Nonlinear modulational instabililty of the Stokes waves
in 2d full water waves
Abstract.
The well-known Stokes waves refer to periodic traveling waves under the gravity at the free surface of a two dimensional full water wave system. In this paper, we prove that small-amplitude Stokes waves with infinite depth are nonlinearly unstable under long-wave perturbations. Our approach is based on the modulational approximation of the water wave system and the instability mechanism of the focusing cubic nonlinear Schrödinger equation.
1. Introduction
In this paper, we establish the nonlinear modulational instability of the small-amplitude Stokes waves under long-wave perturbations in the context of 2d full water waves with infinite depth. The famous Stokes wave refers to a periodic steady wave traveling at a constant speed, which was first studied by Stokes in 1847 [58]. The existence of Stokes waves was rigorously proved in the 1920s for the small-amplitude cases [50, 44, 59], and in the early 1960s for the large-amplitude settings [41, 42]. These periodic traveling waves are of crucial importance in both theoretical and practical studies of water waves.
1.1. Modulational instability
The modulational instability, which is also known as the Benjamin-Feir or the sideband instability, is a very important instability mechanism in a diverse range of dispersive and fluid models. Roughly speaking, this is a phenomenon whereby deviations from a periodic waveform are reinforced by the nonlinearity, leading to the generation of spectral-sidebands and the eventual breakup of the waveform into a train of pulses. This instability mechanism has been wildly observed in experiments and in nature, such as water waves and their asymptotic models. In 1967, by Benjamin and Feir [9, 10], this phenomenon was first discovered for periodic surface gravity waves, i.e. Stokes waves, on the deep water. This is the context of our main interest in this paper.
The modulational instability also exists in various dispersive equations. The literature on this topic is extensive and without trying to be exhaustive, we mention the work by Whitham [69], Benny and Newell [11], Ostrovsky [54], Zakharov [79], Lighthill [45]. We also refer interested reader to the excellent survey by Ostrovsky and Zakharov [80] for more details on the history and physical applications of modulational instability. Moving beyond the linear level, recently the nonlinear modulational instability for a class of dispersive models was proved by Jin, Liao, and Lin in [40].
Returning to the water wave problem, nevertheless, the rigorous proof of the linear modulational instability (spectral instability) for the full water waves was quite recent. In the 1990s, Bridges and Mielke [15] were able to prove the spectral modulational instability for the finite-depth water waves linearized near a small-amplitude Stokes wave. Under long-wave perturbations, i.e. frequencies near zero, recently, Nguyen and Strauss in [51] proved the spectral modulational instability of the Stokes waves in infinite depth case. See also [33] for a simplified proof by Hur. The nonlinear instability of the full water waves remains open and is our main result in this paper. The quasilinear feature and the nonlocality of water wave systems make the nonlinear analysis here exceedingly difficult.
1.2. Water wave system
Now we introduce the full water wave system. Consider the motion of an inviscid and incompressible ideal fluid with a free surface in two space dimensions (that is, the interface separating the fluid and the vacuum is one dimensional). We refer such fluid as water waves. For simplicity, we consider the infinite depth case, that is, without a finite bottom. Denote the fluid region by and the free interface by . The equations of motion are Euler’s equations, coupled to the motion of the boundary, and with the vanishing boundary condition for the pressure. It is assumed that the fluid region is below the air region. Assume that the density of the fluid is , and the gravitational field is normalized as . In two dimensions, if the surface tension is zero, then the motion of the fluid is described by
(1.1) |
Here, is the fluid velocity, and is the pressure. We shall consider the water waves such that as .
This system, along with many variants and generalizations, has been extensively studied in the literature. The so-called Taylor sign condition (also referred as Rayleigh-Taylor sign condition in many literature) on the pressure is an important stability condition for the water waves problem. If the Taylor sign condition fails, the system is, in general, unstable, see, for example, [8, 14, 63, 28, 61]. In the irrotational case without a bottom, the validity of the Taylor sign condition was shown by Wu [72, 73], and was the key to obtain the first local-in-time existence results for large data in Sobolev spaces. In the case of non-trivial vorticity or with a bottom, the Taylor sign condition can fail and the sign condition has to be assumed for the initial data. In the irrotational case, Nalimov [49], Yosihara [78] and Craig [23] proved local well-posedness for 2d water waves equation for small initial data. In S. Wu’s breakthrough works [72, 73] she proved the local-in-time well-posedness without smallness assumptions. Since then, a lot of interesting local well-posedness results were obtained, see for example [5, 7, 18, 20, 38, 43, 46, 52, 57, 81, 1, 2, 47, 3], and the references therein. See also [55, 71, 70, 76] for water waves with non-smooth interfaces. For the formation of splash singularities, see for example [17, 16, 21, 22]. Regarding the local-in-time wellposedness with regular vorticity, see [37, 52, 53, 18, 18, 46, 81] and [60] for water waves with point vortices. In the irrotational case, almost global and global well-posedness for water waves were proved in [74, 75, 29, 39, 6], and see also [32, 34, 68, 82, 4]. In the rotational case, see [35, 13, 30], and [60]. In [77], Wu obtained long-time existence results without imposing size restrictions on the slope of the initial interface and the magnitude of the initial velocity. In particular these allow the interface to have arbitrarily large steepnesses and initial velocities to have arbitrarily large magnitudes.
1.3. Asymptotic models
To understand the behavior of the water waves, one can study the system in various asymptotic regimes. It is well-known that the 1d cubic nonlinear Schrödinger equation (NLS)
(1.2) |
is related to the full water wave system in the sense that asymptotically it is the envelope equation for the free interface of the water waves. Formally speaking, consider the modulational approximation to the solution of 2d water waves equations, i.e., a solution of the parametrized free interface whose leading order is a wave packet of the form
(1.3) |
then from the multi-scale analysis, we obtain that , and solves the 1d focusing cubic NLS. One observes that the envelope is a profile that it travels at the group velocity determined by the dispersion relation of the water wave equations and it evolves according to the NLS on the time scale .
This discovery was derived formally by Zakharov [79] for the infinite-depth case, and by Hasimoto and Ono [31] for the finite-depth case. In [24], Craig, Sulem and Sulem applied the modulation analysis to the finite depth 2D water wave equation. They derived an approximate solution in the form of a wave packet and showed that the modulation approximation satisfies the 2D finite-depth water wave equation to the leading order. In [56], Schneider and Wayne justified the NLS as the modulation approximation for a quasilinear model that captures some of the main features of the water wave equations.
The rigorous justification of the NLS for the full water waves was given by Totz and Wu [66] in infinite-depth case with data in Sobolev spaces. The justification in a canal of finite depth was proved by Düll, Schneider and Wayne [27]. See also [36]. In [62], the second author rigorously justified the NLS from the full 2d infinite-depth water waves with data of the form and therefore justified the Peregrine soliton from the water waves.
As mentioned before, to analyze the instability of Stokes waves using the water wave system directly could be complicated. In this paper, instead of working on the full system directly, we further explore the modulational approximation to the water wave via the NLS with the appearance of Stokes waves and incorporate the instability of the NLS.
Our mativation is that numerical results, for example [26] by Deconinck and Oliveras, showed that the spectrum of the linearized operator given by the Stokes wave in the 2d full water waves was qualitatively resembled by the spectrum of the linearized operator given by the special solution in the 1d cubic NLS. Therefore, it is natural to conjecture that the mechanism of modulational instability in the full water waves is governed by the 1d NLS.
1.4. Basic setting and main results
In this subsection, we formulate the basic setting of the problem and state the main result. More details and estimates are presented in Section 9.
First of all, we fix some constants. Throughout this paper, we assume that the pressure on the interface, the gravity is given by and the density of the fluid is . Denote . We identify with .
1.4.1. Wu’s modified Lagrangian formulation
It implies from and that is holomoprhic in , so is completely determined by its boundary value on . Let the interface be parametrized by , with as the Lagrangian coordinate, i.e., is chosen in such a way that . So we have . Because , we can write , where is a real-valued function. Therefore the momentum equation along can be written as
(1.4) |
Since is the boundary value of , the water wave equations (1.1) is equivalent to
(1.5) |
where by holomorphic, we mean that there is a bounded holomorphic function on with as such that .
While the coordinate above is well-suited to quasilinearize the water wave system and prove the local wellposedness, see [72], it is not convenient to study the long-time behavior of the system due to some quadratic terms appearing in and the quasilinearzation, see [74]. To obtain the nonlinear instability of the Stokes wave, we have to solve the water wave system for sufficiently long time. Following [74], we introduce Wu’s modified Lagrangian coordinate. 111Throughout this paper, we will call it Wu’s coordinate or modified Lagrangian coordinate interchangeably
Let be a diffeomorphism. Denote . We pick the such that is holomorphic in the sense as above. Then composing the equation (1.5) with the diffeomorphism , we know that solves
(1.6) |
where we used the notations
(1.7) |
Such coordinates system was first used in [74] to prove the almost global wellposedness of 2d water waves with small and localized data. Then it has been used in [75, 66, 65, 62, 60] to study the water wave problems on the long-time scale. Once knowing the existence of such coordinates, one can directly work on the water wave system in this coordinate without invoking the Lagrangian coordinates. In §3, by deriving formulae for the quantities and , we formulate the water wave system in the variables directly and avoid using the change of variables .
1.4.2. The Stokes waves
A Stokes wave is a periodic steady wave traveling at a fixed speed to the system (1.6). Under Wu’s coordinate, we can use a triple to represent a Stokes wave with the traveling velocity and velocity field given by . In this paper, we shall consider small-amplitude Stokes waves. Notice that (1.6) is time reversible, that is, if is a solution to (1.6), then is also a solution to (1.6) where
(1.8) |
Without loss of generality, we consider those Stokes waves traveling to the left with period .
Regarding the existence of the Stokes waves, one has the following result in Wu’s coordinate.
Proposition 1.1.
There exists a smooth curve of periodic traveling wave solutions to the water wave system (1.6) parametrized by the a parameter which we call the amplitude of . For each solution on this curve, one can write it as
where and satisfy the following properties:
-
•
and are periodic
-
•
and are odd, and are even.
For the element of the curve with amplitude , we denote it as
(1.9) |
and it has the following asymptotic expansions
(1.10) |
and
(1.11) |
Throughout of this paper, for the sake of simplicity, we will only focus on the case that since the case that can be treated in the same manner.
