Rate of the enhanced dissipation for the two-jet Kolmogorov type flow on the unit sphere
Abstract.
We study the enhanced dissipation for the two-jet Kolmogorov type flow which is a stationary solution to the Navier–Stokes equations on the two-dimensional unit sphere given by the zonal spherical harmonic function of degree two. Based on the pseudospectral bound method developed by Ibrahim, Maekawa, and Masmoudi [15] and a modified version of the Gearhart–Prüss type theorem shown by Wei [48], we derive an estimate for the resolvent of the linearized operator along the imaginary axis and show that a solution to the linearized equation rapidly decays at the rate when the viscosity coefficient is sufficiently small as in the case of the plane Kolmogorov flow.
Key words and phrases:
Navier–Stokes equations, Kolmogorov type flow, enhanced dissipation2010 Mathematics Subject Classification:
35Q30, 35R01, 47A10, 76D051. Introduction
In this paper, as a continuation of [32], we consider the incompressible Navier–Stokes equations on the two-dimensional (2D) unit sphere in :
(1.1) |
Here is the tangential velocity of the fluid, is the scalar-valued pressure, and is a given tangential external force. Also, is the viscosity coefficient, is the covariant derivative of along itself, is the Hodge Laplacian via identification of vector fields and one-forms, and and are the gradient and the divergence on . Here the viscous term is taken to be the twice of the divergence of the deformation tensor :
where is the Ricci curvature of . The Navier–Stokes equations on spheres and more general manifolds with this kind of viscous term have been studied by many authors (see e.g. [43, 36, 33, 31, 9, 21, 5, 7, 39, 40, 35, 22, 37, 38]). There are also several works on the Navier–Stokes equations on manifolds in which the viscous term is taken to be by analogy of the flat domain case (see e.g. [17, 16, 4, 18, 51, 26, 41]). We refer to [12, 1, 11, 44, 6] for the above identity and the choice of the viscous term in the Navier–Stokes equations on manifolds.
Identifying vector fields with one-forms and taking in (1.1), where and are the Hodge star operator and the external derivative, we have the vorticity equation
(1.2) |
for the vorticity (see [32] for derivation). Here is the directional derivative of along and is the Laplace–Beltrami operator on which is invertible on , the space of functions on with zero mean, is the unit outward normal vector field on , and is the vector product in . Note that the vorticity equation (1.2) is equivalent to the Navier–Stokes equations (1.1) since any closed one-form on is exact.
For and let be the spherical harmonics which satisfy with (see Section 2). The vorticity equation (1.2) with external force has a stationary solution with velocity field
(1.3) |
for and , where is the parametrization of by the colatitude and the longitude (see Section 2). The flow (1.3) can be seen as the spherical version of the Kolmogorov flow which is a stationary solution to the Navier–Stokes equations in a 2D flat torus with shear external force (see e.g. [30, 19, 28, 34, 29] for the study of the stability of the plane Kolmogorov flow). Ilyin [18] called (1.3) the generalized Kolmogorov flow and studied its stability for the Navier–Stokes equations on with viscous term . Also, Sasaki, Takehiro, and Yamada [39, 40] called (1.3) an -jet zonal flow and investigated its stability for the Navier–Stokes equations on with viscous term . The stability of (1.3) for the Euler equations on was also studied by Taylor [42]. We call (1.3) the -jet Kolmogorov type flow in order to emphasize both the similarity to the plane Kolmogorov flow and the number of jets.
In this paper we consider the linear stability of the two-jet Kolmogorov type flow. We substitute for (1.2) and omit the nonlinear term with respect to to get
where is the identity operator (see [32] for derivation of the linearized equation). Replacing and by and , we rewrite this equation as
(1.4) |
which is considered in . Then since is nonnegative and self-adjoint in and is -compact in , the operator generates an analytic semigroup in by a perturbation theory of semigroups (see [13]) and thus the solution of (1.4) with initial data is given by . In [32] the second author of the present paper proved the linear stability of the two-jet Kolmogorov type flow for all by getting the exponential decay of towards a (not orthogonal) projection of onto the kernel of . Moreover, he showed that does not have eigenvalues in by making use of the mixing structure of expressed by a recurrence relation for , and applied it to find that the enhanced dissipation occurs for the rescaled flow , which is a solution to with , as in the case of an advection-diffusion equation [8, 52, 48]. More precisely, let
which is a closed subspace of invariant under the actions of and (see Section 3.1), and be the orthogonal projection from onto the orthogonal complement of the kernel of restricted on . Then it was shown in [32, Theorem 1.4] that
This result in particular gives the convergence of to zero in as for each fixed and , but does not give the actual convergence rate. The purpose of this paper is to give an explicit decay rate of the original flow .
In the plane case, Beck and Wayne [3] numerically conjectured that a perturbation of the plane Kolmogorov flow decays at the rate when the viscosity coefficient is sufficiently small. They also verified this enhanced dissipation for a linearized operator without a nonlocal term based on the hypocoercivity method developed by Villani [47]. Lin and Xu [27] proved the enhanced dissipation for a full linearized operator but without an explicit decay rate by using the Hamiltonian structure of a perturbation operator and the RAGE theorem. The decay rate was confirmed by Ibrahim, Maekawa, and Masmoudi [15] based on the pseudospectral bound method, by Wei and Zhang [49] based on the hypocoercivity method, and by Wei, Zhang, and Zhao [50] based on the wave operator method.
In this paper we show that decays at the rate as in the plane case. For let be the orthogonal projection from onto the space of functions on of the form (see (2.8) for the precise definition). Note that is orthogonally decomposed as and
for (see Section 3.1). The main result of this paper is as follows.
Theorem 1.1.
There exist constants such that
(1.5) |
for all , , , , and . Also, if and is sufficiently large, then for all and we have
where are constants independent of , , , , and .
Here we note that the exponent of in (1.5) is in the sphere case, while it is in the plane case (see [15, Corollary 3.14]). This difference comes from the factor appearing in the estimate for the -norm of by the -norm of (see Lemma 2.3) which yields a different coercive estimate for with , , and (see Lemma 3.8).
