1. Introduction
This paper studies the large time decay of solutions to the initial value problem for a linear nonautonomous system arising from fluid dynamics, specifically a viscous incompressible flow past an obstacle. Let
be an exterior domain in
,
, with
-boundary
. Complement
is identified with the obstacle (rigid body) immersed in a fluid, and it is assumed to be a compact set in
with nonempty interior. Towards the understanding of the stability or attainability of time-dependent Navier–Stokes flow past the obstacle whose motion could also be time-dependent, an essential step is to deduce some decay properties of a linearized system of form
in
, where vector field
and scalar function
are unknowns denoting fluid velocity and pressure, while
and
are prescribed functions. When
and
, (
1) is just the well-known Stokes system. We consider perturbed Stokes system (
1) subject to homogeneous Dirichlet boundary condition
and initial condition
at initial time
. The adjoint of the solution operator (evolution operator)
:
to (
1)–(
3) provides solution operator
:
to backward problem
in
, with
being the
identity matrix, where
and
are unknowns, subject to
and final condition
at final time
. In what follows, let us assume that
for simplicity. This condition is actually satisfied for a typical example, (
9) below. Initial and final velocities
f and
g are taken from class
of solenoidal
-vector fields,
, with vanishing normal trace at boundary
.
In [
1,
2], the present author established the
-
estimate
of solution
to (
1)–(
3) and the same estimate for (
4)–(
6) when
(rigid motion consisting of translation and rotation) and
(one half of the Coriolis force) in 3D, where
for
and
for
. We interpret (
7) as the linearized stability with definite rate. Since the only assumption on
is
with some
, the result of [
1,
2] completely recovers Estimate (
7) for the Stokes and Oseen semigroups, those semigroups with rotating effect due to [
3,
4,
5,
6,
7]. In order to study the attainability (relating to the celebrated starting problem raised by Finn [
8]) and stability of the steady (or even time-periodic) Navier–Stokes flow, especially within the
framework, see [
6,
9,
10,
11,
12,
13,
14,
15,
16], it is always crucial to make full use of (
7) when
, whereas even more linear analysis is needed in [
17] by Maekawa for 2D cases. Since the equation is nonautonomous without any specific structure of
, such as time periodicity, one can no longer carry out spectral analysis that is standard as a strategy of obtaining time-decay estimates of semigroups for the autonomous case; see [
3,
5,
6,
7,
11,
18]. Moreover, as observed in [
19,
20], drift operator
with unbounded coefficient arising from the presence of rotation of the obstacle prevents the usual analysis on the basis of the theory of parabolic evolution operators; see, for instance, Tanabe [
21], which corresponds to analytic semigroups for the autonomous case.
The aim of this expository paper is to clarify how one can deduce decay estimate (
7) of the solution to (
1)–(
3) if we adapt the approach developed in [
1,
2]. A typical example that we have in mind is 3D case
with
being time-dependent Navier–Stokes flow induced from rigid motion (
8) of the obstacle. Roughly speaking, we make Assumptions (i)–(iii):
- (i)
Initial value problem (
1)–(
3) and backward adjoint problem (
4)–(
6) are well-posed in
. A unique solution
to (
1)–(
3) satisfies smoothing estimate (
7) only for
and
with some
, where
is arbitrary, constants
C may depend on
, and we set
. See (A3)–(A5) in the next section.
- (ii)
Initial value problem (
1) in
subject to (
3) and backward adjoint problem (
4) in
subject to (
6) are well-posed in
. A unique solution
to the former enjoys (
7) (both smoothing action and decay, in which
is replaced by
) and
and we have the same estimates for the latter (backward problem), where
. See (A6)–(A7) in the next section.
- (iii)
Initial value problem (
1)–(
3) and backward adjoint problem (
4)–(
6) fulfil the energy relations with dissipation
with some
, where the second term of each exhibits the dissipative effect. See (A8) in the next section.
Several comments on Assumptions (i)–(iii) above are in order. The well-posedness in Assumption (i), in other words, the generation of the evolution operator, is never obvious; however, this is a different issue from what we address in this paper. When
and
with (
8), the generation of the evolution operator with (
7) for
was successfully proved by Hansel and Rhandi [
22] for every
; then, it was verified by [
1] that constant
in (
7) can be taken uniformly in
with
, and by [
2] that smoothing estimate (
10) of
near the obstacle holds with
. The latter estimate is closely related to the asymptotic behavior of pressure (in a bounded domain near the obstacle) and very crucial in [
2] on account of the lack of smoothing action exhibited by analytic semigroups since the evolution operator is not parabolic. The remarkable smoothing rate
was already found by [
6,
7,
23] for the Stokes and Oseen semigroups with rotating effect, and it was a slight improvement of the rate deduced by Noll ans Saal [
24] in another context. Assumption (ii) on the
-
estimate (
7) for the whole space problem is nontrivial, but it is the starting point of analysis in this paper. When
and
with (
8), the solution in the whole space can be explicitly described in terms of the heat semigroup in which a change of variable is made, so that Estimates (
7) and (
11) are in fact available. Energy Relations (
12) and (
13) in (iii) are reasonable assumptions that play several roles, especially for deduction of
. When
and
, (
12) and (
13) are obvious on account of the skew–symmetry of
without any smallness condition. Concerning case (
9) as well, we can easily see (
12) and (
13), provided that
V is small enough in
, see, for instance, [
25], where
is the weak-
space (a Lorentz space). Except for Assumption (iii), we do not have useful higher energy estimates, which play an important role in [
4] by Maremonti and Solonnikov for the Stokes semigroup.
