 / p 1 / q ( F ) ε

with some small ε > 0 as well as the compatibility conditions (6.1). Since we will choose ε small eventually, we may suppose 0 < ε < 1 . By Theorem 6.2, for given T 0 > 0 , there exists ε > 0 such that (3.7) admits a unique solution ( u , q , Tr G [ q ] , h ) E δ ( ( 0 , T 0 ) ; F ) . In the following, we may assume ε < ε . We further assume that the system (3.7) admits a solution ( u , q , Tr G [ q ] , h ) E δ ( J 0 ; F ) with J 0 = ( 0 , T 0 ) . We shall show that the solution ( u , q , Tr G [ q ] , h ) can be prolong to the time interval R + . To this end, it suffices to verify the a priori estimate

(6.3) e ε 0 t ( u , q , Tr G [ q ] , h ) E δ ( 0 , T ; F ) C ( ( u 0 , η 0 ) B q , p 2 ( δ 1 / p ) ( F ) × B q , p 2 + δ 1 / p 1 / q ( G ) + e ε 0 t ( u , q , Tr G [ q ] , h ) E δ ( 0 , T ; F ) 2 )

for any T ( 0 , T 0 ] , where a constant C is independent of ε , T , and T 0 . Here, ε 0 is the same constant as in Theorem 5.1. In fact, combining the local existence result and the a priori estimate (6.3), a standard bootstrap argument implies the desired result.

From Theorem 5.1 and Proposition 6.1, we easily find that ( u , q , Tr G [ q ] , h ) enjoys the estimate

e ε 0 t ( u , q , Tr G [ q ] , η ) E δ ( 0 , T ; F ) C [ ( u 0 , η 0 ) B q , p 2 ( δ 1 / p ) ( F ) × B q , p 2 + δ 1 / p 1 / q ( G ) + e ε 0 t ( u , q , Tr G [ q ] , η ) E δ ( 0 , T ; F ) 2 + 0 T ( e ε 0 s ( u ( , s ) , 1 ) F ) p d s 1 / p + α = 1 , 2 0 T e ε 0 s ( u ( , s ) , e α × y ) F ω G h ( , s ) y α y 3 d G p d s 1 / p + 0 T ( e ε 0 s ( u ( , s ) , e 3 × y ) F ) p d s 1 / p + m = 1 4 0 T ( e ε 0 s ( h ( , s ) , φ m ) G ) p d s 1 / p .

A similar argument given in [26, Sec. 6] gives

(6.4) 0 T ( e ε 0 s ( u ( , s ) , 1 ) F ) p d s 1 / p + α = 1 , 2 0 T e ε 0 s ( u ( , s ) , e α × y ) F ω G h ( , s ) y α y 3 d G p d s 1 / p + 0 T ( e ε 0 s ( u ( , s ) , e 3 × y ) F ) p d s 1 / p + m = 1 4 0 T ( e ε 0 s ( h ( , s ) , φ m ) G ) p d s 1 / p C e ε 0 t ( u , q , Tr G [ q ] , η ) E δ ( 0 , T ; F ) 2 .

In fact, by [32, Sec. 2], it follows

0 = Ω ˜ ( t ) d z F d y = G h h 2 2 H G + h 3 3 K G d G , 0 = Ω ˜ ( t ) z d z F y d y = G h h 2 2 H G + h 3 3 K G y + ν G ( ) h 2 2 h 3 3 H G + h 4 4 K G d G ,

which gives the estimate

(6.5) m = 1 4 0 T ( e ε 0 s ( h ( , s ) , φ m ) G ) p d s 1 / p C e ε 0 t ( u , q , Tr G [ q ] , η ) E δ ( 0 , T ; F ) 2 ,

cf., [26, Sec. 6]. Next, it follows from ( 1.4 ) 2 and (1.6) that

( v , e × x ) Ω ( t ) = ( e 3 × y , e × y ) F = ω δ , 3 F y 2 d y = γ δ , 3 , = 1 , 2 , 3 .

Hence, according to the transform explained in Section 3, we see that

(6.6) ( V ˜ , e × z ) Ω ˜ ( t ) + ω ( e 3 × z , e × z ) Ω ˜ ( t ) = γ δ , 3 , = 1 , 2 , 3 .

