The number of traveling wave families in a running water with Coriolis force
Abstract.
In this paper, we study the number of traveling wave families near a shear flow under the influence of Coriolis force, where the traveling speeds lie outside the range of the flow . Under the -plane approximation, if the flow has a critical point at which attains its minimal (resp. maximal) value, then a unique transitional value exists in the positive (resp. negative) half-line such that the number of traveling wave families near the shear flow changes suddenly from finite to infinite when passes through it. On the other hand, if has no such critical points, then the number is always finite for positive (resp. negative) values. This is true for general shear flows under mildly technical assumptions, and for a large class of shear flows including a cosine jet (i.e. the sinus profile) and analytic monotone flows unconditionally. The sudden change of the number of traveling wave families indicates that long time dynamics around the shear flow is much richer than the non-rotating case, where no such traveling wave families exist.
1. Introduction
The earth’s rotation influences dynamics of large-scale flows significantly. Under the -plane approximation, the motion for such a flow could be described by 2-D incompressible Euler equation with rotation
(1.1) |
where is the fluid velocity, is the pressure, is the rotation matrix, and is the Rossby number. Here we study the fluid in a periodic finite channel , with non-permeable boundary condition on :
(1.2) |
The -plane approximation is commonly used for large-scale motions in geophysical fluid dynamics [35, 36]. The vorticity form of (1.1) takes
(1.3) |
where . By the incompressible condition, we introduce the stream function such that . Consider the shear flow , which is a steady solution of (1.3). The linearized equation of (1.3) around takes
(1.4) |
which was derived in [42].
In the study of long time dynamics near a shear flow, the most rigid case is the nonlinear inviscid damping (to a shear flow), a kind of asymptotic stability. This means that if the initial velocity is taken close enough to the given shear flow in some function space, then the velocity tends asymptotically to a nearby shear flow in this space. The existence of nearby non-shear steady states or traveling waves means that nonlinear inviscid damping (to a shear flow) is not true, and long time dynamics near the shear flow may be richer and fruitful. To understand the richer long time dynamics near the shear flow in this situation, an important step is to clarify whether the number of curves of nearby traveling waves with traveling speeds converging to different points is infinite. Indeed, if the number is finite, then the velocity might tend asymptotically to some nonlinear superpositions of finite such non-shear states when the initial data is taken close to the flow, and quasi-periodic nearby solutions are expected, which indicates new but not so complicated dynamics. If the number is infinite, then the evolutionary velocity might tend asymptotically to superpositions of infinite such non-shear states, and almost periodic nearby solutions potentially exist, which predicts complicated even chaotic long time behavior near the flow. Similar phenomena were observed numerically in the study of Vlasov-Poisson system, a model describing collisionless plasmas [9, 4, 5, 25]. This model shares many similarities with the 2D incompressible Euler equation. By numerical simulations, it was found that for some initial perturbation near homogeneous states, the asymptotic state toward which the system evolves can be described by a superposition of BGK modes [9]. This offers a hint for further numerical study in the 2D Euler case. It is very challenging to study long time dynamics near a shear flow in a fully analytic way when such non-shear steady states or traveling waves exist. The first step towards this direction is to construct nonlinear superpositions of traveling waves as in the Vlasov-Poisson case.
When there is no Coriolis force, long time dynamics near monotone flows is relatively rigid in strong topology, while it is still highly non-trivial to give a mathematical confirmation. A first step is to understand the linearized equation. Orr [38] observed the linear damping for Couette flow, and Case predicted the decay of velocity for monotone shear flows. Recently, their predictions are confirmed in [29, 48, 49, 44, 14, 19, 20] and are extended to non-monotone flows in [45, 46]. Meanwhile, great progress has been made in the study of nonlinear dynamics near shear flows. Bedrossian and Masmoudi [3] proved nonlinear inviscid damping near Couette flow for the initial perturbation in some Gevrey space on . Ionescu and Jia [17] extended the above asymptotic stability to a periodic finite channel under the assumption that the initial vorticity perturbation is compacted supported in the interior of the channel. Later, nonlinear inviscid damping was proved near a class of Gevrey smooth monotone shear flows in a periodic finite channel if the perturbation is taken in a suitable Gevrey space, where is compactly supported [18, 34]. It is still challenging to study the long time behavior near general, rough, monotone or non-monotone shear flows. On the other hand, inviscid damping (to a shear flow) depends on the regularity of the perturbation, and the existence of non-shear stationary structures is shown near some specific flows. Lin and Zeng [29] found cats’ eyes flows near Couette for vorticity perturbation in a periodic finite channel, while no non-shear traveling waves near Couette exist if the regularity is , in contrast to the linear level, where damping is always true for any initial vorticity in . For Kolmogorov flows, which is non-monotone, Coti Zelati, Elgindi and Widmayer [8] constructed non-shear stationary states near Kolmogorov at analytic regularity on the square torus, while there are no nearby non-shear steady states at regularity for velocity on a rectangular torus. They also proved that any traveling wave near Poiseuille must be shear for vorticity perturbation in a periodic finite channel.
As indicated in [36], the study of the dynamics of large-scale oceanic or atmospheric motions must include the Coriolis force to be geophysically relevant, and once the Coriolis force is included a host of subtle and fascinating dynamical phenomena are possible. By numerical computation, Kuo [24] found the boundary of barotropic instability for the sinus profile, which is far from linear instability in no Coriolis case. Later, based on Hamiltonian index theory and spectral analysis, Lin, Yang and Zhu theoretically confirmed large parts of the boundary and corrected the rest. New traveling waves, which are purely due to the Coriolis effects, are found near the sinus profile [28]. Barotropic instability of other geophysical shear flows has also attracted much attention. For instance, by looking for the neutral solutions, most of the stability boundary, which is again different from no Coriolis case, of bounded and unbounded Bickley jet is found numerically and analytically in [30, 16, 32, 2, 12]. More fruitful geophysical fluid dynamics, such as Rossby wave and baroclinic instability, could be found in [24, 36, 10, 31, 35]. On the other hand, similar to no Coriolis case, linear inviscid damping is still true for a large class of flows and moreover, the same decay estimates of the velocity can be obtained for a class of monotone flows [47]. Elgindi and Widmayer [11] viewed Coriolis effect as one mechanism helping to stabilize the motion of an ideal fluid, and proved the almost global stability of the zero solution for the -plane equation. Global stability of the zero solution is further to be confirmed in [37].
When Coriolis force is involved, long time dynamics near a shear flow becomes fruitful. One of the main reasons is that, comparing with no Coriolis case, there are new traveling waves with fluid trajectories moving in one direction. This paper is devoted to studying the number of such traveling wave families near a general shear flow under the influence of Coriolis force. Here, a traveling wave family roughly includes the sets of nearby traveling waves with traveling speeds converging to a same number outside the range of the flow, see Definition 2.8 for details. Precisely, we prove that if the flow has a critical point at which attains its minimal (resp. maximal) value, then a unique transitional value (resp. ) exists in the positive (resp. negative) half-line, through which the number of traveling wave families changes suddenly from finite to infinite. The transitional values are defined in (1.11)–(1.12). If the flow has no such critical points, then the number of traveling wave families is always finite for positive (resp. negative) values. This is true for general shear flows under mildly technical assumptions. Based on Hamiltonian structure and index theory, we unconditionally prove the above results for a flow in class , which is defined as follows.
Definition 1.1.
A flow in class means that is not a constant function, and for any , there exists such that is positive and bounded on .