By the translational symmetry of the system (1.6), for any Stokes wave with amplitude , one can find a unique and the solution from the curve in Proposition 1.1 associated with the amplitude , such that we can write
(1.12) |
From Proposition 1.1 and (1.12), using the notation (1.9), we define the family of Stokes waves of small amplitude,
(1.13) |
where is a given small number.
1.4.3. The main result
With preparations above, we are ready to state the main result in this paper.
Theorem 1.1.
There exists a sufficiently small number such that for all , a Stokes wave with amplitude is nonlinearly modulational unstable in the following sense:
Let be a fixed positive number. For any satisfying and any , there exists a solution to (1.6) satisfying the following conditions:
-
•
Its initial data and satisfy
(1.14) -
•
The solution exists on and it satisfies
where is a fixed number.
-
•
The solution satisfies
(1.15) for some constant which is uniform in and .
Before discussing the main ideas of this paper, let us give a few comments on this result.
Remark 1.2.
Notice that (1.14) is an open condition. Therefore, our instability construction is stable under perturbations of size in .
Remark 1.3.
Remark 1.4.
The quantitative estimate (1.15) precisely implies that the waveform of the Stokes wave is broken under long-wave perturbations in . Moreover, by our explicit construction and Sobolev embedding, this phenomenon is also captured pointwisely.
Remark 1.5.
By construction, the solution we constructed here has the fundamental period . See Section 9 for details.
Remark 1.6.
One can translate the instability to the Eulerian coordinate in term of the elevations. Again see Section 9 for details.
Remark 1.7.
The long-time existence of size holds for more general flows around the Stokes wave. See Theorem 9.2 for details.
1.5. Essential ideas and outline of the proof
In this subsection, we highlight the key features and present the outline of our approach.
1.5.1. Choice of coordinates
The first key part of our work is to find good coordinates to perform expansions and the long-time estimates.
In Eulerian coordinates, it is known that the elevation of the free interface of a given Stokes wave, , the following expansion
(1.16) |
holds. See [58] and [51] for details. We should immediately notice that in this setting, up to order , there are three nontrivial frequencies. Putting this expansion into the nonlinear problem, one should expect that due to the interaction of frequencies, the nonlinear analysis would be very complicated.
Moreover, although numerical simulations tell us that the NLS is a good asymptotic model to analyze the modulational instability problem of the deep water wave problem, in the Eulerian coordinates, it is highly nontrival to see the connection between the NLS and the water wave system.
In this paper, we utilize Wu’s modified Lagrangian coordinates. In this setting, the Stokes wave has a remarkable asymptotic expansion:
(1.17) |
Compared with the asymptotic expansion (1.16) in Eulerian coordinates, up to an error of , (1.17) has only one nonzero fundamental frequency. This fact plays an essential role in our work and simplifies the analysis.
Furthermore, using Wu’s modified Lagrangian coordinates, it is relatively clear how to derive the relation between the NLS and the water wave system. The NLS has been derived from water wave system using Wu’s coordinates in other settings, see [66] and [62]. Due to the appearance of Stokes waves as will be explained in §1.5.2, the current situation is quite different than those earlier works.
From the perspective of the long-time existence, under Wu’s coordinates, one can derive structures without quadratic nonlinearities which we call cubic structures for both the Stokes wave and the perturbed flow which are fundamental for us to capture the instability. We will discuss this in details in Section §3.
1.5.2. Derivation of the NLS and its instability
The second main step is to derive the NLS from the water wave system via the modulational approximation.
From the expansion (1.17), the leading order term of the Stokes wave is given as which is in the form of a plane wave. It is natural that when we perturb the Stokes wave, the perturbed flow should be written as a wave packet
(1.18) |
where from the view of the modulational approximation, see §1.3, with and will be chosen appropriately to solve a NLS.
To derive the NLS in our current setting, as mentioned above, is different than earlier works. In the current setting, the parameter in the phase also depends on since the velocity of the Stokes wave depends on its amplitude. Whereas, in earlier works, when proceeding the modulational approximation analysis, it is always assumed that , see (1.3), which is independent of . We need some extra care to handle this dependence when we perform the multi-scale expansion. This extra dependence is also crucial for us to obtain the approximate solutions to Stokes wave and the perturbed flow.
In Section §5.3, we obtain that the NLS for is
(1.19) |
We also note that the coefficient in front of in the expansion (1.17) is a special solution to the NLS above.
Letting , then by a direct computation, it solves
(1.20) |
which is the standard cubic NLS. In this setting, the special solution corresponding to the Stokes wave is .
Therefore, in the scale of the NLS, the stability problem of the Stokes wave is reduced to the corresponding problem for the special solution
to (1.20).
Consider the perturbation of the form
Plugging the ansatz above into (1.20), we have
where is the nonlinear term.
To see the instability at the linear level, we take the real part of , . Then satisfies
Consider the plane wave, . Then direct computations give us
Solving from the expression above,
we can conclude that
-
•
When , the linear solution is dispersive.
-
•
When , the linear solution produces the exponential instability.
1.5.3. Construction of perturbations
The third step of our analysis is to construct the unstable perturbation of the Stokes wave.
Performing the multi-scale analysis in Section §5, we further expand the perturbed flow as
(1.23) |
where from the discussion above, .
Under this setting, suppose that we can control the flow for a sufficiently long time interval such that
then the dominated behavior should be given by which turns out to solve the NLS (1.19).
From the discussion before, the NLS has the strong instability around the special solution given by the Stokes wave. We can always use the initial data of the NLS to construct the initial data to the water wave system using the multi-scale analysis. Taking the initial data constructed via which causes the instability of the NLS, from (1.22), returning back to the water wave system, we obtain that at ,
(1.24) |
which implies the instability. Therefore the problem now is reduced to control the solution for a sufficiently long time interval.
1.5.4. Long-time existence
From the construction above, we notice that in order to see the instability in the water wave system, we need to solve the system on a time interval of size . In the general setting, there is no global-in-time theory for periodic water waves. At this stage, the best lifespan for the general periodic water wave systems with initial data of size is , see [12] and [77]. But the above could be arbitrarily small, say, . So any lifespan independent of will not be sufficient for us to see the instability. In order to achieve our goal, we need to design appropriate perturbations such that the perturbed flow exists at least and it needs to satisfy
(1.25) |
but
(1.26) |
at some . This is another place where this work differs from [66, 62] where the existence were proved for .
To achieve the long-time existence and obtain the corresponding the estimates, the fact that the Stokes wave is a global solution of size plays a pivotal role.
By the multi-scale analysis and expanding the solutions in terms of powers of , we can obtain the asymptotic expansions for the perturbed flow as (1.23) and for the Stokes wave (1.17). Then we define
(1.27) |
and
(1.28) |
With these two notations, we define the approximate solution as
(1.29) |
Notice that the approximate solution defined above is different from those in [66, 62]. This definition of approximate solution together with the modified Lagrangian coordinate allow us to gain extra long-time existence of solutions.
Finally, we define the remainder term as
(1.30) |
To control the remainder term , one uses the following functional
(1.31) |
In Sections §6, §7, §8, we establish that for , one has the following estimate:
(1.32) |
In particular, a direct computation of the time integral gives
(1.33) |
Then the bootstrap argument and the local wellposedness theory for water wave systems will ensure the long-time existence and estimates. The factor on the right-hand side allows us to gain the extended lifespan .
To obtain the conclusion above is far from being standard. Here we briefly illustrate the idea of computations. Formally, the quantity satisfies
with the initial data satisfying
(1.34) |
Ignoring higher order terms, the operator is morally like when acting on anti-holomorphic functions. The nonlinear term on the right-hand side consists of many cubic and higher order terms, for example, some cubic structures in terms of , , and . Although one can bound each of them separately in terms of powers of , it is not sufficient for us. We need to explore the cancellations among cubic terms to recast .
1.5.5. Development of instability
With all the preparations above, the nonlinear instability is produced naturally. Choosing the unstable solution to the NLS to construct the corresponding solution to the water wave system, then solution satisfies
(1.36) |
and
(1.37) |
at time .
By construction, the frequencies of the leading order term of the instability are in completely different scales from the Fourier modes of the family of Stokes waves. Therefore, under the long-wave perturbation, the solution will deviate from the family of Stokes waves and completely change the structure of the wave trains.
The mechanism here gives the dynamical and mathematical description of the modulational instability under long-wave perturbations: the instability of periodic wave trains due to self modulation and the development of large scale structures.
1.5.6. General remarks
As we pointed out before, the quasilinear feature and the nonlocality of water wave systems make the nonlinear analysis exceedingly difficult. One should expect that in the quasilinear setting, the interactions of long-waves and short-waves should be fairly complicated. Consequently, to obtain the long-time existence and estimate could be hard. Our modulational approximation approach and the well-chosen coordinates could get rid of these difficulties.
Our approach is quite general. To study the (in)stability problem directly in quaslinear problems could be very elaborate. On the other hand, since many quasilinear problems can be approximated by semilinear equations, we believe that our method has broader application to other problems. In particular, long-wave perturbations problems fit well into the general idea here.
1.6. Outline of the paper
In §2 we will provide some analytical tools and the basic definitions that will be used in later sections. In §3, we discuss the local wellposedness of the water waves in Wu’s coordinates and derive the corresponding formulas for a few quantities. In §4, the existence of Stokes waves and their expansions are present in Wu’s coordinates. In §5, we use the multiscale analysis to derive the NLS from the full water waves with the desired properties. In §6, we derive the governing equations for the error term and define the energy functionals. In §7, we estimate the important quantities used in the energy estimates. In §8, we obtain the a priori energy estimates. In §9, we prove the modulational instability. In Appendix A, we study the Cauchy integral in the periodic setting bounded by a nonflat curve. In Appendix §B, we provide the proof of some important identities that are used in this paper. In the Appendix §C, we provide the proof for some estimates in the previous sections for the sake of completeness. In the Appendix §D, the instability of the NLS is analyzed in details. Finally we list the frequently used notations in the Appendix E.
1.7. Notations
For positive quantities and , we write for where is some prescribed constant. Throughout, we use , for the derivative in the time variable and for the derivative in the space variable. These two way of denoting are used interchangebly. We use , to represent the real and imaginary part of , respectively.
Acknowledgement
G.C. was supported by Fields Institute for Research in Mathematical Sciences via Thematic Program on Mathematical Hydrodynamics.
