On the spectra of the gravity water waves linearized at monotone shear flows
Abstract.
We consider the spectra of the 2-dim gravity waves of finite depth linearized at a uniform monotonic shear flow , , where the wave numbers of the horizontal variable is treated as a parameter. Our main results include a.) a complete branch of non-singular neutral modes strictly decreasing in and converging to as ; b.) another branch of non-singular neutral modes , for some , with ; c.) the non-degeneracy and the bifurcation at ; d.) the existence and non-existence of unstable modes for near , , and interior inflection values of ; e.) the complete spectral distribution in the case where does not change sign or changes sign exactly once non-degenerately. In particular, is spectrally stable if and unstable if has a non-degenerate interior inflection value or accumulate at or . Moreover, if is an unstable shear flow of the fixed boundary problem in a channel, then strong gravity could cause instability of the linearized gravity waves in all long waves (i.e. ).
1. Introduction
Consider the two dimensional gravity water waves in the moving domain of finite depth
or
The free surface is given by . For , let denote the fluid velocity and the pressure. They satisfy the free boundary problem of the incompressible Euler equation:
(1.1a) | ||||
(1.1b) | ||||
(1.1c) | ||||
(1.1d) | ||||
(1.1e) |
where is the gravitational acceleration and the constant density is normalized to be 1. The kinematic boundary condition (1.1c) is equivalent to that is tangent to at any , which means that moves at the velocity restricted to .
Shear flows are a fundamental class of stationary solutions of laminar flows
(1.2) |
Our goal in this paper is to analyze the spectral distribution (and thus the spectral stability and the linear instability) of the gravity water wave system linearized at a monotonic shear flow satisfying
(H) |
Remark 1.1.
Due to the symmetry of horizontal reflection
the case of is completely identical except the sign assumptions on and in the theorems should be reversed.
1.1. Linearization and the spectral problem
We first derive the linearized system of (1.1) at the shear flow given in (1.2) satisfied and we denote the linearized solutions by . Let be a one-parameter family of solutions of (1.1) with . Following the same procedure as in [17], differentiating the Euler equation system (1.1) with respect to and then evaluating it at yield
(1.3a) | |||
(1.3b) | |||
(1.3c) | |||
(1.3d) |
Observe that the variable coefficients in the linear system (1.3) depend only on and thus the Fourier modes of evolve in independently of each others. The zeroth Fourier mode corresponds to the perturbed shear flow component and can be handled easily, so the main interest lies in the -th Fourier modes with . To find eigenvalues and eigenfunctions, we consider linearized solutions in the form of
(1.4) |
where apparently the eigenvalues take the form with the wave speed . In seeking solutions in the form of (1.4), the wave number is often treated as a parameter. Substituting (1.4) into the linearization (1.3), through straight forward calculations (see [28, 10] as well as [17]), one obtains the boundary value problem of the standard Rayleigh equation
(1.5a) | |||
with the boundary conditions | |||
(1.5b) | |||
(1.5c) |
Here and throughout the paper ′ denotes the derivative with respect to . Apparently the last boundary condition at differs it from that of the linearized channel flow for with the slip boundary conditions at . It actually makes a substantial difference as we shall see below. For , and can be determined by using the divergence free and the kinematic condition, respectively,
So we shall mainly focus on .
The linear system (1.3) is linearly unstable if there exist solutions to (1.5) with , which appear in conjugate pairs. Obviously the Rayleigh equation becomes singular when , otherwise its initial value problem depends on and analytically and thus also even in . We recall the following standard terminology.
Definition 1.1.
is a non-singular mode if and there exists a non-trivial solution to (1.5) (thus also yields a solution to (1.3) in the form of (1.4)). is a singular mode if and
has a non-trivial solution satisfying the boundary conditions (1.5b) and (1.5c). A non-singular mode is a stable (or unstable) mode if (or ). is a neutral mode if it is a singular or non-singular mode and .
We first present our main results and then comment on them along with the literatures.
1.2. Main results
In this paper we focus on the spectral distribution of the gravity waves linearized at a uniformly monotonic shear flow. It serves as the first step for us to understand the linearized flow at monotonic shear flows including the possible linear inviscid damping and the dispersive character (see [17] for capillary gravity waves). This would lay the foundation for the study of the local nonlinear dynamics. The following is a non-technical summary (assuming ) of the main results mostly given in the following three main theorems.
-
•
There is a complete upper branch of neutral modes decreasing in and converging to in the same fashion as the dispersion relation of the free gravity wave as .
-
•
Possibly two more branches of eigen-modes for and , respectively.
-
–
There is a branch of neutral modes increasing in and reaches at some and then, near and for , it either bifurcates into a (possibly broken) branch of unstable modes whose real parts are given by near or the branch disappears completely if near .
-
–
Unless near , there is a (possibly broken) branch of "extremely weakly" unstable modes converging to as , whose real parts are given by near .
-
–
-
•
Inflection values of in always yield one or two singular neutral modes. Bifurcation to unstable modes occurs with on one side of any non-degenerate interior inflection value.
-
•
is necessary and also "almost" sufficient for linear instability.
-
•
If on , then connects to and, along with and , give all the eigen-modes (singular and non-singular modes) and (1.3) is spectrally unstable for all .
-
•
If on , then (1.3) is spectrally stable and for , and , are all the eigen-modes besides those at interior inflection values.
-
•
If changes from negative to positive exactly once non-degenerately, then connects to the inflection value at , and, along with and , give all the eigen-modes. Hence (1.3) is spectrally unstable for all .
-
•
If changes from positive to negative instead, the eigenvalue distribution is also obtained. In particular, in some cases (1.3) with sufficiently strong gravity is spectrally unstable for all long waves (i.e. ).
- •
The following first main theorem is on the above branches and , the singular neutral modes, as well as the bifurcations near singular neutral modes.
Theorem 1.1.
Assume and on , then singular neutral modes occur only if or is an inflection value of . Moreover the following hold for a constant determined only by .
-
(1)
There exists such that there exist an analytic function and a function both defined for and even in and the following hold for any .
-
(a)
and have the same signs (, or ) and
-
(b)
is a non-singular or singular mode iff i.) , ii.) and , or iii.) and .
-
(c)
can be extended to an even and analytic function of with and for . Moreover, for any , is the only non-singular mode in and it corresponds to an eigenvalue of (1.3).
-
(a)
-
(2)
There exist a unique and a (for any ) even function defined for such that the following hold.
-
(a)
is analytic in , , and for all .
-
(b)
is a non-singular mode with outside the disk (1.8) iff and .
-
(c)
There exist such that can be extended to as a complex valued even function satisfying the following.
-
(i)
For , , and have the same signs (, or ), and
-
(ii)
is a singular or non-singular mode of (1.3) with , , and iff ( iff ).
-
(i)
-
(a)
-
(3)
Suppose with and , then the following hold.
-
(a)
There exists such that or , , and is a singular neutral mode of (1.3) iff .
-
(b)
If, in addition, , then, for any , there exist and a complex valued function defined on satisfying the following.
-
(i)
has the same sign as and
-
(ii)
is a singular or non-singular mode of (1.3) with , , and iff and ( iff and ).
-
(i)
-
(a)
-
(4)
The linearized gravity water wave system (1.3) is linearly unstable for some wave number if has a non-degenerate zero point in or there exists a sequence converging to or as such that for all .
Clearly the above , , and are relevant only when their imaginary parts are non-negative, so the subsets of these branches corresponding to eigenvalues are possibly broken. They correspond to unstable modes iff the imaginary parts are positive. From the estimates of the imaginary parts, the strength of the instability (the exponential growth rates) is the strongest near non-degenerate inflection values, and the weakest near .
Remark 1.2.
a.) Due to the symmetry, the case of (and also in the following theorems) is completely identical except the signs of and should be reversed.
b.) More detailed asymptotics of can be found in Lemma 3.15.
c.) See Lemma 3.7 for more details on when or in statement (3) and its relationship to the singular neutral mode of the linearized channel flow. The statement had also been proved in [28, 11] and we gave a different proof here. It actually holds for a wider class of shear flows not necessarily monotonic, see Remark 3.2. When , see Lemma 3.12 for how a closed branch of unstable modes emerges from the smaller wave number and makes the linearized system unstable for .
d.) While is the most easily checked non-degeneracy condition in statement (3b), a more precise condition and more details of the bifurcation (including the case of ) can be found in Lemma 3.11.
e.) Local bifurcation in statement (4) from an interior inflection value, similar to [17], can be compared to that in [10] for a different class of shear flows. Also see comment below.
In the next theorem we consider the case where does not change sign.
Theorem 1.2.
Assume and on , then the following hold.
- (1)
- (2)
Remark 1.3.
Finally, the last theorem is for the case where has exactly one non-degenerate inflection point in .
Theorem 1.3.
Assume
(1.6) |
then the following hold.
- (1)
- (2)
-
(3)
If , , and , then and we have the following.
-
(a)
can be extended evenly for such that and for .
-
(b)
There exists a even complex valued function defined for such that and for all .
