Stability of planar exterior stationary flows with suction
Abstract. We consider the two-dimensional Navier-Stokes system in a domain exterior to a disk. The system admits a stationary solution with critical decay written as a linear combination of the pure rotating flow and the flux carrier. We prove its nonlinear stability in large time for initial disturbances in under smallness conditions, assuming that there is suction across the boundary, namely that the sign of coefficients of the flux carrier is negative. This result partially solves an open problem in the literature.
Keywords. Navier-Stokes system, Two-dimensional exterior domains, Stability of stationary solutions, Scale-critical decay.
2020 MSC. 35Q30, 35B35, 76D05, 76D17.
1 Introduction
We consider the two-dimensional Navier-Stokes system in an exterior disk
(NS) |
The unknown functions and are respectively the velocity field of the fluid and the pressure field. The function is a given initial data. The set denotes the exterior unit disk where . We assume that both and are real number constants. The vector refers to . The system (NS) describes the time evolution of viscous incompressible fluids around the disk rotating at angular velocity on whose surface there is suction in the orthogonal direction when and injection when .
The system (NS) admits an explicit stationary solution where
(1.1) |
and
(1.2) |
This velocity is a linear combination of the vector field denoting the pure rotating flow in and the flux carrier. To lighten notation, in the following, we write
(1.3) |
The solution is invariant under the scaling of the Navier-Stokes equations. A (non-trivial) solution having this property is said to be scale-critical and it represents the balance between the nonlinear and linear parts of the equations. Therefore, investigating the properties of the scale-critical solutions is a fundamental and important issue in understanding the typical behavior of the Navier-Stokes flows. Let us mention that is an element of the family of stationary solutions of (NS) found by Hamel [20]. This family is known to be an example showing the non-uniqueness of the -solutions; see Galdi [15, Section XII.2]. The Hamel solutions are generalized by Guillod and Wittwer [19] in view of rotation symmetries.
In this paper, we study the nonlinear stability of in large time. More precisely, assuming that an initial disturbance around belongs to the Lebesgue spaces, we consider the time evolution of the disturbance in the nonlinear system (NS). Particularly, we are interested in the large-time decay estimate. By using the relation
(1.4) |
and , we see that the pair of new unknown functions
formally solves the nonlinear problem
(NP) |
The linearized problem of (NP) is given by
(LP) |
Our main aim in this paper is to obtain large-time decay estimates of the solutions of (LP), by studying the operators associated with (LP). We will provide the - estimates sufficient to prove the nonlinear stability of the stationary solution in large time.
In order to make the framework clearer, we recall some notations and basic facts about the linear system (LP). We let denote , the closure of in , and the orthogonal projection. The operator is called the Helmholtz projection and satisfies for with . The operator, called the Stokes operator, is defined by
It is well known that is nonnegative and self-adjoint in and that generates the -analytic semigroup; see Sohr [39]. Moreover, the spectrum of is the set of nonpositive real numbers ; see Section 3 for the references. With these notations, we define the operator associated with (LP) by
and write (LP) equivalently with the evolution system
(1.5) |
We aim at proving the properties of solutions of (1.5) by studying the operator .
One basic way to study the properties of is to consider the equation
(R) |
for given and . This equation can be obtained by formal application of the Laplace transform to (1.5). From the general theory of functional analysis, we find the following two facts. First, as the operator is lower order with respect to , from theory for sectorial operators, we see that is also sectorial in and generates the -analytic semigroup, denoted by ; see Lunardi [32, Proposition 2.4.3]. Second, as is relatively compact with respect to , from the perturbation theory of operators, we see that where denotes the discrete spectrum of ; see Section 3 for details. These two facts, however, are not sufficient to obtain the large-time estimate of since contains . We need a precise estimate of the resolvent when is close to the origin.
The fundamental difficulty in analyzing (R) when is that the Hardy inequality
(1.6) |
does not hold in exterior domains when . If (1.6) holds when , the term in (R) can be controlled by the dissipation from if is small. Nevertheless, one needs a logarithmic correction in the left-hand side of (1.6) to obtain the correspondence; see [15, Theorem II.6.1]. This implies that energy method does not work well in general in deriving estimates for (R) when . One way to recover inequalities of the type (1.6) when is to assume symmetries both on and ; see Galdi and Yamazaki [16], Yamazaki [41], and Guillod [18] for the stability results of symmetric flows under symmetries. As we do not assume any symmetries on initial data in (NP), unlike [16, 41, 18], such inequalities are not applicable to (LP) nor (R). This is in stark contrast to the three-dimensional stability results by Heywood [21] and by Borchers and Miyakawa [4, 5] in which the Hardy inequality (1.6) with is an essential tool. As a recent monograph of the three-dimensional results, we refer to Brandolese and Schonbek [7].
Therefore, even for the flow explicitly given in (1.3), the stability analysis in two-dimensional exterior domains requires specific considerations depending on the parameters and . The known results are summarized as follows.
-
•
The case and is treated in Guillod [18]. This case is tractable and similar to the three-dimensional cases if is sufficiently small. In fact, for general exterior domains , Russo [38, Lemma 3] proves the Hardy-type inequality
(1.7) The reader is referred to [15, Remark X.4.2] and [18, Lemma 3] for further discussions. Combining (1.7) with the relation (1.4), we obtain the control
(1.8) This observation implies that, by a simple energy estimate applied to (R), we can obtain the - estimates for the system (LP) and prove the nonlinear stability of . Alternatively, as is done in [18], one can prove the stability by considering -estimates of the semigroup generated by the adjoint of the operator . A similar idea is also used in Karch and Pilarczyk [27].