To prove Theorem 1.1 we follow the argument of the work by Ibrahim, Maekawa, and Masmoudi [15] which verified the enhanced dissipation for the plane Kolmogorov flow by the pseudospectral bound method. By rescaling time as in (1.4) and taking the Fourier series with respect to the longitude , we consider the equation
in for each , where and is the multiplication operator by . By the definition of , the operator is self-adjoint and positive in . Also, the kernel of is when and trivial when . We intend to estimate the semigroup generated by , especially or equivalently , where is the orthogonal projection from onto , the orthogonal complement of the kernel of . To this end, we apply abstract results given in Section 4 to to derive an estimate for the quantity
which was introduced in [14] and called the pseudospectral bound of in [15]. Then we get an estimate for the semigroup generated by by combining the estimate for the pseudospectral bound of with an abstract theorem for an -accretive operator on a weighted Hilbert space which is a version of the Gearhart–Prüss type theorem shown by Wei [48]. In order to use the abstract results, we need to confirm several assumptions for and . The main effort is to verify a coercive estimate for on with (see Section 3.2). This coercive estimate involves a loss of derivative, but thanks to a small factor in front it can be controlled by the smoothing effect of in the estimate for the pseudospectral bound of . When , we easily get the coercive estimate by taking the -inner product of with for and using the coerciveness of on (see Lemmas 3.7 and 3.8). On the other hand, the proof is more involved when (see Lemma 3.9). In this case, noting that is of the form , we analyze an ordinary differential equation (ODE) with variable corresponding to and prove the coercive estimate by a contradiction argument after giving auxiliary statements. The main difficulty comes from the nonlocalities of the orthogonal projection which is nontrivial when and of the inverse operator appearing in . These nonlocalities affect the analysis for different values of . The nonlocality of with is relevant to the case where is closed to zero, which is the eigenvalue of . On the other hand, the nonlocality of causes a difficulty when is close to , which are the critical values of appearing in , i.e. the derivative vanishes when . Hence we can deal with these difficulties separately by using a contradiction argument. Moreover, since is the orthogonal projection onto the kernel of which is spanned by the smooth function , we can handle the nonlocality of by introducing a suitable auxiliary function which involves . Also, when is close to , the use of a contradiction argument enables us to focus on analysis of functions concentrating around the critical points of , i.e. , for which the nonlocal term consisting of essentially becomes a small order by the smoothing effect of . By these facts we can verify the coercive estimate for , but the actual proof requires very long and careful calculations.
The coercive estimate for given in this paper is basically the same as the one in the plane case [15]. In the sphere case, the surface measure on is of the form under the spherical coordinate system. The weight function vanishes at the critical points of appearing in , so one may expect to have a better coercive estimate than the one in the plane case when and is close to , but we cannot get such an estimate. In fact, for with close to , a small parameter , and a function on of the form , we have a better bound due to the weight function when we estimate the -norm of on a narrow band on corresponding to by the -norm of on . However, we also have a factor when we use an interpolation type inequality which bounds the -norm on by the - and -norms of on (see Lemma 2.6), so the resulting coercive estimate is the same as in the plane case.
Lastly, let us explain difference from the plane case [15] in the proof of the coercive estimate. The authors of [15] derived coercive estimates for on with away from zero and for on with close to zero in order to avoid the nonlocality of when is away from zero. In this paper, however, we deal with for all since the nonlocality of can be handled by taking the inner product in when and by introducing an auxiliary function when . Also, we encounter an additional difficulty due to the larger coefficient of the nonlocal operator in . To prove the coercive estimate when and is close to (or ), we analyze an auxiliary function on of the form and derive an estimate for the -norm of on a subset of corresponding to (or ) with by using an ODE for (see Lemma 3.17). In the proof of the estimate, we need to show with a constant , where
Here is always nonnegative but may become negative by the presence of . Moreover, the coefficient of comes from the coefficient of in . In the plane case, one also needs to deal with integrals similar to and , but the coefficient of becomes since the coefficient of in is . Thus one can easily get an estimate just by using the zeroth order term of in . In our case, however, the use of only the zeroth order term does not work for since the coefficient of is too large. To overcome this difficulty, we take into account the first order term of in . In fact, it is natural to use both the zeroth and first order terms since
for the gradient of on . In the actual proof, we split into the integrals and over and with and estimate them separately. To we just use for . Also, we carry out integration by parts for and apply Young’s inequality to get the integral of a function involving the term . Then, taking an appropriate and estimating the integrand, we obtain an estimate for which gives the lower bound of when combined with the estimate for .
The rest of this paper is organized as follows. In Section 2 we fix notations and give auxiliary inequalities. Section 3 is devoted to the study of the linearized operator for the two-jet Kolmogorov type flow. We provide settings and basic results in Section 3.1, verify coercive estimates for in Section 3.2, and derive estimates for the semigroup generated by and prove Theorem 1.1 in Section 3.3. Section 4 gives abstract results used in Section 3. In Section 5 we show basic formulas of differential geometry on .
2. Preliminaries
In this section we fix notations and give auxiliary inequalities. We also refer to Section 5 for some notations and basic formulas of differential geometry.
Let be the unit sphere in equipped with the Riemannian metric induced by the Euclidean metric of . We denote by and the colatitude and the longitude so that is parametrized by for and . For a (complex-valued) function on , we sometimes abuse the notation
when no confusion may occur. Thus the integral of over is given by
where is the Hausdorff measure of dimension . As usual, we set
where is the complex conjugate of , and write , for the Sobolev spaces of functions on with (see [2]).
Let and be the gradient and Laplace–Beltrami operators on . It is well known (see e.g. [46, 45]) that is an eigenvalue of with multiplicity for each and the corresponding eigenvectors are the spherical harmonics
(2.1) |
Here , are the Legendre polynomials defined as
and the associated Legendre functions , , are given by
so that (see [23, 10]). Moreover, the set of all forms an orthonormal basis of , i.e. for each . It is also known that the recurrence relation
holds (see [23, (7.12.12)]) and thus (see also [46, Section 5.7])
(2.2) |
for and , where we consider . Note that the superscript of just corresponds to that of and does not mean the -th power.
Let be the space of functions on with vanishing mean, i.e.
Then is invertible and self-adjoint as a linear operator
Also, for , the operator is defined on by
(2.3) |
We easily observe by a density argument and integration by parts that
(2.4) |
By this relation and Poincaré’s inequality (see e.g. [2, Corollary 4.3])
(2.5) |
we have the norm equivalence
(2.6) |
Let be a function on . We write if is of the form
with some function of the colatitude and . In this case,
(2.7) |
by (5.1), where and is the covariant Hessian of , and
If , then for . In particular, if , then and we can apply (2.4)–(2.6) to . Also, if for , then by (2.7). We use these facts without mention.
For a function on and we define a function on by
(2.8) |
Note that and for if .
The following results are shown in [32, Section 2].
Lemma 2.1.
If with , then .
Lemma 2.2.
For with let
(2.9) |
Then for each we have
(2.10) |
Lemma 2.3.
For let . Then
(2.11) |
Lemma 2.4.
Let us give interpolation type inequalities for .
Lemma 2.5.
Let and with . Then
(2.13) | ||||
(2.14) |
for all , where .
Proof.
By the mean value theorem for integrals, we have
with some . Then we observe that
(2.15) |
by and (2.7). Let . Then
(2.16) |
Moreover, by for , (2.4), and (2.7),
(2.19) |
Hence (2.13) follows from (2.15)–(2.19). Also, for ,
by for and (2.7), and
We apply these estimates, , and the second line of (2.19) to (2.15) and (2.16), and then combine the resulting inequalities to obtain (2.14). ∎
Lemma 2.6.
For let . Also, let and
Then there exists a constant independent of , , , and such that
(2.20) | ||||
(2.21) |
for all , where .
Proof.