A typical example of such a flow is a cosine jet (i.e. the sinus profile), which was studied in geophysical literature [23, 24, 36]. For and a general shear flow , Rayleigh [39] gave a necessary condition for spectral instability that for some , and even under this condition, Fjørtoft [13] provided a sufficient condition for spectral stability that on . For and , the above two conditions can be extended as for some and on , respectively. See, for example, (6.3)–(6.4) in [24]. For a flow in class , the extended Rayleigh’s condition implies that is necessary for spectral instability, but the flow does not satisfy the extended Fjørtoft’s sufficient condition for spectral stability. The sharp condition for spectral stability indeed depends on and the wave number , which was obtained in [27] for and in [28] for .
Consider a class of general shear flows satisfying
A flow in class satisfies the assumption (H1). In fact, it is trivial for ; if and satisfies and , then . Thus, solves , . Then , which is a contradiction.
To state our main results with few restriction, we first consider flows in class , and left the extension to general shear flows satisfying (H1) in Section 2.
Theorem 1.2.
Let and the flow be in class .
-
(1)
If , then there exists such that there exist at most finitely many traveling wave families near for , and infinitely many traveling wave families near for . Moreover, is specified in (1.11).
-
(2)
If , then there exists such that there exist at most finitely many traveling wave families near for , and infinitely many traveling wave families near for . Moreover, is specified in (1.12).
-
(3)
If , then there exist at most finitely many traveling wave families near for .
-
(4)
If , then there exist at most finitely many traveling wave families near for .
Here, the precise description of a traveling wave family near is given in Definition 2.8.
Unless otherwise specified, “near ” always means “in a (velocity) neighborhood of ” in Theorem 1.2 and the rest of this paper, as indicated in Definition 2.4. These traveling wave families do not exist if there is no Coriolis force. By Theorem 1.2, Coriolis force and its magnitude indeed bring fascinating dynamics near the shear flow. On the one hand, for flows having no critical point which is meanwhile a minimal point, the number of traveling wave families is always finite no matter how much magnitude of Coriolis force, which is a mild Coriolis effect. On the other hand, for flows having such a critical point, there is a surprisingly sharp difference, namely, when the Coriolis parameter passes through the transitional point , the number of traveling wave families changes suddenly from finite to infinite. In particular, quasi-periodic solutions to (1.1)–(1.2) can be expected near the shear flow for , while almost periodic solutions potentially exist for . This could be regarded as a strong Coriolis effect and predicts chaotic long time dynamics near these flows.
The same dynamical phenomena are true for general shear flows under some mildly technical assumptions. The explicit result is stated in Theorem 2.2. For , the technical assumption for flows having a critical and meanwhile minimal point is that is not an embedding eigenvalue of the linearized Euler operator for small wave numbers. The assumption for flows having no such critical points is some regularized condition near the endpoints of . Note that the first spectral assumption has only restriction for one point , no matter whether the interior of has embedding eigenvalues. The second assumption is more generic and quite easy to verify. Both the two technical assumptions are used for ruling out eigenvalues’ oscillation for Rayleigh-Kuo boundary value problem (BVP) as the parameter tends to , see Subsection 2.2 for details.
Let us give some remarks on properties of such traveling waves near the flow .
- •
-
•
The traveling waves can be constructed near a smooth shear flow for (including ) velocity perturbation when the Coriolis parameter is large, see Corollary 2.6. In contrast, in the case of no Coriolis force, no traveling waves could be found near Couette flow for velocity perturbation [29] and near Poiseuille flow for velocity perturbation [8].
-
•
Let and . The directions of vertical velocities of the nearby traveling waves might change frequently with small amplitude as the traveling speeds converge to , see Remark 5.2.
We apply the main results to analytic monotone flows (including Couette flow) and the sinus profile. For an analytic monotone flow, there exist at most finitely many nearby traveling wave families for , see Corollary 2.3. For the sinus profile, as mentioned above, it is in class , and so applying Theorem 1.2 (1)-(2) we get that , , , , there exist at most finitely many traveling wave families near the sinus profile for , and infinitely many nearby traveling wave families for . Moreover, we will give a systematical study on the number of isolated real eigenvalues of the linearized Euler operator and traveling wave families near the sinus profile on the whole ’s region in Section 7 (here is the wave number in the -direction), which plays an important role in further study on its long time dynamics. We make a comparison with the previous work in [28]. By Theorem 2.1, the number of isolated real eigenvalues of the linearized Euler operator (i.e. non-resonant modes) determines that of traveling wave families. The explicit number of isolated real eigenvalues in the region can be obtained in [28], but no information can be concluded outside this region, see the discussion below Figure 4 in [28]. Our new contribution for the sinus profile in this paper is that we calculate the explicit number of isolated real eigenvalues in the remaining area , and thus completely get the number of traveling wave families near the sinus profile on the whole ’s region. For the sinus profile, the novelty is that we give the asymptotic behavior of the -th eigenvalue of the Rayleigh-Kuo BVP (2.6) as for and as for , from which we find the transitional values such that the number of traveling wave families changes suddenly from finite to infinite. For general shear flows satisfying , the key is to study whether is unbounded from below as is close to (or ) in Theorem 2.9 and to rule out the oscillation of in Theorems 2.11-2.13. In this paper, we focus on the description of the eigenvalues of the Rayleigh-Kuo BVP (2.6), which in turn, by Theorem 2.1, yields information on traveling wave families.
The rest of this paper is organized as follows. In Section , we extend Theorem 1.2 to general shear flows and give the outline of the proof. In Sections 3-4, we study the asymptotic behavior of the -th eigenvalue of Rayleigh-Kuo BVP, where we determine the transitional values for the -th eigenvalue of Rayleigh-Kuo BVP in Section 3, and rule out oscillation of the -th eigenvalue in Section 4. In Section , we establish the correspondence between a traveling wave family and an isolated real eigenvalue of the linearized Euler operator. In Section , we prove the main Theorems 2.2 and 1.2. As a concrete application, we thoroughly study the number of traveling wave families near the sinus profile in the last section.
Notation
We provide the notations that we use in this paper. Let and for . For a shear flow satisfying , we use the following characteristic quantities of the flow. If we define
(1.5) |
If we define
(1.6) |
Note that and in (1.5)–(1.6). In fact, implies for . Then is an isolated point of . Thus, is a finite set and in (1.5). Similarly, in (1.6). Besides (1.5)–(1.6), we define
(1.7) | ||||
(1.8) |
If we define
(1.9) |
If we define
(1.10) |
Note that and in (1.9)–(1.10). Then we define
(1.11) |
and
(1.12) |
We denote
(1.13) | |||
(1.14) |
where is defined in (2.5). Moreover, we define
(1.17) |
and
(1.20) |
where is the -th eigenvalue of the Rayleigh-Kuo BVP (2.6).
, and denote the set of all the real numbers, integers and positive integers, respectively. or is the cardinality of the set . Let be a linear operator from a Banach space to . is the dual space of . , and are the spectrum, essential spectrum and discrete spectrum of the operator , respectively. For , the Fourier transform of in is denoted by .
2. Extension to general shear flows and outline of the proof
In this section, we first extend the main Theorem 1.2 to general shear flows under mild assumptions, and then discuss our approach in its proof.
2.1. Main results for general shear flows
For a shear flow in , we give the exact number of traveling wave families near the flow.
Theorem 2.1.
Then we state our main theorem for a shear flow satisfying .
Theorem 2.2.
Let and satisfy .
- (1)
- (2)
Assume that and for , there exist and such that for or for or , where .
-
(3)
If , then there exist at most finitely many traveling wave families near for .
-
(4)
If , then there exist at most finitely many traveling wave families near for .