-
(c)
All singular and non-singular modes of (1.3) are given by for , and for , and and for .
-
(a)
-
(4)
If , , and , then and (1.3) is unstable iff . Moreover there exist connected components , , of
such that
and either
-
(a)
and , or
-
(b)
and .
-
(a)
Remark 1.4.
a.) An equivalent condition to the existence of is given in Lemma 3.7.
b.) In this theorem, the case of corresponds to the intersection of increasing and the so-called class (see [10]) shear flows.
c.) In particular, (1.3) becomes unstable for all long waves (i.e. ) if .
d.) Assuming and is an unstable shear flow of the channel flow with fixed boundaries. Letting increase, the system could deform from case (2) to (3) and then possibly to (4), see Lemma 3.7. In Lemma 3.12 which holds under a weaker assumption, one can see a closed branch of unstable modes bifurcates from as . As this branch grows, it may or may not intersect the branch in case (2), resulting in the two possibilities in case (4) where the three subsets may even coincide and be equal to .
1.3. Backgrounds and discussions
Due to its strong physical and mathematical relevance there have been extensive studies of the Euler equation linearized at shear currents. In particular the linear instability is often viewed as the first step in understanding the transition of the fluid motion from the laminar flows to turbulent ones. Much of the existing analysis was on the fluid in a fixed channel with slip boundary conditions
(1.7) |
and some of the results have been extended to free boundary problems.
Linearized channel flows. Classical results on the spectra of the Euler equation (1.7) in a channel linearized at a shear flow include:
-
•
Unstable eigenvalues are isolated for each and do not exist for .
- •
-
•
Howard’s Semicircle Theorem [8]: eigenvalues exist only with in the disk
(1.8) - •
Many classical results can be found in books such as [5, 19] etc. For a class of shear flows, the rigorous bifurcation of unstable eigenvalues was proved, e.g., in [7, 15]. In particular, continuation of branches of unstable eigenvalues were obtained by Lin in the latter.
Linearized gravity water waves. It has been extended to the linearized free boundary problem of the gravity waves at certain classes of shear flows (see [4, 9, 1, 28, 25, 18, 10, 13, 24, 2] etc.) that: a.) the semicircle theorem still holds for unstable modes; b.) the bifurcation and continuation of branches of unstable modes starting from limiting singular neutral modes; c.) the stability if and , etc. Compared to channel flows with fixed boundaries, new phenomena of the linearized gravity waves include:
-
•
In addition to inflection values, critical values of where , and may be limiting singular neutral modes.
-
•
may not ensure the spectral stability.
-
•
There are non-singular neutral modes with in both and which are outside the circle (1.8).
-
•
can be a singular neutral mode for certain wave number.
However, the rigorous bifurcation analysis at or critical values of was still missing. This holds the key how Rayleigh’s stability condition may fail in the gravity wave case. Moreover we have not been aware of rigorous studies of the branch of unstable modes near for . These have been addressed in the above Theorem 1.1.
In particular, we would like discuss some of our results (assuming ) in relation to those in [28, 10, 11, 24].
* Singular neutral modes and branches from bifurcations. Firstly we do not only prove that and internal inflection values are always singular neutral modes for some wave numbers as in [28, 10, 11], but also obtain the exact numbers of those wave numbers (see Theorem 1.1 and Subsection 3.3). For an internal inflection value , it turns out that whether it becomes a singular neutral mode at one or two wave numbers depends on i.) whether it is a singular neutral mode for the linearized channel flow and also on ii.) whether is greater than some threshold. Secondly our bifurcation analysis is carried out in whole neighborhoods of singular neutral modes and allows broken branches, instead of in cones of neighborhoods as in [15, 10]. This is more than a technical improvement, as it gives both the existence and the number of nearby unstable modes which is crucial for the index counting used in the analytic continuation argument. In particular, when the above i) holds, strong gravity could creates a closed branch of unstable modes with for some (Lemma 3.12 and Theorem 1.3). This is contrary to the common expectation that gravity stabilizes long waves. Finally, when changes the sign at most once, these ingredients allow us to obtain the complete eigenvalue distribution by identifying how different branches of unstable modes connect to each other. Some of these branches had been observed in numerics [24].
* Comparison to numerics. Three examples of shear flows were computed in [24] which we shall adapt to our notations to avoid unnecessary confusions. In the last two examples, for , iff , and is the only inflection value. Their second example is and or , where clearly . The former is stable in the channel flow case, while the latter is unstable. For , a branch of unstable modes corresponding to in Theorem 1.3(2) was found. For , two branches of unstable modes corresponding to those in either (3) or (4a) of Theorem 1.3 were observed. In their third example, where , and the branch of unstable modes corresponds to that in Theorem 1.3(1).
Comparison with the linearized capillary gravity waves at monotonic shear flows. In this paper we follow a strategy to study the spectra of the linearized gravity waves (1.3) similar to that of the linearized capillary gravity waves in [17]. The analysis is based on a detailed understanding of the Rayleigh equation (1.5a), particularly near its singularity where . Subsequently, we consider the solution to (1.5a) satisfying the normalized initial condition and , which are analytic in and . Such a solution gives rises to an eigenfunction iff the boundary condition (1.5c) is satisfied. This is formulated into an equation where is defined in (2.23) whose regularity is carefully studied. In Subsections 2.1 and 2.2, we outline those basic analysis developed in [17] needed in this paper. To identify the roots of , the key ingredients are a.) the asymptotic properties of as and ; b.) the bifurcations for near ; and c.) the analytic continuation of the roots of . Here is treated as a parameter in both b.) and c.).
As revealed in Theorems 1.1 and 1.2, the spectral distributions are significantly different when the surface tension is not present. Moreover, the analysis in the current paper also substantially improves some results for capillary gravity waves (see Proposition 3.16).
On the one hand, while one might expect that the water waves tend to be more unstable when there is no surface tension, particularly in high wave numbers, the main differences in the spectral distributions at monotonic shear flows include the following.
-
(i)
For , capillary gravity waves are linearly stable, while the gravity waves are unstable if the set accumulates at .
-
(ii)
While and interior inflection values of may or may not be a singular neutral mode of the linearized capillary gravity waves depending on the combined strength of the gravity and surface tension, they are always singular neutral modes of the linearized gravity waves for one or two (for a unique in the case of ).
-
(iii)
As a consequence, is "almost" sufficient for the instability of the linearized gravity waves (Theorem 1.1(4)), which is not always the case for the capillary gravity waves.
-
(iv)
Contrary to the natural expectation, a.) when the surface tension is very weak with the coefficient , it actually creates instability near non-degenerate inflection values in relatively large wave numbers (Proposition 4.11 in [17]), and b.) strong gravity may also cause instability in all small wave numbers.
Related to (i), one is reminded that, for shear flows, the Rayleigh-Taylor sign at any shear flow implies that the nonlinear gravity waves are locally well-posed for initial data near shear flows and there is no unbounded exponential growth rates in high waves numbers. However, when near , the linearization at the shear flow is still unstable for all , but the exponential growth rates of those unstable linear solutions (Theorem 1.1(1a)) are extremely weak.
When the combination of the surface tension and the gravity is sufficiently strong, the linearized capillary gravity wave at a monotonic shear flow is essentially the superposition of a linearized channel flow with fixed boundaries (corresponding to the continuous spectra) and a dispersive system (corresponding to the non-singular modes and the surface motions) with two branches of the dispersion relation given by (see [17]). Without surface tension, the complete branches of the non-singular modes alway reach and also possibly converge fast to . This interaction between the continuous spectra and eigenvalues would make the the linear inviscid damping (see [17]) a very subtle issue except in the simplest case where near and the period in eliminates all the singular neutral modes. We would address the inviscid damping in a future work.
These differences in the spectra could potential result in substantial differences in the nonlinear local dynamics.
On the other hand, the bifurcation analysis of unstable eigenvalues requires certain regularity and non-degeneracy of , particular in the most subtle case near . The regularity and non-degeneracy at in [17] were verified under the additional assumption on . In Subsections 3.1 and 3.4, we prove these regularity and non-degeneracy without any assumptions additional to the monotonicity of . Therefore the work in this paper also substantially improves the bifurcation analysis in the capillary gravity wave case, see Proposition 3.16 in Subsection 3.8.
Compared to [17], in this paper we also thoroughly analyzed the case where changes sign exactly once.
The spectra of the linearized fluid interface problem.
Another related problem is the two phase fluid problem linearized at a shear flow where the slip boundary condition is assumed at . The famous Kelvin Helmholtz instability was first identified in the simplest setting where [26, 5, 2]. In the case where the upper fluid is much lighter than the bottom one (the air-water interface, for example), Miles [20, 21, 22, 12] proposed the critical layer theory to model how the wind (shear flow) in the air above a stationary water generates waves through linear instability due to the resonance between the shear flow in the air and the temporal frequency of the linear irrotational capillary gravity water waves. This was later rigorously proved in [3]. More recently, the spectra of the linearized fluid interface problem at monotonic shear flows had been studied more comprehensively and in more details in [16].
