Well-posedness of the water-waves equations

By David Lannes

Abstract

We prove that the water-waves equations (i.e., the inviscid Euler equations with free surface) are well-posed locally in time in Sobolev spaces for a fluid layer of finite depth, either in dimension or under a stability condition on the linearized equations. This condition appears naturally as the Lévy condition one has to impose on these nonstricly hyperbolic equations to insure well-posedness; it coincides with the generalized Taylor criterion exhibited in earlier works. Similarly to what happens in infinite depth, we show that this condition always holds for flat bottoms. For uneven bottoms, we prove that it is satisfied provided that a smallness condition on the second fundamental form of the bottom surface evaluated on the initial velocity field is satisfied.

We work here with a formulation of the water-waves equations in terms of the velocity potential at the free surface and of the elevation of the free surface, and in Eulerian variables. This formulation involves a Dirichlet-Neumann operator which we study in detail: sharp tame estimates, symbol, commutators and shape derivatives. This allows us to give a tame estimate on the linearized water-waves equations and to conclude with a Nash-Moser iterative scheme.

1. Introduction

1.1. Presentation of the problem

The water-waves problem for an ideal liquid consists of describing the motion of the free surface and the evolution of the velocity field of a layer of perfect, incompressible, irrotational fluid under the influence of gravity. In this paper, we restrict our attention to the case when the surface is a graph parameterized by a function , where denotes the time variable and the horizontal spatial variables. The method developed here works equally well for any integer , but the only physically relevant cases are of course and . The layer of fluid is also delimited from below by a not necessarily flat bottom parameterized by a time-independent function . We denote by the fluid domain at time . The incompressibility of the fluids is expressed by

where denotes the velocity field ( being the horizontal, and the vertical components of the velocity). Irrotationality means that

The boundary conditions on the velocity at the surface and at the bottom are given by the usual assumption that they are both bounding surfaces, i.e. surfaces across which no fluid particles are transported. At the bottom, this is given by

where denotes the outward normal vector to the lower boundary of . At the free surface, the boundary condition is kinematic and is given by

where , with denoting the outward normal vector to the free surface.

Neglecting the effects of surface tension yields that the pressure is constant at the interface. Up to a renormalization, we can assume that

Finally, the set of equations is closed with Euler’s equation within the fluid,

where is the acceleration of gravity.

Graphic without alt text

Early works on the well-posedness of Eqs. (Equation 1.1)-(Equation 1.6) within a Sobolev class go back to Nalimov Reference 27, Yosihara Reference 38 and Craig Reference 10, as far as 1D-surface waves are concerned. All these authors work in a Lagrangian framework, which allows one to consider surface waves which are not graphs, and rely heavily on the fact that the fluid domain is two dimensional. In this case, complex coordinates are canonically associated to the -coordinates, and the incompressibility and irrotationality conditions (Equation 1.1) and (Equation 1.2) can be seen as the Cauchy-Riemann equations for the complex mapping . There is therefore a singular integral operator on the top surface recovering boundary values of from boundary values of . The water-waves equations (Equation 1.1)-(Equation 1.6) can then be reduced to a set of two nonlinear evolution equations, which can be “quasi-linearized” using a subtle cancellation property noticed by Nalimov. It seems that this cancellation property was the main reason why the Lagrangian framework was used. A major restriction of these works is that they only address the case of small perturbations of still water. The reasons for this restriction are quite technical, but the most fundamental is that this smallness assumption ensures that a generalized Taylor criterion is satisfied, thus preventing formation of Taylor instabilities (see Reference 33Reference 4 and the introduction of Reference 36). Physically speaking, this criterion assumes that the surface is not accelerating into the fluid region more rapidly than the normal acceleration of gravity. From a mathematical viewpoint, this condition is crucial because the quasilinear system thus obtained is not strictly hyperbolic (zero is a multiple eigenvalue with a Jordan block) and requires a Lévy condition on the subprincipal symbol to be well-posed; one can see Taylor’s criterion precisely as such a Lévy condition (see Section 4.1 below). In Reference 3, Beale et al. proved that the linearization of the water-waves equations around a presumed solution is well-posed, provided this exact solution satisfies the generalized Taylor’s sign condition (which is a weaker assumption than the smallness conditions of Reference 27Reference 38Reference 10). Wu’s major breakthrough was to prove in Reference 36 that Taylor’s criterion always holds for solutions of the water-waves equations, as soon as the surface is nonself-intersect. Her energy estimates are also better than those of Reference 3 and allow her to solve the full (nonlinear) water-waves equations, locally in time, and without restriction (other than smoothness) on the initial data, but in the case of a layer of fluid of infinite depth. The only existing theorems dealing with the case of finite depth require smallness conditions on the initial data when the bottom is flat Reference 10, and an additional smallness condition on the variations of the bottom parameterization when the bottom is uneven Reference 38.