2. Preliminaries
In this section, we collect some basic definitions and tools which will be used in the later part of the this paper.
2.1. Functional spaces
We start with the functional spaces we use in this paper.
Definition 2.1.
Let . The Fourier transform or the Fourier coefficient of a function on is defined by
(2.1) |
and the Fourier inversion is given as
(2.2) |
Definition 2.2.
Remark 2.1.
Notice that by the definition above, for , we have
Remark 2.2.
For simplicity, we take .
Lemma 2.1 (Sobolev embedding).
Let , . Then . Moreover,
(2.4) |
where is an absolute constant.
Proof.
By the Fourier inversion formula, we write . From
we conclude
as desired.
∎
Definition 2.3.
2.2. Hilbert transform and double layer potential
Next, we define the Hilbert transform and the double layer potential used in the analysis of water wave systems.
2.2.1. The Hilbert transform
Definition 2.4.
Given . Assume that is periodic and satisfies
(2.6) |
where are constants. The Hilbert transform associates with the curve is defined as
(2.7) |
We define to be the Hilbert transform associated with , that is,
(2.8) |
Lemma 2.2.
It is well-known (see, for example, the celebrated paper by Guy David [25, Theorem 6]) that if satisfies (2.6), then is bounded on .
Lemma 2.3.
We have the following bounds for Hilbert transforms.
- (1)
-
(2)
Assume in addition that , then
(2.11) where depends on .
-
(3)
Assume that and there exists constants () such that
(2.12) Then we have
(2.13)
2.2.2. Double layer potentials and its adjoint
We define the double layer potential operator as follows.
Definition 2.5 (Double layer potential).
Suppose satisfies (2.6) and is periodic. Let be the curve parametrized by , and be the region in bounded above by . Let be the outward normal of . The double layer potential operator is defined by, for ,
(2.14) |
The adjoint of the double layer potential is defined as
(2.15) |
By Lemma 2.3, , is well-defined as an function. Similarly, for , , so is also well-defined as an function. Moreover, we have the following celebrated results due to Verchota [67]. See also [19], [64].
Lemma 2.4.
Let be a Jordan curve parametrized by such that is periodic and satisfies (2.6). Then and their adjoints are invertible, with
(2.16) |
and
(2.17) |
for some constant depending only on and .
2.2.3. Some relevant notations
Throughout this paper, we parametrize the interface in modified Lagrangian coordinates by . So we will frequently use the notation , , .
2.3. Identities
Here we collect some commutator identities which are frequently used later on.
Lemma 2.5.
Let be fixed. Assume that . We have
(2.18) |
(2.19) |
(2.20) |
(2.21) |
(2.22) |
(2.23) |
2.4. Expansion of
In this subsection, we formally derive the expansion of the Hilbert transform associated with a curve of small amplitude. This expansion will one of the basic tools for us to derive the asymptotic expansion of solutions.
Consider a curve of small amplitude. Formally, it has the expansion
(2.24) |
We rewrite as
(2.25) |
Using the expansion (2.24), we have
(2.26) |
By a simple Taylor expansion,
(2.27) |
we obtain (with and )
From the the explicit computations above, we regroup everything with respect to the power of and get
(2.28) |
where
(2.29) |
(2.30) |
(2.31) |
2.5. Commutator estimates
For the following commutator estimates, we refer the interested reader to Propositions 3.2, 3.3 in [74] and Theorem 2.1, Proposition 2.3 in [66] for the versions on the whole real line.
Given such that , , we define
(2.32) |
(2.33) |
For the commutators above, we have the following estimates.
Proposition 2.1.
Assume and () satisfy the conditions
(2.34) |
(2.35) |
Then one has
(2.36) |
(2.37) |
where the constant depends on , and or , or . Moreover, only one of the and norms takes the norm ( if takes the norm and if takes the norm).
Next, we estimate the differences of commutators produced by different curves. These will be helpful when we analyze the differences of different solutions.
Proposition 2.2.
Let satisfy
(2.38) |
for some constants , , . Then we have
(2.39) |
(2.40) |
Furthermore, one has
(2.41) |
for some constant depends on , , , . Moreover, only one of the and norms takes the norm ( if takes the norm and if takes the norm).
Using the same idea, we can also prove the following estimate for the differences of Hilbert transforms assoiated to different curves.
Proposition 2.3.
Suppose we have four curves satisfying
(2.42) |
for some constants , , . Then
(2.43) |
where the norm is either or . Moreover, only one of these -norms takes the norm.
By construction, the commutator defined by (2.32) can be regarded as a trilinear form in terms of the triple . The following estimate for the differences of commutators produced by different triples are useful in our analysis.
Proposition 2.4.
With notations above and the same assumption as Proposition 2.2, we have the following estimate
(2.44) |
3. Local wellposedness, cubic structure
In this section, we first write down the equations for the water waves system in Wu’s coordinates directly and derive explicit formulae for important quantities. Then we record the basic results on the local well-posedness of the water waves system. Finally, we derive a cubic structure for studying the long time existence of the water waves. The derivations of these formulae and the cubic structure were first used by Wu in [74] in the Euclidean setting.
3.1. Water waves in Wu’s coordinates
As discussed in §1.4.1, we formulate the water waves by the following
(3.1) |
where
(3.2) |
for some function .
3.2. Formulae for important quantities
To obtain a closed system in (3.1), we need to derive formulae for and in terms of the unknown .
3.2.1. Formula for
3.2.2. Formula for
From the momentum equation, the first equation in (3.1), we have
(3.7) |
Since is holomorphic, it follows that
(3.8) |
where is a holomorphic function in satisfies as .
3.2.3. Formula for quantities of the form
Next we derive a formula for quantities of the form . In order to simplify the calculation of the commutators, we consider a change of variables defined by
(3.10) |
where is given by (3.6). Set . We also need the quantity which is defined by . Then solves
(3.11) |
Composing with yields . So we obtain
(3.12) |
Going back to the original coordinate by composing on both sides of (3.12), one has
(3.13) |
3.2.4. Formula for
3.3. Local wellposedness
By formulae (3.9) and (3.6), (3.1) is a closed fully nonlinear system. One way to achieve the well-posedness is to quasilinearize the system by differentiating it with respect to . We only state the local well-posedness result here and refer [72] for the proof.
Theorem 3.1 (Local well-posedness).
Let . Given with , there is depending on such that the water waves system (1.6) with initial data has a unique solution for , satisfying
(3.15) |
Moreover, if is the supremum over all such times , then either , or , but
(3.16) |
or
(3.17) |
Remark 3.2.
Throughout this paper, we fix a satisfying the condition in the theorem above.
3.4. Cubic structure
The key advantage to use Wu’s coordinate is that it is well-suited to study the long-time existence of the water wave system. To obtain the long-time existence of , we need to derive a cubic structure for the system (3.1).
Setting . Since , then one has
(3.18) |
Then we have
(3.19) |
where and are cubic or higher-power nonlinearities.
4. Stokes waves in Wu’s coordinates
In this section, we study Stokes waves in Wu’s coordinates. We will first write the equations for the Stokes waves and formulae for some important quantities as last section. The goal here is to introduce some notations specific to Stokes waves. Then we will show the existence of small-amplitude Stokes waves in this coordinate and give the asymptotic expansions of them.
4.1. Equations for Stokes waves
We denote a given Stokes wave as . It is a special solution to (3.1). As in Section 3, we denote
where is given by
(4.1) |
Let be a real valued function given by
(4.2) |
Then by (3.1), we have
(4.3) |
The same as the general setting in Section 3 we have the cubic structure for the Stokes wave . Following (3.19), the quantity
(4.4) |
satisfies the cubic equation:
(4.5) |
4.2. Existence of Stokes waves of small amplitude
In this subsection we prove Proposition 1.1. Our proof is based on the existence and uniqueness of Stokes waves in Eulerian coordinates. To begin with, consider the fluid domain
(4.6) |
and . Let be the velocity field as in (1.1). Denote
(4.7) |
One has the following result on the existence of Stokes waves.
Proposition 4.1.
There exists a curve of smooth solutions to (1.1) parametrized by a small parameter which we call the amplitude of . For each solution on this curve, one can write it as
where and satisfy the following properties:
-
(i)
and are periodic smooth functions.
-
(ii)
is even, is odd and is even.
Other than the trivial solutions (with ), the curve is unique. Moreover, for any given , the following estimates hold
(4.8) |
for some constant depending on only.
Proposition 4.1 has been known for a century, see [50] and [44]. Also see Theorem 2.1 in [51] for the version in the Zakharov-Craig-Sulem formulation.
Remark 4.1.
We shall also use to represent the curve of solutions constructed in Proposition 4.1.
Using Proposition 4.1, we obtain the existence of Stokes waves in the Lagrangian formulation.
Proposition 4.2.
Let be a given Stokes wave of amplitude as given in Proposition 4.1. There exists a unique odd smooth function , such that if we denote
(4.9) |
(4.10) |
then
(4.11) |
Proof.
We prove the existence. The uniqueness follows easily. Suppose we have constructed , since is odd, we have . Hence is determined by .
Differentiating the expression of with respect to , one has
(4.12) |
We want to find to satisfy
(4.13) |
Let . Then (4.13) can be written as
(4.14) |
Integrating both sides of (4.14) yields
(4.15) |
From the expression above, to find is equivalent to obtain a fixed point. We use the standard iteration to find the fixed point.
Set
(4.16) |
Define
(4.17) |
Assume has been defined, then we can define
(4.18) |
Given , we then define
(4.19) |
It is straightforward to check that by construction, the smoothness of and implies that is smooth. Moreover, for each ,
(4.20) |
where is a constant depending on only. For sufficiently small such that , the standard Banach fixed point theorem gives . Using the smoothness of and the estimates (4.20), one obtains that is also smooth. Finally define
(4.21) |
By construction, we have as desired. ∎
Corollary 4.1.
Proof.
The restriction of on can be written as . Since , we have is in the direction normal to . So there is a real-valued function such that . We are done. ∎
With preparations above, we obtain the existence of small-amplitude Stokes waves in Wu’s coordinates.
Theorem 4.2.
There exists such that for all , there is a unique solution to the system (1.6) such that
-
(A)
and for some and some periodic smooth functions and .
-
(B)
and are odd, and are even.
-
(C)
, and .
-
(D)
For each , one has
(4.23) for some constant depending only on .