-
•
The case and is treated in Maekawa [33]. In this case, energy method is not useful for (R). Indeed, [18, Lemma 4] points out that the Hardy-type inequality (1.7) does not hold if is replaced by . To relax the situation, [33] considers the problem in an exterior disk and performs explicit computations. The - estimates for (LP) are obtained when is sufficiently small by an explicit formula for the resolvent . Also, the nonlinear stability of is proved when both and the -norm of initial data in (NP) are sufficiently small. This stability result is extended by the author in [22] to a certain class of non-symmetric domains where the domains are assumed to be small perturbations of the exterior unit disk, and in [24] for three-dimensional initial disturbances around an infinite cylinder.
-
•
The case and is treated in Maekawa [34]. The problem is considered on an exterior disk as in [33]. The idea of the proof is to regard the term in (R) as an external force and to utilize the estimate of in [33]. The - estimates for (LP) are obtained when is sufficiently small, under the restriction that initial data belong to a subcritical space for some . Also, the nonlinear stability of is proved when both and the ()-norm of initial data in (NP) are sufficiently small. This restriction on exponents is essentially needed when estimating . As is mentioned in [34, Remark 2], it is not clear if the condition can be removed in this method.
1.1 Main results
This paper addresses large-time estimates for the system (LP), namely, of the semigroup , when and as in [34]. Our particular interest is the - estimates left as open problems in [34]. The following theorem solves it affirmatively under a condition on the sign of . This condition is discussed in Remark 1.2 (iii) below.
Theorem 1.1
Let satisfy and and let be sufficiently small. For , we have
(1.9) |
The constant depends on .
Remark 1.2
- (i)
-
(ii)
The proof of Theorem 1.1 is based on an analysis of the operator . The estimate (1.9) for is deduced by the Dunford integral of the resolvent . Inspired by [33], we determine the spectrum of and estimate by explicit computations. It is shown in Section 3 that the function characterizing the discrete spectrum of crucially depends on both and . Therefore, it is suggested that, when estimating for , one cannot regard in (R) as an external force even if is small, in spite of the control (1.8).
-
(iii)
It is an open problem whether the same estimate as in (1.9) can be obtained for the case . Actually, by following the argument in Section 4, one can prove (1.9) if is chosen to depend on a given , but the general case is still open. It might be meaningful to recall here that the case corresponds to the situation where there is injection into fluids at the boundary. We mention Drazin and Reid [11, Problem 3.7] and Drazin and Riley [12, Section 3.1] as the references related to this topic.
-
(iv)
It is important to extend the - estimates in (1.10) to general exterior domains. However, this is a difficult problem because of the dependence of constants on . The problem when is tackled in [22] and it is shown that, if the domain is a perturbation from the exterior unit disk in algebraic order of , then the - estimates can be obtained by energy method combined with explicit formulas. The restriction to a class of domains is due to singularity in the operator norm of the resolvent for small . It is observed that, in explicit computations, cancellation of the effects from the two terms and in (R) (with ) occurs for in a certain domain, dubbed the “nearly-resonance regime” in [22]. This cancellation causes the singularity at algebraic order of , which in energy method restricts the shape of domains, more precisely the lengths between domains and the exterior unit disk. Such singularity also appears in the operator norm of for small in the present problem and is an obstacle to the extension.
By using Theorem 1.1, we can prove the nonlinear stability of . Using the semigroup , we consider the mild solutions of (NP) solving
(1.11) |
The following theorem can be shown by a simple application of the Banach fixed point theorem and thus is omitted in this paper. For details, see [33] treating the case .
Theorem 1.3
Let satisfy and and let be sufficiently small. Let belong to and let be sufficiently small depending on . There is a unique mild solution of (1.11) satisfying
(1.12) |
1.2 Related results
Let us refer to the results that are closely related to the present study.
Analysis of (NP) and (LP) when . For (NP), the estimate (1.12) for , which can be viewed as the nonlinear stability of the trivial solution, is classical; see Masuda [37] for the proof when and Kozono and Ogawa [29] when . These results do not require smallness on the initial data in . For (LP), the - estimates of the Stokes semigroup are established by Maremonti and Solonnikov [36] and by Dan and Shibata [9, 10]. We note that all of the results above hold in general exterior domains . It is pointed out in [33, Remark 1.4] that the logarithmic singularity of the resolvent for small , observed in [9, §3], disappears in if . As compensation, however, singularity appears in the operator norm of for small . Such singularity, as discussed in Remark 1.2 (iv), also appears in the operator norm of , and is an obstacle when generalizing the - estimates in (1.10). Let us mention the study of the boundedness of in spaces pioneered by Borchers and Varnhorn [6]. See Abe [1, 2] for the recent progress.
Non-symmetric stationary solutions around . We consider the stationary problem of (NS), which also admits the explicit solution . It is known that, for suitably chosen , the fundamental solution for the linearized problem around , namely for the stationary problem of (LP), has a better spatial decay compared to the one for the problem linearized around the trivial solution . This improvement is due to the vorticity transport by and implies the resolution of the famous Stokes paradox; see [8, 14, 15, 30, 26] for descriptions. Furthermore, these new fundamental solutions allow us to construct non-symmetric solutions for the nonlinear problem decaying in the order . This is done in Hillairet and Wittwer [25] when and for given zero-flux boundary data in a suitable class, and in [23] when and for given external forces with suitable spatial decay. The solutions in [23] are compatible with the Liouville-type theorem in Guillod [17, Proposition 4.6]. We emphasize that the results in [25, 23] do not require any symmetries on the given data. Interestingly, such improvement in the fundamental solutions occurs even for small . Indeed, Maekawa and Tsurumi [35] constructs non-symmetric solutions for the nonlinear problem in the whole space , whose principal part at spatial infinity is with a small but nonzero constant . This result is contrasting with [25] in view of the size of coefficients, and the reason is that, as there are no boundaries in , the terms needed to match the no-slip boundary condition in exterior domains do not appear in the problem.