Very few papers deal with the well-posedness of the water-waves equations in Sobolev spaces in the three-dimensional setting (i.e. for a 2D surface). In Reference 22, the generalization of the results of Reference 3 to the three-dimensional setting is proved. More precisely, the authors show, in the case of a fluid layer of infinite depth, that the linearization of the water-waves equations around a presumed solution is well-posed, provided this exact solution satisfies the generalized Taylor’s sign condition. As in Reference 3, the energy estimates provided are not good enough to allow the resolution of the nonlinear water-waves equations by an iterative scheme. In Reference 37, S. Wu (still in the case of a fluid layer of infinite depth) solved the nonlinear equations. Her proof relies heavily on Clifford analysis in order to extend to the 3D case (some of) the results provided by harmonic analysis in 2D. In the case of finite depth, no results exist.

1.2. Presentation of the results

In this paper, we deliberately chose to work in the Eulerian (rather than Lagrangian) setting, since it is the easiest to handle, especially when asymptotic properties of the solutions are concerned. Inspired by Reference 29Reference 13 we use an alternate formulation of the water-waves equation (Equation 1.1)-(Equation 1.6). From the incompressibility and irrotationality assumptions (Equation 1.1) and (Equation 1.2), there exists a potential flow such that and

the boundary conditions (Equation 1.3) and (Equation 1.4) can also be expressed in terms of :

and

where we used the notation and . Finally, Euler’s equation (Equation 1.6) can be put into Bernouilli’s form

As in Reference 13, we reduce the system (Equation 1.7)-(Equation 1.10) to a system where all the functions are evaluated at the free surface only. For this purpose, we introduce the trace of the velocity potential at the surface

and the (rescaled) Dirichlet-Neumann operator (or simply when no confusion can be made on the dependence on the bottom parameterization ), which is a linear operator defined as

Taking the trace of (Equation 1.10) on the free surface and using the chain rule shows that (Equation 1.7)-(Equation 1.10) are equivalent to the system

which is an evolution equation for the elevation of the free surface and the trace of the velocity potential on the free surface . Our results in this paper are given for this system.

The first part of this work consists in developing simple tools in order to make the proof of the well-posedness of the water-waves equations as simple as possible. It is quite obvious from the equations (Equation 1.11) that the Dirichlet-Neumann operator will play a central role in the proof; we give here a self-contained and quite elementary proof of the properties of the Dirichlet-Neumann operator that we shall need. A major difficulty lies in the dependence on of the operator . It is known that such operators depend analytically on the parameterization of the surface. Coifman and Meyer Reference 9 considered small Lipschitz perturbations of a line or plane, and Craig et al. Reference 12Reference 13 perturbations of hyperplanes in any dimension. Seen as an operator acting on Sobolev spaces, is of order one. In Reference 13, an estimate of its operator norm is given in the form:

for all integer (estimates in -based Sobolev spaces are also provided). In order to obtain this estimate, the authors give an expression of as a singular integral operator (inspired by the early works of Garabedian and Schiffer Reference 17 and Coifman and Meyer Reference 9 on Cauchy integrals) and use a multiple commutator estimate of Christ and Journé Reference 6. Estimate (Equation 1.12) has the interest of being “tame” (in the sense of Hamilton Reference 21; i.e., the control in the norms depending on the regularity index is linear), but is only proved for flat bottoms and requires too much smoothness on : a control of is needed in (Equation 1.12), and hence of , with , if one works in a Sobolev framework. A rapid look at equations (Equation 1.11) shows that one would like to allow only a control of in (i.e., and should have the same regularity). Using an expression of involving tools of Clifford Algebras Reference 18 and deep results of Coifman, McIntosh and Meyer Reference 8 and Coifman, David and Meyer Reference 7, S. Wu obtained in Reference 37 another estimate with a sharp dependence on the smoothness of :

for all real numbers large enough. If estimate (Equation 1.13) is obviously better than (Equation 1.12), it has two drawbacks. First, it is not tame, and hence not compatible for later use in a Nash-Moser convergence scheme. Second, its proof requires very deep results, which make its generalization to the present case of finite and uneven bottom highly nontrivial. In this paper, we prove in Theorem 3.6 the following estimate:

for all , and where is a fixed positive real number. This estimate has the sharp dependence on of (Equation 1.13) and is tame as (Equation 1.12). Moreover, it is sharper than the above estimates in the sense that only the gradient of is involved; this will prove very useful here. Estimate (Equation 1.14) also holds for uneven bottoms and its proof uses only elementary tools of PDE: since the fluid layer is diffeomorphic to the flat strip , we first transform the Laplace equation (Equation 1.7) with Dirichlet condition at the surface and homogeneous Neumann condition at the bottom into an elliptic boundary value problem (BVP) with variable coefficients defined in the flat strip . The Dirichlet-Neumann operator can be expressed in terms of the solution to this new BVP (see Prop. 3.4). We give sharp tame estimates for a wide class of such elliptic problems in Theorem 2.9. Choosing the most simple diffeomorphism between the fluid domain and as in Reference 12Reference 2 and applying Theorem 2.9 to the elliptic problem thus obtained, we can obtain, via Prop. 3.4, a tame estimate on . However, this estimate is not sharp since instead of as in (Equation 1.14), one would need a control of . We must therefore gain half a derivative more to obtain (Equation 1.14). The trick consists in proving (see Prop. 2.13) that there exists a “regularizing” diffeomorphism between the fluid domain and the flat strip .

We also need further information on the Dirichlet-Neumann operator. In Theorem 3.10, we give the principal symbol of : for all ,

where , and where the constant involves the -norm of a finite number of derivatives of . Note in particular that for 1D surfaces, , while for 2D surfaces it is a pseudo-differential operator (and not a simple Fourier multiplier). We then give tame estimates of the commutator of with spatial (in Prop. 3.15) and time (in Prop. 3.19) derivatives. Finally, we give in Theorem 3.20 an explicit expression of the shape derivative of , i.e. the derivative of the mapping , and tame estimates of this and higher derivatives are provided in Prop. 3.25.

Note that all the above results are proved for a general constant coefficient elliptic operator instead of in (Equation 1.7). This is useful if one wants to work with nondimensionalized equations. This first set of results consists therefore in preliminary tools for the study of the water-waves problem; we would like to stress the fact that they are sharp and only use the classical tools of PDE.

We then turn to investigate the water-waves equations (Equation 1.11). The first step consists of course in solving the linearization of (Equation 1.11) around some reference state , and in giving energy estimates on the solution. Using the explicit expression of the shape derivative of the Dirichlet-Neumann operator given in Theorem 3.20, we can give an explicit expression of the linearized operator . Having the previous works on the water-waves equations in mind, it is not surprising to find that is hyperbolic, but that its principal symbol has an eigenvalue of multiplicity two (i.e., it is not strictly hyperbolic). In the works quoted in the previous section, this double eigenvalue is zero. Due to the fact that we work here in Eulerian, as opposed to Lagrangian, variables, this double eigenvalue is not zero anymore, but , being the dual variable of , and being the horizontal component of the velocity at the surface of the reference state . It is natural to seek a linear change of unknowns which transforms the principal part of into its canonical expression consisting of an upper triangular matrix with double eigenvalue and a Jordan block. Prop. 4.2 gives a striking result: this a priori pseudo-differential change of unknown is not even differential, and the commutator terms involving the Dirichlet-Neumann operator that should appear in the lower-order terms all vanish! This simplifies greatly the sequel.