Here, the precise description of a traveling wave family near is given in Definition 2.8.
As mentioned in Introduction, or is “one spectral point” assumption for small wave numbers. Note that if , then , and is not needed in Theorem 2.2 (1)–(2). One of the conditions (i)–(iii) is the “good” endpoints assumption and rather generic. For example, if , and for some , then is true for . Thus, for analytic flows, (ii) holds if for some and (i) holds otherwise. Applying Theorem 2.2 (3)-(4) to analytic monotone flows, we have the following result.
Corollary 2.3.
Let be an analytic monotone flow: for . Then there exist at most finitely many traveling wave families near for .
2.2. Outline and our approach in the proof
Non-parallel steady flows or traveling waves may be bifurcated from a shear flow if the linearized Euler operator has an embedding or isolated real eigenvalues [1, 29, 28]. Based on the existence of an embedding eigenvalue for a class of monotone shear flows near Couette flow, cat’s eyes steady states are bifurcated from these flows [29]. When the Coriolis force is involved, non-parallel traveling waves are bifurcated from the sinus profile on account of the existence of an isolated real eigenvalue [28]. The traveling speeds lie outside the range of the sinus profile and are contiguous to the isolated real eigenvalue. Now, we consider such bifurcation theorem for general shear flows, namely, using an isolated real eigenvalue of the linearized Euler operator, we prove that such traveling waves can be bifurcated from general shear flows. We use the following concept.
Definition 2.4.
Then we give the bifurcation result for general shear flows.
Lemma 2.5.
Let , and . Assume that , where is defined in (2.5). Then there exists a set of traveling wave solutions near with traveling speeds converging to . Moreover, we have .
Here, we mention some differences from the construction of traveling waves in the literature. First, the horizontal period of constructed traveling waves in Proposition 7 of [28] is not the given period , and for the sinus profile, the period of traveling waves is modified to by adjusting the traveling speed in Theorem 7 of [28]. But the price is an additional condition, namely, the isolated eigenvalue can not be an extreme point of (i.e. in Theorem 7 (ii) of [28]), where is the -th eigenvalue of (2.6). In Lemma 2.5, we can construct traveling waves for general flows no matter whether is an extreme point of , and thus improve the result in Theorem 7 of [28] even for the sinus profile. Second, it is possible that , which makes it subtle to guarantee that the bifurcated solutions near the flow is not a shear flow. Thus, the extension of the bifurcation result for the sinus profile in [28] to general shear flows in Lemma 2.5 is still non-trivial, since we have to treat the unsolved case that is an extreme point of for some or . To overcome the difficulty, we carefully modify the flow to a suitable shear flow, which satisfies that is locally monotone near and , and then study the bifurcation at the suitable shear flow. Finally, the minimal horizontal periods of constructed traveling waves are possibly less than if In fact, the Sturm-Liouville operator could indeed have more than one negative eigenvalues (e.g., if , and is close to ), where is defined in (2.6). In this case, we give sufficient condition to guarantee that the minimal period is in Lemma 5.3. In contrast, the minimal period must be in Theorem 5.1 of [26] and Theorem 1 of [29], since the corresponding Sturm-Liouville operator has only one negative eigenvalue.
Since the isolated real eigenvalue lies outside the range of the flow , by a similar proof of Lemma 2.5 we can improve the regularity of traveling waves as follows.
Corollary 2.6.
One naturally asks whether the assumption in Lemma 2.5 is necessary. By studying the asymptotic behavior of traveling speeds and normalized vertical velocities for nearby traveling waves, we confirm that it is true.
Lemma 2.7.
Let , and . Assume that is a set of traveling wave solutions near with traveling speeds converging to . Then where is defined in (2.5). Moreover, if , then there exists such that
(2.2) |
where the operator is defined by
(2.3) |
with periodic boundary condition in and Dirichlet boundary condition in , and .
The limit function in Lemma 2.7 is a superposition of finite normal modes, see Remark 5.1. If in Lemma 2.7, the vertical velocities of the nearby traveling waves have simple asymptotic behavior as seen in (2.2). However, if , then the asymptotic behavior of vertical velocities might be complicated, see Remark 5.2. The proofs of Lemmas 2.5-2.7 are given in Section 5.
By Lemma 2.7, for any set of traveling waves near with traveling speeds converging to , must be an isolated real eigenvalue of the linearized Euler operator (besides and ). By Lemma 2.5, every isolated real eigenvalue is contiguous to the speeds of nearby traveling waves. As the minimal periods of traveling waves in can be less than , there might be two or more sets of traveling wave solutions near with traveling speeds converging to a same isolated real eigenvalue. For example, if , is monotone near for , and for and , then an application to Lemma 2.5 (see Case 1 in its proof) gives two sets of traveling wave solutions, which has minimal periods and respectively, near with traveling speeds converging to . Moreover, traveling wave solutions could be bifurcated from nearby shear flows, which might induce more sets of traveling wave solutions near with traveling speeds converging to . This suggests us to define a traveling wave family near by an equivalence class as follows.
Definition 2.8.
A traveling wave family near is defined by an equivalence class under , where if , , are two sets of traveling wave solutions near with traveling speeds converging to , then and are equivalent, , if
By Lemma 2.7, there exists such that in , where and are given in Definition 2.8. By Lemmas 2.5 and 2.7, we obtain the exact number of traveling wave families near a flow in Theorem 2.1.
Thus, the number of isolated real eigenvalues of the linearized Euler operator plays an important role in counting the traveling wave families near the shear flow. In terms of the stream function , (1.4) can be written as By taking Fourier transform in , we have
(2.4) |
For and , the linearized Euler operator is given by
(2.5) |
Then (2.4) becomes Recall that . Then the set of isolated real eigenvalues . Moreover, it is well-known that if ; if ; and if , see [23, 43, 36, 28]. Therefore, we only need to study for and for . We mainly study for , since the other is similar. if and only if its corresponding eigenfunction satisfies the Rayleigh-Kuo BVP:
(2.6) |
where and . This equation is formulated by Kuo [23]. For , it follows from [40] that the -th eigenvalue of (2.6) is
(2.7) | ||||
In this way, we have
To determine whether is finite, we need to study the number of solutions such that for . Since by Proposition 4.2 in [28] and is real-analytic on , the only possibility such that is that there exists a sequence such that . Thus, the key is to study the asymptotic behavior of as .
We divide it into two steps.
Step 1. We study how many ’s such that as . We determine a transitional value such that the number
changes suddenly from finite to infinite when passes through it.
Theorem 2.9.
The transitional value is illustrated in Figure 1. We give a simple example to explain why such a transitional value exists. Consider the flow on and . If is very close to , then the energy quadratic form in (2.7) roughly looks like
Thus, if , by Hardy type inequality (Lemma 3.1) we have is bounded from below for any test functions with . From this formal observation, we may expect is bounded from below. If , is unbounded from below by looking at the test functions with . We will construct test functions motivated by the function to show that all the eigenvalues are unbounded from below.
![]() ![]() |
Figure 1.
For general flows, the main idea in the proof of Theorem 2.9 (1)-(2) is to control using the norm of near a singular point (see Lemma 3.2), which involves very delicate and careful localized analysis. The transitional values are essentially due to the optimal Hardy type inequality (3.1). The idea in the proof of Theorem 2.9 (3)-(4) is to construct suitable test functions such that the functional in (2.7) converges to as or , see (3.22). This is inspired by the “eigenfunction” for the optimal Hardy type equality and a support-separated technique. The proof of Theorem 2.9 is given in Section 3.