As J ( h ) describes the Jacobian of the transform Ξ h , which is introduced in Section 3, we have

( V ˜ , e × z ) Ω ˜ ( t ) = F u ( e × Ξ h ) J ( h ) d y = ( u , e × y ) F + ( u , e × ξ h ) F + J 0 ( h ) ( u , e × Ξ h ) F

with = 1 , 2 , 3 . Following [32, Sect. 2], for y G , we introduce J ^ ( y , h ) defined by

J ^ ( y , h ) i , j = 1 3 ν G ( i ) ( y ) ν G ( j ) ( y ) J ( h ) ( I [ M 1 ( h ) ] ) i , j ,

where M 1 is the matrix given in Section 3. Then, by [32, p. 1772], for = 1 , 2 , 3 , we may write

( e 3 × z , e × z ) Ω ˜ ( t ) = ( e 3 × y , e × y ) F + 0 1 G ( ( e 3 × Ξ r h ) ( e × Ξ r h ) ) h J ^ ( y , r h ) d G d r ,

which yields the expression

ω ( e 3 × z , e × z ) Ω ˜ ( t ) = γ δ , 3 ω G h y y 3 d G + ω N ˜ ( y , h ) .

Here, N ˜ ( y , h ) is a nonlinear term that is given by

N ˜ ( y , h ) = G h y y 3 d G + 0 1 G ( ( e 3 × Ξ r h ) ( e × Ξ r h ) ) h J ^ ( y , r h ) d G d r .

Hence, (6.6) turns into

( u , e × y ) F ω G h y y 3 d G = ( u , e × ξ h ) F J 0 ( h ) ( u , e × Ξ h ) F ω N ˜ ( y , h ) ,

where the right-hand side is nonlinear. Thanks to [39, Lem. 5.3], we have the estimate

(6.7) 0 T ( e ε 0 s ( u ( , s ) , 1 ) F ) p d s 1 / p + α = 1 , 2 0 T e ε 0 s ( u ( , s ) , e α × y ) F ω G h ( , s ) y α y 3 d G p d s 1 / p C e ε 0 t ( u , q , Tr G [ q ] , η ) E δ ( 0 , T ; F ) 2 .

Combining (6.5) and (6.7), we observe (6.4). Thus, we obtain (6.3).

Finally, we deal with the original problem (1.1). Notice that the compatibility condition (1.14) is valid if and only if (6.1) is satisfied. Define Ξ h 0 ( y ) y + ξ h 0 ( y ) with replacing h by h 0 in the definition of ξ h . Then, the mapping Ξ h 0 defines a C 2 -diffeomorphism from F onto Ω ( 0 ) with inverse Ξ h 0 1 . This gives the existence of a constant C h 0 depending on h 0 such that

C h 0 1 v 0 v B q , p 2 ( δ 1 / p ) ( Ω ( 0 ) ) u 0 B q , p 2 ( δ 1 / p ) ( F ) C h 0 v 0 v B q , p 2 ( δ 1 / p ) ( Ω ( 0 ) ) .

Hence, there exists ε > 0 such that the smallness condition of Theorem 1.2 yields (6.2). Recalling the discussion in Section 3, we see that there exists a unique global classical solution ( v , π , Γ ) to (1.1), especially, the unique global solution ( v , π , Γ ) to (1.1) is real analytic. Noting B q , p 2 ( δ 1 / p ) ( Ω ( t ) ) B q , p 2 ( δ 1 / p ) ( F ) , we observe the asymptotic behavior of solutions. This completes the proof of Theorem 1.2.

  1. Funding information: This research was partly supported by JSPS KAKENHI Grant Numbers 20K22311 and 21K13826 and the Waseda University Grant for Special Research Projects (Project number: 2021C-583).

  2. Conflict of interest: Author states no conflict of interest.

References

[1] G. Allain, Small-time existence for the Navier-Stokes equations with a free surface, Appl. Math. Optim. 16 (1987), no. 1, 37–50. 10.1007/BF01442184Search in Google Scholar

[2] P. Appell, Traité de mécanique rationnelle, Tome IV, Fascicule 1: Figures daéquilibre daune masse liquide homogéne en rotation, Gauthier-Villars, Paris, 1932, Available at: https://gallica.bnf.fr/ark:/12148/bpt6k290921/.Search in Google Scholar

[3] H. Bae, Solvability of the free boundary value problem of the Navier-Stokes equations, Discrete Contin. Dyn. Syst. 29 (2011), no. 3, 769–801, 10.3934/dcds.2011.29.769Search in Google Scholar

[4] J. T. Beale, Large-time regularity of viscous surface waves, Arch. Rational Mech. Anal. 84 (1983/84), no. 4, 307–352. 10.1090/conm/017/04Search in Google Scholar