Notations. Throughout the paper, wave numbers are always denoted by "" and is also used. For complex quantities, the subscripts "R" and "I" denote their real and imaginary parts, respectively. Sometimes "Re" and "Im" are also used when the subscripts make the notations too cumbersome.
2. Preliminary analysis
In this section we review some basic results on the homogeneous Rayleigh equation and the equation satisfied by the eigenmodes . Most of these results have already been obtained in [17] and the readers are referred there for more details.
2.1. Results on the homogeneous Rayleigh equation
(2.1) |
where
had been thoroughly analyzed in Section 3 of [17]. In this subsection, we summarize those results which we shall use here. When necessary some modifications will be outlined to address the different needs in this paper. Throughout this subsection we assume
(2.2) |
As pointed out in Remark 1.1, the case of can be reduced to the above one. Hence all results under (2.2) hold for all uniformly monotonic , namely those satisfying on . The solutions to the Rayleigh equation (2.1) are obviously even in and thus will be assumed mostly. Similarly complex conjugate of solutions also solve (2.1) with replaced by , hence we shall consider the case of . Most of the estimates are uniform in with and thus hold in the limiting case as . The dependence on will also be carefully tracked.
Recall . For convenience we extend to be a function defined in a neighborhood of , where
(2.3) |
so that
(2.4) |
In the analysis of the most singular case of near , we let be such that
(2.5) |
Consider the solution to (2.1) satisfying the initial conditions
(2.6) |
where, as often in the rest of the paper, the dependence on and are skipped to make the notations simpler when there is no confusion in the context. Apparently this is a non-trivial solution and the following lemma holds.
Lemma 2.1.
For any and , is analytic in and .
To give some basic results on , we first introduce some subintervals of ,
(2.7) |
Clearly where any of the three subintervals may be empty. In particular, if , then we consider as in the statement of the following lemma. The choice of the above constant and the fact ensure
(2.8) |
The following is a part of Lemma 3.9 in [17].
Lemma 2.2.
For any , there exists depending only on , , and , such that, for any , it holds
(2.9) |
for all . In addition, if , then for all ,
(2.10) |
Otherwise if , then
(2.11) |
and for ,
(2.12) |
Remark 2.1.
Even though is assumed in the lemma, the estimates are uniform in and hence they also hold for the limits of solutions as , while the limits as are the conjugates of those as .
To study the limits as converges to the singular set , let
(2.13) |
The following lemma (a combination of parts of Lemmas 3.10, 3.12, 3.13, 3.14, 3.15, 3.16, and 3.19 in [17]) gives some basic properties of including their dependence in .
Lemma 2.3.
Assume , . For and , the following hold, where is determined only by .
-
(1)
As , uniformly in and , locally uniformly in .
-
(2)
For each , if . Moreover for any and is in .
-
(3)
There exists a unique continuous-in- real matrix valued determined only by such that
Here,
and is in , , and and
-
(4)
For any , for any and .
-
(5)
For any , it holds
(2.14) for satisfying and moreover, for ,
(2.15) -
(6)
is locally in both and for any and are in at any except for at .
-
(7)
are locally in both and for any at any satisfying .
-
(8)
are in both and at any satisfying and . Moreover, the following estimates hold for in a neighborhood of (uniform in ), where is given in (2.7).
-
(a)
For any with and , where is any interval satisfying , we have
(2.16) where depends only on , , , and .
-
(b)
Suppose and , then, for satisfying ,
(2.17) and for ,
-
(a)
-
(9)
is in and for . In addition, for , it holds
(2.18)
Remark 2.2.
In many cases, to simplify the notation, we denote for . In (2.14), is understood as when . The regularity of , the analyticity of and in and with , and its convergence as imply is also in and when restricted to .
Finally, the following quantity related to the Reynolds stress is crucial for the linearized water wave problem:
(2.19) |
with the domain
Those excluded points (except ) exactly correspond to the eigenvalues of the linearized Euler equation in the fixed channel at the shear flow . The end point is also excluded due to the singularity of at . The following lemma (Lemmas 3.20, and 3.22 and parts – some in the proofs – of Lemmas 3.23, 3.24, and 4.5 in [17]) summarizes some basic properties of .
Lemma 2.4.
Assume , . It holds that and is a.) analytic in ; and, when restricted to , b.) in , and c.) in and locally in for any . Moreover,
-
(1)
for all and .
-
(2)
There exists depending only on such that
-
(3)
For any , there exist and depending only on , , and such that,
-
(4)
For any and , there exists depending on and such that
if and satisfy
-
(5)
for and
Moreover, the above formula implies is in and in .
-
(6)
Assume , then for any , , , and , is locally in .
-
(7)
Assume satisfies for all , then, for any ,
(2.20) and for ,
(2.21) -
(8)
For any and
where and the equal sign in the second inequality happens if and only if .
Here we recall that denotes the Hilbert transform, namely,
for any function defined on , where P.V. is the principle value of the singular integral.
2.2. Basic properties of eigenvalues
From (1.4), , with , is an eigenvalue of (1.3) in the -th Fourier modes iff
(2.22) |
In the limit as , let
Clearly, also generates the corresponding eigenfunction of (1.3) if . The roots of with are often referred to as unstable modes, while those roots as neutral modes. We recall that Yih proved that the semicircle theorem also holds for (1.3), i.e., (1.8) holds for all unstable modes [28].
Since may not be in near (see Lemma 2.3) which would turn out to be a key bifurcation point, we also consider an actually equivalent equation
(2.23) |
with defined in (2.19), and
Clearly it holds
(2.24) |
(2.25) |
According to the following Lemma 3.1, which is an improvement of Lemma 3.24 in [17], is near (without additional assumptions, unlike Lemma 3.24 in [17]). Obviously and are analytic in their domains except for .
Our strategy to analyze the eigenvalue distribution includes the following key ingredients:
-
•
asymptotic analysis of eigenvalues for , which turn out to accumulate at ;
-
•
the existence and bifurcation of (possibly unstable) eigenvalues for near singular neutral modes at or interior inflection values of ;
-
•
an analytic continuation argument on the extension of the non-singular modes away from the above bifurcation points.
This strategy had been successfully applied to analyze the eigenvalue distribution of the linearized capillary gravity water waves at monotonic shear flows in Section 4 in [17]. The absence of the surface tension does not affect some basic properties of and , some of the bifurcation analysis of unstable eigenvalues, or the continuation argument. In the rest of this subsection, we shall outline the basic properties and the continuation argument obtained in [17]. The bifurcation analysis will be given in Section 3 with a similar approach, but under considerably relaxed assumptions. The existence of singular neutral modes and the eigenvalues for will also be studied in Section 3, where we shall see the phenomena would turn out to be substantially different from the capillary gravity waves.
We first give some elementary properties of and for a monotonic shear flow , starting with some relatively qualitative properties.
Lemma 2.5.
Assume , , then for any , the following hold.
-
(1)
is well defined for . When restricted to , is in and is in both and and is also in both and with . The same holds for except at where .
-
(2)
if and thus .
- (3)
-
(4)
For any and ,
and thus if and .
These statements are mainly contained in Lemma 4.1 in [17] where one could easily see that the lack of surface tension (namely, ) does not affect the proof. The next lemma is on some quantitative properties of and .
Lemma 2.6.
Suppose . The following hold for any .
-
(1)
is in and is well-defined for close to and , near , and
-
(2)
Let , we have
where “=” occurs only at .
Again, as the missing surface has no impact on the proofs, the results in this lemma are mostly contained in the statements and the proofs of Lemmas 4.1 and 4.5 in [17], except in (2) follows directly from Lemma 2.4(8). In particular, the form of for was obtained in Section 4 in [17] based on the explicit formulas
(2.26) |
As and are analytic functions in their domains outside , the analytic continuation argument is a standard tool in the study of the spectra of the linearized Euler equation at shear flows. The following lemma also applies to due to Lemma 2.5(2).
Lemma 2.7.
Assume . Suppose and satisfy and , then the following hold.
-
(1)
There exists an analytic function defined on a max interval such that and .
-
(2)
for all if and only if .
-
(3)
If (or ), then
-
(a)
(or if ), or
-
(b)
(or if ).
-
(a)
Remark 2.3.
This is exactly Lemma 4.3 in [17] where the absence of the surface tension does not affect the proof. It is based on the non-negative integer valued index of (or any general complex analytic functions) on various appropriate bounded piecewise smooth domains satisfying that is holomorphic in , in , and on ,
(2.27) |
which is equal to the total number of zeros of inside , counting their multiplicities. In particular, given a bounded domain with piecewise smooth and such that for any and , Ind is a constant in . If Ind, then the unique root of in is simple and analytic in . In addition, if , then is even in due to the evenness of in and the uniqueness of its root in . The proof of statement also needs Lemma 3.2(2) which ensures the roots to be bounded for in any compact interval.