1.3 Outlined proof
We describe the proof of Theorem 1.1. However, the estimate (1.9) is almost a direct consequence of the estimate of the resolvent in Proposition 5.1. Hence we give in Appendix C the proof that derives Theorem 1.1 from Proposition 5.1, and outline here the proof of Proposition 5.1. As noted in Remark 1.2 (ii), it consists of two steps:
(I) Spectral analysis of . Recall that . Thus we identify the location of the discrete spectrum to obtain the large-time decay of . For this purpose, we consider the homogeneous equation of (R) and its general solutions, by using the streamfunction-vorticity equations. We see that the no-slip boundary condition imposes that belongs to if and only if belongs to
Here is the analytic function defined in (3.7) in Section 3. For , one can show that has no zeros in the sector by energy method if is sufficiently small; see Propositions 3.3 and 3.4. However, for , we need to deal with the function directly to determine the location of its zeros, which reflects the fact that the Hardy inequality does not hold in two-dimensional exterior domains. We will prove that with has no zeros in sectors for if and is sufficiently small depending on . The proof is the most tricky part of this paper and will take the whole of Section 4. We perform an asymptotic analysis of that refines the methods in [33, 22]. Interestingly, the analysis is highly dependent on the sign of being positive or negative. Furthermore, we observe that the condition , which is also an assumption of Theorem 1.1, provides a certain stabilizing effect compared to the case ; see Remark 4.8.
(II) Estimate of the resolvent . In the next step, we estimate the solution of (R) for belonging to the resolvent set. We derive and estimate an explicit formula for the solution using the streamfunction-vorticity equations. The computations are lengthy ones estimating the formulas involving the modified Bessel functions, but the approach itself is broadly the same as that used in [33, 22]. Thus we omit some details; see Section 5.
This paper is organized as follows. In Section 2, we collect the items used in this paper. In Sections 3 and 4, we study the spectrum of the operator . We apply the perturbation theory of operators in Section 3 and perform an asymptotic analysis of with in Section 4. In Section 5, we provide the estimate of the resolvent. Some facts about the modified Bessel functions and technical supplements are given in Appendices A, B and C.
Notations. We let denote a constant and the constant depending on . Both of these may vary from line to line. We denote , , and . For , let and denote the real and imaginary parts of , respectively. For , let denote where , and . We take the square root so that . We use the function spaces
and
Not to burden notation, we use the same symbols to denote the quantities for scalar-, vector- or tensor-valued functions, e.g., is the inner product on , or .
2 Preliminaries
This section collects the items used throughout the paper.
2.1 Vectors in the polar coordinates
The polar coordinates on the exterior unit disk are written as
Let a vector field on be given. We set
and for a given ,
(2.1) |
We will use the formulas
(2.2) |
and
(2.3) |
2.2 Fourier series and decomposition
Let and be defined in (2.1). We set, for a vector field on ,
(2.4) |
for a scalar function on ,
(2.5) |
and for a function space or ,
The definition of differs according to whether is vectorial or scalar. The former and latter are defined in (2.4) as and in (2.5) as , respectively.
By definition, any vector field is expanded into the convergent series
and is an orthogonal projection of onto . Moreover, the following orthogonal decomposition of the subspace holds:
(2.6) |
From (2.2), we have
In particular,
Therefore, if , the Hardy-type inequality
holds. Thus it is convenient to set
for which we have
(2.7) |
Again from (2.2), we have, for ,
and, from (2.3), for ,
Since the condition on is preserved under , it can be shown that
We refer to Farwig and Neustupa [13, Lemma 3.1] for a more detailed proof. Although the proof in [13] is for the three-dimensional cases, a similar argument is applicable.
Now we define the closed linear operator on in (2.6) by
It is not hard to see that is nonnegative and self-adjoint. Also, keeping the relation
in mind, we define the closed linear operator on by
2.3 Equations in the polar coordinates
To study the operator , we consider
(Rn) |
for given and . The equation is equivalent to the system
(2.8) |
with some pressure . Operating to the first line, we see that solves
(2.9) |
In the polar coordinates on where is written as
we see from (2.8) that and satisfy
(2.10) | |||
(2.11) |
and the divergence-free and the no-slip boundary conditions
(2.12) |
Moreover, from (2.9), satisfies
(2.13) |
2.4 Biot-Savart law
To simplify the explanation, only in this subsection, we use the function space
For a given , we consider the Poisson equation
Let with and let be the decaying solution, called the streamfunction. Applying the notation in (2.5), we find that satisfies
(2.14) |
By elementary computation, we see that is given by
(2.15) |
The following vector field
(2.16) |
is called the Biot-Savart law. It is straightforward to see that
(2.17) |
If additionally with some , we can check that .
Here are useful two propositions in the subsequent sections. The reader is referred to [33, Proposition 2.6 and Lemma 3.1], [23, Proposition 2.1] for the proof of the first proposition and [33, Corollary 2.7], [23, Proposition 2.2] for the proof of the second.