When , we set . Then since
we have . Also, by Taylor’s theorem for at ,
On the other hand, when ,
Hence we set to get and
Therefore, in both cases and , we have (2.20) and (2.21) by applying (2.13) and (2.14) with the above and using (2.6). Similarly, we can prove (2.20) and (2.21) when by considering the cases and separately. ∎
3. Analysis of the linearized operator
In this section we study the linearized operator for the two-jet Kolmogorov type flow.
3.1. Settings and basic results
Let be the identity operator and the multiplication operator by a function on , i.e. for a function on . We define linear operators and on by
where on . Then is self-adjoint in . Also, is densely defined in since its domain contains the dense subspace of . Since is a bounded operator on (note that it does not map into itself) and is a closed operator from into , we see that is closed in . Moreover, is -compact in since is compactly embedded in and
(3.1) |
by the elliptic regularity theorem. For and we have
(3.2) |
By this equality and (2.2), we see that (here )
(3.3) |
In particular, for . Let
Then is a closed subspace of and each is expressed as
(3.4) |
and we observe by (3.3) (in particular ) and (3.4) that is invariant under the actions of and . For the sake of simplicity, we write and for their restrictions on with domains and .
The follows lemmas are proved in [32, Section 5].
Lemma 3.1.
The operator is self-adjoint in and satisfies
(3.5) |
Also, is densely defined, closed, and -compact in .
Lemma 3.2.
For let . Then
(3.6) |
and the kernel of in is . Thus is the orthogonal projection from onto .
By Lemma 3.1 and a perturbation theory of semigroups (see [13]), the operator with domain generates an analytic semigroup in for each . Our aim is to study the decay rate of in terms of .
For , let and with given by (2.8). Then and are closed subspaces of and , respectively, and the orthogonal decompositions
hold. Moreover, since is of the form
(3.7) |
the subspace is invariant under the actions of and by (3.3) (in particular ). Hence is diagonalized as
Moreover, we see by (2.2), (3.2), and (3.7) that is expressed as
where and are linear operators on defined by
Note that, by , (3.2), (3.7), and for ,
(3.8) |
Also, is a closed subspace of and each is of the form
(3.9) |
Let be the orthogonal projection from onto . Then
(3.10) |
for by (3.7) and (3.9). We intend to derive decay estimates for the semigroup generated by in by applying abstract results given in Section 4. To this end, we verify Assumptions 4.1–4.3 and 4.6 for and .
Note that maps into itself by (3.2) and (3.7). Also, as in (2.3), we set
(3.11) |
for and of the form (3.7).
Proof.
Since is a closed subspace of , Lemma 3.1 implies that satisfies Assumption 4.1 in . The operator is densely defined, closed, and -compact in since it is a bounded operator on , the inequality (3.1) holds, and the embedding is compact. If for , then since is injective on . Thus, by (3.6), (3.7), and for , we see that
(3.12) |
and we get by (3.9) and (3.12). Also, we have in since is smooth on , the relations (3.3) and (3.10) hold, and is self-adjoint in . Since is of the form (3.9), we see by (3.2), (3.3), (3.11), and for that the inequalities (4.2) and (4.3) hold for and with constant . Hence Assumptions 4.2 and 4.3 are valid. ∎
Lemma 3.4.
For and we have
(3.13) | ||||||
(3.14) |
Proof.
The next result is crucial for the proof of Lemma 3.9 below.
Lemma 3.5.
Let . Then in has no eigenvalues in .
Proof.
The statement follows from [32, Theorem 4.1]. ∎
3.2. Verification of Assumption 4.6
Next we show that and satisfy Assumption 4.6 in several long steps. In what follows, we frequently use the fact that each function in is of the form (see (2.8)).
Lemma 3.6.
Let . For we have
(3.15) |
Also, let on . Then for we have
(3.16) |
Proof.
Next let and . Then
with on . Moreover, we have
by applying integration by parts once or twice, where we also used by (5.1) for and in the second line. Hence
Moreover, since
we have
(3.17) |
Now we recall that and is of the form (3.7) to get
(3.18) |
by (2.3), (2.4), and for . Also, since on by and (5.1), we have with . Thus we get (3.16) by this equality, (3.15), (3.17), and (3.18). ∎
Let us verify (4.8) and (4.10) in Assumption 4.6. We consider three cases for :
Here is a small constant given in Lemma 3.8 below and . In what follows, we write for a general positive constant independent of , , , and . Also, we frequently use the expressions on and for .
Lemma 3.7.
There exists a constant such that
(3.19) | ||||
(3.20) |
for all , , and satisfying .
Proof.
For let and . Then
since by (3.2) and (3.9). Moreover,
by and integration by parts. Hence
and we divide both sides by and take the real part to get
(3.21) |
Moreover, since is of the form (3.9) and for , we observe by , (2.3), and (2.4) that
(3.22) |
Applying (3.22) and on to the left-hand side of (3.21), and using Hölder’s and Young’s inequalities to the right-hand side, we obtain
and thus (note that )
(3.23) |
Hence we get (3.19) by (3.14) and (3.23). To prove (3.20), we see that
by . Taking the real part of this equality, we get
(3.24) |
Moreover, since on by and (5.1),
(3.25) |
by (2.6) and . Noting that
we deduce from (3.24) and (3.25) that
and applying (3.19) and (3.23) to the last line we obtain (3.20). ∎
Lemma 3.8.
There exist constants and such that
(3.26) | ||||
(3.27) |
for all , , , and satisfying
(3.28) |
Proof.
Let be a constant which will be determined later, and let and satisfy (3.28). We may assume since the case is similarly handled. Let and . Since , we can use (3.21) in the proof of Lemma 3.7 to get
Then, by , , (3.22), and Young’s inequality,
(3.29) |
Since , we can write . Then
by (2.7). Moreover, since for and
by (2.7), (2.11), and , it follows that
We further observe that by Taylor’s theorem for at and . Therefore,
We apply this inequality to (3.29) and use (3.13) with to find that
(3.30) |
with some constants independent of , , , , and . Now we define
(3.31) |
Then since , it follows from (3.30) that
(3.32) |
and we get (3.26) by (3.14) and (3.32). Also, since
by , (3.24), and (3.25), and since on ,
Thus we apply (3.26) and (3.32) to the last line to get (3.27). ∎
Lemma 3.9.
The proof of (3.33) relies on a contradiction argument. In order to focus on the part of getting a contradiction in that proof, we derive auxiliary estimates in the following lemmas. We assume in the sequel, since the case can be handled similarly.
In Lemmas 3.10–3.17 below, let be the constant given by (3.31) and , , and . Note that
(3.36) |
by and . For let
(3.37) |
i.e. in (2.9). Also, let
(3.40) |
and and be given by (3.35) for with . Hereafter we write for a general positive constant independent of , , , and .
Lemma 3.10.
Let be a positive constant satisfying
(3.41) |
Also, let . Then
(3.42) | |||
(3.43) |
Note that . Also,
(3.44) |
Proof.
Lemma 3.11.
We have
(3.45) | on | |||||
(3.46) | on |
Proof.
Lemma 3.12.
We have
(3.47) |
Proof.