3. Eigenvalue distribution of the linearized gravity water waves
In this section, we study the eigenvalues of the linearized gravity waves in details. A complete eigenvalue distribution will be obtained under certain conditions. On the one hand, we shall focus on those aspects which are different from the linearized capillary gravity waves, e.g. the instability in wave number . On the other hand, some similar results as in [17] will be obtained under substantially weaker assumptions, which requires some further detailed analysis on and given in the next two subsections.
3.1. Further analysis of
In this subsection, we extend the analysis of defined in (2.19). In particular, instead of using Lemma 3.24 in [17], we shall obtain the same improved regularity of near without any assumption additional to the monotonicity of .
Lemma 3.1.
Suppose . Assume , then for any there exists depending on and such that for any and satisfying
(3.1) |
it holds
(3.2) |
Moreover, if then for any , , , and , are locally in and in the domain .
Proof.
We first work on the regularity of near . Fix . Our proof is based on an integral formula of for near and .
The first step is to identify a domain where there are only finitely many zero points of . Due to the Semi-circle Theorem for the linearized Euler equation at a shear flow in a fixed 2-dim channel , its unstable and stable modes, which correspond to the zero points of , are contained inside the disk (1.8) with a diameter . The analyticity of in with yields all zero points of with are isolated. Due to the continuity of in restricted to , Lemma 2.3(4)(5), and the Semi-circle Theorem, the set of all accumulation points of the zero points of is compact and
Therefore, for any , there exists a function such that
Roughly is supported inside and the region bounded by the graphs of contains all except finitely many roots of . Clearly can be constructed so that it is supported in any prescribed neighborhood of . Let
which is a compact subset of with its smooth upper and lower boundaries given by the graph of restricted to . For close to , is holomorphic in , where denotes the open ball in centered at with radius .
Our next step is to derive an integral formula of . For any , let
(3.3) |
which is a neighborhood of roughly with the margin . For
the Cauchy Integral Theorem yields
Since the term in the Rayleigh equation (2.1) as , one may prove and as (see Lemma 3.3 and the proof of Lemma 3.21 in [17]), which also holds even if . Therefore the outer integral along converges to as and we obtain
We observe that is the union of the two graphs of over , the left half of the boundary of the square centered at with the horizontal and vertical side length , and the right half boundary of such a square centered at . As , due to the continuity of at and its logarithmic upper bound near (Lemma 2.4(2)), the Cauchy integrals along the half boundaries of the squares converge to zero as . Therefore the above integral formula yields
(3.4) |
for
where, for ,
(3.5) |
Here we used the regularity of (Lemma 2.4) and the property . For and , the above formula applies to . By taking , we obtain
(3.6) |
For any such that (namely, ), by taking sufficiently small in the choice of , the above formulas (3.4)–(3.6) apply to for close to .
While analytic in and, when restricted to , is in (Lemma 2.4), so we only need to focus on near restricted to . In the above formulas, the integrals along and its conjugate are smooth in near and thus we only need to consider the regularity of the term . We recall is an interior point of according to the definition of . From Lemma 2.4(6), even though is locally in if , when viewed as a function of in a whole neighborhood of , only holds due to the jump of at . The desired regularity of follows from that of , the above representation formula of , and the boundedness in of the convolution by uniform in the parameter .
Finally, it remains to prove inequality (3.2). According to (2.9), there exists such that for any , it holds
and thus is well-defined for all . In this case, (2.20) for becomes
Again, (2.9) implies, for and with ,
Therefore
where the substitution was used in the above integration. Hence inequality (3.2) follows for . For , let be the open disk centered at with radius . From the Semi-circle Theorem, for all and and is analytic. From the same procedure in deriving (3.4), but replacing the inner contour by , we obtain
Since , due to the Semi-circle Theorem and the regularity of , (3.2) follows immediately. ∎
The following corollary is a direct consequence of the above lemma, the definition (2.23) of , and Lemma 2.4.
Corollary 3.1.1.
Assume , , then, when restricted to and near , is in and is for any , , and .
3.2. Basic properties of and
Recall that eigenvalues of the linearized gravity waves at the shear flow are given in the form of where satisfy with and defined in (2.22) and (2.23). In this subsection, we derive some basic estimates of them. In particular, in Lemma 3.10 we prove a non-degenerate sign property of at , crucial for the bifurcation of eigenvalues, without assuming the convexity/concavity of as in Lemma 4.9 in [17].
Lemma 3.2.
Assume , then we have the following for any .
-
(1)
There exists depending on , , and , such that
where we recall .
-
(2)
For any , there exists depending only on , , and , such that, for any and satisfying ,
Statement (1) is mainly applied for while statement (2) mainly for .
Proof.
From Lemma 2.2, one observes that the logarithmic singularity in is significant only when and it becomes bounded when multiplied by . Therefore
which implies statement (1) in the lemma.
When , , and , it is easy to estimate the Rayleigh equation (2.1) and derive that and are uniformly bounded by some depending on , , and . Hence by regarding in (2.1) as a perturbation term, from the variation of parameter formula, it is straight forward to estimate
Statement (2) follows immediately. ∎
The next lemma states that becomes a singular neutral mode for a unique .
Lemma 3.3.
Assume , then for any and there exists , unique among , such that .
3.3. Interior singular neutral modes
Lemma 2.7 on the analytic continuation and the continuity of (Lemma 2.5) when restricted to imply that any branch of roots of can be continued until it either collides with another branch or approaches singular neutral modes – roots of with which are at the boundary of analyticity of . While is impossible (Lemma 2.6(1)), besides obtained in lemma 3.3, in the following we give some basic properties of singular neutral modes with .
Lemma 3.4.
Assume . The following are equivalent for .
-
(1)
There exists such that .
-
(2)
There exists a such that .
-
(3)
and the Sturm-Liouville operator has a non-positive eigenvalue, where with boundary conditions
Moreover, it is also satisfied that for such and any with .
We observe that the assumption implies that along with its boundary conditions is a self-adjoint operator with coefficient.
Remark 3.1.
The last statement also applies to . Namely is the eigenvalue of at .
Proof.
Before we prove the equivalence of the above statements, we first extend the monotonicity, as well as the concavity under certain conditions, of and with respect to (originally for in Lemma 2.4(8) and 2.6(2), respectively) also to any satisfying . Here we observe that implies and is even for near for such and any and thus as well. The proof is similar to that of Lemma 4.5 in [17], starting with the following claim.
Claim. Assume or with and are solutions to
then we have
(3.7) |
Moreover, if on which in particular is true if due to lemma 2.3(4), then it also holds
(3.8) |
Here the assumption on insures that has coefficients over . The claim follows through direct computations using .
If , then apparently and . It is straight forward to compute by differentiating with respect to ,
and and satisfy the zero Dirichlet boundary conditions assumed in the claim. Applying (3.7) to and , along with Lemma 2.3(4)(5), implies, for or with as long as ,
(3.9) |
Moreover, if on , then
(3.10) |
We are ready to prove the lemma. According to Lemmas 2.5(2)(3) and 2.6(1), iff where and are necessary. Consequently (3.9) immediately yields the equivalence of statements (1) and (2). Moreover Lemmas 2.5(4) and 2.3(5) imply or , where in both cases. We obtain the equivalence with statement (3) simply by observing the associated Sturm-Liouville problem structure for the neutral modes. ∎
Next we prove that any interior inflection value of is a singular neutral mode for some wave number, which is through a different proof as in [28, 11].
Lemma 3.5.
Suppose and satisfy , then there exists such that with its eigenfunction for all .
Proof.
According to Lemma 3.4, iff is an eigenfunction with the eigenvalue of the Sturm-Liouville operator along with boundary conditions in (1.5). It corresponds to the variational functional
for with . Consider
which is clearly a qualified test function. It is straight forward to verify
It yields the negative sign of the first eigenvalue of with its eigenfunction . ∎
Remark 3.2.
The proof of Lemma 3.5, which is different from that in [28, 10, 11], actually implies the existence of singular neutral modes at inflection values without assuming the monotonicity of . More precisely, it holds that
"Suppose , , , and for all , then has at least distinct roots in and at least roots in ."
To see this, one may consider the test functions where is understood. Clearly they are -orthogonal and and . Hence the non-positive subspace is at least -dim. The statement follows immediately since each eigenvalue of this Sturm-Liouville problem has only one linear independent eigenfunction. The above assumptions on is more general than those on the so-called class in [10, 11].
Given an interior inflection value , the property , whenever , does not imply the uniqueness of the root of due to the possibility of for some . In the following we address the number of wave numbers which make a singular neutral mode. We first prove a lemma on the relationship among several critical wave numbers.
Lemma 3.6.
Suppose and satisfy and is the maximal root of , then the following hold.
-
(1)
If satisfies , then .
-
(2)
If on , then where is the unique root of given in Lemma 3.3.
Clearly the above coincides with a singular neutral mode for the linearized channel flow at the monotonic .
Proof.
It is clear that iff is an eigenfunction with the eigenvalue of the Sturm-Liouville problem
(3.11) |
which corresponds to the variational functional
Since obviously and has one more restriction on its domain, so the first eigenvalue of must satisfy .
