Existence of steady Navier-Stokes flows
exterior to an infinite cylinder
Abstract. We consider the 3D steady Navier-Stokes system on the exterior of an infinite cylinder under the action of an external force. We are concerned with the class of solutions in which the velocity field is vertically uniform and at rest at horizontal infinity. This configuration includes the 2D exterior problem for the steady Navier-Stokes system known to have characteristic difficulties. We prove the existence of solutions in the above class for a given nonsymmetric external force with suitable decay. The proof can be adapted to the 2D problem and gives a generalization of H. (2023) in view of the regularity of solutions.
Keywords. Navier-Stokes system; Two-dimensional exterior domains; Scale-critical decay.
2020 MSC. 35Q30, 35B35, 76D05, 76D17.
1 Introduction
We consider the steady Navier-Stokes system on
(NS) |
Here denotes a exterior disk, and thus is the exterior of an infinite cylinder. The velocity field and the pressure field are unknown functions, while the external force and the boundary condition are given. We use standard notation for derivatives: , , , , and .
Our interest is to construct a solution for that does not decay in the vertical direction. This configuration includes the 2D exterior problem for the steady Navier-Stokes system if one takes . It is well-known that this problem has characteristic difficulties due to the lack of certain embeddings in 2D unbounded domains and to the famous Stokes paradox [1, 10, 2, 15, 3]. There are many open problems although mathematically fundamental. For an overview of the field, see the recent survey [9] by Korobkov-Ren.
Because of the absence of general theory, it is an important subject to construct 2D steady Navier-Stokes flows for given data belonging to an appropriate class, and to study their properties at infinity. There are two approaches in this context. One is based on symmetry. We impose symmetry on the given data to improve decay of solutions by cancellation in integrals; see [2, 14, 3, 16, 13, 17, 18] for the work in this direction.
The other approach is more constructive and is based on perturbation. We perturb the Navier-Stokes system around an exact solution invariant under the scaling symmetry
and construct a solution, regarded as a remainder at infinity, to the perturbed system. The underlying idea is that the transport by a scale-invariant flow improves decay structure of solutions to the perturbed system. We typically use an exact solution found by Hamel [6]:
where and we remark that is a linear combination of the rotating flow and the flow carrying flux . Indeed, solutions around are constructed by Hillairet-Wittwer [8] for when and by [7] for any when . As related work, we refer to [4] analyzing the exterior problem around the fast rotating flows and Maekawa-Tsurumi [12] constructing the Navier-Stokes flows in around rotating flows. Besides, Guillod-Wittwer [5] generalizes the Hamel solutions.
In this paper, we are interested in using the perturbation method to construct solutions around to the three-dimensional system (NS). There are at least two difficulties. First, since the equations are three-dimensional, the vorticity-streamfunction formulation useful in 2D settings [8, 7, 12] cannot be applied directly. Second, it is not trivial whether the transport by is also present in the three-dimensional part of the flows.
Let us introduce notation for the main result. For , we define a Banach space by
(1.1) |
In the cylindrical coordinates on
(1.2) |
we denote a three-dimensional vector field by
For a scalar function on and , we set
Then, for a vector field on , we define the operator by
We also define acting on scalar- and tensor-valued mappings in an obvious manner.
Our main result states the existence of solutions to (NS) which are vertically uniform.
Theorem 1.1
For , and with , there is a constant such that the following holds. Suppose that the boundary data is given by
and that the external force is a distribution on and given by
where and satisfy
Then there is a weak solution of (NS) unique in a suitable set (defined in the proof in Section 3). Moreover, the solution has the asymptotic behavior
(1.3) |
Remark 1.2
-
(i)
The precise definition of weak solutions is given in Subsection 1.1.
- (ii)
-
(iii)
Even though the given data are independent of the variable , the solvability of (NS) cannot be reduced to that of the 2D Navier-Stokes system on . Indeed, for a smooth external force , the vertical component is subject to
If one linearizes this equation, the 2D Laplace operator appears whose fundamental solution has logarithmic growth. Thus the proof of (1.3) needs careful computation.