Then we give sharp criteria for as if . By Theorem 2.1, the number of traveling wave families is to count the union of for all . Thus, the number of traveling wave families is infinity provided that as . By Theorem 2.9 (1)-(2), we get the sharp criteria for as .
Corollary 2.10.
Here, a key point for Corollary 2.10 (1) and (3) is that if and only if and
Step 2. We rule out the oscillation of as (or ). By Theorem 2.9 (1), we get that for , has only finite number of solutions on . Moreover, if , no solutions exist for on . Now, we consider whether for . Indeed, we rule out the oscillation of under the spectral assumption , or under the “good” endpoints assumption (i.e. one of the conditions (i)–(iii) in Theorem 2.2), or for flows in class . The oscillation of is illustrated in Figure 2.
![]() |
Figure 2.
Case 1. Under the spectral assumption, the main argument to rule out oscillation is to prove uniform bound for corresponding eigenfunctions, and the proof is in Subsection 4.1. In this case, are also transitional values for the number of isolated real eigenvalues of the linearized Euler operator if .
Theorem 2.11.
In fact, by Theorem 2.9 we have for or . Here, we focus on sufficient conditions of (2.8) and (2.9), it is unclear whether (2.8) is true for the case with , or the case with but no assumption .
Note that Theorem 2.11 (3)-(4) is a direct consequence of Theorem 2.9 (3)-(4).
Case 2.
Under the “good” endpoints assumption (i.e. one of the conditions – in Theorem 2.2), a delicate analysis near the endpoints is involved
to rule out oscillation, and the proof is in Subsection 4.2. In this case, we get that no transitional values exist if .
Theorem 2.12.
Let and satisfy . Assume that , and one of the conditions – in Theorem 2.2 holds. Then
-
(1)
for all if and only if ;
-
(2)
for all if and only if .
Consequently, for all if and only if .
Note that if , then (2.8) and (2.9) are true, and the “good” endpoints assumption (i.e. one of the conditions – in Theorem 2.2) is not needed in Theorem 2.12. Consequently, if , then Theorem 2.2 (3)–(4) hold true without this assumption (see their proof).
Let be an analytic monotone flow and . Then for . This is a corollary of Theorem 2.12, and can also be deduced by the method used in Lemma 3.2 and Theorem 4.1 of [41].
Case 3. For flows in class , the main tools to rule out oscillation are Hamiltonian structure and index formula, and the proof is in Subsection 4.3. This is also the main reason why the spectral and “good” endpoints assumptions can be removed in Theorem 1.2.
Theorem 2.13.
The idea of the proof is as follows. The linearized equation has Hamiltonian structure and the energy quadratic form has finite negative directions. The key observation is that oscillation of brings infinite times of sign-changes of . This contributes infinite negative directions of quadratic form for non-resonant neutral modes, which is a contradiction to the index formula. Thus, the oscillation of can be ruled out unconditionally for flows in class .
3. Transitional values for the -th eigenvalue of Rayleigh-Kuo BVP
We begin to study the asymptotic behavior of the -th eigenvalue of Rayleigh-Kuo BVP. In this section, we focus on the number . We prove that the number is finite for and it is infinite for , which is stated precisely in Theorem 2.9.
3.1. Finite number for
The optimal constant in the following Hardy type inequality plays an important role in discovering the transitional values .
Lemma 3.1.
Let and for some . Then
(3.1) |
Here the constant is optimal.
Proof.
Suppose that is real-valued without loss of generality. Let . First, we consider the integration on (if ). Since
(3.2) |
we have
Here we used and Thus, Similarly, . This gives (3.1). Letting , and sending , we see that the constant is optimal. ∎
For other versions of Hardy type inequality, the readers are referred to [15, 33]. To study the lower bound of the -th eigenvalue of Rayleigh-Kuo BVP for close to , it is important to estimate the energy expression (2.7) near singular points. To this end, we need the following lemma.
Lemma 3.2.
Assume that and . Then there exists a constant (depending only on and ) such that for ,
-
(1)
if or , then
(3.3) -
(2)
if , then
(3.4) Here is a positive constant depending only on and .
Proof.
First, we assume (i). Then we have and thus . Without loss of generality, we assume that . In this case, . Choose small enough such that for , and thus, there exists such that for . Note that for ,
Now we take to be small enough such that Then for ,
which implies (3.3) since .
Now we assume (ii), then . Let be small enough such that for . Since , we have and for . Then there exists such that for and , and thus
Now we take . For , we have
(3.5) | ||||
which implies (3.3). Here we used Lemma 3.1 in the last step.
Next, we assume (iii). In this case, . Let . Then . Let be small enough such that and for . Then for and . Now we assume . Direct computation implies
Since , it follows from the proof of Lemma 3.7 in [47] that . By interpolation, we have , and thus
and . Then
Note that for . Then and for and . Let and . Then , and
for . Thus,
(3.6) |
for , where
Note that
where and . Therefore .
Next, we claim that , , is uniformly bounded in for . The proof is similar as that in Lemma 3.7 of [47]. Note that for and . Therefore,
Now, we are ready to prove Theorem 2.9 (1)-(2).
Proof of Theorem 2.9 (1)-(2).
We first give the proof of (1), and (2) can be proved similarly. Consider . It suffices to show that . Let
(3.8) |
and
(3.11) |
where is a constant such that . Then . Define
where , and is small enough such that for and for all . Then and . Thus, in the sense for . Let Then . By (2.7), there exist , , with such that with , and
(3.12) |
Next, we prove (ii). Let be a sufficiently small constant such that for , and for and . There are four cases for zeros of as follows:
Case 1. and ;
Case 2. , (thus );
Case 3. and ;
Case 4. and .
Then we divide our proof into four cases as above.
In fact, for Cases 1–2, by Lemma 3.2 (1) there exists such that for , and ,
(3.13) |
For Case 3, by Lemma 3.2 (2) there exists such that for , and ,
(3.14) |
Here depends only on . Moreover, if , then by Lemma 3.2 (1),
(3.15) |
For Case 4, we have By Lemma 3.2 (1), there exists such that for , , and ,
(3.16) |
Now let . Define
Then there exists such that for ,
For and ,
Let us first consider . For and , we have
(3.17) |
for . We proceed to consider .
(3.18) | ||||
Recall that are defined in (3.8). For any -dimensional subspace in , there exists such that Define Then , . We normalize such that . Then by (3.13), (3.14), (3.16) and (3.18), we have
This, along with (2.7) and (3.17), yields that This proves (ii).
Finally, we prove (iii). Let and be defined as in (ii). Let be the principal eigenvalue of
Then we have and
(3.19) |
Let . Then we have for . Let
Then is a finite set, and we can write its elements in the increasing order
Then and for . Let
Then we have
3.2. Infinite number for
In this subsection, we prove Theorem 2.9 (3)-(4). The proof is based on construction of suitable test functions such that the energy in (2.7) converges to as or .
Proof of Theorem 2.9 (3)-(4).
We only prove Theorem 2.9 (3), since (4) can be proved similarly. Let . Then there exists such that , and . If , our analysis is completely on for small enough. If , the analysis is only on and the proof is similar as . Now we assume that . Then and there exists such that and
for and small enough. Let and where is defined in (3.11). Define
where , and is large enough such that . Then and , . Thus, in the sense for . Note that for . For , we define
Choose such that . Then . We shall show that for
(3.22) |
Assume that (3.22) is true. Similar to (3.12), there exist with such that
(3.23) |
Now we prove (3.22). Direct computation gives
for and , where and . Then
(3.24) | ||||
Note that for ,
where . Since and for , we get
Then we infer from (3.24) that
(3.25) | ||||
when is large enough. Direct computation gives
(3.26) | ||||
Combining (3.25) and (3.26), we have
as . This proves (3.22). ∎
4. Rule out oscillation of the -th eigenvalue of Rayleigh-Kuo BVP
Let . By Theorem 2.9 (1)-(2), has only finite number of solutions outside the range of for , and no solutions exist for . It is non-trivial to study whether the number of solutions is finite for . Recall that is obtained in Theorem 2.9 such that for , for . The main difficulty is that might oscillate when is close to or . In this section, we rule out the oscillation in the following three cases.