One may verify
Hence the assumption on implies . Since , we obtain that the first eigenvalue of satisfy . ∎
Finally we show how whether an interior inflection value has either one or two critical wave numbers depends on and .
Lemma 3.7.
Suppose and satisfy , then the following hold.
-
(1)
There exists at most one such that .
-
(2)
Such exists iff , where
(3.12) -
(3)
Let if exists or otherwise, there exists a unique such that . It also holds that the corresponding eigenfunction on and .
-
(4)
If , then
-
(a)
if , then there exists such that , and
-
(b)
if , then is the only root of on .
-
(a)
Proof.
Recall that iff is an eigenfunction with the eigenvalue of (3.11) corresponding to the variational functional on . Suppose is an eigenvalue with an eigenfunction , then Lemma 2.3(5) applied to initial conditions at both and implies on and , and thus on both . Hence it is the first eigenvalue and therefore the only non-negative one of (3.11), which also implies is unique if it exists. This proves statement (1).
On the one hand, suppose such does not exist, then is well-defined for all . The fact for all (Lemma 3.4) and with (Lemma 3.5) imply if it is well-defined. Since this conclusion is independent of , we obtain if is well-defined. On the other hand, assume exists. From the Sturm-Liouville theory (or one may easily prove it directly), has a root with . Moreover Lemma 2.3(5) yields . As in (2.26), it is straight forward to verify, for ,
Consequently, one may compute
This completes the proof of statement (2).
3.4. Non-degeneracy of at .
As indicated in Lemmas 2.5(4) and 2.6(1), is the only singular neutral mode which is not an inflection value of . It would be one of the key bifurcation points where the instability occurs. Our goal in the following is to verify a non-degeneracy of at to be used in the bifurcation analysis. We need the following two lemmas of technical preparation. The first is an estimate related to the fundamental solution .
Lemma 3.8.
Suppose , , and let if , then, when restricted to and , for , , , and , it holds that locally in and . Moreover, at , it holds,
and there exists depending on and such that, for ,
According to Lemmas 2.2 and 2.3(4)(5), it holds for any and in an open set containing , and thus is well defined for such . For , Lemma 2.3(3) and imply is in and vanishes at . The above lemma shows that the latter is actually of quadratic order and gives the leading order coefficient.
Proof.
From the Rayleigh equation (2.1), satisfies
(3.13) |
For , obeys the above Rayleigh equation with smooth coefficients and initial values given at . Lemma 3.1 implies the regularity of and .
The desired estimates of and at follow immediately from Lemma 2.3(3)(4) along with . With the differentiability in near had been obtained, differentiating (3.13) in yields
(3.14) |
For , the claim in the proof of Lemma 3.4 applies and Lemma 2.3(4) and (3.8) imply
(3.15) |
Using the estimates on and we obtain, at ,
To complete the proof of the lemma, for , we consider the property of near . From (3.13), it holds
Hence
which is clearly of the quadratic order in . We obtain the desired asymptotics using the leading order expansions of and near . ∎
The next lemma is a property of families of real analytic functions whose roots obey certain properties related to the Semi-circle Theorem (1.8).
Lemma 3.9.
Let and be a real analytic function of (i.e. additionally it satisfies ) and is also a function of for , , and . Moreover assume
(3.16) |
then there exists and a function defined for such that, for any such that ,
which clearly implies
Note that being a real analytic function of and up to yields for , hence for any . The first assumption in (3.16) is crucial. Otherwise a counter example is where clearly it has to be , but has another root for .
Proof.
It is easy to see how should be defined. In fact, from and the Implicit Function Theorem applied to , there exists and a function defined for and such that and
(3.17) |
Let . As we assumed iff and only if , by the compactness of , there exist and such that
(3.18) |
It remains to prove if and , which we proceed by an argument by contradiction.
Assume
(3.19) |
The definition of yields . When is sufficiently small, the function of is not identically zero, which, along with its analyticity, implies that it is not identically zero on . Therefore is not identically equal to for . Without loss of generality, suppose
Let
From (3.17), (3.18), and , the index Ind is well defined and takes a constant non-negative integer value for . Since with and , it holds Ind for . However, for any and . It implies Ind for , which is a contradiction. Therefore (3.19) can not be true and we complete the proof of the lemma. ∎
We are ready to prove the main lemma of the non-degeneracy of at . In order for this result to be also applicable to the capillary gravity waves to improve results in [17], we include the surface tension in the consideration. Let
(3.20) |
whose zero points correspond to the eigenvalues of the capillary gravity waves linearized at the shear flow if (see [17]).
Lemma 3.10.
Suppose , , and satisfy
then
Proof.
According to Lemma 2.6(1), it must hold . Due to the evenness of in , without loss of generality, we assume . Essentially we need to prove , which will be done by an argument by contradiction based on Lemma 3.9 and a carefully constructed localized perturbation to . We shall first prove the lemma under the assumption
and then remove it at the end of the proof.
Let be an auxiliary function satisfying
(3.21) |
which implies
(3.22) |
where is the characteristic function. For and
(3.23) |
to be determined later, let
which coincides with in the -neighborhood of . Sometimes we skip some of the variables to prevent the notations from being overly complicated. Clearly,
Let be the solution to the homogeneous Rayleigh equation (2.1) at the shear flow with the initial condition (2.6) and and be defined in terms of as in (2.19) and (3.20), respectively, which also depend on and .
To see the regularity of and in as well as near , let be the solution to the homogeneous Rayleigh equation (2.1) with the initial conditions
which clearly implies
Recall that is independent of for due to the definition of . In particular, as (Lemma 2.3(4)), which implies that is well-defined for near . Through the same proof of Lemma 3.1 and Corollary 3.1.1 (where being replaced by does not affect the arguments), when restricted to , locally in and , for , , , and . Apparently is smooth in , , , and near . Therefore and are smooth in and satisfy the same regularity in and . Due to the assumptions on and and the semi-circle Theorem of the unstable modes of the water wave problems, the hypotheses of Lemma 3.9 are satisfied. Hence there exist and a function defined for such that
(3.24) |
To prove the lemma by an argument by contradiction, we assume
(3.25) |
and then prove that there exist and satisfying (3.23) such that
(3.26) |
This would immediately lead to a contradiction to (3.24) for some small .
The definition (3.20) and yield
(3.27) |
where we skipped some arguments. Subsequently we can compute
(3.28) |
The negation assumption (3.25) also yields
(3.29) |
Meanwhile, from the Implicit Function Theorem, the definition of , (3.21), and (3.23),
(3.30) |
Moreover, from (3.28) one may compute
where and should be substituted by (3.30) and (3.29), respectively. According to (3.22), can be approximated by . Along with (3.30) and (3.29), we obtain
(3.31) |
where
and is proportional to depending on , , , , , but independent of and satisfying (3.23). In the rest of the proof, we will estimate and carefully to show that there exists such that (3.26) holds for based on (3.31). In the following the generic constant depends on , , , , and , but always independent of and .
Estimates of . For , let
which are well-defined and positive due to Lemma 2.3(4) and satisfies (3.13) and those estimates given in Lemma 3.8. Differentiating (3.13) with respect to , we have
For , the claim in the proof of Lemma 3.4 applies and yields
(3.32) |
One may compute
(3.33) |
along with (3.22) we obtain an estimate on
(3.34) |
where we used the fact that is supported in with total integral equal to 1.
Estimates of . Differentiating (3.13) with respect to and yields
with zero boundary values at . Again from (3.7), we have
(3.35) |
where and were given in (3.15) and (3.32), respectively. Using (3.15) and through an integration by parts, it follows
evaluated at . Through another integration by parts in a similar fashion applied to , we obtain .
Like (3.34) for , we will also identify the leading terms in of . From Lemma 3.8, we have
which implies the continuity of in for any . Since (3.23) implies , with the above inequality, (3.22), and (3.33), we are ready to obtain the leading order term of
(3.36) |
The last term is handled similarly starting with
Moreover we can estimate using Lemma 3.8
Again, using (3.22) we obtain
Therefore (3.35) implies
(3.37) |
where
From (3.31), (3.34), and (3.37) we obtain
(3.38) |
where
(3.39) |
and , and are defined right bellow (3.31). Clearly for (3.26) to hold for some and , it suffices to show , which is independent of , is not identically zero. This will be achieved by computing .
Direct calculations and using the Rayleigh equation (3.13) yield
Similar direct calculations lead to
Using the Rayleigh equations (3.13) and (3.14) it follows
Reorganizing the terms we obtain
As are independent of , Lemma 3.8 and the above computations imply
and thus is not a constant of .
Therefore, according to (3.38), (3.26) holds if is close to and . This contradicts (3.24) and thus (3.25) can not occur. This prove the lemma under the assumption .