Lemma 3.13.
We have
(3.49) |
Proof.
Let , which satisfies (3.41) and
(3.50) |
by (3.36). Also, let be given by (3.37). Since , we can write . Then we see by (2.20) and (3.13) that
for . By this inequality and (3.43),
(3.51) |
Also, since for by (3.45),
(3.52) |
Moreover, since , we have
by (2.7), (3.43), and . Also, we see by (3.44) that
We apply these inequalities to (3.52) and use (2.11) and (3.47) to to get
(3.53) |
By (3.14), (3.51), (3.53), and Young’s inequality, we find that
Proof.
Lemma 3.15.
We have
(3.55) |
and . Moreover, for all ,
(3.56) |
Proof.
Since by the Sobolev embedding theorem (see [2]), and since , we can write (3.55). Also, since and are continuous near and for by (3.45), we have at , otherwise does not belong to . Hence .
Let us show (3.56). For let , which satisfies (3.41) and (3.50) by (3.36). By the mean value theorem for integrals, there exists such that
Then since by (3.42), we have
(3.57) |
by (2.7). Also, we use Hölder’s inequality, (2.7), and (3.42) to get
(3.58) |
where is given by (3.37). Moreover, since
by (3.36) and (3.42), we see by (3.49), (3.50), and (3.54) that
(3.59) |
By , (3.57), and (3.58), we have
Lemma 3.16.
Proof.
By Lemma 3.15 we have (3.55) and . Since
by (3.45), we take the -inner products of both sides with and carry out integration by parts for to get
(3.62) |
Let us estimate the right-hand side. We may assume since
For let and . Then by (3.36). Also, and thus satisfy (3.41) by . Let
Then we can write
(3.63) |
where
We estimate , , and separately. For , we have
Since , we can use (2.12) to get
where is given by (3.37). Also, we see by (2.11) and (3.43) that
We use these inequalities, (3.47), (3.59), and by (3.36) to deduce that
(3.64) |
Note that we can use (3.59) since satisfies (3.41) and (3.50) by (3.36).
To estimate , we decompose as
We apply Hölder’s inequality and (2.12) to to get
Moreover, by (2.11), (3.43), (3.47), and ,
(3.65) |
Similarly, by Hölder’s inequality, (2.4), (2.12), and ,
(3.66) |
Here is the length of . For , since
by (3.44), it follows from (2.11) and that
(3.67) |
To estimate , we see that
Moreover, since and satisfy (3.41), we can use (3.43) to get
Hence we deduce from this inequality and (2.11) that
(3.68) |
For , it follows from (2.7) and (2.11) that
Moreover, since ,
(3.69) |
by (3.43) and . Hence
(3.70) |
Noting that , we apply (3.65)–(3.70) to to get
(3.71) |
Let us estimate . We write with
(3.72) |
To estimate , we split into
We use Hölder’s inequality, (2.4), (2.7), and (2.12) to to get
Also, if , i.e. , then it follows from (3.44) that
By this inequality, Hölder’s inequality, and (2.7), we have
These estimates, (3.36), and (3.47) imply that
(3.73) |
Also, as in (3.69), we see by (3.43) and that
We deduce from this inequality, (2.7), (2.11), and by (3.36) that
(3.74) |
Applying (3.73) and (3.74) to , we get
(3.75) |
Now we deduce from (3.62)–(3.64), (3.71), (3.75), and that
In this inequality, we further use (3.56) with and apply (2.4) to (note that we assume with ). Then we obtain (3.60) since .
Let us show (3.61). We set for (see (3.72)) to get
Moreover, since for and
for by (3.36) and (3.44), we see by (2.7) that
(3.76) |
where is given by (2.9). Also, since ,
Note that when . By this fact and ,
Thus, setting in (3.62) and using (3.63), we have
and we use (3.64), (3.71), (3.74), (3.76), and to find that
(3.77) |
Moreover, since , we can use (3.22) in the proof of Lemma 3.7 to get
We apply this inequality, (3.47) for , and (3.56) with to (3.77), and then use to obtain (3.61). ∎
Lemma 3.17.
Proof.
Since by (3.46), we use (3.55), (5.1), and to rewrite this equation as
for . We multiply both sides by and integrate them over . Then we carry out integration by parts and use and to get
We take the real part of this equality and use
by integration by parts and to get
(3.79) |
where
(3.80) |
Now we claim that there exist and such that
(3.81) |
for all . If this claim is valid, then we apply Hölder’s and Young’s inequalities, (2.7), (3.81), , and to (3.79) to get
and we subtract from both sides, dividing them by , and use (3.36) and (3.56) with to obtain (3.78).
Let us show (3.81). We split into
where is fixed later. By for ,
(3.82) |
Since and , we have by Lemma 2.1. By this fact, , and integration by parts, we have
where . Hence
(3.83) |
by Young’s inequality and (see (3.80)). To estimate , let
where . Then
(3.84) |
Now let and . Then for we see by the mean value theorem for with and that
(3.85) |
and thus by . Hence, by (3.84) and ,
Moreover, by , , (3.85), and ,
Hence and
(3.86) |
with by (3.83). Moreover,
(3.87) |
by , and for , and . Thus, noting that , we combine (3.82), (3.86), and (3.87) and apply (2.7) to to get (3.81). ∎
Proof of (3.33).
As we mentioned above, we assume and prove (3.33) by contradiction. Assume that for each there exist , , , and such that, if we define , , , and by (3.35) and (3.40) with , , , and replaced by , , , and , then
(3.88) | |||
(3.89) |
Taking subsequences, we may assume that
Suppose first that . Then for sufficiently large . Moreover, we see by (3.47) and the boundedness of on that
By these facts, , (2.6), and (3.89), we find that
as , which contradicts (3.88).
Now let . We see by (3.49) for , (3.88), and (3.89) that
(3.90) |
Let . Then we see by Lemma 3.15 that
(3.91) |
Also, by (3.56) for , (3.88), and (3.89),
for all and thus
(3.92) |
In what follows, we consider two cases.
Case 1: . If , then for sufficiently large . Then since is of the form (3.9) with , it follows from (2.3), for , and (3.88) that
(3.93) |
We set in (3.60) for and use (2.4) and (3.88) to find that
for all . Then we send and use (3.89), (3.93), and to get
Hence . Also, by (3.47) and (3.93). By these facts and we get a contradiction with (3.90).
Next suppose that . Taking a subsequence, we may assume and thus for all . Since is bounded in by (2.6) and (3.88), and since is compactly embedded into , we may assume that
(3.94) |
with some by taking a subsequence again. Then since for all and is closed in . If , i.e. as , then we can get a contradiction with (3.90) as in the case of . Suppose that . Since and , where we consider if ,
Hence by (3.35) for , and
(3.95) |
by (3.94), where is given by (2.2) if and if . For each and , we see by (3.60) for and (3.88) that
We send , use (3.89), (3.94), and , and then let to get
for all , which shows that since is dense in . Hence by the elliptic regularity theorem. Now let satisfy
(3.96) |
Then, as in the proof of Lemma 3.16, we can get (see (3.62))
We send and use (3.89), (3.94), (3.95), and uniformly on by (3.96). Then we get
for all satisfying (3.96). Since , this equality shows that
(3.97) |
Let . Then and since , , and maps into itself and is invertible on by (3.2) and (3.9). Moreover, we see by (3.95), (3.97), and that satisfies
(3.98) |
where .