Having transformed the linearized operator into an operator whose principal part exhibits the Jordan block structure inherent to the water-waves equations, we turn to study this operator . The Lévy condition needed on the subprincipal symbol of in order for the associated Cauchy problem to be well-posed is quite natural, due to the peculiar structure of : a certain function depending only on the reference state must satisfy for some positive constant (this is almost a necessary condition, since the linearized water-waves equations would be ill-posed if one had ). It appears in Prop. 4.4 that this sign condition is exactly the generalized Taylor’s sign condition of Reference 3Reference 22Reference 36Reference 37. Assuming for the moment that this condition holds, we use the tools developed in the first sections to show, in Prop. 4.5, that the Cauchy problem associated to is well-posed in Sobolev spaces, and to give energy estimates on the solution. There is a classical loss of information of half a derivative on this solution due to the Jordan block structure, but also a more dramatic loss of information with respect to the reference state , which makes a Picard iterative scheme inefficient for solving the nonlinear equation. Fortunately, the energy estimates given in Prop. 4.5 are tame, and Nash-Moser theory will provide a good iterative scheme. Inverting the change of unknown of Prop. 4.2, tame estimates are deduced in Prop. 4.14 for the solution of the Cauchy problem associated to the linearized operator . The last step of the proof consists in solving the nonlinear equations (Equation 1.11) via a Nash-Moser iterative scheme. This requires proving that Taylor’s sign condition holds at each step of the scheme (and of course that the surface elevation remains positive!). It is quite easy to see that it is sufficient for this condition to be satisfied that the first iterate satisfies it. Wu proved that this is always the case in infinite depth. We prove in Prop. 4.15 that this result remains true in the case of flat bottoms. For uneven bottoms, however, we must assume that the generalized Taylor’s sign condition holds for the initial data. This can be ensured by smallness conditions on the initial data, but we also give a sufficient condition stating that Taylor’s sign condition can be satisfied for initial data of arbitrary size provided that the bottom is “slowly variable” in the sense that

where is the bottom parameterization, the second fundamental form associated to the surface , and the tangential component of the initial velocity field evaluated at the bottom.

Our final result is then given in Theorem 5.3. For flat bottoms (i.e. ), it can be stated as:

Theorem 1.1.

Let and be such that , with ( depending only on ). Assume moreover that

Then there exists and a unique solution to the water-waves equations Equation 1.11 with initial conditions and such that .

Organization of the paper. Section 2 is devoted to the study of the Laplace equation (Equation 1.7) in the fluid domain, or more precisely to the equation , where is a constant coefficient, symmetric and coercive matrix. In Section 2.1, we show that this equation can be reduced to an elliptic boundary problem with variable coefficients on a flat strip, and sharp tame elliptic estimates for such problems are given in Section 2.2. We then show in Section 2.3 that among the various diffeomorphisms between the fluid domain and the flat strip, there are some that are particularly interesting, which we call “regularizing diffeomorphisms” and which allow the gain of half a derivative with respect to the regularity of the surface parameterization.

Section 3 is entirely devoted to the properties of the Dirichlet-Neumann operator. Basic properties (including the sharp estimate (Equation 1.14) mentioned above) are gathered in Section 3.1. In Section 3.2, we are concerned with the derivation of the principal part of the Dirichlet-Neumann operator, and in Section 3.3 with its commutator properties with space or time derivatives. Finally its shape derivatives are studied in Section 3.4.