[5] J. T. Beale and T. Nishida, Large-time behavior of viscous surface waves, North-Holland Mathematical Studies. vol. 128, North-Holland, Amsterdam, 1985, pp. 1–14. 10.1016/S0304-0208(08)72355-7Search in Google Scholar

[6] R. A. Brown and L. E. Scriven, The shape and stability of rotating liquid drops, Proc. Roy. Soc. London Ser. A 371 (1980), 331–357. 10.1098/rspa.1980.0084Search in Google Scholar

[7] S. Chandrasekhar, The stability of a rotating liquid drop, Proc. Roy. Soc. London Ser. A 286 (1965), 1–26. 10.1098/rspa.1965.0127Search in Google Scholar

[8] R. Denk, M. Hieber, and J. Prüss, R-boundedness, Fourier multipliers and problems of elliptic and parabolic type, Mem. Amer. Math. Soc. 166 (2003), no. 788. 10.1090/memo/0788Search in Google Scholar

[9] R. Denk, M. Hieber, and J. Prüss, Optimal Lp-Lq -estimates for parabolic boundary value problems with inhomogeneous data, Math. Z. 257 (2007), no. 1, 193–224. 10.1007/s00209-007-0120-9Search in Google Scholar

[10] M. Haase, The functional calculus for sectorial operators, Operator Theory: Advances and Applications, 169, Birkhäuser Verlag, Basel, 2006. 10.1007/3-7643-7698-8Search in Google Scholar

[11] Y. Hataya, A remark on Beale-Nishidaas paper, Bull. Inst. Math. Acad. Sin. (N.S.) 6 (2011), no. 3, 293–303. Search in Google Scholar

[12] M. Köhne, J. Prüss, and M. Wilke, Qualitative behaviour of solutions for the two-phase Navier-Stokes equations with surface tension, Math. Ann. 356 (2013), no. 2, 737–792. 10.1007/s00208-012-0860-7Search in Google Scholar

[13] N. Lindemulder, Maximal regularity with weights for parabolic problems with inhomogeneous boundary conditions, J. Evol. Equ. 20 (2020), no. 1, 59–108. 10.1007/s00028-019-00515-7Search in Google Scholar

[14] M. Meyries and M. Veraar, Sharp embedding results for spaces of smooth functions with power weights, Studia Math. 208 (2012), no. 3, 257–293, 10.4064/sm208-3-5Search in Google Scholar

[15] M. Meyries and M. Veraar, Traces and embeddings of anisotropic function spaces, Math. Ann. 360 (2014), no. 3–4, 571–606. 10.1007/s00208-014-1042-6Search in Google Scholar

[16] I. Sh. Mogilevskiĭ and V. A. Solonnikov, On the solvability of an evolution free boundary problem for the Navier-Stokes equations in Hölder spaces of functions, in: Mathematical Problems Relating to the Navier-Stokes Equation, Series on Advances in Mathematics for Applied Sciences, vol. 11, World Scientific Publishing, River Edge, NJ, 1992, pp. 105–181. 10.1142/9789814503594_0004Search in Google Scholar

[17] T. Nishida, Y. Teramoto, and H. Yoshihara, Global in time behavior of viscous surface waves: horizontally periodic motion, J. Math. Kyoto Univ. 44 (2004), no. 2, 271–323. 10.1215/kjm/1250283555Search in Google Scholar

[18] M. Padula, On stability of a capillary liquid down an inclined plane, Discrete Contin. Dyn. Syst. Ser. S 6 (2013), no. 5, 1343–1353. 10.3934/dcdss.2013.6.1343Search in Google Scholar

[19] M. Padula and V. A. Solonnikov, On the global existence of nonsteady motions of a fluid drop and their exponential decay to a uniform rigid rotation, in: Topics in Mathematical Fluid Mechanics, Quad. Mat., vol. 10, Department of Mathematics, Seconda Università degli Studi di Napoli, Caserta, 2002, pp. 185–218. Search in Google Scholar

[20] A. Pazy, Semigroups of linear operators and applications to partial differential equations, Applied Mathematical Sciences, vol. 44, Springer-Verlag, New York, 1983. 10.1007/978-1-4612-5561-1Search in Google Scholar

[21] J. Prüss, Vector-valued Fourier multipliers in Lp-spaces with power weights, Studia Math. 247 (2019), no. 2, 155–173. 10.4064/sm170307-19-10Search in Google Scholar