4.1. Rule out oscillation under the spectral assumption
We rule out the oscillation of under the spectral assumption , which is stated in Theorem 2.11 (1)-(2). To this end, we first consider the compactness near a class of singular points.
Lemma 4.1.
Let , , and so that on and . Assume that Let and so that , , in and
holds on . Then in .
Here . The proof of Lemma 4.1 is the same as that of Lemma 3.4 in [47], where we only used the condition rather than the stronger condition:
Since otherwise, we can construct such that and . Recall that all the conditions and conclusions depend only on . Then we prove the uniform bound for the eigenfunctions. More precisely, we have the following result.
Proposition 4.2.
Proof.
Suppose that (4.2) is not true. Up to a subsequence, we can assume that Let on . Then on , and . Thus, in .
Following Definition 3.10 in [47], we call (or ) to be an embedding eigenvalue of if there exists a nontrivial such that for any and ,
Equivalently, (or ) is an embedding eigenvalue of the linearized operator of (1.1) (in velocity form) defined on . In fact, is the corresponding eigenfunction.
We are now in a position to prove Theorem 2.11 (1)-(2).
Proof of Theorem 2.11 (1)-(2).
First, we prove Theorem 2.11 (1). Suppose . Then by Theorem 2.9, there exist and with such that is the -th eigenvalue of (4.1) with the normalized eigenfunction . By the definition of we have , which implies the second statement of Theorem 2.11 (1). To prove the first statement, we now assume that and . By Proposition 4.2, up to a subsequence, there exists such that in . Similar to (3.28) in [47],
for any and . Since in , taking limits in
for any and , we get
If is nontrivial, is an embedding eigenvalue of , which is a contradiction. Therefore, in , which contradicts that , . Thus, . Theorem 2.11 (2) can be proved similarly. ∎
4.2. Rule out oscillation under “good” endpoints assumption
We rule out the oscillation of under the “good” endpoints assumption (i.e. one of the conditions (i)–(iii) in Theorem 2.2). The statement is given in Theorem 2.12. To this end, we need the following two lemmas.
Lemma 4.3.
Let , and . For fixed there exist constants and such that if and , then
(4.3) | ||||
(4.4) | ||||
(4.5) | ||||
(4.6) |
where .
Proof.
Since and we have for Let and for . Let be small enough such that for Then
for fixed and , and thus
(4.7) |
Similarly, for fixed and , we have
(4.8) |
Since and for we have for fixed
(4.9) | ||||
For fixed using (4.7) and the definition of , we have for ,
and
Then by the definition of , we have
(4.10) |
Similarly, for fixed using (4.9) and the definition of , we have for ,
This implies
(4.11) |
Using (4.11) for and (4.10) for , we have
which implies (4.3) by recalling the definition of . Using (4.11) for and (4.10) for , we have
which implies (4.4) by recalling the definition of and .
For fixed using (4.8) and the definition of , we have for ,
and
where we used and Thus,
(4.12) |
Using (4.12) for and (4.10) for , we have
which implies (4.5) by recalling the definition of . Using (4.9) for and the definition of , we have for ,
Then by the definition of , we have
(4.13) |
Using (4.12) for and (4.13), we have
which implies (4.6) by recalling the definition of . ∎
Lemma 4.4.
Proof.
Let and be given in Lemma 4.3. Then on . By Lemma 4.3, (4.3)–(4.6) are still true with replaced by . As and , we have for . Thus, for , we have
(4.15) | ||||
Using (4.3) with replaced by and (4.15) for , we have
(4.16) | ||||
Using (4.4) with replaced by and (4.15) for , we have
(4.17) | ||||
Using (4.6) with replaced by and (4.15) for , we have
(4.18) | ||||
Here, is a constant depending only on (and independent of ). Taking small enough such that in (4.16)–(4.18), we obtain (4.3), (4.4) and (4.6).
We are now in a position to prove Theorem 2.12.
Proof of Theorem 2.12.
We only prove (1), and the proof of (2) is similar. If , then . By Theorem 2.11 (3), for . If , then (i.e. can be achieved only at the endpoints). We assume that . Then and . By taking smaller, we can assume that on . Let , be the solution of
(4.20) |
Note that for , if and only if . Suppose that . Then . Note that if (iii) is true for , then and thus for all . So we divide the discussion into two cases.
Case 1. and (i) holds for .
Case 2. and (ii) holds for ; or .
If Case 1 is true, then on . By (4.20), can be extended to an analytic function in Since has a finite number of zeros in a neighborhood of , which contradicts that .
Now, we assume Case 2 is true. If , define if and (ii) is true, define Then by taking smaller, we can assume that
(4.21) |
As is also true for . Since , we have , here and . If , we have and . If , we have . Thus, in and there exists a constant such that which implies
(4.22) |
(4.23) |
Let and By Rolle’s Theorem, there exists such that and as . For fixed let be large enough such that Then and
(4.24) |
Moreover, satisfies
(4.25) |
where and . Note that is continuous on , and analytic in . Let . Then can be extended to a continuous function on still satisfying (4.20) in Moreover, is increasing and continuous on and for By standard theory of ODE, and is real-valued for . Using this extension, is well-defined and satisfies (4.20) for For fixed and , we have
(4.26) |
since a continuous function is bounded in a compact set. Let be such that and . Then and We claim that the following uniform bounds
(4.27) |
hold for and Assume that the uniform bounds (4.27) are true, which will be verified later. Let Then we get by (4.25) that By (4.27) and using , we have for ,
and thus
which implies Since for and , we have and thus
Since and recall that we have and for Thus, Note that
If , by (4.27) and (4.21), we have for and ,
and thus using and (4.23), we have
Integrating it on and using , we have for ,
and as , we have
Since is continuous and real-valued, and , it does not change sign on . Then for ,
As , for and , we have
Thus,
Since , we have for fixed ,
Thus, by Fatou’s Lemma, we have
By (4.22), we have
for , where we used . Thus,
Now we take . Then is real-valued and for , it satisfies
(4.28) |
Thus, By (4.22), we have for ,
Thus, and by defining . If then there exists such that for . If and then there exists such that for , and for and . Therefore, if or then there exists and such that for , and
which contradicts (4.28). Thus, we must have Then by the proof of Lemma 3 in [28], we have on , which contradicts . This proves (1) for Case 2.
It remains to prove
(4.27).
Let be fixed such that Lemma 4.4 is true. Recall that and
By (4.26) we know that (4.27) is true for . Now we assume that and that (4.3)–(4.6) are used for satisfying (4.14) (i.e. the condition in Lemma 4.4).
The proof of (4.27) for is divided into 7 steps as follows.
Step 1. for and
For by (4.25), (4.3), (4.26) and , we have , and thus for Now, we prove the result by induction. Assume that and for and Then by (4.25), (4.21), (4.23) and , we have for ,
(4.29) | ||||
(4.30) |
By (4.3), (4.26) and (4.30), we have
(4.31) |
which means for Thus, the result in Step 1 is true.
Step 2. for and
Let By Step 1, we know that for and . Thus, (4.29) is still true and for ,
which, along with (4.4) and (4.26), implies that
Then , and thus for
Step 3. for .