Proposition 2.1
Let and . Set . If and for some , we have and in (2.15).
Proposition 2.2
Let and . If in the sense of distributions, we have for some .
3 Spectral analysis
In this section, we study the spectrum of the operator . The main result is Proposition 3.4 which characterizes the discrete spectrum as zeros of certain analytic functions. We are aware that the presentation in Subsections 3.1 and 3.2 has similarity to [33] treating the case . This is quite natural because, in analysis in the -framework, especially in computation of the numerical ranges, one can control terms involving by using the Hardy-type inequality (1.7). Consequently, for example, the statement of Proposition 3.3 holds independently of sufficiently small . However, the difference from the case appears when one studies the spectrum of . Indeed, in Proposition 3.4, the functions characterizing the discrete spectrum depend both on and . These functions will be studied in detail quantitatively in the next section.
3.1 Notation
Let us recall the standard notation in the perturbation theory. Our main reference is Kato [28]. Let be a Banach space and be a closed linear operator. We let denote the null space of , its range, and the quotient space of by . Moreover, denotes the resolvent set of , its spectrum, and its discrete spectrum, namely, the set of isolated eigenvalues of with finite multiplicity. The operator is said to be semi-Fredholm if is closed and at least one of or is finite. If is semi-Fredholm, the index of
is well-defined, taking values in . Finally, let us set
and call the semi-Fredholm domain of and the essential spectrum of , respectively.
Generally, is the union of a countable (at most) family of connected open sets. From the argument in [28, Chapter §5 6], we see that is a constant function of in each component of . Moreover, both and are constants in each except for an isolated set of values of . Therefore, if these constants are zero in particular, then is contained in with possible exception of isolated points of , which are, isolated eigenvalues of finite algebraic multiplicity.
3.2 Perturbation theory
We start with the perturbation theory of operators.
Proposition 3.1
Let . We have the following.
-
(1)
and .
-
(2)
The same statement with replaced by holds for .
-
(3)
and .
Proof.
(1) The fact that is well-known and essentially due to Ladyzhenskaya [31]. Based on this fact, one can prove that by showing the non-existence of eigenvalues in a similar manner as in [13, Lemma 2.6], or by using the property of the index as is done in [33, Proof of Proposition 2.12]. Because of the regularity and decay of , the operator is relatively compact with respect to . The proof is quite similar to the one in [33, Section 2.4] for the case and thus we omit the details. Hence, from [28, Chapter , Theorem 5.35], we see that and have the same essential spectrum. This implies the first statement.
For the second statement, we first observe that the equality
holds by [28, Chapter , Theorems 5.26 and 5.35]. Hence, since has only one component, by the argument in Subsection 3.1, we only need to prove that for at least one point . For this purpose, we consider
which is called the numerical range of ; see [28, Chapter §3 2].
Let . From the relation
we have
(3.1) |
Now let and . The term is estimated as
We have used for in the second line and the Young inequality in the third line. Hence we obtain
which leads to the inclusion
From [28, Chapter , Theorem 3.2], we know that for any belonging to the complement of the right-hand side
This set is obviously a subset of and thus the second statement follows.
(2) The fact that can be proved in a similar manner as in [13, Lemma 3.3], and follows by the property of . Hence the first statement follows from the relative compactness of with respect to . The second statement can be deduced from the same discussion as above with replaced by .
(3) It suffices to prove the first statement . If , there is a nonzero such that . Choosing such that , we have and . Then we see that and hence . Oppositely, if , then there are and nonzero such that . Then we have and and hence . The proof is complete. ∎
The estimate of the numerical range in the proof of Proposition 3.1 is quite rough. We consider its refinement in Lemma 3.2 to prove Proposition 3.3 below.
Lemma 3.2
Let . For , we have
(3.2) |
Moreover, for any ,
(3.3) |
Here the function is defined by
which satisfies
(3.4) |
Proof.
Proposition 3.3
Let be sufficiently small. We have the following.
-
(1)
The set
is contained in .
-
(2)
The same statement with replaced by holds for each .
-
(3)
The set is contained in for each .
Proof.
Let us consider the numerical range as in the proof of Proposition 3.1.
(1) We first estimate . Let . Using Lemma 3.2, we have
Let us choose . From (3.4), we see that
and that
Hence we obtain
(3.5) |
Note that up to this point the smallness of is not needed.
Now let and . From (3.1) and (3.5), we have
Therefore, for sufficiently small , we obtain the inclusion
Then the statement follows from the same argument as in the proof of Proposition 3.1 (2).
(2) A similar proof as above leads to the statement.
3.3 Analysis by explicit computation
Proposition 3.3 does not provide information on the discrete spectrum of near the origin. This is a consequence of the fact that the Hardy inequality fails to hold in two-dimensional exterior domains. Therefore, we investigate the homogeneous equation of (Rn) by a more explicit computation, exploiting the symmetry of the exterior disk .
For , we define
(3.6) |
and
(3.7) |
Proposition 3.4
Let . We have the following.
-
(1)
.
-
(2)
for .
Proof.
(1) Let . In view of Proposition 3.1 (2) and , we will show that the equation has only the trivial solution in . Put in (LABEL:eq.polar.vr)–(2.12) with . The conditions in (2.12) imply that and hence that . From (LABEL:eq.polar.vtheta)–(2.12), we see that satisfies
By summability, the solution is given by, with some constant ,
Then the boundary condition leads to since has no zeros in if ; see [40, Chapter 157]. Hence we obtain that , which is to be shown.