Proof.
Given a diffeomorphism , we denote . Define by
(4.24) |
where is given as in (3.6), and define by . Then any solution to (4.22) gives to a solution to
(4.25) |
In particular, the given Stokes wave from Proposition 4.2 gives a solution to (4.25). It is straightforward to check (A)-(B)-(C)-(D) of the proposition. We are done. ∎
As a direct consequence of Theorem 4.2, we obtain the following.
4.3. Asymptotic expansion of the Stokes waves in Wu’s coordinates
Finally, we present the asymptotic expansion of the Stokes wave. It is worthwhile to point out that in up to , the expansion only has one non-trivial frequency.
Proposition 4.3.
Let be a Stokes wave of amplitude and period . Then we have
(4.27) |
and
(4.28) |
In the remaining part of this subsection, we give detailed computations to show the proposition above.
Given a Stokes wave of amplitude with velocity , we assume that the Stokes wave has an expansion of the form
(4.29) |
where we can further write
(4.30) |
with some -periodic function .
For the velocity , we assume that it has the expansion
(4.31) |
Our goal now is to compute and .
4.3.1. Computations for
First of all, expanding the momentum equation, the first equation of (4.3), in terms of powers of , at the level of , we obtain
(4.32) |
Then in terms of , in the leading order, one has
(4.33) |
Expanding using its Fourier series on , we get
(4.34) |
where is the th Fourier coefficient of .
Writing (4.33) in terms of the Fourier series above, we notice that it only allows one non-trivial Fourier mode. Since we consider the Stokes wave with fundamental period , we take . Then (4.33) implies
(4.35) |
To determine , we recall that from Theorem 4.2, is odd and is even, so has to be purely imaginary. By a Stokes wave of amplitude , we really mean . So . Without loss of generality, let’s take
(4.36) |
Next, we expand and as
(4.37) |
and
(4.38) |
where we used fact that and are at least quadratic in from formulae (4.2) and (4.1).
We need to compute and in order to find .
Proposition 4.4.
Given notations above, we have
(4.39) |
Proof.
Expanding (4.1) in terms of powers of , at the level, we compute
(4.40) | ||||
(4.41) | ||||
(4.42) | ||||
(4.43) | ||||
(4.44) |
So we get Using the same calculations, we obtain . ∎
4.3.2. Computations for
At the level, the holomorphic condition from the equation (4.3) implies
(4.45) | ||||
(4.46) | ||||
(4.47) |
Therefore, we can conclude that
(4.48) |
Note that such choice guarantees the terms of which is also part of (4.3).
From the equation (4.3), and Proposition 4.4, we know
Expanding the equations above in terms of powers of and using (4.29), we have
(4.49) |
Plugging (4.48) back into the equation (4.49), we obtain that
(4.50) |
To find , we will analyze the expansion of the equation (4.3) in the level. After choosing and , in order to find , we need to compute , and which are the levels of , and . We have the following conclusion.
Proposition 4.5.
Using notations above, we have
Proof.
The results for and follow from direct inspection.
4.3.3. Computations for
The constraint implies
(4.56) | ||||
(4.57) |
Then we compute each term on the right-hand side of the equation above.
First of all, note that and the choice of implies
Next, by the explicit formula of (4.36), we have
(4.58) |
Finally, invoking the explicit formula for again, one has
(4.59) | ||||
(4.60) | ||||
(4.61) | ||||
(4.62) |
Putting things together, we obtain
(4.63) |
Therefore we can take
(4.64) |
To find , we notice that from (4.36), it follows
(4.65) |
Using (4.5), and , at the leading level, we have
(4.66) | ||||
(4.67) | ||||
(4.68) |
Therefore, one has
(4.69) |
Plugging (4.64) into the equation (4.69), duo the holomorphicity, we obtain
(4.70) |
From all computations above, we conclude
(4.71) |
and
(4.72) |
We finish the proof of Proposition 4.3.
4.3.4. Approximation of the Stokes wave
Using the obtained above, we are able to defined an approximate form of the Stokes wave.
Define the approximation of the Stokes wave as
(4.73) |
Then by construction, we know that
Here we note that in only one non-trivial fundamental frequency is involved. This fact will be crucial in the analysis.
5. Multiscale analysis and derivation of the NLS from the full water waves
In this section, we shall use the water waves system (3.1) to perform the multiscale analysis and derive the NLS.
5.1. Basic setting
Given a Stoke wave of amplitude from Proposition 1.1 and Section §4, by Proposition 4.3, it has the following expansion in terms of :
(5.1) | ||||
Consider the solution to the system (3.1) as a perturbation of the Stokes wave above in . Expanding the solution in terms of the power of , we seek for of the form
(5.2) |
where satisfies the ansatz
(5.3) |
for some periodic function on wtih from the view of the modulational approximation and the long-wave perturbation.
Since is the perturbation of , from the coefficient of in (5.1), we will choose
Eventually, we will show that in order make the expansion (5.2) valid, will admit the form with , and solve
In the remaining part of this section, we first give some estimates for almost holmorphicity of wavepackets which will be useful to handle functions with slow variables. With these preparations, we then analyze the expansion for , from (3.1) and find , , in (5.2).
5.2. Almost holomorphicity of wavepackets
Let be an integer, we know that is the boundary value of the holomorphic function in the lower half plane. If , then as which implies that . In general, if , can be comparable with . However, given , if , then is as small as . 222Recall that .
Lemma 5.1.
Let be given. Let and . Then
(5.4) |
where depends on only.
Proof.
We consider the case that . By the Fourier series on , we write
(5.5) |
where
(5.6) |
Therefore, one has
(5.7) |
By Parseval’s identity, it follows
(5.8) |
Notice that for , we have . Therefore, we get
as desired. Here in the last step, we used the estimate
(5.9) |
for . We are done. ∎
5.3. Multi-scale expansion
In this subsection, we compute the multi-scale expansion. The general strategy here is similar to [66, 62] (§3 of [66], §4 of [62]). We note that the nonlinear Schrödinger equation has not been derived in the setting with Stokes wave. Also notice that under the current setting, in the leading order , can depend on , which is different from previous works, for example, [66, 62].
Our goal is to choose such that
and 333See Proposition 4.3.
(5.10) | ||||
and the followings hold
-
(1)
solves (3.19).
-
(2)
.
-
(3)
.
-
(4)
.
By Corollary A.1, and imply
(5.11) |
In this section, we will use the notation .
5.3.1. Level
5.3.2. Level
Using the system (3.1) again, we need
(5.13) |
Note that by (2.30) and the explicit choice of (5.3), one has
where in the last step we applied Lemma 5.1. To avoid secular terms, we choose such that
(5.14) |
Using Lemma 5.1 again, we have . So plugging the ansatz, (5.3), into (5.14), it follows
So we choose , with
(5.15) |
To choose , we use . The terms give
We pick
(5.16) |
where .
5.3.3. Expansion of
We expand as
(5.17) |
Since is quadratic, we have . For , one has
(5.18) |
Since is real, we get
(5.19) |
5.3.4. Expansion of
We need also to expand .
Since is quadratic by the formula (3.9), clearly, , and .
5.3.5. Expansions of and
5.3.6. Level
We first note that
(5.21) |
At the level of , we have . Now we expand the momentum equation, the first equation of (3.1), in term of powers of as before. At the level of , we have
(5.22) |
Noticing that is slowly varying, one has
(5.23) |
By our choice of , (5.16), and applying (3) of Lemma 2.2, we have,
(5.24) |
Applying (3) of Lemma 2.2 again, we obtain
(5.25) |
Since is slowly varying in , we obtain
(5.26) |
Using Lemma 5.1, we have
So we obtain
Also we have
(5.27) | ||||
(5.28) | ||||
(5.29) |
Overall, from all computations above, we obtain
(5.30) |
Since , we have
(5.31) |
To avoid the secular growth, we choose such that
(5.32) |
or equivalently,
(5.33) |
Remark 5.1.
Note that is an exact solution to (5.33), which justifies our assumption that
(5.34) |
in the long-wave perturbation setting.
For , since slowly varying, using Lemma 9.4,
(5.36) |
For , one has
(5.37) |
For , it is easy to obtain
(5.38) |
For , we have
(5.39) | ||||
(5.40) | ||||
(5.41) | ||||
(5.42) | ||||
(5.43) |
Therefore we conclude that
(5.44) |
Now we can choose
(5.45) |
5.4. The approximate solution
5.5. NLS estimates
In the final part of this section, we discuss the behavior of the function coming from the expansion (5.2) and (5.3). From our multi-scale analysis, solves
(5.49) |
and the Stokes wave gives a special solution . We perturb the special solution by considering solution of the form . We have the following result on its instability.
Proposition 5.1.
6. The error equation
In this section, we derive governing equations for the remainder term. Let’s denote
(6.1) |
Then
(6.2) |
Define the error term as
(6.3) |
6.1. Notations
We first introduce some notations here.
Denote
(6.4) |
(6.5) |
(6.6) |
(6.7) |
(6.8) |
(6.9) |
(6.10) |
With notations above, we immediately conclude the following:
Lemma 6.1.
We have
(6.11) |
6.2. Governing equation for
From the cubic structure for , (3.19), we have
(6.12) |
And similarly, one has
(6.13) |
Consider the quantity defined by
(6.14) |
Then is holomorphic in . In Lemma 7.5, we will show that is equivalent to .
To control and therefore control , we need to derive a nice structure for . To achieve the nonlinear instability, it is necessary to obtain the control of for , where . So we need to derive an equation for of the form
(6.15) |
such that
(6.16) |
By the definition of , using the notation above, we have
We regroup as
To sum up, we obtain
(6.17) |
where
(6.18) |
(6.19) |
(6.20) |
(6.21) |
6.3. Equation governing
We next derive an equation for . Define
(6.22) |
Applying and using the equation for obtained above, we have
6.4. Energy functional
Lemma 6.2 (Basic lemma, Lemma 4.1 in [74]).
Let satisfy the equation
Suppose that for some . Define
(6.23) |
Then
(6.24) |
Moreover, if is the boundary value of a holomorphic function in , then
(6.25) |
Notations: Denote
(6.26) |
Because and are not necessarily holomorphic in , we decompose them as
(6.27) |
and define
(6.28) |
6.5. Evolution of and
To show that remains small (in the sense of some appropriate norm), we need to show that the energy remains small for a long time. To achieve this goal, we analyze the evolution of and . Note that
(6.31) |
Similarly, we derive the governing equation for . By direct computations, one has
(6.32) |
7. Preparations for energy estimates
In this section, we estimate the quantities which will be used in the energy estimates in the next section. We bound these quantities in terms of an auxiliary quantity , which is essentially equivalent to the energy .