-
(iv)
It is physically more natural to consider solutions of (NS) for given data that are neither vertically uniform nor decaying. However, the problem is difficult even for the vertically periodic data. Indeed, in addition to the analysis of the linearized problem in Section 2, we need to deal with the linear problems corresponding to the non-constant periodic data, which cannot be regarded as small perturbations from the Stokes system due to . New machinery is needed for this family of linear problems. Let us mention Kozono-Terasawa-Wakasugi [11] providing asymptotic behavior for the axisymmetric steady Navier-Stokes system in the above vertically periodic setting.
Let us outline the proof of Theorem 1.1 whose details will be given in Section 3. To adapt the above exact solution to a three-dimensional system (NS), we set
(1.4) |
and . Then the new pair solves
(1.5) |
By the formula
(1.6) |
where is the cross product in , and by , we rewrite (1.5) as
(NP) |
with the pressure
We analyze the perturbed nonlinear system (NP) to prove Theorem 1.1. As will be seen in Section 2, the fundamental solution for the linearized problem has better decay compared with the one for the non-perturbed case when . Consequently, an improvement of decay can be seen similar to the 2D case [8, 7, 12]. However, it should be emphasized, that a new analysis is necessary because the system (NP) is three-dimensional.
This paper is organized as follows. In Section 2, we study the linearized problem of (NP) and prove decay estimates of the solutions. In Section 3, we prove Theorem 1.1.
1.1 Notation and Terminology
We summarize the notation and terminology used throughout this paper.
Notation. We denote by the constant and by the constant depending on . Both of these may vary from line to line. We use the function spaces on
and which is the completion of in the -norm. If there is no confusion, we use the same notation to denote the quantities concerning scalar-, vector- or tensor-valued mappings. For example, denotes to the inner product on , or .
Weak solutions. Let us clarify the definition of weak solutions of (NS) in Theorem 1.1. Let and satisfy the assumption in Theorem 1.1. Then a three-dimensional vector field is called a weak solution of (NS) if satisfies in the sense of distributions, in the sense of trace, and it holds that
for any . |
Axisymmetricity. A scalar function on with variable is said to be axisymmetric if it is independent of . A vector field on with variable is said to be axisymmetric if the scalar functions are axisymmetric. The axisymmetricity of tensor fields on is defined in a similar manner.
2 Linearized Problem
In this section, we consider the linearized problem of (NP)
(LP) |
Assume that this system does not depend on the variable . Set
(2.1) |
Then, by computation
we separate (LP) into two systems; one is for the horizontal components
(LPh) |
where
(2.2) |
and the other is for the vertical component
(LPv) |
Notice that both problems are imposed on the 2D exterior domain . The problem (LPh) is studied in [7] for the case when external forces are more regular. We revisit the results and generalize the class of solutions in view of regularity. On the other hand, (LPv) for the vertical component requires new consideration which is not covered in [7].
2.1 Preliminaries
This subsection collects notation for mappings on independent of the variable . We denote by the variables in the polar coordinates on ; see (1.2). Moreover, we identify and respectively with the two-dimensional vectors and defined on if there is no confusion. Some of the notation in this subsection are special cases of those in the introduction, but are duplicated for clarity of explanation.
In the polar coordinates, we denote a two-dimensional vector field by
Then the operators , and in (2.1)–(2.2) are represented by
(2.3) |
If a tensor field on
is independent of , then we have
(2.4) |
where
In particular, using in (2.2), we see that
(2.5) |
For a tensor field on
we have
where
Let . For a scalar function on , we define the projection
(2.6) |
on the Fourier mode . For a vector field on
using (2.6), we define the operator by
(2.7) |
Applying the operator to (2.4), we obtain
where
We also define acting on scalar- and tensor-valued mappings in an obvious manner. For vector-valued or tensor-valued , we simply denote by or by . We do not identify with when is scalar-valued to avoid confusion with (2.6).
2.2 Horizontal Components
Let . Applying to and
and inserting these to (LPh), we see that and satisfy
(2.8) | |||
(2.9) |
the divergence-free and boundary conditions
(2.10) |
and the condition at infinity
(2.11) |
2.2.1 Axisymmetric Part
Proposition 2.1
Let , and . Suppose that is given by for some . Then there is a unique weak solution of (LPh) satisfying
and
(2.12) |
The constant is independent of , and .