(2) Let . In view of Proposition 3.1 (2) and , we will show that the equation admits a nontrivial solution in if and only if . Let be nontrivial and solve . Notice that is smooth by the elliptic regularity of the Stokes system. Setting , we see that satisfies the homogeneous equation of (2.13). Its linearly independent solutions are (B.1) in Appendix B. By the summability of , we must have, with some constant ,
Since decays exponentially as , Proposition 2.1 leads to that
with the notations in (2.15)–(2.16). The former condition implies that is nonzero since is assumed to be nontrivial. The latter one can be written equivalently to
Thus we have that since . This completes the proof of the only if part.
For the if part, let . Then, for any nonzero , the vector field
gives a nontrivial solution of . Indeed, from the proof of the only if part, we ensure that is smooth and belongs to . Note that the no-slip condition is verified by the assumption that . Moreover, setting
from (2.17), we see that
Thus Proposition 2.2 yields that there is a function such that . Operating the Helmholtz projection to this equality, we find that . This completes the proof of the if part. The proof of Proposition 3.4 is complete. ∎
Corollary 3.5
Let be sufficiently small. We have
4 Quantitative analysis of discrete spectrum
In this section, keeping Corollary 3.5 in mind, we analyze zeros of the analytic function with defined in (3.7). Thanks to Proposition 3.3, it suffices to consider the zeros in disks centered at the origin with radius exponentially small in . The main result is Proposition 4.7. The proof is based on asymptotic analysis under the smallness of .
Note that one can recover the results in [33, 22] by putting in the statements of this section. However, this observation is not useful in the proof since we need to describe precisely the zeros of functions having multiple parameters. A continuity argument is not enough and quantitative analysis is needed. In fact, it is revealed that situations are different depending on the sign of , and that the case seems to be more delicate.
4.1 Expansion of the order
We need the following expansion of in the next subsection.
Lemma 4.1
Let . For sufficiently small , we have
(4.2) | ||||
(4.3) |
All the implicit constants in are independent of .
4.2 Asymptotic analysis
We consider in (3.7) with , namely, the function
(4.4) |
Lemma 4.2
Let . For , we have
(4.5) |
Proof.
Using the relation (4.5), we investigate zeros of near the origin. We perform asymptotic analysis when is sufficiently small. Since the asymptotics of is already obtained in Lemma A.2 (1), we focus on the second term on the right-hand side of (4.5). In what follows in this section, we assume smallness of . Although some estimates can be proved under weaker assumptions, we will not give the details for simplicity.
Lemma 4.3
Let . For sufficiently small , we have
(4.6) |
Here is defined by
(4.7) |
where is the Gamma function, and is the remainder and satisfies
(4.8) |
The constant is independent of .
Proof.
If we show that
(4.9) |
the assertion follows. Indeed, it is not hard to check that
Corollary 4.4
Let . For sufficiently small , we have
(4.10) |
Here is the remainder and satisfies
(4.11) |
The constant is independent of .
Proposition 4.7 below, giving a lower bound of , is proved based on the expansion (4.10). In the proof, we need precise estimates of the coefficients appearing in (4.10).
Lemma 4.5
Let . For sufficiently small , we have
(4.12) | ||||
(4.13) |
where is the Euler constant. Moreover, if ,
(4.14) |
All the implicit constants in are independent of .
Proof.
The following is the key technical lemma in the proof of Proposition 4.7 below.
Lemma 4.6
Let . Suppose that with satisfies
(4.16) |
and
(4.17) |
with some constant independent of . Then, by defining
(4.18) |
one has
(4.19) |
The constant depends only on and .
Proof.
By setting
(4.20) | ||||
(4.21) |
we denote
From
we compute
(4.22) |
Before going into details, let us explain the difficulties. When is close to zero, one essentially needs to provide lower bounds of . However, such bounds require good control of , since vanishes when with . The reason why the conditions (4.16)–(4.17) are needed is to control the range of when is close to zero.
We will consider two cases:
(i) Case . In this case, we have
In addition, by (4.22) and the assumption (4.17),
Thus is equal to if and only if . If , the imaginary part gives
(4.23) |
If , the real part gives
(4.24) |
Hence we estimate . Combining with (4.22), we have
Combining with (4.20),
By these two estimates, we obtain
Therefore, from (4.23) and (4.24), we see that
(4.25) |
Proposition 4.7
Let and . Let be defined in (4.18). For sufficiently small , we have
(4.28) |
The constant depends only on .
Remark 4.8
One observes a sort of stabilizing effect by the flow from this proposition. By the definition (4.18) and Lemma 4.1, we have a simple (but rough) estimate from below
(4.29) |
The second inequality is valid for sufficiently small . Therefore, the radius of the disks on which has no zeros is greater than that for . This is interpreted as a stabilizing effect by in time frequency near zero related to large-time behavior of flows.
Proof.
Let first. Using Corollary 4.4, we write
(4.30) |
Here
is the remainder and satisfies
(4.31) |
The condition and Lemma 4.5 imply
where is the function satisfying
Setting
we will derive a lower bound of
To apply Lemma 4.6, we check that all the conditions are fulfilled by . We have
for sufficiently small . We also have (4.16) by Lemma 4.1. By the same lemma, there are constants independent of such that
Thus, for sufficiently small , we have
and, with a constant independent of ,
which is (4.17) with replaced by . Now, applying Lemma 4.6, we see that
for sufficiently small . The constant depends only on .