7.1. An auxiliary quantity for the energy functional and a priori assumptions
The energy functional is not very convenient in the energy estimates, so we introduce the quantity
(7.1) |
Let . We make the following a priori assumptions.
-
1.
(Bootstrap assumption 1)
(7.2) -
2.
(Assumption 2) We assume satisfies
(7.3)
Here, is a constant depending on only.
Remark 7.1.
We will control in terms of and . Then we can obtain energy estimates on a lifespan of length . For this purpose, we control the quantities appear in the energy estimates in terms of and .
Convention. In this and the next sections, if not specified, we assume
and the bootstrap assumption (7.2) holds. Here is an arbitrary given number and is fixed, independent of and . These and are given as in Proposition 5.1. If not specified, is a constant depending on only.
7.1.1. Consequences of the a priori assumptions
Lemma 7.1.
Assume (7.3), then we have
(7.4) |
Proof.
This is a direct consequence of the definitions of and and the assumption (7.3). ∎
Lemma 7.2.
Assuming the bootstrap assumption (7.2), we have
-
1.
(7.5) -
2.
(7.6) -
3.
(7.7) -
4.
(7.8)
Proof.
Remark 7.2.
Using (7.69), one has
So the bootstrap assumption can be easily justified once we establish the estimate .
Lemma 7.3.
Lemma 7.3 is the direct consequence of Lemma 2.4 and the bootstrap assumption (7.2). Alternatively, we can also prove this lemma as follows.
Proof.
Lemma 7.4.
Proof.
We consider the case only, the case for follows in a similar manner. Since is real, taking the real parts on both sides of , we obtain
(7.20) |
By Lemma 7.3, we get
(7.21) |
as desired. ∎
7.1.2. The equivalence of and
Lemma 7.5.
The proof is similar to that of Lemma 8.3 in [62]. However, we need an additional gain of the factor . For the sake of completeness, we provide the proof in Appendix C.
Corollary 7.1.
Assuming the a priori assumption (7.2), we have
(7.25) |
7.2. Bound , , , , , and
7.2.1. Estimate
7.2.2. Estimate
7.2.3. Estimate
7.2.4. Estimate
7.2.5. Estimate and
Recall that
Also,
(7.34) |
Taking the difference of two expressions above, after some simple manipulations, we get
(7.35) |
By (3) of Lemma 2.3, (7.6), and (7.27), we have
(7.36) |
Applying (2.40) of Proposition 2.2 with , , , , , and , together with (7.2) and (7.3), one has
From computations above, we conclude that
which implies
(7.37) |
As a consequence, we obtain
(7.38) |
(7.39) |
(7.40) |
7.2.6. Estimate
Note that by the explicit formula of and ,
(7.41) |
where
(7.42) |
Applying Proposition 2.4 with , , , , , , , , , , , , from Proposition 2.2 and Proposition 2.4 , Proposition 2.3, Lemma 7.2, it follows
Therefore, we obtain the following results:
(7.43) |
(7.44) |
7.3. Bound , , , , ,
7.3.1. Estimate
Since , , and
we directly obtain
(7.45) |
and
(7.46) |
7.3.2. Estimate , and
We simply have
(7.47) |
due to . Since
(7.48) |
and , we have (noting that )
(7.49) |
Therefore,
(7.50) |
Corollary 7.2.
Assuming the bootstrap assumption (7.2), we have
(7.51) |
Proof.
Writing , one has
Using
(7.52) |
one obtains
(7.53) |
By the similar algebraic manipulations, one has
(7.54) |
So we conclude the proof of the corollary. ∎
7.3.3. Estimate
Applying on both sides of we obtain (for the derivation in the Euclidean setting, see Proposition 2.7 of [74].)
(7.55) |
Similarly, we have
(7.56) |
Taking the difference of two expressions above, we obtain
Similarly,
where
(7.57) |
Applying (3) of Lemma 2.3, one has
(7.58) |
Similarly,
(7.59) |
Using the same argument as for estimating , together with the estimates (7.7), (7.37), one has
(7.60) |
Similarly,
(7.61) |
Taking the differences and using Proposition 2.1, we obtain
(7.62) |
Similarly,
(7.63) |
Applying Proposition 2.4, Lemma 7.2, (7.37), we have
(7.64) |
By Proposition 2.4 and using Corollary 7.2, we obtain
(7.65) |
So we obtain
which implies
Here, we used the fact
Therefore, we can use Lemma 7.4 to conclude
(7.66) |
Since , and
(7.67) |
(7.66) implies
(7.68) |
Since , we also conclude
(7.69) |
Using the same argument, we can also conclude
(7.70) |
7.4. Bounds for
First of all, by construction, one has . Using the same arguments as for , , respectively, we obtain
(7.71) |
which also implies
(7.72) |
by the Sobolev embedding. As a consequence, for ,
(7.73) |
(7.74) |
and
(7.75) |
Using the same argument as for , we can conclude that
(7.76) |
which also implies
(7.77) |
by the Sobolev embedding.
7.5. Estimate sums of the form
Define
(7.78) |
The expression defined above can be regarded as an alternating sums of quadrilinear form in terms of . In the remaining part, we estimates the differences of associated with different quadruples.
In the view of our bootstrap assumptions, we assume that
-
(H1)
For all , ,
(7.79) -
(H2)
For ,
(7.80) -
(H3)
(7.81)
Our goal is to estimate .
Lemma 7.6.
Under the assumptions (H1)-(H2)-(H3), we have
(7.82) |
7.6. Estimate
We can write , where
(7.83) |
and
(7.84) |
The terms , , and are defined similarly. Applying Lemma 7.6 with
(7.85) |
(7.86) |
(7.87) |
(7.88) |
one has
We rewrite as
(7.89) |
Note that
(7.90) |
Denoting , we have
(7.91) |
With the computations above, can be written as
(7.92) |
Noticing that
(7.93) |
from which we see that is a cubic term. Moreover, using the same argument as for , we obtain
So we conclude that
(7.94) |
7.7. Estimate
Recall that , and . Taking the difference, we have
By a direct computation we can write
From computations from previous sections, we have the following:
(7.95) |
(7.96) |
(7.97) |
(7.98) |
(7.99) |
Therefore, one has
(7.100) |
(7.101) |
and
(7.102) |
Hence, we conclude
(7.103) |
7.8. Estimate
We can write
Taking the difference, we have
Using the similar argument to that for , we obtain
(7.104) |
7.9. Estimate
Taking the differences, one has
and
We estimate each term on the right-hand side of the expression above.
7.10. Estimate
Recall that
(7.112) |
Recall that . We first estimate . The estimate for follows from the similar yet even easier calculations.
Using (7.89), we have
Let be the diffeomorphism be defined by and let . Composing444The advantage of the original coordinate is that commutes with in this coordinate. with the diffeomorphism , one has
Using
we obtain
(7.113) |
Changing of variables, we get
Since
(7.114) |
so is essentially of the same type of in (7.78). Using the same argument for , we obtain
(7.115) |
We can write
and
Then we can bound
Using
(7.116) |
(7.117) |
and
(7.118) |
we conclude that
(7.119) |
7.11. Estimate
7.12. Estimate , , and
7.13. Estimate
Recall that . We decompose
By Proposition 2.1 and the estimates for :
(7.129) |
we obtain
(7.130) |
For and , we have
For , by (2.23), we obtain
For , using Proposition 2.1, one has
(7.131) |
For , we decompose as
(7.132) |
Then we can write
(7.133) |
For , using555Note that is the boundary value of a bounded holomorphic function in , so is also the boundary value of a bounded holomorphic function in . So is the boundary value of a bounded holomorphic function in which approaches zero as .
(7.134) |
we have
Then we can use arguments as for in §7.6 to conclude that
(7.135) |
For , we decompose it as
For the first term, note that by Corollary A.1,
where
(7.136) |
For and , it’s easy to obtain
(7.137) |
From the computations above, it follows
(7.138) |
So we obtain
(7.139) |
7.14. Summary of the estimates
8. Energy estimates
The goal of this section is to obtain the following energy estimates.
Proposition 8.1.
Recall that
(8.2) |
where
(8.3) |
and
(8.4) |
From the computations in previous section, we have obtained the estimates for . To close the energy estimates for , we still need to obtain the estimates for , and .
Recall also that
(8.5) |
where
(8.6) |
8.1. Estimate and
8.2. Estimate
8.3. Estimates , , and
8.3.1. Estimate and
Recall that . We have for ,
Here, we used the fact that for some constant , and therefore666Here and after, we use to denote constants depending on only.
(8.17) |
So we obtain for ,
(8.18) |
and for , one has
Then one obtain
(8.19) |
For , we have
(8.20) |
8.3.2. Estimate
Recll that . For , we have
For , using and the fact , one has
(8.21) |
We rewrite and as
(8.22) |
and
(8.23) |
Therefore, we rewrite
For , using (8.19) and Proposition 2.1, we have
(8.24) |
For , one has
Using the same argument as for , we have
(8.25) |
For , we split it into two pieces
Note that since
(8.26) |
one has
(8.27) |
For , we further decompose it as
Clearly, using (8.19), we have
from which it follows
(8.28) |
Finally, we analyze . Note that , and is the boundary value of a holomorphic function in , so is the boundary value of . Notice that as . Therefore,
is the boundary value of a holomorphic function , with
(8.29) |
Applying Cauchy’s theorem, one has
(8.30) |
Therefore we conclude that
(8.31) |
8.3.3. Estimate
Integration by parts, the estimate for is similar to that for . Hence we obtain
(8.32) |
8.3.4. Estimate
8.4. Estimate for
Putting all computations from previous subsections together, we conclude that
(8.34) |
8.5. Estimate
8.6. Estimate
By direct computations, we have
(8.38) |
8.7. Conclude the proof of Proposition 8.1
8.8. Equivalence of and
8.9. Control the growth of the error term
9. Modulational instability of the Stokes waves
In this final section, we prove the nonlinear modulational instability of the Stokes wave. To achieve this goal, the essential step is to use the energy estimates obtained from the previous section to establish the long-time existence of the reminder term.