As the substitution of analysis of a parametrix of the resolvent in exterior domains for the autonomous case, the key of our approach is how we make use of energy Relations (
12) and (
13) to deduce
, see Proposition 1. Here and in what follows, by
we denote estimate (
7) with
. Case
(uniform boundedness) for large
is our main task, yielding the other cases by use of the energy inequality of the differential form. It is reasonable under Assumption (ii) to regard the solution to (
1)–(
3) as a perturbation from (a modification of)
–flow by means of a cut-off procedure. The desired uniformly boundedness of the perturbation is discussed by a bootstrap argument and by duality argument with the aid of Assumption (iii) on the energy relations, and the use of duality is why we need to simultaneously study adjoint evolution operator
with
. Unfortunately, this step does not work when
. If Dirichlet Condition (
2) is replaced by another boundary condition, a core part of this step does not follow, even for
.
With
at hand, we are able to proceed to the decay estimates of
and
near the obstacle that we call the local energy decay, as in several papers for the autonomous case. Among other papers, the method of local energy decay is traced back to [
3] by Iwashita on the Stokes semigroup in the context of mathematical fluid dynamics, and the origin would be even back to studies of hyperbolic equations with dissipation by Shibata. The final step with another cut-off procedure is to derive the decay estimate of
near spatial infinity by the use of Assumption (ii) combined with the local energy decay obtained in the preceding step. The cut-off remainder consists of several terms; among them, two terms are delicate: one is
, and the other is pressure. What we need is both decay for
and smoothing rate for
of those terms; see (
78). The latter of the temporal derivative is the assumption (
10) in (i), and the deduction of (
10) was actually one of the main tasks in [
2] on case
and
with (
8). For more general case, as in [
2], we have to look into details about the construction of a parametrix of the evolution operator to verify (
10). If it is constructed with the use of evolution operators for the whole space problem and for the interior one near the obstacle by a cut-off technique as in [
22], smoothing rate (
10) of
is determined by the one of pressure for the interior problem.
To sum up, we claim that some decay properties of solutions to the same system in whole space
together with the energy relation imply the desired estimates in exterior domains provided
, and that we need to find (
10) through analysis of pressure to justify this statement. Let us close the introductory section with a remark on Case (
9). With the results for case
obtained in [
1,
2] at hand, it is actually possible to show the stability or attainability of scale-critical Navier–Stokes flow
by an interpolation technique due to Yamazaki [
26] as long as it is small enough; see, for instance, Takahashi [
15] on the attainability of steady flow for the purely rotating case. However, we have less information about the asymptotic behavior of disturbance. If we intend to show some decay properties of gradient of the disturbance, we have to know
for Problem (
1)–(
3) with (
9). If we adapt the approach developed in ([
25], Section 4) to case (
9) with the aforementioned
V that is sufficiently small, we could verify Assumption (ii), but only partially.
In the next section, we precisely formulate the problem and provide the main theorem.
Section 3 is devoted to the proof. We close the paper with a conclusion in the final section.
2. Result
Let us fix the notation. Given a domain , and integer , the standard Lebesgue and Sobolev spaces are denoted by and by . We abbreviate norm and even , where is the exterior domain with -boundary under consideration. We assumed that , where denotes the open ball centered at the origin with radius . We set for . By we denote the class of all functions with compact support in D, and by the completion of in . We set , where and . By we denote various duality pairings over domain D. In what follows, we use the same symbols for denoting the scalar and vector functions if there is no confusion. Let X be a Banach space. Then, stands for the Banach space consisting of all bounded linear operators from X into itself.
Let
, where
is the exterior domain under consideration. Class
consists of all solenoidal vector fields in
. Let
. We define space
by the completion of
in
. For
, it is characterized as
where
stands for the outer unit normal to
. When
, the boundary condition in the sense of a normal trace is absent. The space of the
-vector fields admits Helmholtz decomposition
see [
27,
28,
29]; Simader and Sohr [
29] established decomposition under condition
when
. By
we denote the Fujita–Kato projection associated with the decomposition above. We then observe that
as well as
. Note the duality relation
and
, where
. We simply write
for exterior domain
under consideration. We easily see that
, where
is the Riesz transform.