[22] J. Prüss and G. Simonett, Moving interfaces and quasilinear parabolic evolution equations, Monographs in Mathematics, vol. 105, Birkhäuser/Springer, Cham, 2016. 10.1007/978-3-319-27698-4Search in Google Scholar

[23] H. Saito and Y. Shibata, On the Global Wellposedness for Free Boundary Problem for the Navier-Stokes Systems with Surface Tension, 1912, Available at arXiv:1912.10121. Search in Google Scholar

[24] Y. Shibata, Local well-posedness of free surface problems for the Navier-Stokes equations in a general domain, Discrete Contin. Dyn. Syst. Ser. S 9 (2016), no. 1, 315–342. 10.3934/dcdss.2016.9.315Search in Google Scholar

[25] Y. Shibata, On the ℛ-bounded solution operator and the maximal Lp-Lq regularity of the Stokes equations with free boundary condition, in: Mathematical Fluid Dynamics, Present and Future, Springer Proceedings in Mathematics & Statistics, vol. 183, Springer, Tokyo, 2016, pp. 203–285. 10.1007/978-4-431-56457-7_9Search in Google Scholar

[26] Y. Shibata, Global well-posedness of unsteady motion of viscous incompressible capillary liquid bounded by a free surface, Evol. Equ. Control Theory 7 (2018), no. 1, 117–152. 10.3934/eect.2018007Search in Google Scholar

[27] Y. Shibata, MathcalR boundedness, maximal regularity and free boundary problems for the Navier Stokes equations, in: Mathematical Analysis of the Navier-Stokes Eequations, Lecture Notes in Mathematics, vol. 2254, Springer, Cham, 2020, pp. 193–462. 10.1007/978-3-030-36226-3_3Search in Google Scholar

[28] Y. Shibata and S. Shimizu, On the Lp -Lq maximal regularity of the Neumann problem for the Stokes equations in a bounded domain, J. Reine Angew. Math. 615 (2008), 157–209. 10.1515/CRELLE.2008.013Search in Google Scholar

[29] V. A. Solonnikov, Solvability of the problem of evolution of an isolated amount of a viscous incompressible capillary fluid, Russian, with English summary, Zap. Nauchn. Sem. Leningrad. Otdel. Mat. Inst. Steklov. (LOMI) 140 (1984), 179–186. Search in Google Scholar

[30] V. A. Solonnikov, Unsteady motions of a finite isolated mass of a self-gravitating fluid, Algebra i Analiz 1 (1989), no. 1, 207–249, (Russian); Leningrad Math. J. 1 (1990), no. 1, 227–276. Search in Google Scholar

[31] V. A. Solonnikov, Solvability of a problem on the evolution of a viscous incompressible fluid, bounded by a free surface, on a finite time interval, Algebra i Analiz 3 (1991), no. 1, 222–257 (Russian); St. Petersburg Math. J. 3, (1992), no. 1, 189–220. Search in Google Scholar

[32] V. A. Solonnikov, A generalized energy estimate in a problem with a free boundary for a viscous incompressible fluid, Zap. Nauchn. Sem. S.-Peterburg. Otdel. Mat. Inst. Steklov. (POMI) 282 (2001), Issled. po Lineĭn. Oper. i Teor. Funkts. 29, 216–243, 281, J. Math. Sci. (N.Y.) 120 (2004), no. 5, 1766–1783. 10.1023/B:JOTH.0000018874.92754.31Search in Google Scholar

[33] V. A. Solonnikov, On linear stability and instability of equilibrium figures of uniformly rotating liquid, in: Recent Advances in Elliptic and Parabolic Problems, World Scientific Publishing, Hackensack, NJ, 2005, pp. 231–257. 10.1142/9789812702050_0017Search in Google Scholar

[34] V. A. Solonnikov, On the stability of axisymmetric equilibrium figures of a rotating viscous incompressible fluid, Algebra i Analiz 16 (2004), no. 2, 120–153 (Russian); St. Petersburg Math. J.16 (2005), no. 2, 377–400. Search in Google Scholar

[35] A. Tani, Small-time existence for the three-dimensional Navier-Stokes equations for an incompressible fluid with a free surface, Arch. Rational Mech. Anal. 133 (1996), no. 4, 299–331. 10.1007/BF00375146Search in Google Scholar

[36] A. Tani and N. Tanaka, Large-time existence of surface waves in incompressible viscous fluids with or without surface tension, Arch. Rational Mech. Anal. 130 (1995), no. 4, 303–314. 10.1007/BF00375142Search in Google Scholar