Finally, we shall complete the proof of the lemma in the case of . From Lemma 2.6, this can happen only if , namely, in the case of the linearized capillary gravity waves. Our strategy is to consider the problem with modified parameters
The corresponding function associated to the eigenvalue problem becomes
which satisfies
Therefore, for any , satisfies the assumption on along with . Hence the above proof implies . It completes the proof of the lemma. ∎
3.5. Bifurcation analysis
With the technical preparations of the previous subsections, we shall consider the bifurcation of the unstable eigenvalues from limiting neutral modes. In the following lemma we incorporate the bifurcation analysis of near and inflection values of , which often leads to linear instability. The lemma is stated for defined in (3.20) with so that it also applies to linearized capillary gravity waves.
Lemma 3.11.
Suppose and satisfy
then there exist , , and for any if and , or if and , such that ,
and
(3.40) |
Moreover, the following properties hold for and some determined by , and .
-
(1)
Suppose , then
-
(2)
Suppose , then either i.) or ii.) , , and . Moreover the following hold.
-
(a)
If , then
-
(b)
If , then
where , if , and any if .
-
(a)
According to (3.40), clearly is relevant if and only if . The formula for indicates the behavior of for if . The above statements (1) and (2) in combination are useful to provide such information in the degenerate cases including when .
Remark 3.3.
a.) Due to Lemma
2.6(1),
and thus and
are actually excluded.
b.) Assume , which in particular holds if due to Lemma 3.10. Statement (1) yields that has the opposite sign as . Therefore, Lemma 2.4(5) implies that if or . More importantly, (3.40) implies that whether has zero points near for is determined by the sign of . A sufficient test for the latter is obviously the signs of and .
c.) From Lemma 2.5(4), it holds that either , where , or with . In the latter case, is equivalent to which is sufficient for .
d.) According to Lemmas 2.6(2) and 3.6 and (3.10), if a.) , or b.) and is the greatest solution to , then . Hence, when , we have and
Hence in the above statement (2a), gives the leading order term of for near .
Proof.
Since , when restricted to the upper half plane , according to Lemma 2.5(1)(2) and Corollary 3.1.1, is in and near (actually if due to Lemma 2.5(1)). Since is not continuous at in general, let be a extension of into a neighborhood of which coincides with for (or extension if ). The real Jacobian matrix of satisfies
where we also used the Cauchy-Riemann equation satisfied by when restricted to . From
the Implicit Function Theorem implies that all roots of near form the graph of a complex-valued function which contains (or if ). This and (2.25) prove the existence and the basic properties of and (3.40). In the rest of the proof, we study the properties of .
Suppose . For any , from the Mean Value Theorem, there exists between and such that
The regularity of and and the Cauchy-Riemann equation imply
The desired estimate on in statement (1) follows immediately.
In the rest of the proof, we consider the case of . In this situation, Lemma 2.6(1) and 3.4 (and (3.9) as well) implies either i.) or ii.) , , and , where we used . The analysis relies on the following equality obtained from the Implicit Function Theorem
Let us first consider the case of . Without loss of generality, we may assume . The regularity of and yields
and thus
In the case of , where must hold. Since the dependence of and on is actually through , the same conclusions in Lemmas 2.3(8) and 2.5(1) still hold that is in both and near when restricted to . Therefore can also be viewed as a function of and we have
Much as in the above, we first obtain
where if and for any if . Along with the evenness of and in , it implies, for ,
It completes the proof of the lemma. ∎
In the following lemma, we consider a special case of bifurcation of unstable modes from an interior non-degenerate inflection value.
Lemma 3.12.
Suppose and satisfy
then there exists determined by and such that for any , there exist and depending on and only and , , in and and even in , such that
where , hence for , and moreover
iff or . Here is defined as in (2.23) in terms of and .
According to Lemma 3.7(2), the assumption holds only if there exists such that . This lemma shows that the branch of the unstable eigenvalues bifurcating from the interior inflection value , starting at the smaller wave number which makes a singular neutral mode, connect back to .
Proof.
Apparently and , and thus and as well if , are smooth in . From given by (3.9), there exists such that , , and is smooth in . Hence there exists depending only on and such that . Lemma 3.11 and the assumption imply that, for any , has roots near with and . In order to obtain a global branch for each small , let
According to Lemma 2.4, is in (when restricted to near ) and , while is smooth in , is also in such , , and (also even in) . We also extend and as functions defined on a whole neighborhood of . Since and the Jacobian satisfies
the Implicit Function Theorem yields the (even in ) roots of . Clearly due to the definition of . One may compute, for and ,
where is used. Hence for . Obviously satisfies the desired properties. ∎
3.6. Eigenvalues for
A major difference between the eigenvalue distributions of the linearized gravity waves and that of the capillary gravity waves is when . We shall work on and of the linearized gravity waves. The analysis is divided into several steps starting with some rough bounds of the zeros of .
Lemma 3.13.
Assume , here exists depending only on , such that, for any , solutions to (2.22) (which is ) belong to where
Proof.
From Lemma 2.2,
where we recall . Hence we are able to work on (2.23) with and . Due to the evenness of the problem in , we only consider .
From Lemma 2.4(3), for any , there exist depending only on and , such that for any ,
On the one hand, if where , then
Taking and sufficiently large and determined by , we obtain if and .
On the other hand, if , then
Again, taking and sufficiently large determined by , we obtain , either.
The above analysis implies
(3.41) |
which also yields the desired upper bound of . Let us take the real part of , namely,
For and satisfying (3.41), a simple bound of can be derived from Lemma 2.4(3),
which implies the desired lower bound of . Taking the imaginary part of , we obtain
(3.42) |
and thus the estimate on follows from (3.41), the above bounds on and and letting and . ∎
Due to the evenness and conjugacy property (2.25), we shall only need to consider
Before we proceed to obtain the roots of , we first establish some estimates on and for .
Lemma 3.14.
Assume . There exist depending only on , such that, for any , where , and , it holds
Proof.
Due to the evenness we shall focus on . We first consider .
One may compute, for ,
and can be computed in a similar fashion, where we recall . Using Lemmas 2.2 and 2.3(8b)(9), we obtain that there exist depending only on such that, for and ,
Here we recall that was given in (2.7). Due to the singularity in the integral representation of given in Lemma 2.4(7), we use a cut-off function satisfying
For , let
We rewrite Lemma 2.4(7) for and ,
where
and, for , is understood. Using the above estimates on and the definition of , it is straight forward to see, for ,
It implies, for all with ,
where the substitution was also used in the last step. From interpolation, we obtain, for ,
Concerning , we have, for , , and ,
where the substitution was used as well. Therefore we obtain the desired estimate on in .
The estimate of for is similar, but simpler. Actually in this case, the norm of denominator in the integrand of the integral representation in Lemma 2.4(7) of is bounded below by . Hence can be estimated much as the above and we omit the details. ∎
In the following we analyze possible zeros of for in the rectangles .
Lemma 3.15.
Assume . There exist depending only on , such that there exist an analytic function and a function both defined and even for and the following hold.
-
(1)
For any ,
Namely, for , the roots of are either only if , or if .
-
(2)
satisfy the estimates
Moreover, the following inequality holds for if and ,
Proof.
Again we only need to consider for . As in the bifurcation analysis in the proof of Lemma 3.11, we consider a extension of from the domain to , which is different from the original defined on . The extension can be defined as, e.g.,
where can be chosen so that all derivatives up to the order of are matched at . Extend to accordingly and consider , namely,
(3.43) |
Treating it as a quadratic equation of , its two branches of roots must satisfy
For , the argument of the last square root is close to and the signs correspond to the square roots close to . From Lemma 3.14 and the definition of the extension , there exist some determined by such that, for and , it holds
The other term can be handled similarly and thus for such we obtain,
Moreover, from Lemma 3.14 we also have the derivative estimates of
Therefore, for , is a contraction on and an analytic contraction on . Let denote their unique fixed points, which satisfy the same above leading order asymptotics as and are in .
It is straight forward to compute
From Lemma 3.14 and the leading order estimates of which yields
we obtain the estimate on
To complete the proof, we consider the imaginary parts of . Using (3.43) we have
From the Mean Value Theorem, there exists between and such that
which, along with Lemma 3.14, yields, at ,
Therefore we obtain
which, along with Lemma 3.14 and the asymptotics of , implies the desired estimates on in term of . In particular, since , we have and . As belongs to the domain of analyticity of , it is also an analytic function of . Finally, the upper and lower bounds of follow from Lemmas 2.4(5) and 2.2 and . ∎
3.7. Eigenvalue distribution
With the above preparations, we are ready to prove the main theorems on the eigenvalue distribution of the linearized gravity water waves.
Proof of Theorem 1.1. Throughout the proof, we recall that and have the same zero sets (Lemma 2.5(2)) and we often mix them in the arguments.
Most of statement (1) has been proved in Lemma 3.15 except for the global extension etc. of in (1c) which will be proved here.