Proof.
To simplify, we omit the subscript “” for and . Put in (LABEL:eq.LPh.polar.vr)–(2.11). The radial part is identically zero since both and hold by (2.10). Besides, the angular part solves the ordinary differential equation
(2.13) |
As the uniqueness of solutions is clear, we may only consider the existence and estimates.
At first we assume that , which gives
(2.14) |
As is shown in [7, Proof of Proposition 3.1], the solution of (2.13) is represented by
By integration by parts based on (2.14), we rewrite this formula as
(2.15) |
An associated pressure is obtained from the equation (LABEL:eq.LPh.polar.vr). Set . Then the pair is smooth and satisfies (LABEL:eq.LPh.polar.vr)–(2.11) in the classical sense.
Next we let . By computation
and
one can verify that the desired estimate (2.12) follows from the formula (2.15).
Finally, we show that with defined in (2.15) is a weak solution of (LPh) for general . By the definition of , we have . Let be such that in . Let be given by (2.15) replacing by , and let be the associated pressure. Then, using the estimate for
which implies
we have, by integration by parts,
Hence we see that is a weak solution of (LPh). This completes the proof. ∎
2.2.2 Non-axisymmetric Part
Let . We set
(2.16) |
and
(2.17) |
By direct computation, we have
(2.18) |
with independent of , and .
Proposition 2.2
Let and let , and with . Suppose that is give by for some . Then there is a unique weak solution of (LPh) satisfying
(2.19) |
The constant is independent of , , and .
Proof.
To simplify, we omit the subscript “” for and . As the uniqueness of solutions can be shown as in [7, Proof of Proposition 3.2], we only prove the existence and estimates.
At first we assume that , which gives
(2.20) |
Set
From [7, Proof of Proposition 3.2], we see that solves
(2.21) |
and, recalling in (2.16), that it can be represented by
(2.22) |
Here 111Due to a typographical error, the denominator in is written as in [7, (3.19)–(3.20)]. It should be emphasized, however, that this error does not affect the main results in [7].
and
(2.23) |
By integration by parts based on (2.20), we rewrite the formula of as
(2.24) |
where
Then, using in (2.22), we see that defined by the Biot-Savart law
(2.25) |
is a smooth solution of (LPh) with some smooth associated pressure . The reader is referred to [7, Proof of Proposition 3.2] for the details when is a regular function.
2.3 Vertical Component
Let . Applying to and
and inserting these to (LPv), we see that satisfies
(2.26) |
and the boundary conditions
(2.27) |
2.3.1 Axisymmetric Part
Proposition 2.3
Proof.
We may only consider the existence and estimates of solutions.
(1) Putting in (2.26)–(2.27), we see that solves
(2.30) |
The linearly independent solutions of the homogeneous equation of (2.30) are
and their Wronskian is . Hence the solution of (2.30) vanishing at infinity is
(2.31) |
Thus the estimate (2.28) follows from
(2.32) |
One easily checks that defined in (2.31) is a weak solution of (LPv).
2.3.2 Non-axisymmetric Part
Recall that is defined in (2.17).
Proposition 2.4
Let and let , and with .
Proof.
We may only consider the existence and estimates of solutions.
(1) Since satisfies (2.21) with replaced by , it is given by
(2.36) |
Thus the estimate (2.34) follows from (2.17)–(2.18) and
(2.37) |
One easily checks that with defined in (2.36) is a weak solution of (LPv).
3 Proof of Theorem 1.1
We prove Theorem 1.1 in this section. As the proof is similar to [7, Proof of Theorem 1.1], we give only the outline to avoid duplication. For , we define the Banach space
The following is a corollary to Propositions 2.1–2.4 and the results in [7] for 2D problems.
Corollary 3.1
Proof.