Now suppose that . Then in by (3.98) and . Thus, if , then is a nonzero eigenvalue of since , but this contradicts Lemma 3.5. Also, if , then is in the kernel of and thus by (3.12), which contradicts .
Let and . Then, by (3.98) and ,
Moreover, since and by . Hence is a nonzero eigenvalue of , which contradicts Lemma 3.5.
Suppose that and . Since by (2.2) with and , we apply this equality and to (3.98) to get
Then since is of the form (3.9), we have
by (3.2) (note that ). Since , we find by this equality and (3.14) that , which contradicts .
Let and . Since is of the form with a constant by (2.1), we see by (3.98) and that
and thus , otherwise does not belong to . Hence by the above equality. Since , this fact and (3.14) imply that , which contradicts . This completes the proof in Case 1.
Case 2: . In this case, for sufficiently large , where is given in Lemma 3.17. Then we see by (3.78) for , (3.88), and (3.89) that
for all , where is given by (2.9). Hence
(3.99) |
Suppose . Then
(3.100) |
Also, for each we see by (3.61) for and (3.88) that
We send and use (3.89) and to get
for all . Thus and we combine this with (3.100) to get a contradiction with (3.90). Now we assume that
(3.101) |
Since and , we have
Moreover, by Taylor’s theorem for at ,
By the above relations, , and , we find that
Hence as . Based on this observation, we set
(3.102) |
Then
(3.103) |
and thus and for sufficiently large . Also,
(3.104) |
by for . We define
(3.105) | |||
Then, by , , (3.91), and (3.92),
(3.106) |
Also, we use (2.7), set , and apply (3.102) and (3.104) to get
(3.107) |
Similarly, since , we have
(3.108) |
By these inequalities, (3.47), and (3.88), we find that
(3.109) |
Also, as in (3.107), we use (2.7), (3.102), (3.104), and to get
By this inequality and (3.101), we obtain
(3.110) |
Suppose that . Then since
by (2.7) and , it follows from (3.109) that
which contradicts (3.110).
Now let . Taking a subsequence, we may assume that
(3.111) |
Since is bounded in by (3.109) and is compactly embedded into , we may assume, by taking a subsequence again, that
(3.112) |
with some . Then by (3.110) and (3.112). For each we see by (2.10) and (3.112) that converges to strongly in . Moreover, and for by (3.111). Hence
Let us show . First we prove
(3.113) |
We see that by , (2.7), and
(3.114) |
Moreover, by (2.7), (3.112), and (3.114),
(3.115) |
For each , the linear functional
is bounded on . Indeed, by Hölder’s inequality, (3.114), and (5.1),
Hence by (3.112). By this fact and
we find that
(3.116) |
To get , we derive an ODE for , . By (3.105) we have for , where and are the -th derivatives of and . We use this relation and (3.114) and calculate as in (3.107) to get
Thus by . Also, by (3.46),
Using , (3.91) with , and (5.1), we rewrite this equation as
We multiply both sides by , set , and use (3.105) to get
(3.117) |
for , where
Since and
for by Taylor’s theorem and for ,
(3.118) |
Also, since and
by the Taylor series for at , it follows that
(3.119) |
for . Here
where we rearranged the double summation and defined
Since as by , we may assume that for sufficiently large . By this fact and , we have
Hence we easily find that converges absolutely for all and that and its derivative are bounded on uniformly in . Then since
by (3.119), there exists a constant independent of such that
(3.120) | |||
(3.121) |
by (3.103), , and the uniform boundedness of and .
Now let . We take the -inner product of (3.117) with and carry out integration by parts for the left-hand side to get
(3.122) |
We send , use (3.115), (3.116), and (3.118), and then apply Hölder’s inequality and and for to get
(3.123) |
Let us estimate the last term as in the proof of Lemma 3.16. We define
and extend it to by zero outside . Then
(3.124) |
by (3.114) and (3.120) with a constant independent of . Let
for . Then
by , , and (3.102). Hence
(3.125) |
and for sufficiently small since . Let
Then
(3.126) |
where
Let us estimate . For let and
Since by (3.106), we can use Hardy’s inequality to get
(3.127) |
Noting that , , and , we set in the second line of (3.127) and calculate as in (3.108) by using (2.7) and (3.104) to get
(3.128) |
where (see (2.9)). Moreover, noting that
by (3.42), we apply (3.47) and (3.54) to (3.128) to deduce that
Also, since ,
Hence by the above inequalities, (3.49) for , and (3.88) (recall that is given by (3.40)). By this inequality, , (3.124), and (3.125), we find that
(3.129) |
Next we estimate . We split into
Also, let
As in (3.127) and (3.128), we apply Hardy’s inequality to , set , and use (2.7) and (3.104). Then we further use (3.47) and (3.88) to get
(3.130) |
Also, by Hardy’s inequality, , and (3.124),
(3.131) |
We use Hölder’s inequality, , (3.124), (3.125), and (3.130) to get
(3.132) |
Also, by Hölder’s inequality, , , and (3.131),
(3.133) |
We have since
For , we see that
As in (3.107), we use , (2.7), and (3.104) to get
(3.134) |
Also, we observe by (3.125) that
We deduce from the above inequalities and (3.124) that
Thus, by this inequality, , (3.132), (3.133), and ,
(3.135) |
To estimate , we decompose into
We apply Hölder’s inequality to and use (3.124) and (3.130) to get
When , we have by (3.125). Thus
by , , and (3.124). Similarly,
by Hölder’s inequality, (3.124), (3.134), and . Hence
(3.136) |
Now we deduce from (3.126), (3.129), (3.135), and (3.136) that
for all , where we also used and . Then we send , use (3.89) and (3.106), and let to find that
Therefore, applying this inequality to (3.123), we obtain
for all , which gives (3.113) since is dense in .
Next we derive an ODE for . For , let satisfy near . Then we observe by (3.122) that
We send in this equality. Then since
as by (3.89), (3.114), (3.134), and , and since
by (3.103), (3.121), and near , we find by (3.115), (3.116), and (3.118) that
Since this equality is valid for all and vanishing near , and since by (3.113), we obtain
(3.137) |
Now we observe that for since . We apply this inequality and calculate as in (3.107) by using (3.104) and . Then
Moreover, we apply (2.7) and (3.102) to the last term and use (3.99) to get
By this fact, (3.103), and (3.115) with , we find that
Thus on and, since by (3.113), we further get
(3.138) |
Then we easily find that any solution of (3.137)–(3.138) in must be trivial by the standard theory of ODEs, since , , and given by (3.118) and (3.121) are smooth on and the singularity of at is of order one. Hence , i.e. , which contradicts . This completes the proof in Case 2, and we conclude that (3.33) is valid. ∎
Proof of (3.34).