The linearized water-waves equations are the object of Section 4. We first show in Section 4.1 that the linearized equations can be made trigonal and prove in Section 4.2 that the Cauchy problem associated to the trigonal operator is well-posed in Sobolev spaces, assuming that a Lévy condition on the subprincipal symbol holds. We also provide in this section tame estimates on the solution. The link with the solution of the original linearized water-waves equations is made in Section 4.3, and the Lévy condition is discussed in Section 4.4.

The fully nonlinear water-waves equations are solved in Section 5. A simple Nash-Moser implicit function theorem is first recalled in Section 5.1 and then used in Section 5.2 to obtain our final well-posedness result.

Finally, a technical proof needed in Section 2.1 has been postponed to Appendix A.

1.3. Notation

Here is a set of notation we shall use throughout this paper:

always denotes a numerical constant which may change from one line to another. If the constant depends on some parameters , we denote it by .

For any , we write .

For all , we write ; similarly, we write , and, for all , .

We denote by the set of functions continuous and bounded on together with their derivatives of order less than or equal to , endowed with its canonical norm . We denote also .

We denote by the usual scalar product on .

We denote by , or , the Fourier multiplier with symbol .

For all , we denote by the space of distributions such that , where denotes the Fourier transform of . We also denote .

If , we write .

If is a Banach space and if , then we write . If , then .

For all , denotes the first integer strictly larger than (so that ).

2. Elliptic boundary value problems on a strip

Throughout this section, we work on a domain defined as

where and satisfy the following condition:

(this assumption means that we exclude beaches or islands for the fluid domain, either perturbed or at rest).

We also consider a constant coefficients elliptic operator , where is a symmetric matrix satisfying the following condition:

Finally, we consider boundary value problems of the form

where is a function defined on and are functions defined on . Moreover, denotes the conormal derivative associated to of at the boundary ,

where denotes the outwards normal derivative at the bottom.

Notation 2.1.

For all open sets , we denote by , and the canonical norms of , and respectively. When no confusion is possible on the domain , we write simply , and .

2.1. Reduction to an elliptic equation on a flat strip

Throughout this section, we denote by any diffeomorphism between and the flat strip , which we assume to be of the form

and we denote its inverse by ,

We always assume the following on :

Assumption 2.2.

One has with and . Moreover, there exists such that on .

Finally, we need the following definition:

Definition 2.3.

Let . The mapping , given by (Equation 2.6), is called -regular if it satisfies Assumption 2.2 and can moreover be decomposed into with and , and if on .

Remark 2.4.

The most simple diffeomorphism between and is given by

and hence . If and , it is clear that is -regular, with , , and .

To any distribution defined on one can associate, using the diffeomorphism and its inverse given by (Equation 2.5)-(Equation 2.6), a distribution defined on as

and vice-versa,

The following lemma shows that the constant coefficients elliptic equation on can equivalently be formulated as a variable coefficients elliptic equation on .

Lemma 2.5.

Suppose that the mapping , given by Equation 2.6 satisfies Assumption 2.2. Let with satisfying Equation 2.2. Then the equation holds in if and only if the equation holds in , where and are deduced from and via formula Equation 2.7, and , with

Moreover, one has, for all ,

Proof.

By definition, in if and only if

By definition of , one also has

Integrating by parts yields therefore that is equal to

and thus to

By definition of and , one has for all . Differentiating this identity with respect to and respectively yields

Using these expressions in the above expressions gives the equality

where is as given in the statement of the lemma. Since one clearly has

the first claim of the lemma follows from (Equation 2.9) and (Equation 2.10).

We now prove the coercivity of . One has, for all ,

and owing to (Equation 2.2) we have therefore

The matrix is invertible, and its inverse is given by

so that can be bounded as

Together with (Equation 2.11), this estimate yields the result of the lemma.

The next lemma shows how the boundary conditions are transformed by the diffeomorphism .

Lemma 2.6.

Suppose that the mapping , given by Equation 2.6, satisfies Assumption 2.2. For all , one has

Proof.

The first assertion of the lemma is straightforward. We now prove the second. By definition,

Replacing by its expression given in Lemma 2.5, one obtains easily that

which ends the proof of the lemma.