By [7, Lemma 4.1], for solutions of the 2D problem
there is a constant independent of , and such that
(3.2) |
Hence the desired estimate (3.1) is a consequence of the estimates in Propositions 2.1–2.4 combined with (2.5), (2.17)–(2.18) and (3.2). In addition, the existence and uniqueness of solutions follow from Propositions 2.1–2.4 and [7, Lemma 4.1]. The proof is complete. ∎
Proof of Theorem 1.1: We consider the Banach space
equipped with the norm . By Corollary 3.1, for any , there is a unique weak solution to (LPh)–(LPv) with the external force
and the solution satisfies
We have used the Young inequality for sequences to estimate in the last line. Hence the mapping is well-defined, and will be denoted by .
It is not hard to check that is a contraction on the closed subset
if are small enough depending on . Thus the existence of a weak solution of (NP) in Introduction unique in follows from the Banach fixed-point theorem.
Let us set by . Then one can check that is a weak solution of (NS) with by using (1.6). Moreover, is unique in the set
and satisfies the limit (1.3). This completes the proof of Theorem 1.1.
Acknowledgements
MH is partially supported by JSPS KAKENHI Grant Number JP 20K14345.
References
- [1] 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.
- [2] 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.
- [3] 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.
- [4] Isabelle Gallagher, Mitsuo Higaki, and Yasunori Maekawa. On stationary two-dimensional flows around a fast rotating disk. Math. Nachr., 292(2):273–308, 2019.
- [5] 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.
- [6] Georg Hamel. Spiralförmige Bewegungen zäher Flüssigkeiten. Jahresbericht der Deutschen Mathematiker-Vereinigung, 25:34–60, 1917.
- [7] Mitsuo Higaki. Existence of planar non-symmetric stationary flows with large flux in an exterior disk. J. Differential Equations, 360:182–200, 2023.
- [8] 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.
- [9] Mikhail Korobkov and Xiao Ren. Stationary solutions to the Navier-Stokes system in an exterior plane domain: 90 years of search, mysteries and insights. J. Math. Fluid Mech., 25(3):Paper No. 55, 22, 2023.
- [10] 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.
- [11] Hideo Kozono, Yutaka Terasawa, and Yuta Wakasugi. Asymptotic behavior and Liouville-type theorems for axisymmetric stationary Navier-Stokes equations outside of an infinite cylinder with a periodic boundary condition. J. Differential Equations, 365:905–926, 2023.
- [12] Yasunori Maekawa and Hiroyuki Tsurumi. Existence of the stationary Navier-Stokes flow in around a radial flow. J. Differential Equations, 350:202–227, 2023.
- [13] Konstantin Pileckas and Remigio Russo. On the existence of vanishing at infinity symmetric solutions to the plane stationary exterior Navier-Stokes problem. Math. Ann., 352(3):643–658, 2012.
- [14] Antonio Russo. A note on the exterior two-dimensional steady-state Navier-Stokes problem. J. Math. Fluid Mech., 11(3):407–414, 2009.
- [15] Remigio Russo. On Stokes’ problem. In Advances in mathematical fluid mechanics, pages 473–511. Springer, Berlin, 2010.
- [16] Masao Yamazaki. Unique existence of stationary solutions to the two-dimensional Navier-Stokes equations on exterior domains. In Mathematical analysis on the Navier-Stokes equations and related topics, past and future, volume 35 of GAKUTO Internat. Ser. Math. Sci. Appl., pages 220–241. Gakkotosho, Tokyo, 2011.
- [17] Masao Yamazaki. Two-dimensional stationary Navier-Stokes equations with 4-cyclic symmetry. Math. Nachr., 289(17-18):2281–2311, 2016.
- [18] Masao Yamazaki. Existence and uniqueness of weak solutions to the two-dimensional stationary Navier-Stokes exterior problem. J. Math. Fluid Mech., 20(4):2019–2051, 2018.
M. Higaki
Department of Mathematics, Graduate School of Science, Kobe University, 1-1 Rokkodai, Nada-ku, Kobe 657-8501, Japan.
Email: [email protected]
R. Horiuchi
Department of Mathematics, Graduate School of Science, Kobe University, 1-1 Rokkodai, Nada-ku, Kobe 657-8501, Japan.
Email: [email protected]