9.1. Existence of initial data with desired properties
Notice that the approximate solution , (6.2), we obtained in Section 5 could not be taken as the initial data for the system (3.1) since it does not satisfy the holomorphic conditions in the equation.
We should first construct initial data to the water wave system (3.1) for long-wave perturbations of the Stokes wave .
Proposition 9.1.
Suppose that , and for some absolute constant . In addition, we assume that
(9.1) |
Then there exist and such that and satisfying
(9.2) |
(9.3) |
and
(9.4) |
(9.5) |
for some constant depending on and .
Proof.
The initial data here is obtained by iteration. We start by defining
Let be the Hilbert transform associated with . Define by
(9.6) |
Assuming that has been constructed, we define
(9.7) |
Since
(9.8) |
and by Lemma 5.1,
(9.9) |
one has
(9.10) |
It is straightforward to obtain
(9.11) |
Therefore, after applying the fixed point theorem, converges in to a function . Define as
(9.12) |
By the definition of and the iteration procedure above, there hold
(9.13) |
(9.14) |
One has
(9.15) |
By Lemma A.1, we know that
(9.16) |
where
Note that because . So one has
By (9.14), we have
(9.17) |
By the orthogonal condition (9.1), one has
(9.18) |
Define . Noting that , we have
Moreover,
(9.19) |
and
(9.20) |
Define as
(9.21) |
where
(9.22) |
By Corollary A.1, we have
(9.23) |
Note that we can rewrite
Since is holomorphic and vanishes as , we have
(9.24) |
and
(9.25) |
Therefore, by (9.19),
By Lemma 5.1, we can bound
Therefore, it follows that
Also, using , we split
For , we have
(9.26) |
For , by the orthogonality condition (9.1), one has
(9.27) |
By (9.4), we finally conclude that
We are done. ∎
9.2. Extended lifespan
Now we are ready to conclude the result on the long-time existence.
Given a solution to the NLS
(9.30) |
we define
(9.31) |
9.2.1. Long-time estimates
By Theorem 3.1, Proposition 8.2, Proposition 9.1, and the standard bootstrap argument, we obtain the following.
Theorem 9.2.
Let be given and . Let be a Stokes wave of period and amplitude . Let be an arbitrarily small but fixed number. For any given with , and any solution to the NLS (9.30) satisfying
(9.32) |
and the orthogonality condition (9.1), there exist and such that and they satisfy the estimate
(9.33) |
where is constructed as (9.31). For all such data , the water wave system (1.6) admits a unique solution on with
satisfying the following estimate
(9.34) |
for all where is a fixed number.
Proof.
For given , and satisfying (9.32), the existence of satisfying (9.33) is guaranteed by Proposition 9.1. By Proposition 5.1, there exists a fixed number uniform in and such that for all solution to the NLS (9.30) satisfying (9.32), one has
(9.35) |
So the existence of satisfying (7.3) is guaranteed.
Assume the bootstrap assumption (7.2) with constant . Clearly, the bootstrap assumption (7.2) is satisfied at . Using the a priori energy estimates provided in Proposition 8.2, the estimate (8.46), the constant appearing the bootstrap assumption is improved since . Therefore together with the blowup criterion (3.16) and (3.17), we are able to use the bootstrap argument to prove that the solution , and (9.34) holds. ∎
Remark 9.3.
Note that by construction, the initial data is a long-wave perturbation (with fundamental period ) of the Stokes wave .
Remark 9.4.
This theorem also shows the validity of the modulational approximation via NLS. It might be interesting to point out that due to the fact that the Stokes wave is a global solution to the water wave system, the valid time scale for the modulational approximation of the perturbed flow is longer than other settings. See for example [66, 62] where the valid time scale is of order .
9.3. Nonlinear instability
With the long-time existence and estimates, we now analyze the instability of Stokes waves under long-wave perturbations.
9.3.1. Growth of large scales
By Theorem 9.2, with estimates between the difference of and , it suffices to analyze the growth of . In this setting, from the proof of the NLS instability, Appendix §D, for , we can take the initial data
(9.36) |
where and are defined as
(9.37) |
and
with is given as Theorem 9.2. Note that by construction, (9.37), and . Clearly, satisfies the orthogonality condition (9.1).
Furthermore, with initial data above, the solution can be written as
(9.38) |
where and Then clearly by construction, the above satisfies the perturbation condition (9.32). Denoting
(9.39) |
then satisfies additional estimate: for
(9.40) |
Moreover, at , one has
(9.41) |
for some constant .
The instability mechanism above is precisely the deriving force of the instability of the Stokes wave.
9.3.2. The modulational instability
Let , and be given. We use to denote the Stokes wave with period , the amplitude , and the phase translation . The following result gives the nonlinear instability of Stokes wave.
Corollary 9.1.
In particular, we conclude that the Stokes wave given in Theorem 9.2 is modulationally unstable under the long-wave perturbation in .
Proof.
First of all, clearly, by construction, the solution given in Theorem 9.2 is a long-wave perturbation of the Stokes wave in the same theorem.
We first prove that
(9.43) |
Recalling that for a Stokes wave of period and amplitude and phase shift , we have the asymptotic expansion
(9.44) |
By estimate (9.33) from Theorem 9.2, to prove (9.43), it suffices to show
(9.45) |
From the explicit formula of , it suffices to prove
(9.46) |
From the leading order term of the expansion (9.44), we note that is orthogonal to the leading order term of . This orthogonality and (9.41) together imply (9.46) from which (9.43) follows.
Remark 9.6.
From our explicit construction , the instability here also holds pointwisely.
Remark 9.7.
Here, we just picked one special and (9.38). We should point out that there are plenty of choices to construct the unstable perturbations since each unstable solution from the scale of NLS can produce a corresponding unstable solution for the water wave system.
9.3.3. Nonlinear modulational instability in Eulerian coordinates
We take from Corollary 9.1. Then denote
(9.47) |
Since , defines a diffeomorphism. We can find the inverse of as satisfying
(9.48) |
In Eulearian coordinates, the elevation of the perturbed flow is given by
(9.49) |
and the elevation of Stokes waves is defined as
(9.50) |
Corollary 9.2.
With notations above, we have
(9.51) |
for some constant which is uniform in and .
Proof.
We again start with the instability in . We claim that
(9.52) |
Given the from Corollary 9.1 and a fixed Stokes wave , we first observe that from estimate (9.33) and the construction of , we know that when ,
On the other hand . Therefore, when or , then clearly, (9.52) holds.
It remains to consider . In this case we write
(9.53) |
whose leading order terms are given
(9.54) |
Applying to (9.54) and then we take the norm. The lower bound for the first term
(9.55) |
follows from (9.46) after applying the change of variable with (9.48).
For the second part,when , we have the upper bound
(9.56) |
from (9.48) and the similar one for . This is of higher order in .
Therefore, taking the leading order terms and the applying a simple triangle inequality, from (9.55) and (9.56), we get (9.52).
Finally, by the same argument as for Corollary 9.1, using the support of Fourier modes, the version (9.51) follows from (9.52). Indeed, we can write and as . Then we expand and in terms of powers of . At the level of , the Fourier modes of (9.54) are supported around . The remaining pieces are of higher orders of . We omit the details since it is routine. For the expansion of the Stokes wave, we also refer to the formula (1.16).
∎
Appendix A The Hilbert transform and the Cauchy integral
In this appendix, we provide some detailed analysis of the Hilbert transform and the Cauchy integral used in this paper. We start with some basic definitions.
Definition A.1.
Let and . We define as the cone
(A.1) |
Definition A.2.
Given a chord arc parametrized by such that is periodic. Define the Cauchy integral as
(A.2) |
where means integrating over a fundamental period of .
With preparations above, we have the following properties of the Cauchy integral.
Lemma A.1.
Let and be sufficient nice functions, and has endpoints . Assume that , and . Let be the region below . We have the following conclusions:
-
(1)
Let .Then one has that
(A.3) as nontangentially.
-
(2)
If , for some bounded holomorphic function in such that , then
(A.4) as nontangentially.
Proof.
Let be fixed. Given , we denote and . Again here, we abuse of notation that here means restricting onto a fundamental period of after parameterizing the curve. Let . For the integral over , by the continuity of , we have
(A.5) |
On the other hand, for the integral over , we split it into two pieces:
For , as in , we observe the followings:
-
1.
On , we have
(A.6) -
2.
.
-
3.
For , we have
(A.7) where and depend continuously on .
These facts imply
for some constant depending on and , but with no dependence on . Therefore as , (as ).
Next we analyze . Assume that has the starting point and the ending point . For , since , for (which is valid as ), we have
It is straightforward to verify that for any fixed,
(A.8) |
For , we observe that as , one has . Therefore
Hence we obtain
(A.9) |
Let , then (A.3) is proved.
To prove (A.4), suppose is the boundary value of the holomorphic function in . Take . Denote the left and right endpoints of by and , respectively. Taking , we set
Let be the segment from to , be the segment from to , and be the segments from to . Let be , oriented clockwisely. By Cauchy’s Theorem, since is the boundary value of in , we have
(A.10) |
Note that
(A.11) |
by the periodicity assumption and the orientation.
As , we have
Note that here for , the orientation is clockwise.
Notice that , one has . So for large,
By Cauchy’s Theorem again, we have
Putting everything together, one has
(A.12) |
Therefore we conclude that
(A.13) |
as desired. ∎
As a consequence of Lemma (A.1), we obtain the following conclusion.
Corollary A.1.
Let and be sufficient nice functions, and suppose that has endpoints . Let be the region below . Assume that , and .
-
(a)
is the boundary value of a periodic holomorphic function on , with as .
-
(b)
is the boundary value of a holomorphic function on satisfying for all and as if and only if
(A.14) where .
Proof.
For (a), by Lemma A.1, has the boundary value , and as .
For (b), in one direction, if is the boundary value of a bounded holomorphic function in , then by Lemma A.1, we have
(A.15) |
which implies
(A.16) |
On the other hand, if , then . Define by . The boundary value of is
Moreover, using that fact
we obtain
So as , and has the boundary value . ∎
Appendix B Identities
In this appendix, we provide the proof of Lemma 2.5 in details.
Proof of Lemma 2.5.
For (2.18), performing integration by parts, we rewrite the Hilbert transform as
(B.1) |
By direct computations, one has
Therefore, we can conclude that
which gives (2.18) and then (2.19) is an easy consequence of (2.18).