If then and By Step 2, we have for . If then by Step 1, we have for . Thus, is always true for . Then by (4.24) and , we have . By (4.5), (4.26) and , we have . Then , and thus for
Step 4. for and
Let By (4.25), (4.21), (4.22), Step 1 and Step 3, we have for ,
Here, we used and . Thus, (4.31) is still true for i.e. for .
Step 5. for .
By Step 1 and Step 4, we have for . Then by (4.24), we have . By (4.6), (4.26) and , we have , which gives and for
Step 6. and for and
Since and , we have by (4.25), (4.21), (4.23), Step 1 and Steps 4-5 that for ,
Then by (4.4) and (4.26), we have
Therefore, and for
Step 7. for .
4.3. Rule out oscillation for flows in class
We rule out the oscillation of for flows in class , which is stated in Theorem 2.13. The proof is based on Hamiltonian structure and index theory.
Proof of Theorem 2.13.
The assumption (H1) is satisfied for a flow in class . By Theorem 2.11, it suffices to prove for and . Similar proof is valid for . First, we consider . Define the non-shear space
Note that as , is equivalent to constant. Thus, implies .
The linearized equation (1.4) has a Hamiltonian structure in the traveling frame :
where Let and on . It follows from Theorem 3 in [28] that
where is the Morse index of , is the sum of algebraic multiplicities of positive eigenvalues of , is the sum of algebraic multiplicities of eigenvalues of in the first and the fourth quadrants and is the total number of non-positive dimensions of restricted to the generalized eigenspaces of nonzero purely imaginary eigenvalues of .
Suppose that . Then it follows from Theorem 2.9 that there exists such that . Let be a solution of with eigenfunction . can be chosen sufficiently close to . Then is a purely imaginary eigenvalue of with eigenfunction . By Theorem 4 in [28],
Note that does not change sign when is sufficiently close to . Then
Hence, . This contradicts that
where . Therefore, .
5. Relations between a traveling wave family and an isolated real eigenvalue
In this section, we establish the correspondence between a traveling wave family near a shear flow and an isolated real eigenvalue of . For a given isolated real eigenvalue , we prove that there exists a set of traveling wave solutions near with traveling speeds converging to , which is stated precisely in Lemma 2.5. We assume .
Proof of Lemma 2.5.
We assume that , and the case for is similar. Since for some , we have and we choose such that . By (1.3), is a solution of (1.1)–(1.2) if and only if solves
(5.1) |
and takes constant values on , where , and . Let be a stream function associated with the shear flow , i.e., . Since for , is increasing on . Let for , and then we can define a function such that
(5.2) |
Moreover,
for . We extend to such that on and is continuous for and . Taking as the bifurcation parameter, we now construct steady solutions near by solving the elliptic equations
(5.3) |
with the boundary conditions that takes constant values on , . Define the perturbation of the stream function by
Define the spaces
and
where . Consider the mapping
Then for . We study the bifurcation near the trivial solution of the equation in , whose solutions give steady flows of (5.1).
For fixed , by linearizing around , we have
where is the restriction of in and is defined in (2.3). Then we divide the discussion of bifurcation near of the equation into three cases. Since , there exists such that , where is the -th eigenvalue of and is defined in (2.6). Let
(5.4) |
Then exists by our assumption and . Now we denote
Case 1. (the transversal crossing condition)
and .
In this case, we have . Let and . Consider the restriction and . Then by the definition of , we have
(5.5) |
where is a real-valued eigenfunction of . Note that
Then by Lemma in [28], we have
where we used that is real-valued. By (5.5), we have and thus, . Then by Theorem in [7], there exist and a nontrivial bifurcating curve of , which intersects the trivial curve at , such that
So the stream functions take the form
Let the velocity Since , we have
(5.6) | ||||
and
(5.7) |
when
is small. Moreover, for some constant large enough. Thus, we can take smaller and such that for , satisfies that , , and .
By (5.7),
in , where .
Case 2.
and
.
In this case, there exist and such that is increasing in , and thus,
(5.8) |
Let be a positive function, be a solution of the regular ODE
(5.9) |
and be such that for . Since and , we have . Let denote the -th eigenvalue of defined by
for and . Then by (5.9) and the fact that is a positive function, we have
where is a normalized eigenfunction of . By the definition of , for . Since , we have . By the continuity of and the small perturbation of , we can take and smaller such that and
(5.10) |
for . By taking smaller and the Implicit Function Theorem, there exists such that and for . By (5.8), we have and for . Then for fixed , there exists such that and . Note that . Let . Then we have
Fix any . Then we can choose and smaller such that for ,
(5.11) |
By (5.10), and , we can apply Case 1 to the shear flow : there exists a traveling wave solution to (1.1)–(1.2) which has period in ,
(5.12) |
In this case, and there exists such that . There exist and such that both and are decreasing in .
Since is linearly independent of , there exists such that
(5.13) |
Let be a solution of (5.9) with , and be such that for . Then by (5.13), we have and As in Case 2, we can take and smaller such that
for . Note that . By the continuity of , and , we can choose and smaller such that
(5.14) |
for .
As and is decreasing in , we can choose such that . Then there exists such that
To prove Corollary 2.6, we only need to modify the spaces and from and to and in the proof of Lemma 2.5. We also use the fact that , and due to the assumption that .
Conversely, for a set of traveling wave solutions near with traveling speeds converging to , we show that is an isolated real eigenvalue besides and , which is given in Lemma 2.7.
Proof of Lemma 2.7.
It suffices to show that if , then and (2.2) holds. Note that solves
(5.15) |
Moreover,
By taking smaller,
(5.16) |
for and . Note that . By Sobolev embedding, we have
(5.17) | ||||
(5.18) |
Since , we get by (5.15) that
(5.19) |
where . By (5.16), we have for and
where we used (5.17)–(5.18) and . Since for , we have Thus, and
(5.20) | ||||
(5.21) |
Up to a subsequence, there exists such that in , in and . Taking derivative in (5.19) with respect to and , we get by (5.18) and (5.20)–(5.21) that
and
which implies that and thus, in . For any with periodic boundary condition in and Dirichlet boundary condition in , we have
(5.22) |
Since , we have by (5.16) and (5.18) that
Noting that in and sending in (5.22), we have
Thus, is a weak solution of
(5.23) |
Since , we have for . Then by elliptic regularity theory, we have is a classical solution of (5.23). Thus, . Since has no nontrivial solutions satisfying the boundary conditions, we have . Since solves (5.23), there exists such that solves
with . Now we show that . Let for . Then is a bounded linear operator on . Since , we have . Taking limit as , we have and thus, , which implies that . Thus, , where we used the fact that ∎
We give two remarks to Lemma 2.7: the first is to study the Fourier expansion of the limit function , and the second is to show that the asymptotic behavior of normalized vertical velocities might be complicated if .
Remark 5.1.
The function in Lemma 2.7 is a superposition of finite normal modes. In fact, since and , we have . Let . Then and is an eigenfunction of .
Remark 5.2.
Consider a flow satisfying , , and . By Theorem 2.9 , there exists such that , and for . By Lemma 2.5, we can choose and nearby traveling wave solutions with period in such that , and there exists such that is small enough for large , where . Then . A possible case is that there exists a subsequence such that for . In this case, since it is odd in (see the construction in Lemma 2.5), where is a normalized eigenfunction of . Since has sign-changed zeros in , oscillates frequently in the -direction for large .
The minimal period of any nearby traveling wave solution in can be determined under the following condition.
Lemma 5.3.