We state two corollaries to this proposition. The first one gives a simpler version of (4.28) useful for later calculation. The second one uses the results in Section 3.
Corollary 4.9
Let and . For sufficiently small , we have
In particular,
The constant depends only on .
Corollary 4.10
Let . For sufficiently small , we have
Proof.
In view of Proposition 3.3 and Corollary 4.9, we set
From Propositions 3.1 (1) and 3.3, and Corollary 3.5, we see that both and are contained in for sufficiently small . For a given , by an easy geometric consideration, we find that is contained in if is small enough depending on . This implies the assertion. ∎
5 Resolvent estimate
In this section, we estimate the solutions of
(R) |
for given and . The main result is the following.
Proposition 5.1
Let and let be sufficiently small. We have, for and ,
(5.1) |
and, for ,
(5.2) |
The constant depends only on .
Once Proposition 5.1 is proved, it is routine to prove Theorem 1.1 by representing in the Dunford integral of the resolvent. Thus the detail will be given in Appendix C.
We prove Proposition 5.1 in Subsection 5.3 by a combination of energy method and explicit formulas for the solution. Note that the estimate (5.1) cannot be obtained by energy method alone, due to the absence of the Hardy inequality. However, this is not the case when belongs to sectors shifted exponentially small in ; see Proposition 5.3 for details. Therefore, all that remains is to prove the estimate when belongs to the intersection of sectors and the disks centered at the origin whose radius is exponentially small in . This proof is done by explicit formulas; see Proposition 5.4 for details.
5.1 Energy method
We start with a priori estimates for (R) using energy method.
Lemma 5.2
Let . For , and , suppose that there is a solution of (R). Then we have the following.
-
(1)
For with ,
-
(2)
For ,
The constant depends only on .
Proof.
Lemma 5.2 gives the following estimate of the resolvent.
Proposition 5.3
Let and let be sufficiently small. Set
For and , we have
(5.8) |
The constant depends only on .
5.2 Explicit formulas
Energy method can not lead to Proposition 5.1 due to the absence of the Hardy inequality. Instead, we employ explicit formulas and prove the following proposition.
Proposition 5.4
Let and and let be sufficiently small. Set
We have, for and ,
(5.9) |
and, for ,
(5.10) |
The constant depends only on .
The derivation of the formula is as follows. Let and assume first in (R). Then the solution is smooth in thanks to the elliptic regularity of the Stokes system, and solves the equation (2.13) in Subsection 2.3. Since the linearly independent solutions of its homogeneous equation are (B.1) in Appendix B and the Wronskian is , we see that is given by
(5.11) |
The constant is determined later and is defined by
Using integration by parts and setting
(5.12) |
we have
(5.13) |
Since decays exponentially, we see from Proposition 2.1 that is uniquely represented by the Biot–Savart law as, with the notations in (2.15)–(2.16),
(5.14) |
This formula is implemented with the constraint , which we write
(5.15) |
by using in (3.7). This relation determines . We set
(5.16) |
Collecting (5.11)–(5.16), we find that
(5.17) |
and that
(5.18) |
For general , one should understand the formulas (5.17)–(5.18) by density argument. This understanding is possible thanks to the estimates in Proposition 5.4. Note that the uniqueness of representation is guaranteed by Proposition 2.1.
Lemma 5.5
Let and let . For and , we have
(5.19) | ||||
(5.20) |
and
(5.21) |
where
Remark 5.6
Proof of Lemma 5.5: The equalities (5.19)–(5.20) can be proved by change of the order of integration, the recurrence relations (see [40, Chapter 371 (3), (4)])
and integration by parts. We omit the details since they are analogous to those in the proof of [33, Lemmas 3.6 and 3.9] corresponding to the case . The equality (5.21) follows from the definition (5.16) and (5.20) with . The proof is complete.
Lemma 5.7
Let and and let . We have the following.
-
(1)
Let . For and ,
-
(2)
Let . For ,
The constant depends only on .
Proof.
Lemma 5.8
Let and and let . We have the following.
-
(1)
Let . For or and ,
-
(2)
For and ,
The constant depends only on .
Proof.
Lemma 5.9
Let and and let . We have the following.
-
(1)
For ,
-
(2)
For ,
and for ,
-
(3)
For and ,
The constant depends only on .
Proof.
(1) The estimate can be proved by Lemma A.3 and interpolation theorems. We omit the details since they are analogous to those in the proof of [33, Proposition 3.17] and [22, Proposition 3.9] corresponding to the case .
Proof of Proposition 5.4: Since by Corollary 4.10, we see that exists for any . Let . By density argument, it suffices to prove (5.9)–(5.10) for . From Corollaries 4.9 and 5.8 (2), we have
Thus, from (5.17) and Lemma 5.5, putting in Lemmas 5.8 and 5.9, we see that
which is (5.9). Also, from (5.18), putting in Lemma 5.9, we see that
which leads to (5.10) since for . All the constants above are independent of . This completes the proof.
5.3 Proof of Proposition 5.1
Proof of Proposition 5.1: Let be given. The same consideration as in the proof of Corollary 4.10 shows that for sufficiently small depending on . In view of Proposition 5.3, the desired estimates (5.1)–(5.2) follow if we prove
(5.22) |
for and . Let and set . Thanks to Proposition 5.4, we only need to estimate . Lemma 5.2 (2) implies that
with a constant , and hence that
Hence the proof is complete.