We assume for simplicity, since the case can be handled similarly. Let be the constant given by (3.31) and , , and . Also, for with , let , , and be given by (3.35) and (3.40). We set , which satisfies (3.41) by (3.36). Since , we can write
For , we observe by (2.21) and (3.42) that
By this inequality, (3.43), and Young’s inequality,
(3.139) |
Since for by (3.45),
By this equality and (see (3.44)),
Moreover, if , then . By these inequalities, , and (2.7), we find that
(3.140) |
To estimate the last term, we see that
where
We use Hardy’s inequality, (2.7), and (3.47) to get
Also, by . Hence
We combine this inequality, (3.139), and (3.140) and use (2.6) to to get
(3.141) |
Let us estimate the last term. By Lemma 3.15 we have . Fix . Since satisfies (3.41), there exists such that
as in (3.57) and (3.58), where is given by (3.37). Thus
where we also used . We apply this inequality to (3.141) and then use (3.54), , and to get
In this inequality, we fix so that , apply (3.33) to , and use and (3.36) to obtain (3.34). ∎
Now we are ready to verify Assumption 4.6 for and . Let be the constant given by (3.31). For , , and , we define
Note that and for .
Lemma 3.18.
Let . Then and satisfy Assumption 4.6 with the functions and given above and the operator on .
Proof.
Since is a bounded operator on , it is closed in and
Also, (4.9) holds by (3.16). Let us verify (4.8) and (4.10). If and , then (4.8) is valid by (3.19) and . Also, if and , then (4.8) follows from (3.15), (3.26), and (3.33) with . Let and . For of the form (3.9), we have
by (2.3), for , and (3.15). Moreover,
since , , and . Hence (4.8) is valid.
3.3. Estimates for the semigroup
By Lemmas 3.3 and 3.18, and satisfy Assumptions 4.1–4.3 and 4.6. Moreover, by (3.13) and the constants in (4.2), (4.3), and (4.8)–(4.10) can be taken independently of . Hence we can apply the abstract results in Section 4 to to get the following results.
Theorem 3.19.
Let . Then
(3.142) |
for all . Moreover, for all and we have
(3.143) |
Here is a constant independent of , , and . Also,
(3.144) |
for and .
Proof.
By Lemma 4.4 and Theorem 4.7 with replaced by we have (3.142) for all and
(3.145) |
for all and . Here is a constant independent of , , and . Also,
(3.146) |
for . Let us estimate . In what follows, we write for general positive constant depending only on the constant given by (3.31).
When , let . Then and thus for . Also,
By these relations and , we find that
(3.147) |
When , we set . Then
for . Hence , , and
By these equalities and , we find that
(3.148) |
When , we take
Then for we see by , , and that
Hence , , and
Therefore,
(3.149) |
Hence we get (3.143)–(3.144) by (3.145) and (3.147)–(3.149). ∎
Theorem 3.20.
Let . Then generates an analytic semigroup in for all . Moreover,
(3.150) |
for all , where and are positive constants independent of , , and .
Proof.
When , on by (3.10) and thus (3.150) gives a decay estimate for . On the other hand, when , does not vanish for , so we need to estimate the -part of .
Theorem 3.21.
Let . Then for all , , and such that is sufficiently large, we have
(3.152) |
where are constants independent of , , , and .
Proof.
Let (see (3.12)). Then
(3.153) |
for since by (3.3) and . Also, by (4.5),
(3.154) |
Since is self-adjoint and satisfies (3.5) in and is -compact in , we see by a perturbation theory of semigroups (see [13]) that is expressed by the Dunford integral
where is any piecewise smooth curve in going from to with any and located on the right-hand side of . We apply to the above expression to get
(3.155) |
by (4.6) with replaced by , where
Noting that is located on the left-hand side of , we use (3.153) and the residue theorem (after replacing by a small circle centered at ) to deduce that
(3.156) |
Next we estimate . For and let
Then since , where is the function given by (3.144), we see by (3.143) with replaced by that
(3.157) |
with a constant independent of , , and . Also,
(3.160) |
by (3.157) and a standard Neumann series argument. Let
We take sufficiently large so that and set
Then we define a piecewise smooth curve
For we have by . Hence
(3.161) |
and
which shows . Similarly, for and we have
(3.162) |
and thus . Hence it follows from (3.154) and (3.160) that
and we can take this as the path of the integral for to get , where
Let us estimate each . In what follows, we write and for general positive constant depending only on . For we have
and and , where
By these relations, (3.8), (3.153), (3.160), (3.161), we find that
Moreover, since , we have and thus when is large. Then since and ,
for sufficiently large and thus
by the mean value theorem for with . Hence
(3.163) |
For we have and
by . It follows from these inequalities, (3.8), (3.153), (3.160), (3.162), and that
(3.164) |
Similarly, since , , and
by , we see by (3.8), (3.153), (3.160), and (3.162) that
(3.165) |
where is the length of . For we have
Note that here . By these inequalities, (3.8), (3.153), (3.160), and (3.162),
(3.166) |
where we also used in the last inequality. Now, noting that is sufficiently large, we deduce from (3.163)–(3.166) that
Hence we get (3.152) by applying this inequality and (3.156) to (3.155). ∎
Now we recall that on is diagonalized as
where is the operator given by (2.8). Moreover, for we have
where is the orthogonal projection from onto (see Lemma 3.2). Hence by Theorems 3.20 and 3.21 we get the next result which implies Theorem 1.1.
Theorem 3.22.
There exist constants such that
for all , , , and (note that on if ). Also, if and is sufficiently large, then for all and we have
where are constants independent of , , , and .
4. Abstract results
In this section we present abstract results for a perturbed operator.
For a linear operator on a Banach space , we denote by , , and the domain, the resolvent set, and the spectrum of in . Also, let and be the kernel and range of in .
Let be a Hilbert space and and linear operators on . We make the following assumptions.
Assumption 4.1.
The operator is self-adjoint in and satisfies
(4.1) |
with some constant .
Assumption 4.2.
The following conditions hold:
-
(i)
The operator is densely defined, closed, and -compact in .
-
(ii)
Let be the orthogonal complement of in and the orthogonal projection from onto . Then in .
Assumption 4.3.
There exist a Hilbert space , a closed symmetric operator on , and a bounded self-adjoint operator on such that the following conditions hold:
-
(i)
The inclusion holds and for all .
-
(ii)
The relation holds in and
-
(iii)
There exists a constant such that
(4.2) (4.3)
Note that is a linear operator on the original space , not on the auxiliary space . Also, the operator on does not necessarily map into itself.
By in Assumption 4.2 we can consider as a linear operator
In what follows, we use the notation for simplicity. Let be the orthogonal projection from onto (note that is closed in since is closed). Then and we can also consider as a linear operator
Note that and are closed in and in , respectively. Also, is -compact in . For we define a linear operator on by
and consider on with domain .