Suppose that
(A1) The coefficients of (
1)
are measurable in
and continuous in
with the property
where
denotes the
–norm. For simplicity,
is assumed to be a solenoidal vector field, that is,
in the sense of distributions for each
.
Let
. For
and
given above, we introduce linear operator
on
by
Then initial value Problem (
1)–(
3) and backward adjoint problem (
4)–(
6) (with
) are formulated, respectively, as
It is easily seen that
for all
and
, where
. In fact, we fix
such that
for
and
for
, and set
for
; then, we find
By (
14) and by
, passing to the limit as
justifies
for
and
, yielding (
18) and, therefore,
, where the adjoint is well-defined since
is densely defined. At this point, we need the maximality in the sense that
(A2) For each there is such that both and are surjective, where .
This means the solvability of the associated elliptic problem, which implies that
is injective. Under this condition, (
18) leads to duality relation
Hence,
is a closed operator. In what follows, given
, we abbreviate
and even
. For case
and
, Condition (A2) follows from [
23] by Shibata along with the Mozzi–Chasles transform ([
30], Chapter VIII) (to reduce the problem to the particular case when the translational and angular velocities are parallel each other).
As in [
2,
22], auxiliary spaces
for
play a role to describe the regularity of solutions, since domain
varies as
t goes on. Note that
for every
, and that the homogeneous Dirichlet condition at
is not involved in the space
.
We further make Assumptions (A3)–(A5):
(A3) Let
. Operator family
generates an evolution operator
on
such that
is a bounded linear operator from
into itself with the semigroup property
in
and that the map
is continuous for every
. Moreover, we have the following properties:
Let
, where
. For every
and
, we have
and
with
in
(condition
is consistent with Lemma 3);
Let
. For every
, we have
with
in
.
(A4) Let
and
. Given
and
, there is a constant
, such that evolution operator
satisfies
for all
,
and
whenever
, where
In addition, for every
, we have
and there is a constant
with the following property: Given
and
, there is a constant
such that
for all
and
whenever
.
(A5) Let
. Given
, operator family
(see (
20)) generates an evolution operator
on
in the same sense as in Assumption (A3) (with obvious change).
Assumption (A5) is related to the solution operator to backward adjoint problem (
17). In fact, we set
then, for every
and
, we have
where
, as well as
with
in
. Given
, we compute
with use of (
24) and (
31) to find
where
should be understood for the pair of
and
with
in this computation; nevertheless, once we have (
32), continuity argument justifies (
32) for every
,
and
as well. We thus verified duality relation
in
for
, which, together with (
22), leads to backward semigroup property
in
. By duality, it follows from
that
for all
, see (
27), and
whenever
, where
and
. If one wishes to claim (
43) below for
as well, additional assumptions
and (
28) in which
is replaced by
are needed.
We next make the following assumption on the evolution operators for the whole space problem.
(A6) Let
. Operator family
generates an evolution operator
on
in the same sense as in (A3) (with obvious change). Given
, operator family
(see (
20)) generates an evolution operator
on
in the same sense as in Assumptions (A3) and (A5) (with obvious change).
In the same manner as for the exterior problem mentioned above, we observe
We need the following estimates of as well as .
(A7) Let
and
(
is not needed for case
below). Given
, there is a constant
, such that evolution operator
satisfies
for all
,
and
whenever
. In addition, for every
, we have
and, given
, there is a constant
, such that
for all
whenever
. We have the same estimates, (
36) and (
37), for adjoint
as well.
Lastly, we suppose
(A8) There is a constant
, such that
for all
and
.
This, combined with (
24) and (
31), implies energy relations
and, thereby,
as long as
. On account of (
19), condition
with some
independent of
is sufficient for (
38) with
. This is accomplished for case (
9) with small
by the Lorentz-Hölder inequality and the Lorentz–Sobolev embedding relation.
We are now in a position to give our main theorem.
Theorem 1. Let . Suppose (A1)–(A8). Given , there exists a constant , such that for all with and whenever , see (14), where and is also allowed for .
Remark 1. When and (Stokes semigroup), decay rate (44) is optimal even for case ; see [4,31]. Remark 2. For case and with (8) studied in [1,2], the global Hölder continuity condition is needed to prove Assumptions (A3)–(A5), so that constants and (see (26) and (28)) involve the Hölder seminorm of , which is thereby hidden in (43) through . If we consider of which the Hölder seminorm (and -norm) is bounded from above by , constant C in (43) can be taken uniformly with respect to such . Remark 3. It is clearly hopeless to verify (A2)–(A8) for , which merely satisfies (A1). The verification of those conditions for with more properties such as Hölder continuity (in ) mentioned in Remark 2 is an important but different issue, and it very much depends on how a parametrix of evolution operator is constructed. For case and with (8), Conditions (A1)–(A8) are met due to [1,2,22].