For (2.20), using (B.1), we get
by direct differentiation. So we obtain
which gives (2.20). And then (2.19) and (2.20) together implies (2.21).
Appendix C Supplementary proofs
In this section, we provide supplementary proofs of some lemmata used in the main part of the paper.
C.1. Proof of Lemma 7.5
Proof of Lemma 7.5.
Recall that
where
and
where by the explicit construction of and ,
(C.1) |
Taking the difference, it follows
where
(C.2) |
By manipulating the differences, it is easy to obtain
(C.3) |
Combing computations and estimates above, one has
(C.4) |
Finally by the same argument we obtain
(C.5) |
and
(C.6) |
We are done. ∎
C.2. Proof of Proposition 2.2
Proof.
To suppress notations, we write as .
Using elementary trigonometric identities
we can write the difference as
By (2.37), we have
(C.7) |
Next we analyze (2.41). We first introduce the notation:
(C.8) |
Then we perform some elementary computations. Using trigonometric identities again, we have
where we define
(C.9) |
(C.10) |
We write as
(C.11) |
and similarly
(C.12) |
Taking the difference of to expressions above, we have
We further decompose as
(C.13) |
where
(C.14) |
and
(C.15) |
Now we can write the difference of commutators from (2.41) as
(C.16) | ||||
(C.17) |
By the explicit form of , , and and (2.37), we have
(C.18) |
Secondly, one can estimate
(C.19) | ||||
(C.20) | ||||
(C.21) |
Finally, we have
(C.22) | ||||
(C.23) | ||||
Putting three estimates above together, we obtain the desired result. ∎
C.3. Proof of Proposition 2.4
Proof.
The proof of this proposition is purely algebraic. Firstly, we regroup the expression we are interested as
Then we rewrite as
We can also rewrite as
So we can write
We rewrite as
Next, can be rewritten as
We expand as
Finally, we notice that all and can be estimated separately. These estimates give the desired result. ∎
C.4. Proof of Lemma 7.6
Proof.
We first regroup the expression we are interested in as
(C.24) |
Note that we can estimate by the same way as .
Explicitly, We have
(C.25) |
We still need to compute
and explore the cancellations.
Note that
Moreover, we have
So we can conclude that
(C.26) |
Appendix D Instability of the NLS
In this appendix, we provide the details on the instability of the Stokes wave in the setting of the NLS problem. Here to abuse the notation, we consider the NLS on with . Recall that in the main body of the article, the function which solves the NLS is defined on .
D.1. Basic setting
Consider the focusing cubic NLS in one dimension
(D.1) |
We are interested in the instability of the special solution given by the Stokes wave
(D.2) |
Consider the perturbation of the form
(D.3) |
where .
D.2. First order system
Next we analyze the linear part of the equation (D.4):
(D.7) |
Working on the circle , we can also rewrite (D.7) as a first order system
(D.8) |
where and are given as
(D.9) |
Denoting , the equation (D.4) can be rewritten as
(D.10) |
where
(D.11) |
and
(D.12) |
where and are defined by
By the formalism above, our problem is written as a canonical Hamiltonian system. Using the Duhamel formula, we can write the solution to (D.10) as
(D.13) |
Recall that for a function on , the Fourier series of is given by
(D.14) |
where
For a nice function , we define the Fourier multiplier as
(D.15) |
With explicit formulae above, we can compute explicitly using the Fourier series.
Lemma D.1.
Given notations above, the evolution operator can be written as the following:
For , one has
and for , we have
Remark D.1.
The problem in the full line was computed in Muñoz [48] using a slightly different formalism.
Proof.
Consider the linear problem
Using the notation (D.8), we can rewrite the problem as
(D.16) |
with
From the equation (D.16), we have
which implies
(D.17) |
and
(D.18) |
To solve (D.17) and (D.18) using the Fourier series on now are standard. We only illustrate the idea by solving (D.17). Expanding by the Fourier series as (D.14), we obtain that
(D.19) |
Solving the ODE above, we conclude that
(D.20) |
After solving the problem for in a similar manner, we obtain that for
(D.21) |
and for
(D.22) |
Using the multiplier notation (D.15), we conclude the desired results. ∎
From the explicit computations in Lemma D.1, we can read off the growth rate of the linear flow directly.
Corollary D.1.
Consider the first order system
Define
(D.23) |
Then for any , we have for
(D.24) |
Proof.
This follows from the explicit computations above. After taking the Fourier series, by a direct inspection, from formulae (D.21) and (D.22), when , the linear flow will result in the exponential growth. More precisely with notations from (D.21), we have
and
These Fourier coefficients have the exponential growth rate provided that
(D.25) |
The desired result follows from the Fourier representation of the solution . ∎
D.3. Nonlinear problem
Given and fixed, we consider the nonlinear equation
(D.26) |
on with initial data given in the complex form by
(D.27) |
where is defined as
and , . It is straightforward to check that (D.27) satisfies
and the condition (D.25) for .
Recall that by construction, consists of quadratic and cubic terms in .
Theorem D.2.
Proof.
By the Duhamel formula, one has
(D.30) |
(D.31) |
Then clearly by construction, it follows that
(D.32) |
Define as
(D.33) |
and as
(D.34) |
where is to be determined later.
Note that if is small. Otherwise, by contradiction, we assume that . Taking the Sobolev norms of both side of (D.30) and applying Corollary D.1, we have
(D.35) |
where we used with is an algebra in . When is small, clearly, the estimate above contradicts the definition of . Therefore indeed, we have .
Evaluating at , we have
(D.36) |
provided that is small. In particular, we know that
(D.37) |
as desired. ∎
Remark D.3.
The instability argument above holds for all initial data such that the corresponding linear flow satisfies (D.32).
Appendix E List of notations
The free boundary at time | |
The fluid region at time | |
The pressure at the position at time | |
The velocity field of the water waves at the position at time | |
The labeling of the free interface in Wu’s coordinates | |
The labeling of the free surface of a Stokes wave in Wu’s coordinates | |
The leading order of the amplitude of a given Stokes wave. | |
The wave speed of a given Stokes wave | |
A given solution to the NLS, | |
The Hilbert transform associated with a curve parametrized by | |
The double layer potential associated with a curve parametrized by | |
The adjoint of | |
A real-valued function associated with and | |
A real-valued function associated with and | |
A given positive number. |
, the wavelength of the perturbation of the water waves | |
---|---|
The standard torus | |
has the expansion | |
---|---|
The flat Hilbert transform on | |
, | |
, | |
References
- [1] Albert Ai. Low regularity solutions for gravity water waves. Water Waves, 1(1):145–215, 2019.
- [2] Albert Ai. Low regularity solutions for gravity water waves ii: The 2d case. Annals of PDE, 6:4, 2020.
- [3] Albert Ai, Mihaela Ifrim, and Daniel Tataru. Two dimensional gravity waves at low regularity i: Energy estimates. arXiv preprint arXiv:1910.05323, 2019.
- [4] Albert Ai, Mihaela Ifrim, and Daniel Tataru. Two dimensional gravity waves at low regularity ii: Global solutions. arXiv preprint arXiv:2009.11513, 2020.
- [5] Thomas Alazard, Nicolas Burq, and Claude Zuily. On the cauchy problem for gravity water waves. Inventiones mathematicae, 198(1):71–163, 2014.
- [6] Thomas Alazard and Jean-Marc Delort. Global solutions and asymptotic behavior for two dimensional gravity water waves. Ann. Sci. Éc. Norm. Supér.(4), 48(5):1149–1238, 2015.
- [7] D Ambrose and Nader Masmoudi. The zero surface tension limit two-dimensional water waves. Communications on pure and applied mathematics, 58(10):1287–1315, 2005.
- [8] J Thomas Beale, Thomas Y Hou, and John S Lowengrub. Growth rates for the linearized motion of fluid interfaces away from equilibrium. Communications on Pure and Applied Mathematics, 46(9):1269–1301, 1993.
- [9] T Brooke Benjamin and JE Feir. The disintegration of wave trains on deep water. J. Fluid mech, 27(3):417–430, 1967.
- [10] Thomas Brooke Benjamin. Instability of periodic wavetrains in nonlinear dispersive systems. Proceedings of the Royal Society of London. Series A. Mathematical and Physical Sciences, 299(1456):59–76, 1967.
- [11] DJ Benney and AC Newell. The propagation of nonlinear wave envelopes. Journal of mathematics and Physics, 46(1-4):133–139, 1967.
- [12] Massimiliano Berti, Roberto Feola, and Fabio Pusateri. Birkhoff normal form and long time existence for periodic gravity water waves. arXiv preprint arXiv:1810.11549, 2018.
- [13] Lydia Bieri, Shuang Miao, Sohrab Shahshahani, and Sijue Wu. On the motion of a self-gravitating incompressible fluid with free boundary. Communications in Mathematical Physics, 355(1):161–243, 2017.
- [14] Garrett Birkhoff. Helmholtz and taylor instability. In Proc. Symp. Appl. Math, volume 13, pages 55–76, 1962.
- [15] Thomas J Bridges and Alexander Mielke. A proof of the benjamin-feir instability. Archive for rational mechanics and analysis, 133(2):145–198, 1995.
- [16] Angel Castro, Diego Córboda, Charles Fefferman, Francisco Gancedo, and Javier Gómez-Serrano. Finite time singularities for the free boundary incompressible euler equations. Annals of Mathematics, pages 1061–1134, 2013.
- [17] Angel Castro, Diego Córdoba, Charles Fefferman, Francisco Gancedo, and Javier Gómez-Serrano. Finite time singularities for water waves with surface tension. Journal of Mathematical Physics, 53(11):115622, 2012.
- [18] Demetrios Christodoulou and Hans Lindblad. On the motion of the free surface of a liquid. Communications on Pure and Applied Mathematics, 53(12):1536–1602, 2000.
- [19] Ronald Raphaël Coifman, Alan McIntosh, and Yves Meyer. L’intégrale de cauchy définit un opérateur borné sur l2 pour les courbes lipschitziennes. Annals of Mathematics, pages 361–387, 1982.
- [20] Daniel Coutand and Steve Shkoller. Well-posedness of the free-surface incompressible euler equations with or without surface tension. Journal of the American Mathematical Society, 20(3):829–930, 2007.