Our purpose is to derive a decay estimate for the -part of the semigroup generated by in terms of . The following results were obtained in [32, Lemmas 6.4–6.6].
Lemma 4.4.
Moreover, the next result was shown in [32, Theorem 6.7] by application of the Gearhart–Prüss type theorem given by Wei [48] to the -accretive operator on the Hilbert space equipped with the weighted inner product .
Theorem 4.5.
In order to get an upper bound of by a function of , we make an addition assumption on and .
Assumption 4.6.
There exist bounded nonnegative functions
a constant , a Banach space , and a closed operator on such that the following conditions hold:
-
(i)
For all and we have
(4.8) -
(ii)
The inclusions and hold and
(4.9) Moreover, for all and we have
(4.10)
Let us give an estimate for the resolvent of along the imaginary axis which was originally shown in [15, Theorem 2.9] under assumptions on coercive estimates for on with away from zero and for on with close to zero.
Theorem 4.7.
Proof.
Let and . Then
by (4.5) and (4.6). By these relations, we easily find that the first equality of (4.11) holds. Let us prove the second inequality of (4.11). Without loss of generality, we may assume . We set and abbreviate to for and . In what follows, we denote by a general positive constant depending only on and the constants appearing in (4.2), (4.3), and (4.8)–(4.10). Let and
We take the inner products of both sides with in . Then
(4.13) |
Since , , and is self-adjoint in , we have
Thus, taking the imaginary part of (4.13), we obtain
By this equality and (4.9) we get
and therefore
(4.14) |
Moreover, for each it follows from (4.10) and (4.14) that
Hence, subtracting from both sides and taking the square root of the resulting inequality, we obtain
We apply this inequality to (4.14) to find that
For each , we deduce from (4.8) and the above inequality that
(4.15) |
To estimate , we observe by that
We take the real part of this equality and use (4.4) and by the self-adjointness of in to get . Hence
by (4.3) and the boundedness of . Applying this inequality to (4.15) we obtain
We further subtract from both sides and take the square root of the resulting inequality to find that
for all . Since , the above inequality shows that
for all with . Therefore, the second inequality of (4.11) is valid. ∎
Remark 4.8.
Contrary to [15, Assumption 1], we do not assume that has a compact resolvent in Assumption 4.1. Thus the essential spectrum of in may be not empty, but we must have for the essential spectrum of in under Assumptions 4.1–4.3 and 4.6. Indeed, and for all by Theorem 4.7. Thus, by a standard Neumann series argument,
On the other hand, since is -compact in and the essential spectrum is invariant under a relatively compact perturbation (see [20, Theorem IV.5.35]),
Hence, if contains some , then and thus for all , but this is impossible since by for . Thus . In particular, if .
Theorem 4.9.
5. Appendix: basic formulas of differential geometry
This section gives some notations and basic formulas of differential geometry. We refer to e.g. [24, 25] for details.
Let be a two-dimensional Riemannian manifold. For a local coordinate system of , let and be the coordinate frame and its dual coframe. We set for and denote by the inverse matrix of so that the inner products of one-forms and -tensor fields on are given by and
Let be the Christoffel symbols of given by
For a (complex-valued) function on , we write and for the gradient and the covariant Hessian of , respectively, which are locally expressed as
with for . Then
Also, the Laplace–Beltrami operator on is locally given by
We use these expressions under the spherical coordinate system of .
Lemma 5.1.
Let be the spherical coordinate system
Then for a function on we have
(5.1) |
Proof.
We use the index instead of . Since
we observe by direct calculations that
and the other , , and vanish identically. Hence
and we obtain (5.1) by the above expressions. ∎
Acknowledgments
The work of the first author was supported by JSPS KAKENHI Grant Numbers 20K03698, 19H05597, 20H00118. Also, the work of the second author was supported by Grant-in-Aid for JSPS Fellows No. 19J00693.
References
- [1] R. Aris. Vectors, tensors and the basic equations of fluid mechanics. Mineola, NY: Dover Publications, reprint of the 1962 original edition, 1989.
- [2] T. Aubin. Some nonlinear problems in Riemannian geometry. Springer Monographs in Mathematics. Springer-Verlag, Berlin, 1998.
- [3] M. Beck and C. E. Wayne. Metastability and rapid convergence to quasi-stationary bar states for the two-dimensional Navier-Stokes equations. Proc. Roy. Soc. Edinburgh Sect. A, 143(5):905–927, 2013.
- [4] C. Cao, M. A. Rammaha, and E. S. Titi. The Navier-Stokes equations on the rotating -D sphere: Gevrey regularity and asymptotic degrees of freedom. Z. Angew. Math. Phys., 50(3):341–360, 1999.
- [5] C. H. Chan and M. Czubak. Non-uniqueness of the Leray-Hopf solutions in the hyperbolic setting. Dyn. Partial Differ. Equ., 10(1):43–77, 2013.
- [6] C. H. Chan, M. Czubak, and M. M. Disconzi. The formulation of the Navier-Stokes equations on Riemannian manifolds. J. Geom. Phys., 121:335–346, 2017.
- [7] C. H. Chan and T. Yoneda. On the stationary Navier-Stokes flow with isotropic streamlines in all latitudes on a sphere or a 2D hyperbolic space. Dyn. Partial Differ. Equ., 10(3):209–254, 2013.
- [8] P. Constantin, A. Kiselev, L. Ryzhik, and A. Zlatoš. Diffusion and mixing in fluid flow. Ann. of Math. (2), 168(2):643–674, 2008.
- [9] M. Dindoš and M. Mitrea. The stationary Navier-Stokes system in nonsmooth manifolds: the Poisson problem in Lipschitz and domains. Arch. Ration. Mech. Anal., 174(1):1–47, 2004.
- [10] NIST Digital Library of Mathematical Functions. http://dlmf.nist.gov/, Release 1.1.1 of 2021-03-15. F. W. J. Olver, A. B. Olde Daalhuis, D. W. Lozier, B. I. Schneider, R. F. Boisvert, C. W. Clark, B. R. Miller, B. V. Saunders, H. S. Cohl, and M. A. McClain, eds.
- [11] L. R. Duduchava, D. Mitrea, and M. Mitrea. Differential operators and boundary value problems on hypersurfaces. Math. Nachr., 279(9-10):996–1023, 2006.
- [12] D. G. Ebin and J. Marsden. Groups of diffeomorphisms and the motion of an incompressible fluid. Ann. of Math. (2), 92:102–163, 1970.
- [13] K.-J. Engel and R. Nagel. One-parameter semigroups for linear evolution equations, volume 194 of Graduate Texts in Mathematics. Springer-Verlag, New York, 2000. With contributions by S. Brendle, M. Campiti, T. Hahn, G. Metafune, G. Nickel, D. Pallara, C. Perazzoli, A. Rhandi, S. Romanelli and R. Schnaubelt.
- [14] I. Gallagher, T. Gallay, and F. Nier. Spectral asymptotics for large skew-symmetric perturbations of the harmonic oscillator. Int. Math. Res. Not. IMRN, (12):2147–2199, 2009.