Proof.
Let the minimal horizontal period of the traveling wave solution be for , where and . Fix and let . Since for and , we have . Thus, if , then and for .
Remark 5.4.
Let and be defined in (5.4). It follows from Lemma 5.3 that under the assumption of Lemma 2.7, if has period in for , then the period is minimal for small enough.
In Proposition of [28], it should be corrected that the minimal period of constructed traveling wave solutions in might be less than , since it is possible that is a non-resonant neutral mode for some and . Consequently, the minimal period of constructed traveling wave solutions near the sinus profile in Theorem of [28] might be less than , see Example 7.1 for systematic study of traveling wave families near the sinus profile. If (5.24) holds true for , then the minimal period of these traveling wave solutions in Proposition and Theorem of [28] is .
6. The number of traveling wave families near a shear flow
In this section, we prove the main theorems-Theorems 2.2 and 1.2. The proof is based on the study on the number of isolated real eigenvalues of the linearized Euler operator in Sections 3-4, and correspondence between a traveling wave family near the shear flow and an isolated real eigenvalue in Section 5. We only prove Theorem 2.2, since the other is similar.
Proof of Theorem 2.2..
Let the number of traveling wave families near be denoted by . By Theorem 2.1,
. Here .
Proof of (1): Since , we have . First, let and we divide the discussion into two cases.
Case 1a. .
By Corollary 2.10 (1),
(6.1) |
Thus, there exists such that
(6.2) |
By Theorem 2.11 (1), we have
(6.3) |
Case 1b. .
By Corollary 2.10 (1) for and Theorem 2.9 (3) for , we have
(6.4) |
Thus, there exists such that for every , and . Then .
Next, let and we separate the proof into two cases.
Case 2a. .
By Corollary 2.10 (3), we obtain (6.1).
Thus, there exists such that (6.2) holds.
By Theorem 2.11 (1), we obtain
(6.3).
Case 2b. .
By Theorem 2.9 (3), we have
(6.5) |
Using (6.5) for and the fact that , it can be proved that by a similar way as in Case 1b. This completes the proof of (1). The proof of (2) is similar.
Proof of (3): Since , we have and .
Fix .
By Corollary 2.10 (3), we obtain (6.1), and thus,
there exists such that (6.2) holds.
By Theorem 2.12 (1), we obtain
(6.3).
This completes the proof of (3). The proof of (4) is similar.
∎
Remark 6.1.
In Cases and of the above proof, the infinitely many traveling wave families are produced by the asymptotic behavior of the first eigenvalue of , see (6.4). There could be many other traveling wave families in general, which are produced by the asymptotic behavior of for , see (6.5). In fact, if , for fixed , there exists such that for every . If , then a simple application of Theorem 2.1 yields other traveling wave families.
7. Application to the sinus profile
In this section, we apply our main results to the sinus profile. Moreover, we calculate the explicit number of isolated real eigenvalues of and traveling wave families near sinus profile.
Example 7.1.
The sinus profile is . We determine and the number of traveling wave families for the sinus profile on the ’s region. For the sinus profile, we have , , , , and . We divide the plane into nine parts as follows.
Figure 3.
The explicit number is given as follows:
(7.1) | ||||
In addition, if . Now we fix .
The number (denoted by ) of traveling wave families near the sinus profile is given by
(7.2) | ||||
In addition, if . Moreover,
(7.3) |
If , then
(7.4) |
for small enough, where satisfies the assumption of Lemma 2.7.
for ;
and is decreasing on for ;
, , and is decreasing on for ;
, , on , there exist such that for , is decreasing on , and is increasing on for ;
, , on , there exists such that for and is decreasing on for ;
on for ;
, , on , there exist such that for , is decreasing on , and is increasing on for ;
, , and is increasing on for ;
, , , and is increasing on for ;
and is increasing on for ,
where , for ,
for , and for by Proposition in [28].
Assertions – provide pictures of the negative eigenvalues of for fixed . Assume that – are true. Note that if and , and if and . Then we claim that
(7.5) | ||||
(7.6) |
for , and . In fact, (7.6) implies (7.5). If , then by – we have on , which gives (7.6). If with , then by we have on and thus, has at most one negative eigenvalue for , which gives (7.6). Similarly, we can prove (7.6) for with by –. If with , then by we have on , which gives (7.6).
By applying Theorem 2.1, we get (7.1)–(7.3). (7.4) is a direct consequence of Lemma 5.3. Here, (7.5) is used in the proof of (7.3)–(7.4).
Using (7.3) we can evaluate for as follows.
(7.7) | ||||
The case is more complicated. By , we have with . Then by and the expression of , , we have , , and , i.e. . In fact, , but it seems difficult to give an explicit formula if
Now, we prove –. and – are a summary of spectral results in Section of [28]. Monotonicity of for is due to Corollary in [28]. Asymptotic behavior of in and is obtained by Corollary 2.10. Signatures of in and – are due to Proposition in [28] and simple computation.
The rest is to prove the asymptotic behavior of in and –. First, we consider . We only prove that for and . Note that and for . Let . Then
as . Similarly, . Clearly, . Then . It follows from Theorem in [22] that for .
We then consider . We use the eigenfunctions of in Proposition of [28]. Here, we rewrite to indicate its dependence on if necessary. There exist , , satisfying
Here, , and is a polynomial with order . Moreover, is real-valued, and we normalize it such that Then we have for ,
Note that has zeros in , and we denote . For any -dimensional subspace in , there exists such that Define Then , , i.e. We normalize such that . Since , we have . Similar to (3.2), we have as or or , where . Integration by parts gives
If , then using and for , we have
This, along with (2.7), yields that Now, we consider the upper bound. Let Then . By (2.7), there exist , , with such that with , and
Combining the upper and lower bounds, we have for
Now, we consider . By Corollary in [28], we have for fixed , if . As if (see Corollary in [28]) and , we have for fixed , if . Then
Here, we used the left continuity of at Thus,
Next, we consider . By Proposition in [28], there exists if is odd; there exists if is even; and for ,
Here, and . Moreover, is real-valued and we normalize it such that Then for ,
If is odd, then has zeros in , and . For any -dimensional subspace in , there exists such that Define Then , , i.e. We normalize such that . Since , we have . Integration by parts gives
If , then using and for , we have
This, along with (2.7), yields that If is even, then . Thus, is always true. Now, we consider the upper bound. Let here is defined in (3.11), and are normalized eigenfunctions. Then . By (2.7), for there exists with , and
Since is finite dimensional, there exist and such that in Then and
Since and for and , by Fatou’s Lemma, we have
In particular, if , then
If , then and
As there exist , , with such that and
Therefore, if , then if , this is clearly true since the limit is in this case. By monotonicity of , we have Combining the upper and lower bounds, we have for .
For , the limits , , can be proved similarly as in the case Finally, the limit for follows from Theorem 2.9 .
Acknowledgement
Z. Lin is partially supported by the NSF grants DMS-1715201 and DMS-2007457. Z. Zhang is partially supported by NSF of China under Grant 11425103.
References
- [1] A. Barcilon, P. G. Drazin, Nonlinear waves of vorticity, Stud. Appl. Math., 106 (2001), 437-479.
- [2] N. J. Balmforth, C. Piccolo, The onset of meandering in a barotropic jet, J. Fluid Mech., 449 (2001), 85-114.
- [3] J. Bedrossian, N. Masmoudi, Inviscid damping and the asymptotic stability of planar shear flows in the 2D Euler equations, Publ. Math. Inst. Hautes Études Sci., 122 (2015), 195-300.