Since for which can be verified directly using its integral formula given in Lemma 2.6(1), there exists a unique such that
(3.44) |
Let
Because is also strictly increasing in for any (Lemma 2.6(2)), we have for all and . Therefore, according to Lemma 2.6(1) and the semicircle theorem [28], for any and and is the only root of in , which is also simple. It implies Ind for all . On the one hand, from Lemma 2.7 and Remark 2.3, the unique root of in is simple and depends on analytically and evenly. On the other hand, Lemma 3.15 implies that the root for . Therefore coincides with for and thus serves as its extension for all as the only root of in , which is simple. It is also the only root of in since for all and obtained in the above. Finally, since , does not change sign as is a simple root for all , and where (Lemma 2.6(2)), we obtain for . This completes the proof of statement (1c).
In statement (2), the existence and uniqueness of has been obtained in Lemma 3.3. From Lemma 2.6, there exist unique in , such that
(3.45) |
Due to the semi-circle theorem, has exactly two roots and in , both of which are simple, and the only one in , where
The defintions of as well as that of , the monotonicity of in for , the local monotonicity of in , and the semi-circle theorem also imply that a.) for any , the roots of in have to belong to and b.) the only root of for is . Consequently, the total number of roots (counting the multiplicity) of in can change only at . From the same argument as in the above proof of statement (1c), we obtain that, for , the only root of in , which is also the only one in , is given by some . It is analytic and even in and satisfies
Here the last property is due to the monotonicity of in for (Lemma 2.6(2)) and that of . Hence and are the only roots of in for all . Moreover, for , the only root of in (and as well) is implies that this also holds for all . The limit exists due to its boundedness and monotonicity. It has to be equal to , otherwise, if , by Lemma 2.7 could be continued beyond into and then would have at least two roots and in for , which contradicts the above analysis. Finally, statement (2c) follows from Lemma 3.11(1) and the signs and at (Lemma 3.10).
Statement (3) regarding near the inflection values of has been proved in Lemmas 3.7 and 3.11, see also Remark 3.3.
Statement (4) on the linear instability is a direct corollary of statements (1)–(3).
We proceed to prove Theorem 1.2 assuming either or on .
Proof of Theorem 1.2. Suppose on , then Theorem 1.1(2) and Lemmas 2.3(4) and 2.5(4) imply that with iff . From Lemma 3.15 (or Theorem 1.1(1)), there exists such that, for any , there are exactly three roots of , which are simple and given by (already extended for all in Theorem 1.1(1)), , and , where due to . For any , for all and . The conjugacy property (2.24) and the continuity of (restricted to ) imply that there exists such that if and . For any , let
and be its (disjoint) upper and lower connected components with . Clearly , , and for all and due to the semi-circle theorem and the choice of . Therefore
From Lemma 2.7 and Remark 2.3, the simple root of can be extended as a simple root of for all , which is the only root of in . Hence, by taking all , can be extended for all as the only root (counting the multiplicity) of in . From the semi-circle theorem and Theorem 1.1(1b), and are the only roots of for , which are also simple.
Recall the branch of the root of for obtained in Theorem 1.1(2). The assumption yields . Choose sufficiently small so that for all and . Hence the semi-circle theorem implies where is defined in the same form as in the above. While Ind, both and are roots of in . Therefore we obtain for .
A similar argument based on the index, continuation, and the Semi-circle Theorem starting at also yields that the roots of for are either or close to . Hence Theorem 1.1(2c) implies that are all the roots for . There exists such that for all and . Hence for all , , and where
and and have been given in (3.44) and (3.45), respectively. Clearly, for all , IndInd and the two roots are and . For , has no other roots outside due to the semi-circle theorem and the choice of and . Therefore are also the only roots of for all . It completes the proof of the case of .
Assume on . Firstly, we show the spectral stability following the most standard technique, which had actually been obtained in [28] (see also [10]). Suppose is an unstable mode (i.e. ) with the eigenfunction . Multiplying the Rayleigh equation by , integrating by parts over , taking the imaginary part, and using the boundary condition of (1.5c), we have
The semi-circle theorem implies and thus Lemma 2.5(2) yields which is a contradiction. Hence the spectral stability follows.
Due to Lemmas 2.5(2)(4) and 2.6(1), the only singular or non-singular modes inside the circle (1.8) have to be at inflection values of . The remaining statements in this case has been proved in Theorem 1.1.
Remark 3.4.
Finally we prove Theorem 1.3 assuming has exactly one non-degenerate inflection point on .
Proof of Theorem 1.3. Suppose (1.6) holds with and . The same type of argument as above based on the index and the continuation will be employed extensively below on the semi-disk domain
Let us start with the case of which implies on . Lemma 2.3 yields over for any . Multiplying the Rayleigh equation by , integrating by parts over , and using , we obtain for all . According to Lemma 3.7, there exists , unique among , such that . Along with the Semi-circle Theorem111From the standard proof of the Semi-circle Theorem, one could see that unstable modes can not occur on the boundary semi-circle of (1.8) when on ., Lemmas 2.5(4) and 3.3, it implies that and are the only roots of with . On the one hand, since near , Lemma 3.11 implies that there are never singular or non-singular modes with near for any . Hence the index Ind remains a constant for . On the other hand, since near , Lemma 3.15 implies that with is the only unstable mode for , which implies Ind for all . Again from Lemma 3.11, there exists a branch of unstable modes with and for , which are the only singular or non-singular modes near with . From Lemma 2.7, this branch can be continued for all . Since has exactly one root with for , we obtain that the continuations of and must coincide and form the only singular or non-singular modes with . The local bifurcation given by Lemma 3.11 implies that Ind for and thus there are no unstable modes if . Along with Theorem 1.1(2b), it completes the proof of this case.
In the rest of the proof we assume . Lemma 3.6 yields . Much as the above, the index Ind remains a constant except for (no if it does not exist). Note (1.6) and implies on and on . From Lemmas 3.15
(3.46) |
By a similar argument to that in the proof of Theorem 1.2, Lemma 3.11, which gives all the unstable modes near the singular neutral modes at and , implies that the index increases by 1 as increases through and decreases by 1 as increases through , as well as through if exists. Therefore we have
(3.47) |
where is understood if does not exist. In the case does not exist, we also have
(3.48) |
and otherwise
(3.49) |
The linearized system (1.3) is unstable when the above index is greater than zero. By a similar argument to that in the proof of Theorem 1.2 bases on Lemma 2.7 and the index, the branches of unstable modes bifurcating from the singular neutral modes can be continued in unless approaches , , and or it collides with another branch, the latter of which happens only possibly for . Let us consider the different cases (2)–(4) in the theorem separately.
Suppose does not exist and thus are the only wave numbers which make a singular neutral mode. From Lemma 2.7 and (3.47), the branch of unstable modes obtained in Theorem 1.1(2) (or Lemma 3.11 more directly bifurcating from ) can be continued for all . Due to (3.47), this branch has to coincide with the branch bifurcating from (again by Lemma 3.11) for and with . From the Semi-circle Theorem and (3.48), is the only singular or non-singular modes with and statement (2) is proved.
Suppose exists and . By continuity, the branch satisfies for . The same argument as above also yields that the branch connects to and is the only singular or non-singular modes with and . Due to (3.49), there are no singular or non-singular modes with and . From Lemmas 3.11 and 2.7, both branches of unstable modes bifurcating from for can be continued for all and have to coincide due to (3.49). This branch is apparently even in and the only singular or non-singular modes with and . It completes the proof of statement (3).
To prove the last case, suppose exists and . From the Semi-circle Theorem and the assumptions, it is clear
Claim. For any connected component of , it holds
We shall prove the infimum case and the supremum case is similar. In fact, since the compactness of yields that of , there exists such that . If where is analytic, then is an isolated zero of . Since, as varies locally, the set of zero points of analytic functions can always be extended continuously beyond (even though splitting may occur if is not simple), it contradicts with the definition of . Therefore and thus . Lemma 3.11 implies that is the end point of a branch of unstable modes with , therefore can not be the infimum. Another two possibilities or can be excluded in the same way as . The claim is proved.
Let be the connected component of containing and . From (3.49),
(3.50) |
Applying the above claim to the supremum to , there are two possibilities.
Case A. and thus . In this case, let be the connected component of containing . As is the supremum of and , we have . This implies that , otherwise which would yield . Again the above claim implies either or . From (3.50) and the connectedness of , it holds . By the even symmetry of in , immediately we derive the same even symmetry of about and thus . Since all points with belong to , it follows from the above claim that these three sets are all the connected component of and thus . The desired statement (4) in case (4b) is obtained.
Case B. and thus . Accordingly we let be the connected component of containing . Again we consider the two possibilities separately.
* Case B.1. , too. In this case, we have . Again since all points with belong to , it follows from the above claim . Due to the symmetry, we obtain and thus . Hence the desired statement (4) in case (4b) holds.
* Case B.2. , which yields . The above claim implies and thus or . The desired statement (4) in case (4a) follows from the same argument as in the above Case A. The proof of the theorem is complete.
According to Lemma 3.7, if exists, i.e. is unstable for the channel flow, then exists if defined in (3.12). From Lemma 3.12 which is under a weaker condition, the closed branch of unstable modes in case (3) bifurcates from . As this branch grows, we can not exclude the possibility that it intersects with the branch connecting and obtained in case (2).