[37] I. Tice, Asymptotic stability of shear-flow solutions to incompressible viscous free boundary problems with and without surface tension, Z. Angew. Math. Phys. 69 (2018), no. 2, Paper No. 28. 10.1007/s00033-018-0926-9Search in Google Scholar

[38] H. Triebel, Interpolation Theory, Function Spaces, Differential Operators, 2nd edition, Johann Ambrosius Barth, Heidelberg, 1995. Search in Google Scholar

[39] K. Watanebe, Local well-posedness of incompressible viscous fluids in bounded cylinders with 90∘-contact angle, Nonlinear Anal. Real World Appl. 65 (2022), 103489. 10.1016/j.nonrwa.2021.103489Search in Google Scholar

[40] K. Watanebe, Stability of rotating liquid drops with surface tension, submitted. Search in Google Scholar

Received: 2021-12-13
Revised: 2022-09-02
Accepted: 2022-09-04
Published Online: 2023-01-03

© 2023 Keiichi Watanabe, published by De Gruyter

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

Articles in the same Issue

  1. Regular Articles
  2. On sufficient “local” conditions for existence results to generalized p(.)-Laplace equations involving critical growth
  3. On the critical Choquard-Kirchhoff problem on the Heisenberg group
  4. On the local behavior of local weak solutions to some singular anisotropic elliptic equations
  5. Viscosity method for random homogenization of fully nonlinear elliptic equations with highly oscillating obstacles
  6. Double-phase parabolic equations with variable growth and nonlinear sources
  7. Logistic damping effect in chemotaxis models with density-suppressed motility
  8. Bifurcation diagrams of one-dimensional Kirchhoff-type equations
  9. Standing wave solution for the generalized Jackiw-Pi model
  10. Weak-strong uniqueness and energy-variational solutions for a class of viscoelastoplastic fluid models
  11. Eigenvalue inequalities for the buckling problem of the drifting Laplacian of arbitrary order
  12. Homoclinic solutions for a differential inclusion system involving the p(t)-Laplacian
  13. Equivalence between a time-fractional and an integer-order gradient flow: The memory effect reflected in the energy
  14. Bautin bifurcation with additive noise
  15. Small solitons and multisolitons in the generalized Davey-Stewartson system
  16. Nonstationary Poiseuille flow of a non-Newtonian fluid with the shear rate-dependent viscosity
  17. A global compactness result with applications to a Hardy-Sobolev critical elliptic system involving coupled perturbation terms
  18. On a strongly damped semilinear wave equation with time-varying source and singular dissipation
  19. Singularity for a nonlinear degenerate hyperbolic-parabolic coupled system arising from nematic liquid crystals
  20. Stability of stationary solutions to the three-dimensional Navier-Stokes equations with surface tension
  21. Uniform decay estimates for the semi-linear wave equation with locally distributed mixed-type damping via arbitrary local viscoelastic versus frictional dissipative effects
  22. Symmetry and nonsymmetry of minimal action sign-changing solutions for the Choquard system
  23. Periodic and quasi-periodic solutions of a four-dimensional singular differential system describing the motion of vortices
  24. Multiple nontrivial solutions of superlinear fractional Laplace equations without (AR) condition
  25. Existence and blow-up of solutions in Hénon-type heat equation with exponential nonlinearity
  26. Existence and concentration behavior of positive solutions to Schrödinger-Poisson-Slater equations
  27. On the fractional Korn inequality in bounded domains: Counterexamples to the case ps < 1
  28. Gradient estimates for nonlinear elliptic equations involving the Witten Laplacian on smooth metric measure spaces and implications
  29. Dynamical analysis of a reaction–diffusion mosquito-borne model in a spatially heterogeneous environment
  30. Existence and multiplicity of solutions for a quasilinear system with locally superlinear condition
  31. Nontrivial solution for Klein-Gordon equation coupled with Born-Infeld theory with critical growth
  32. Modeling Wolbachia infection frequency in mosquito populations via a continuous periodic switching model
  33. Infinitely many localized semiclassical states for nonlinear Kirchhoff-type equation
  34. Fujita-type theorems for a quasilinear parabolic differential inequality with weighted nonlocal source term
  35. Approximations of center manifolds for delay stochastic differential equations with additive noise
  36. Periodic solutions to a class of distributed delay differential equations via variational methods
  37. Groundstate for the Schrödinger-Poisson-Slater equation involving the Coulomb-Sobolev critical exponent
  38. Existence of nontrivial solutions for the Klein-Gordon-Maxwell system with Berestycki-Lions conditions