- [15] S. Ibrahim, Y. Maekawa, and N. Masmoudi. On pseudospectral bound for non-selfadjoint operators and its application to stability of Kolmogorov flows. Ann. PDE, 5(2):Paper No. 14, 84, 2019.
- [16] A. A. Il′in. Navier-Stokes and Euler equations on two-dimensional closed manifolds. Mat. Sb., 181(4):521–539, 1990.
- [17] A. A. Il′n and A. N. Filatov. Unique solvability of the Navier-Stokes equations on a two-dimensional sphere. Dokl. Akad. Nauk SSSR, 301(1):18–22, 1988.
- [18] A. A. Ilyin. Stability and instability of generalized Kolmogorov flows on the two-dimensional sphere. Adv. Differential Equations, 9(9-10):979–1008, 2004.
- [19] V. I. Iudovich. Example of the generation of a secondary stationary or periodic flow when there is loss of stability of the laminar flow of a viscous incompressible fluid. Journal of Applied Mathematics and Mechanics, 29(3):527–544, 1965.
- [20] T. Kato. Perturbation theory for linear operators. Springer-Verlag, Berlin-New York, second edition, 1976. Grundlehren der Mathematischen Wissenschaften, Band 132.
- [21] B. Khesin and G. Misioł ek. Euler and Navier-Stokes equations on the hyperbolic plane. Proc. Natl. Acad. Sci. USA, 109(45):18324–18326, 2012.
- [22] M. Kohr and W. L. Wendland. Variational approach for the Stokes and Navier-Stokes systems with nonsmooth coefficients in Lipschitz domains on compact Riemannian manifolds. Calc. Var. Partial Differential Equations, 57(6):Paper No. 165, 41, 2018.
- [23] N. N. Lebedev. Special functions and their applications. Dover Publications, Inc., New York, 1972. Revised edition, translated from the Russian and edited by Richard A. Silverman, Unabridged and corrected republication.
- [24] J. M. Lee. Introduction to smooth manifolds, volume 218 of Graduate Texts in Mathematics. Springer, New York, second edition, 2013.
- [25] J. M. Lee. Introduction to Riemannian manifolds, volume 176 of Graduate Texts in Mathematics. Springer, Cham, 2018. Second edition of [ MR1468735].
- [26] L. A. Lichtenfelz. Nonuniqueness of solutions of the Navier-Stokes equations on Riemannian manifolds. Ann. Global Anal. Geom., 50(3):237–248, 2016.
- [27] Z. Lin and M. Xu. Metastability of Kolmogorov flows and inviscid damping of shear flows. Arch. Ration. Mech. Anal., 231(3):1811–1852, 2019.
- [28] C. Marchioro. An example of absence of turbulence for any Reynolds number. Comm. Math. Phys., 105(1):99–106, 1986.
- [29] M. Matsuda and S. Miyatake. Bifurcation analysis of Kolmogorov flows. Tohoku Math. J. (2), 54(3):329–365, 2002.
- [30] L. D. Mešalkin and J. G. Sinaĭ. Investigation of the stability of a stationary solution of a system of equations for the plane movement of an incompressible viscous liquid. J. Appl. Math. Mech., 25:1700–1705, 1961.
- [31] M. Mitrea and M. Taylor. Navier-Stokes equations on Lipschitz domains in Riemannian manifolds. Math. Ann., 321(4):955–987, 2001.
- [32] T.-H. Miura. Linear stability and enhanced dissipation for the two-jet kolmogorov type flow on the unit sphere. arXiv:2105.07964.
- [33] T. Nagasawa. Construction of weak solutions of the Navier-Stokes equations on Riemannian manifold by minimizing variational functionals. Adv. Math. Sci. Appl., 9(1):51–71, 1999.
- [34] H. Okamoto and M. Shōji. Bifurcation diagrams in Kolmogorov’s problem of viscous incompressible fluid on -D flat tori. Japan J. Indust. Appl. Math., 10(2):191–218, 1993.
- [35] V. Pierfelice. The incompressible Navier-Stokes equations on non-compact manifolds. J. Geom. Anal., 27(1):577–617, 2017.
- [36] V. Priebe. Solvability of the Navier-Stokes equations on manifolds with boundary. Manuscripta Math., 83(2):145–159, 1994.
- [37] J. Prüss, G. Simonett, and M. Wilke. On the Navier-Stokes equations on surfaces. J. Evol. Equ., 2020.
- [38] M. Samavaki and J. Tuomela. Navier-Stokes equations on Riemannian manifolds. J. Geom. Phys., 148:103543, 15, 2020.
- [39] E. Sasaki, S. Takehiro, and M. Yamada. Linear stability of viscous zonal jet flows on a rotating sphere. Journal of the Physical Society of Japan, 82(9):094402, 2013.
- [40] E. Sasaki, S. Takehiro, and M. Yamada. Bifurcation structure of two-dimensional viscous zonal flows on a rotating sphere. J. Fluid Mech., 774:224–244, 2015.
- [41] Y. N. Skiba. Mathematical problems of the dynamics of incompressible fluid on a rotating sphere. Springer, Cham, 2017.
- [42] M. Taylor. Euler equation on a rotating surface. J. Funct. Anal., 270(10):3884–3945, 2016.
- [43] M. E. Taylor. Analysis on Morrey spaces and applications to Navier-Stokes and other evolution equations. Comm. Partial Differential Equations, 17(9-10):1407–1456, 1992.
- [44] M. E. Taylor. Partial differential equations III. Nonlinear equations, volume 117 of Applied Mathematical Sciences. Springer, New York, second edition, 2011.
- [45] G. Teschl. Mathematical methods in quantum mechanics, volume 157 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, second edition, 2014. With applications to Schrödinger operators.
- [46] D. A. Varshalovich, A. N. Moskalev, and V. K. Khersonskiĭ. Quantum theory of angular momentum. World Scientific Publishing Co., Inc., Teaneck, NJ, 1988. Irreducible tensors, spherical harmonics, vector coupling coefficients, symbols, Translated from the Russian.
- [47] C. Villani. Hypocoercivity. Mem. Amer. Math. Soc., 202(950):iv+141, 2009.
- [48] D. Wei. Diffusion and mixing in fluid flow via the resolvent estimate. Sci. China Math., 64(3):507–518, 2021.
- [49] D. Wei and Z. Zhang. Enhanced dissipation for the Kolmogorov flow via the hypocoercivity method. Sci. China Math., 62(6):1219–1232, 2019.
- [50] D. Wei, Z. Zhang, and W. Zhao. Linear inviscid damping and enhanced dissipation for the Kolmogorov flow. Adv. Math., 362:106963, 103, 2020.
- [51] D. Wirosoetisno. Navier-Stokes equations on a rapidly rotating sphere. Discrete Contin. Dyn. Syst. Ser. B, 20(4):1251–1259, 2015.
- [52] A. Zlatoš. Diffusion in fluid flow: dissipation enhancement by flows in 2D. Comm. Partial Differential Equations, 35(3):496–534, 2010.