3.8. The linearization of the capillary gravity water waves at monotonic shear flow revisited
When the surface tension is also considered, the boundary condition (1.1d) is replaced by
(3.51) |
where , is the mean curvature of at . Shear flow under the flat surface is also a stationary solution. The linearization of the capillary gravity water waves at such a shear flow is given by a system similar to (1.3) only with the linearized boundary condition (1.3d) replaced accordingly by
(3.52) |
Following the same procedure, we obtain that corresponds to an eigenvalue in the -th Fourier modes of if and only if , or equivalently, , where was defined (3.20) and .
In [17], we obtained results on the eigenvalue distribution and the inviscid damping of the linearized capillary gravity waves. In particular, the semi-circle theorem still holds for unstable modes, namely, all non-singular modes outside the circle (1.8) belong to . Moreover, there exists such that has exactly two roots for , which are simple roots, even and analytic in , and satisfy
The branch can be continued as simple roots for each with and even and analytic in . The continuation of the other branch may or may not reach depending on whether
If it does, the bifurcation at was analyzed under the assumption on , which was only used to ensure a.) the regularity of near and b.) the sign if . More details can be found in Theorem 1.1 in [17]. Thanks to Corollary 3.1.1 and Lemma 3.10, these two key properties of also hold without assuming . Therefore the following proposition on the eigenvalue distribution of the linearized capillary gravity waves holds with weakened assumptions.
Proposition 3.16.
Assume and on , then the following hold.
-
(1)
If , then there exist , unique in , and an even function (for any ) defined for all such that
-
(a)
iff ;
-
(b)
is analytic in , and ;
-
(c)
and are simple roots of which include all roots of outside the disk (1.8); and
-
(d)
there exists that with and iff .
-
(a)
-
(2)
If , then there exist , and even functions (for any ) defined for and for satisfying the following.
-
(a)
iff .
-
(b)
is analytic in if and is analytic in if . Moreover, , with , and with are the only roots of outside the disk (1.8), which are all simple.
-
(c)
, , .
-
(d)
with and with have the same sign (, or ) as and , respectively, and for such ,
-
(e)
with , , iff or .
-
(a)
Remark 3.5.
a.) The bifurcation near inflection values of can be found in Lemma 3.11.
b.) If , it had been proved in [17] that can be extended for all as a simple root of which is the only root other than outside the disk (1.8)). It also satisfies . An explicit necessary and sufficient condition for was also given there. Under the assumption a complete picture of the eigenvalue distribution in the fashion of Theorem 1.1(5) was also obtained in [17], where the continuations of and coincide.
c.) Following from the same proof as in statement (6) of Theorem 1.1, one can prove that the spectral stability of the linearized capillary gravity waves.
Proof.
Due to the concavity of in (Lemma 2.6(2)), for , the existence of or as the only zero points of follows from the definition of . Much in the proof of Theorem 1.1(2), the branch starts with the simple root of . For , another branch had been constructed in Theorem 1.1 in [17]. Applying Lemma 2.7 and the semi-circle theorem, and can be continued in as the only roots of outside the disk (1.8) for all and , respectively ( is understood if ), where for the first in the continuation. It suffices to prove and the rest of the proposition would follow from the same proof as of Theorem 1.1.
Our strategy is to vary the parameter starting from . From the definitions of and , if , then it is strictly decreasing in and there exists such that . We use the notations for and for which occurs only when . From the concavity of in , we have
They also clearly satisfy
where is given in Theorem 1.1(2) satisfying .
Let us use and to denote the branches of the roots in of for , while we skip if there is no confusion. According to Theorem 1.1(2), can be extended as the only root in of for all and . The continuation in the direction of through the Implicit Function Theorem applied in a neighborhood of implies that, for each , can be extended until it reaches which has to occur at .
Let
Lemma 3.2(2) implies that has no roots in for . For any , the only roots of on are due to the semi-circle theorem. Hence from an index argument applied to sufficiently large compact subsets of , the total number of roots (counting the multiplicity) of in are constants for in , , and , respectively. Since is the only root in for and the only one for and , the total numbers of roots of in is for in both and . As , by the bifurcation analysis based on Lemma 3.11 and the signs (Lemma 3.10) and (from the concavity of in ), has no roots in for . Hence has no roots in for . For the total number of roots in of to change from to as increases through , it must hold , otherwise would be continued to a root in for which is a contradiction.
Finally, follows from taking the limit as . ∎
A direct corollary of the proposition is a sufficient condition for the linear instability.
Corollary 3.16.1.
If and there exists a sequence converging to as such that for all , then the linearized capillary gravity wave system is linearly unstable.
Remark 3.6.
For the linearized capillary gravity wave problem, a non-degenerate inflection value of does not necessarily leads to instability as a strong surface tension may prevent from becoming a neutral mode at all.
References
- [1] T. B. Benjamin. The solitary wave on a stream with an arbitrary distribution of vorticity. J. Fluid Mech., 12:97–116, 1962.
- [2] D. Bresch and M. Renardy. Kelvin-Helmholtz instability with a free surface. Z. Angew. Math. Phys., 64(4):905–915, 2013.
- [3] O. Buhler, J. Shatah, S. Walsh, and C. Zeng. On the wind generation of water waves. Arch. Ration. Mech. Anal., 222:827 – 878, 2016.
- [4] J. C. Burns. Long waves in running water. Proc. Cambridge Philos. Soc., 49:695–706, 1953. With an appendix by M. J. Lighthill.
- [5] P. Drazin and W. Reid. Hydrodynamic stability. Cambridge: Cambridge University Press., 2004.
- [6] R. Fjø rtoft. Application of integral theorems in deriving criteria of stability for laminar flows and for the baroclinic circular vortex. Geofys. Publ. Norske Vid.-Akad. Oslo, 17(6):52, 1950.
- [7] S. Friedlander and L. Howard. Instability in parallel flows revisited. Stud. Appl. Math., 101(1):1–21, 1998.
- [8] L. Howard. Note on a paper of john w. miles. J. Fluid. Mech., 10:509–512, 1961.
- [9] J. N. Hunt. Gravity waves in flowing water. Proc. Roy. Soc. London Ser. A, 231:496–504, 1955.
- [10] V. Hur and Z. Lin. Unstable surface waves in running water. Comm. Math Phys., 282:733 – 796, 2008.
- [11] V. M. Hur and Z. Lin. Erratum to: Unstable surface waves in running water [mr2426143]. Comm. Math. Phys., 318(3):857–861, 2013.
- [12] P. Janssen. The interaction of ocean waves and wind. Cambridge University Press, 2004.
- [13] A. Kaffel and M. Renardy. Surface modes in inviscid free surface shear flows. ZAMM Z. Angew. Math. Mech., 91(8):649–652, 2011.
- [14] C. C. Lin. The theory of hydrodynamic stability. Cambridge, at the University Press, 1955.
- [15] Z. Lin. Instability of some ideal plane flows. SIAM J. Math. Anal., 35(2):318–356, 2003.
- [16] X. Liu. Instability and spectrum of the linearized two-phase fluids interface problem at shear flows. arXiv:2208.11159, 2022.
- [17] X. Liu and C. Zeng. Capillary gravity water waves linearized at monotone shear flows: eigenvalues and inviscid damping. arXiv:2110.12604, 2021.
- [18] M. S. Longuet-Higgins. Instabilities of a horizontal shear flow with a free surface. J. Fluid Mech., 364:147–162, 1998.
- [19] C. Marchioro and M. Pulvirenti. Mathematical theory of incompressible nonviscous fluids, volume 96 of Applied Mathematical Sciences. Springer-Verlag, New York, 1994.
- [20] J. Miles. On the generation of surface waves by shear flows. J. Fluid Mech., 3:185–204, 1957.
- [21] J. W. Miles. On the generation of surface waves by shear flows. II. J. Fluid Mech., 6:568–582, 1959.
- [22] J. W. Miles. On the generation of surface waves by shear flows. III. Kelvin-Helmholtz instability. J. Fluid Mech., 6:583–598. (1 plate), 1959.
- [23] L. Rayleigh. On the stability or instability of certain fluid motions. Pro. London Math. Soc., 9:57–70, 1880.
- [24] M. Renardy and Y. Renardy. On the stability of inviscid parallel shear flows with a free surface. Math. Fluid Mech., 15:129–137, 2013.
- [25] V. I. Shrira. Surface waves on shear currents: solution of the boundary value problem. J. Fluid Mech., 252:565–584, 1993.
- [26] W. L. K. Thomson. Hydrokinetic solutions and observations. Philosophical Magazine Series 4, (42):362–377, 1871.
- [27] W. Tollmien. Ein allgemeines kriterium der instabititat laminarer geschwindigkeitsverteilungen. Nachr. Ges. Wiss. Gottingen Math. Phys., 50:79–114, 1935.
- [28] C.-S. Yih. Surface waves in flowing water. J. Fluid. Mech., 51:209–220, 1972.