  39. Global Sobolev regular solution for Boussinesq system
  40. Normalized solutions for the p-Laplacian equation with a trapping potential
  41. Nonlinear elliptic–parabolic problem involving p-Dirichlet-to-Neumann operator with critical exponent
  42. Blow-up for compressible Euler system with space-dependent damping in 1-D
  43. High energy solutions of general Kirchhoff type equations without the Ambrosetti-Rabinowitz type condition
  44. On the dynamics of grounded shallow ice sheets: Modeling and analysis
  45. A survey on some vanishing viscosity limit results
  46. Blow-up for logarithmic viscoelastic equations with delay and acoustic boundary conditions
  47. Generalized Liouville theorem for viscosity solutions to a singular Monge-Ampère equation
  48. Front propagation in a double degenerate equation with delay
  49. Positive solutions for a class of singular (pq)-equations
  50. Higher integrability for anisotropic parabolic systems of p-Laplace type
  51. The Poincaré map of degenerate monodromic singularities with Puiseux inverse integrating factor
  52. On a system of multi-component Ginzburg-Landau vortices
  53. Existence and concentration of solutions to Kirchhoff-type equations in ℝ2 with steep potential well vanishing at infinity and exponential critical nonlinearities
  54. Multiplicity results for fractional Schrödinger-Kirchhoff systems involving critical nonlinearities
  55. On double phase Kirchhoff problems with singular nonlinearity
  56. Estimates for eigenvalues of the Neumann and Steklov problems
  57. Nodal solutions with a prescribed number of nodes for the Kirchhoff-type problem with an asymptotically cubic term
  58. Dirichlet problems involving the Hardy-Leray operators with multiple polars
  59. Incompressible limit for compressible viscoelastic flows with large velocity
  60. Characterizations for the genuine Calderón-Zygmund operators and commutators on generalized Orlicz-Morrey spaces
  61. Non-local gradients in bounded domains motivated by continuum mechanics: Fundamental theorem of calculus and embeddings
  62. Noncoercive parabolic obstacle problems
  63. Touchdown solutions in general MEMS models
  64. Existence and nonexistence of nontrivial solutions for the Schrödinger-Poisson system with zero mass potential
  65. Short-time existence of a quasi-stationary fluid–structure interaction problem for plaque growth
  66. Fine bounds for best constants of fractional subcritical Sobolev embeddings and applications to nonlocal PDEs
  67. Symmetries of Ricci flows
  68. Asymptotic behavior of solutions of a second-order nonlinear discrete equation of Emden-Fowler type
  69. On the topological gradient method for an inverse problem resolution
  70. Supersolutions to nonautonomous Choquard equations in general domains
  71. Uniform complex time heat Kernel estimates without Gaussian bounds
  72. Global existence for time-dependent damped wave equations with nonlinear memory
  73. Normalized solutions for a critical fractional Choquard equation with a nonlocal perturbation
  74. Reversed Hardy-Littlewood-Sobolev inequalities with weights on the Heisenberg group
  75. Lamé system with weak damping and nonlinear time-varying delay
  76. Global well posedness for the semilinear edge-degenerate parabolic equations on singular manifolds
  77. Blowup in L1(Ω)-norm and global existence for time-fractional diffusion equations with polynomial semilinear terms
  78. Boundary regularity results for minimisers of convex functionals with (p, q)-growth
  79. Parametric singular double phase Dirichlet problems
  80. Special Issue on Nonlinear analysis: Perspectives and synergies
  81. Editorial to Special issue “Nonlinear analysis: Perspectives and synergies”
  82. Identification of discontinuous parameters in double phase obstacle problems
  83. Global boundedness to a 3D chemotaxis-Stokes system with porous medium cell diffusion and general sensitivity
  84. On regular solutions to compressible radiation hydrodynamic equations with far field vacuum
  85. On Cauchy problem for fractional parabolic-elliptic Keller-Segel model
  86. The evolution of immersed locally convex plane curves driven by anisotropic curvature flow
  87. Stability of combination of rarefaction waves with viscous contact wave for compressible Navier-Stokes equations with temperature-dependent transport coefficients and large data
  88. On concave perturbations of a periodic elliptic problem in R2 involving critical exponential growth
  89. Existence of solutions for a double-phase variable exponent equation without the Ambrosetti-Rabinowitz condition
Downloaded on 9.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/anona-2022-0279/html
Scroll to top button