Appendix A Modified Bessel function
We summarize the facts about the modified Bessel functions. Our main references are [40, 3]. The modified Bessel function of the first kind of order is defined by
(A.1) |
where is the Gamma function, the second kind of order is by
(A.2) |
and of order is by the limit of in (A.2) as . In this paper, we exclusively consider the case where the order satisfies and .
The functions and are linearly independent solutions of
with the Wronskian
(A.5) |
It is well known that grows exponentially and decays exponentially as in ; see [3, Section 4.12]. As an integral representation useful in Section 4, we have
(A.6) |
which can be deduced by the formula [40, Chapter 622 (5)] and change of variables.
Collected below are the estimates involving and used in this paper. Each of them can be found in the references [3, 40] or follows from a simple calculation using Lemma A.1. and hence we omit the proof. For the details when , we refer to [33, Lemma 3.31 and Appendix A] and to [22] studying the dependence on in the estimates.
Lemma A.1
Let , and . We have
The constant depends on .
Lemma A.2
Let . We have the following.
-
(1)
For sufficiently small ,
Here is the remainder and satisfies
-
(2)
For sufficiently small ,
Here is the remainder and satisfies
The constant is independent of .
Lemma A.3
Let and and let . For , we have the following.
-
(1)
For and ,
-
(2)
For and ,
-
(3)
For and ,
-
(4)
For and ,
-
(5)
For and ,
The constant depends only on .
Lemma A.4
Let and and let . For , we have the following.
-
(1)
For and ,
-
(2)
For and ,
-
(3)
For and ,
-
(4)
For and ,
-
(5)
For and ,
The constant depends only on .
Lemma A.5
Let and and let . For , we have the following.
-
(1)
For ,
-
(2)
For ,
-
(3)
For ,
-
(4)
For ,
The constant depends only on .
Appendix B Homogeneous equation for vorticity
For , we consider the homogeneous equation of (2.13)
We will prove that its linearly independent solutions are, with defined in (3.6),
(B.1) |
and the Wronskian is . The proof is as follows. Applying the transformation
(B.2) |
we find that solves
By Appendix A, its linearly independent solutions are
Hence, by the inverse transformation of (B.2), we see that the desired solutions are (B.1). One can easily compute the Wronskian using (A.5). The proof is complete.
Appendix C Proof of Theorem 1.1
Let and fix a number . Taking and a curve in
oriented counterclockwise, we use a representation of in the Dunford integral
From (5.1) for in Proposition 5.1, we see that is bounded in , which implies the first line of (1.9). From (5.2), letting and ,
which implies the second line of (1.9). This completes the proof of Theorem 1.1.
Acknowledgements
The author is partially supported by JSPS KAKENHI Grant Number JP 20K14345.
References
- [1] Ken Abe. Liouville theorems for the Stokes equations with applications to large time estimates. J. Funct. Anal., 278(2):108321, 30, 2020.
- [2] Ken Abe. On the large time -estimates of the Stokes semigroup in two-dimensional exterior domains. J. Differential Equations, 300:337–355, 2021.
- [3] George E. Andrews, Richard Askey, and Ranjan Roy. Special functions, volume 71 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge, 1999.
- [4] Wolfgang Borchers and Tetsuro Miyakawa. -decay for Navier-Stokes flows in unbounded domains, with application to exterior stationary flows. Arch. Rational Mech. Anal., 118(3):273–295, 1992.
- [5] Wolfgang Borchers and Tetsuro Miyakawa. On stability of exterior stationary Navier-Stokes flows. Acta Math., 174(2):311–382, 1995.
- [6] Wolfgang Borchers and Werner Varnhorn. On the boundedness of the Stokes semigroup in two-dimensional exterior domains. Math. Z., 213(2):275–299, 1993.
- [7] Lorenzo Brandolese and Maria E. Schonbek. Large time behavior of the Navier-Stokes flow. In Handbook of mathematical analysis in mechanics of viscous fluids, pages 579–645. Springer, Cham, 2018.
- [8] I-Dee Chang and Robert Finn. On the solutions of a class of equations occurring in continuum mechanics, with application to the Stokes paradox. Arch. Rational Mech. Anal., 7:388–401, 1961.
- [9] Wakako Dan and Yoshihiro Shibata. On the – estimates of the Stokes semigroup in a two-dimensional exterior domain. J. Math. Soc. Japan, 51(1):181–207, 1999.
- [10] Wakako Dan and Yoshihiro Shibata. Remark on the - estimate of the Stokes semigroup in a -dimensional exterior domain. Pacific J. Math., 189(2):223–239, 1999.
- [11] P. G. Drazin and W. H. Reid. Hydrodynamic stability. Cambridge Mathematical Library. Cambridge University Press, Cambridge, second edition, 2004. With a foreword by John Miles.
- [12] P. G. Drazin and N. Riley. The Navier-Stokes equations: a classification of flows and exact solutions, volume 334 of London Mathematical Society Lecture Note Series. Cambridge University Press, Cambridge, 2006.
- [13] Reinhard Farwig and Jiří Neustupa. On the spectrum of a Stokes-type operator arising from flow around a rotating body. Manuscripta Math., 122(4):419–437, 2007.
- [14] Giovanni P. Galdi. Stationary Navier-Stokes problem in a two-dimensional exterior domain. In Stationary partial differential equations. Vol. I, Handb. Differ. Equ., pages 71–155. North-Holland, Amsterdam, 2004.