- [4] M. Buchanan, J. J. Dorning, Superposition of nonlinear plasma waves, Phys. Rev. Lett., 70 (1993), 3732-3735.
- [5] M. Buchanan, J. J. Dorning, Near equilibrium multiple-wave plasma states, Phys. Rev. E, 50 (1994), 1465-1478.
- [6] K.M. Case, Stability of inviscid plane Couette flow, Phys. Fluids 3, (1960), 143-148.
- [7] M. Crandall, P. Rabinowitz, Bifurcation from simple eigenvalues, J. Funct. Anal., 8 (1971), 321-340.
- [8] M. Coti Zelati, T. M. Elgindi, K. Widmayer, Stationary Structures near the Kolmogorov and Poiseuille Flows in the 2d Euler Equations, Preprint, arXiv:2007.11547.
- [9] L. Demeio, P. F. Zweifel, Numerical simulations of perturbed Vlasov equilibria, Phys. Fluids B, 2 (1990), 1252-1255.
- [10] P. G. Drazin, Introduction to hydrodynamic stability, Cambridge Texts in Applied Mathematics, Cambridge University Press, Cambridge, 2002.
- [11] T. M. Elgindi, K. Widmayer, Long time stability for solutions of a -plane equation, Comm. Pure Appl. Math., 70 (2017), 1425-1471.
- [12] L. Engevik, A note on the barotropic instability of the Bickley jet, J. Fluid Mech., 499 (2004), 315-326.
- [13] R. Fjørtoft, Application of integral theorems in deriving criteria of stability of laminar flow and for baroclinic circular vortex, Geofys. Publ. Norske Vid.-Akad. Oslo, 17 (1950), 1-52.
- [14] E. Grenier, T.T. Nguyen, F. Rousset and A. Soffer, Linear inviscid damping and enhanced viscous dissipation of shear flows by using the conjugate operator method, J. Funct. Anal., 278 (2020), 108339.
- [15] G. H. Hardy, J. E. Littlewood, G. Pólya, Inequalities, Cambridge Mathematical Library, Cambridge University Press,Cambridge (1988); reprint of the 1952 edition.
- [16] L. N. Howard, P. G. Drazin, On instability of a parallel flow of inviscid fluid in a rotating system with variable Coriolis parameter, J. Math. Phys., 43 (1964), 83-99.
- [17] A. Ionescu, H. Jia, Inviscid damping near the Couette flow in a channel, Comm. Math. Phys., 374 (2020), 2015-2096.
- [18] A. Ionescu, H. Jia, Nonlinear inviscid damping near monotonic shear flows, Preprint, arXiv:2001.03087.
- [19] H. Jia, Linear inviscid damping in Gevrey spaces, Arch. Ration. Mech. Anal., 235 (2020), 1327-1355.
- [20] H. Jia, Linear inviscid damping near monotone shear flows, SIAM J. Math. Anal., 52 (2020), 623-652.
- [21] T. Kato, Perturbation theory for linear operators, second edition, Springer-Verlag, Heidelberg, 1980.
- [22] Q. Kong, H. Wu, A. Zettl, Dependence of the -th Sturm-Liouville eigenvalue on the problem, J. Differential Equations, 156 (1999), 328-354.
- [23] H. L. Kuo, Dynamic instability of two-dimensional non-divergent flow in a barotropic atmosphere, J. Meteor., 6 (1949), 105-122.
- [24] H. L. Kuo, Dynamics of quasi-geostrophic flows and instability theory, Adv. Appl. Mech., 13 (1974), 247-330.
- [25] C. Lancellotti, J. J. Dorning, Nonlinear Landau damping, Transport Theor. Stat., 38 (2009), 1-146.
- [26] C. Y. Li, Z. Lin, A resolution of the Sommerfeld Paradox, SIAM J. Math. Anal., 43 (2011), 1923-1954.
- [27] Z. Lin, Instability of some ideal plane flows, SIAM J. Math. Anal., 35 (2003), 318-356.
- [28] Z. Lin, J. Yang, H. Zhu, Barotropic instability of shear flows, Stud. Appl. Math., 144 (2020), 289-326.
- [29] Z. Lin, C. Zeng, Inviscid dynamic structures near Couette flow, Arch. Rat. Mech. Anal., 200 (2011), 1075-1097.
- [30] F. B. Lipps, The barotropic stability of the mean winds in the atmosphere, J. Fluid Mech., 12 (1962), 397-407.
- [31] A. Majda, Introduction to PDEs and waves for the atmosphere and ocean, Courant Lecture Notes in Mathematics, vol. 9, New York University, Courant Institute of Mathematical Sciences, New York; American Mathematical Society, Providence, RI, 2003.
- [32] S. A. Maslowe, Barotropic instability of the Bickley jet, J. Fluid Mech., 229 (1991), 417-426.
- [33] N. Masmoudi, About the Hardy inequality. In: Schleicher D., Lackmann M. (eds) An Invitation to Mathematics. Springer, Berlin, Heidelberg.
- [34] N. Masmoudi, W. Zhao, Nonlinear inviscid damping for a class of monotone shear flows in finite channel, Preprint, arXiv:2001.08564.
- [35] J. C. McWilliams, Fundamentals of geophysical fluid dynamics, Cambridge University Press, 2006.
- [36] J. Pedlosky, Geophysical fluid dynamics, 2nd edn. Springer, New York (1987).
- [37] F. Pusateri, K. Widmayer, On the global stability of a -plane equation, Anal. PDE, 11 (2018), 1587-1624.
- [38] W. Orr, Stability and instability of steady motions of a perfect liquid, Proc. Ir. Acad. Sect. A: Math Astron. Phys. Sci., 27(1907), 9-66.
- [39] L. Rayleigh, On the stability or instability of certain fluid motions, Proc. London Math. Soc., 9 (1880), 57-70.
- [40] M. Reed, B. Simon, Methods of modern mathematical physics. IV. Analysis of operators. Academic Press [Harcourt Brace Jovanovich, Publishers], New York-London, 1978. xv+396 pp.
- [41] S. I. Rosencrans, D. H. Sattinger, On the spectrum of an operator occuring in the theory of hydrodynamic stability, J. Math. Phys., 45 (1966), 289-300.
- [42] C. G. Rossby, Relation between variations in the intensity of the zonal circulation of the atmosphere and the displacements of the semi-permanent centers of action, J. Mar. Res., 2 (1939), 38-55.
- [43] K. K. Tung, Barotropic instability of zonal flows, J. Atmos. Sci., 38 (1981), 308-321.
- [44] D. Wei, Z. Zhang and W. Zhao, Linear inviscid damping for a class of monotone shear flow in Sobolev spaces, Commun. Pure Appl. Math., 71 (2018), 617-687.
- [45] D. Wei, Z. Zhang and W. Zhao, Linear inviscid damping and vorticity depletion for shear flows, Ann. PDE, 5 (2019), 3, 101 pp.
- [46] D. Wei, Z. Zhang and W. Zhao, Linear inviscid damping and enhanced dissipation for the Kolmogorov flow, Adv. Math., 362 (2020), 106963, 103 pp.
- [47] D. Wei, Z. Zhang, H. Zhu, Linear inviscid damping for the -plane equation, Comm. Math. Phys., 375 (2020), 127-174.
- [48] C. Zillinger, Linear inviscid damping for monotone shear flows, Trans. Am. Math. Soc., 369 (2017), 8799-8855.
- [49] C. Zillinger, Linear inviscid damping for monotone shear flows in a finite periodic channel, boundary effects, blow-up and critical Sobolev regularity, Arch. Ration. Mech. Anal., 221 (2016), 1449-1509.