- [21] Daniel Coutand and Steve Shkoller. On the finite-time splash and splat singularities for the 3-d free-surface euler equations. Communications in Mathematical Physics, 325(1):143–183, 2014.
- [22] Daniel Coutand and Steve Shkoller. On the impossibility of finite-time splash singularities for vortex sheets. Archive for Rational Mechanics and Analysis, 221(2):987–1033, 2016.
- [23] W. Craig. An existence theory for water waves and the boussinesq and korteweg-devries scaling limits. Comm. in P.D.E, 10(8):787–1003, 1985.
- [24] Walter Craig, Catherine Sulem, and Pierre-Louis Sulem. Nonlinear modulation of gravity waves: a rigorous approach. Nonlinearity, 5(2):497, 1992.
- [25] Guy David. Opérateurs intégraux singuliers sur certaines courbes du plan complexe. In Annales scientifiques de l’École Normale Supérieure, volume 17, pages 157–189, 1984.
- [26] Bernard Deconinck and Katie Oliveras. The instability of periodic surface gravity waves. J. Fluid Mech., 675:141–167, 2011.
- [27] Wolf-Patrick Düll, Guido Schneider, and C Eugene Wayne. Justification of the nonlinear schrödinger equation for the evolution of gravity driven 2d surface water waves in a canal of finite depth. Archive for Rational Mechanics and Analysis, 220(2):543–602, 2016.
- [28] David G Ebin. The equations of motion of a perfect fluid with free boundary are not well posed. Communications in Partial Differential Equations, 12(10):1175–1201, 1987.
- [29] Pierre Germain, Nader Masmoudi, and Jalal Shatah. Global solutions for the gravity water waves equation in dimension 3. Annals of Mathematics, pages 691–754, 2012.
- [30] Daniel Ginsberg. On the lifespan of three-dimensional gravity water waves with vorticity. arXiv preprint arXiv:1812.01583, 2018.
- [31] Hidenori Hasimoto and Hiroaki Ono. Nonlinear modulation of gravity waves. Journal of the Physical Society of Japan, 33(3):805–811, 1972.
- [32] John K Hunter, Mihaela Ifrim, and Daniel Tataru. Two dimensional water waves in holomorphic coordinates. Communications in Mathematical Physics, 346(2):483–552, 2016.
- [33] Vera Mikyoung Hur. The benjamin-feir instability in the infinite depth. arXiv preprint arXiv:2009.01593, 2020.
- [34] Mihaela Ifrim and Daniel Tataru. Two dimensional water waves in holomorphic coordinates ii: global solutions. arXiv preprint arXiv:1404.7583, 2014.
- [35] Mihaela Ifrim and Daniel Tataru. Two dimensional gravity water waves with constant vorticity: I. cubic lifespan. arXiv preprint arXiv:1510.07732, 2015.
- [36] Mihaela Ifrim and Daniel Tataru. The nls approximation for two dimensional deep gravity waves. Science China Mathematics, 62(6):1101–1120, 2019.
- [37] T Iguchi, N Tanaka, and A Tani. On a free boundary problem for an incompressible ideal fluid in two space dimensions. Adavances in Mathematical Sciences and Applications, 9:415–472, 1999.
- [38] Tatsuo Iguchi. Well-posedness of the initial value problem for capillary-gravity waves. Funkcialaj Ekvacioj Serio Internacia, 44(2):219–242, 2001.
- [39] Alexandru D. Ionescu and Fabio Pusateri. Global solutions for the gravity water waves system in 2d. Inventiones mathematicae, 199(3):653–804, 2015.
- [40] Jiayin Jin, Shasha Liao, and Zhiwu Lin. Nonlinear modulational instability of dispersive pde models. Archive for Rational Mechanics and Analysis, 231(3):1487–1530, 2019.
- [41] Yu P Krasovskii. The theory of steady-state waves of large amplitude. SPhD, 5:62, 1960.
- [42] Yu P Krasovskii. On the theory of steady-state waves of finite amplitude. USSR Computational Mathematics and Mathematical Physics, 1(4):996–1018, 1962.
- [43] David Lannes. Well-posedness of the water-waves equations. Journal of the American Mathematical Society, 18(3):605–654, 2005.
- [44] T. Levi-Civita. Determination rigoureuse des ondes permanentes d’ampleur finie. Math. Ann, 93(1):264–314, 1925.
- [45] MJ Lighthill. Contributions to the theory of waves in non-linear dispersive systems. In Hyperbolic Equations and Waves, pages 173–210. Springer, 1970.
- [46] Hans Lindblad. Well-posedness for the motion of an incompressible liquid with free surface boundary. Annals of mathematics, pages 109–194, 2005.
- [47] Shuang Miao, Sohrab Shahshahani, and Sijue Wu. Well-posedness for free boundary hard phase fluids with minkowski background. arXiv preprint arXiv:2003.02987, 2020.
- [48] Claudio Muñoz. Instability in nonlinear Schrödinger breathers. Proyecciones, 36(4):653–683, 2017.
- [49] V. I. Nalimov. The cauchy-poisson problem (in russian). Dynamika Splosh, Sredy(18):104–210, 1974.
- [50] Aleksandr I Nekrasov. On steady waves. Izv. Ivanovo-Voznesensk. Politekhn. In-ta, 3:52–65, 1921.
- [51] Huy Q Nguyen and Walter A Strauss. Proof of modulational instability of stokes waves in deep water. arXiv preprint arXiv:2007.05018, 2020.
- [52] Masao Ogawa and Atusi Tani. Free boundary problem for an incompressible ideal fluid with surface tension. Mathematical Models and Methods in Applied Sciences, 12(12):1725–1740, 2002.
- [53] Masao Ogawa and Atusi Tani. Incompressible perfect fluid motion with free boundary of finite depth. Advances in Mathematical Sciences and Applications, 13(1):201–223, 2003.
- [54] LA Ostrovskii. Propagation of wave packets and space-time self-focusing in a nonlinear medium. Sov. Phys. JETP, 24(4):797–800, 1967.
- [55] Sijue Wu Rafe Kinsey. A priori estimates for two-dimensional water waves with angled crests. preprint 2014, arXiv1406:7573.
- [56] Guido Schneider and C Eugene Wayne. Justification of the nls approximation for a quasilinear water wave model. Journal of differential equations, 251(2):238–269, 2011.
- [57] Jalal Shatah and Chongchun Zeng. Geometry and a priori estimates for free boundary problems of the euler’s equation. arXiv preprint math/0608428, 2006.
- [58] George G Stokes. On the theory of oscillatory waves. Transactions of the Cambridge philosophical society, 1880.
- [59] Dirk J Struik. Détermination rigoureuse des ondes irrotationelles périodiques dans un canal à profondeur finie. Mathematische Annalen, 95(1):595–634, 1926.
- [60] Qingtang Su. Long time behavior of 2d water waves with point vortices. Communications in Mathematical Physics, pages 1–94, 2020.
- [61] Qingtang Su. On the transition of the rayleigh-taylor instability in 2d water waves. arXiv preprint arXiv:2007.13849, 2020.
- [62] Qingtang Su. Partial justification of the peregrine soliton from the 2d full water waves. Archive for Rational Mechanics and Analysis, pages 1–97, 2020.
- [63] Geoffrey Ingram Taylor. The instability of liquid surfaces when accelerated in a direction perpendicular to their planes. i. Proc. R. Soc. Lond. A, 201(1065):192–196, 1950.
- [64] Michael E Taylor. Tools for PDE: pseudodifferential operators, paradifferential operators, and layer potentials. Number 81. American Mathematical Soc., 2007.
- [65] Nathan Totz. A justification of the modulation approximation to the 3d full water wave problem. Communications in Mathematical Physics, 335(1):369–443, 2015.
- [66] Nathan Totz and Sijue Wu. A rigorous justification of the modulation approximation to the 2d full water wave problem. Communications in Mathematical Physics, 310(3):817–883, 2012.
- [67] Gregory Verchota. Layer potentials and regularity for the dirichlet problem for laplace’s equation in lipschitz domains. Journal of Functional Analysis, 59(3):572–611, 1984.
- [68] Xuecheng Wang. Global infinite energy solutions for the 2d gravity water waves system. Communications on Pure and Applied Mathematics, 71(1):90–162, 2018.
- [69] GB Whitham. Non-linear dispersion of water waves. Journal of Fluid Mechanics, 27(2):399–412, 1967.
- [70] Sijue Wu. A blow-up criteria and the existence of 2d gravity water waves with angled crests. preprint 2015, arXiv:1502.05342.
- [71] Sijue Wu. On a class of self-similar 2d surface water waves. preprint 2012, arXiv1206:2208.
- [72] Sijue Wu. Well-posedness in sobolev spaces of the full water wave problem in 2-d. Inventiones mathematicae, 130(1):39–72, 1997.
- [73] Sijue Wu. Well-posedness in sobolev spaces of the full water wave problem in 3-d. J. Amer. Math. Soc., 12:445–495, 1999.
- [74] Sijue Wu. Almost global wellposedness of the 2-d full water wave problem. Inventiones mathematicae, 177(1):45, 2009.
- [75] Sijue Wu. Global wellposedness of the 3-d full water wave problem. Inventiones mathematicae, 184(1):125–220, 2011.
- [76] Sijue Wu. Wellposedness of the 2d full water wave equation in a regime that allows for non- interfaces. Inventiones mathematicae, 217(2):241–375, 2019.
- [77] Sijue Wu. The quartic integrability and long time existence of steep water waves in 2d. arXiv preprint arXiv:2010.09117, 2020.
- [78] H. Yosihara. Gravity waves on the free surface of an incompressible perfect fluid of finite depth. RIMS Kyoto, 18:49–96, 1982.
- [79] Vladimir E Zakharov. Stability of periodic waves of finite amplitude on the surface of a deep fluid. Journal of Applied Mechanics and Technical Physics, 9(2):190–194, 1968.
- [80] Vladimir E Zakharov and LA Ostrovsky. Modulation instability: the beginning. Physica D: Nonlinear Phenomena, 238(5):540–548, 2009.
- [81] Ping Zhang and Zhifei Zhang. On the free boundary problem of three-dimensional incompressible euler equations. Communications on Pure and Applied Mathematics, 61(7):877–940, 2008.
- [82] Fan Zheng. Long-term regularity of 3d gravity water waves. arXiv preprint arXiv:1910.01912, 2019.