- [15] Giovanni P. Galdi. An introduction to the mathematical theory of the Navier-Stokes equations. Springer Monographs in Mathematics. Springer, New York, second edition, 2011. Steady-state problems.
- [16] Giovanni P. Galdi and Masao Yamazaki. Stability of stationary solutions of two-dimensional Navier-Stokes exterior problem. In The proceedings on Mathematical Fluid Dynamics and Nonlinear Wave, volume 37 of GAKUTO Internat. Ser. Math. Sci. Appl., pages 135–162. Gakkōtosho, Tokyo, 2015.
- [17] Julien Guillod. Steady solutions of the Navier–Stokes equations in the plane. arXiv:1511.03938, 2015.
- [18] Julien Guillod. On the asymptotic stability of steady flows with nonzero flux in two-dimensional exterior domains. Comm. Math. Phys., 352(1):201–214, 2017.
- [19] Julien Guillod and Peter Wittwer. Generalized scale-invariant solutions to the two-dimensional stationary Navier-Stokes equations. SIAM J. Math. Anal., 47(1):955–968, 2015.
- [20] Georg Hamel. Spiralförmige bewegungen zäher flüssigkeiten. Jahresbericht der Deutschen Mathematiker-Vereinigung, 25:34–60, 1917.
- [21] John G. Heywood. On stationary solutions of the Navier-Stokes equations as limits of nonstationary solutions. Arch. Rational Mech. Anal., 37:48–60, 1970.
- [22] Mitsuo Higaki. Note on the stability of planar stationary flows in an exterior domain without symmetry. Adv. Differential Equations, 24(11-12):647–712, 2019.
- [23] Mitsuo Higaki. Existence of planar non-symmetric stationary flows with large flux in an exterior disk. arXiv:2207.11922, 2022.
- [24] Mitsuo Higaki. Stability of a two-dimensional stationary rotating flow in an exterior cylinder. Math. Nachr., 296(1):314–338, 2023.
- [25] Matthieu Hillairet and Peter Wittwer. On the existence of solutions to the planar exterior Navier Stokes system. J. Differential Equations, 255(10):2996–3019, 2013.
- [26] Toshiaki Hishida. Asymptotic structure of steady Stokes flow around a rotating obstacle in two dimensions. In Mathematical fluid dynamics, present and future, volume 183 of Springer Proc. Math. Stat., pages 95–137. Springer, Tokyo, 2016.
- [27] Grzegorz Karch and Dominika Pilarczyk. Asymptotic stability of Landau solutions to Navier-Stokes system. Arch. Ration. Mech. Anal., 202(1):115–131, 2011.
- [28] Tosio Kato. Perturbation theory for linear operators. Grundlehren der Mathematischen Wissenschaften, Band 132. Springer-Verlag, Berlin-New York, second edition, 1976.
- [29] Hideo Kozono and Takayoshi Ogawa. Two-dimensional Navier-Stokes flow in unbounded domains. Math. Ann., 297(1):1–31, 1993.
- [30] Hideo Kozono and Hermann Sohr. On a new class of generalized solutions for the Stokes equations in exterior domains. Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4), 19(2):155–181, 1992.
- [31] O. A. Ladyzhenskaya. The mathematical theory of viscous incompressible flow. Mathematics and its Applications, Vol. 2. Gordon and Breach Science Publishers, New York-London-Paris, 1969. Second English edition, revised and enlarged, Translated from the Russian by Richard A. Silverman and John Chu.
- [32] Alessandra Lunardi. Analytic semigroups and optimal regularity in parabolic problems, volume 16 of Progress in Nonlinear Differential Equations and their Applications. Birkhäuser Verlag, Basel, 1995.
- [33] Yasunori Maekawa. On stability of steady circular flows in a two-dimensional exterior disk. Arch. Ration. Mech. Anal., 225(1):287–374, 2017.
- [34] Yasunori Maekawa. Remark on stability of scale-critical stationary flows in a two-dimensional exterior disk. In Mathematics for nonlinear phenomena—analysis and computation, volume 215 of Springer Proc. Math. Stat., pages 105–130. Springer, Cham, 2017.
- [35] Yasunori Maekawa and Hiroyuki Tsurumi. Existence of the stationary navier-stokes flow in around a radial flow. Journal of Differential Equations, 350:202–227, 2023.
- [36] P. Maremonti and V. A. Solonnikov. On nonstationary Stokes problem in exterior domains. Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4), 24(3):395–449, 1997.
- [37] Kyūya Masuda. Weak solutions of Navier-Stokes equations. Tohoku Math. J. (2), 36(4):623–646, 1984.
- [38] Antonio Russo. On the existence of D-solutions of the steady-state navier-stokes equations in plane exterior domains. arXiv:1101.1243, 2011.
- [39] Hermann Sohr. The Navier-Stokes equations: An elementary functional analytic approach. Modern Birkhäuser Classics. Birkhäuser/Springer Basel AG, Basel, 2001.
- [40] G. N. Watson. A Treatise on the Theory of Bessel Functions. Cambridge University Press, Cambridge, England; The Macmillan Company, New York, 1944.
- [41] Masao Yamazaki. Rate of convergence to the stationary solution of the Navier-Stokes exterior problem. In Recent developments of mathematical fluid mechanics, Adv. Math. Fluid Mech., pages 459–482. Birkhäuser/Springer, Basel, 2016.
M. Higaki
Department of Mathematics, Graduate School of Science, Kobe University, 1-1 Rokkodai, Nada-ku, Kobe 657-8501, Japan. Email: [email protected]