Lemmas 2.5 and 2.6 show that the study of the boundary problems (Equation 2.3) can be deduced from the study of elliptic boundary value problems on a flat strip:

Proposition 2.7.

Suppose that the mapping , given by Equation 2.6, satisfies Assumption 2.2. Then is a (variational, classical) solution of Equation 2.3 if and only if given by Equation 2.7 is a (variational, classical) solution of

where is as given in Lemma 2.5.

The next section is therefore devoted to the study of the well-posedness of such variable coefficients elliptic boundary value problems on a flat strip. Before this, let us state a lemma dealing with the smoothness of the coefficients of . Its proof is given in Appendix A.

Lemma 2.8.

Let and assume that the mapping , given by Equation 2.6, is -regular. Then one can write with , and

2.2. Variable coefficients elliptic equations on a flat strip

We have seen in the previous section that the theory of elliptic equations on a general strip of type (Equation 2.3) can be deduced from the study of elliptic equations on a flat strip, but with variable coefficients. In this section, we study the following generic problem:

where we recall that denotes the conormal derivative associated to ,

We also assume that satisfies the following coercivity assumption:

The main result of this section is the following theorem.

Theorem 2.9.

Let , . Let , and .

i.

If satisfies Equation 2.15, then there exists a unique solution to Equation 2.13. Moreover,

ii.

If and are such that satisfies Equation 2.15, then there exists a unique solution to Equation 2.13. Moreover, when ,

where .

Remark 2.10.
i.

The proof below shows that the quantity can be estimated more precisely than . Namely, one can replace the quantities and in both estimates of the theorem by and respectively. This remark is very useful when giving estimates on the Dirichlet-Neumann operator.

ii.

The second estimate of the theorem remains of course valid when , but in that case, the first estimate of the theorem is more precise.

Proof.

Even though is unbounded, the proof follows the same lines as the usual proofs of existence and regularity estimates of solutions to elliptic equations on regular bounded domains (Reference 26Reference 19), but special care must be paid to use the specific Sobolev regularity of the coefficients of . We only prove the second point of the theorem since the first one can be obtained by skipping the fourth step of the proof below.

Step 1. Construction of a variational solution to (Equation 2.13). We first introduce , where is a smooth compactly supported function such that . Classically, one has and

It follows that is a variational solution of (Equation 2.13) if and only if is a variational solution to

where .

Define the space as , where the closure is taken relative to the -norm. It is a classical consequence of Lax-Milgram’s theorem that there exists a unique such that

Step 2. Regularity of the variational solution. We show that using the classical method of Nirenberg’s tangential differential quotients. For all and , one has , where is defined as

we also recall that the adjoint operator of is and that one has the product rule . Using (Equation 2.18) with instead of , one gets therefore

By the trace theorem and Poincaré’s inequality, we get , so that the r.h.s of the above inequality can be bounded from above by

Taking as a test function in (Equation 2.19), using condition (Equation 2.15), and letting , one gets therefore

for all such that . The missing term is obtained as usual using the equation,

where , from which it follows easily that

From (Equation 2.20) and (Equation 2.22), it follows that and satisfies

Replacing and by their expressions in the above inequality, and using the estimates (Equation 2.16) yields

Step 3. Further regularity. We show by finite induction on that for all , , one has

By Step 2, this assertion is true when . Let and assume it is also true for . For all , we apply (Equation 2.24) with to the function :

We now estimate the four terms which appear in the r.h.s. of (Equation 2.25). Since , one has

The second and third terms of (Equation 2.25) are very easily controlled. For the fourth one, we use the explicit expression of , use the trace theorem, and proceed as for the derivation of (Equation 2.26) to obtain

From (Equation 2.25), (Equation 2.26) and (Equation 2.27) (and letting ), it follows that is bounded from above by the right-hand side of (Equation 2.24). In order to complete the proof, we still need an estimate of in . As in Step 2, such an estimate is obtained using (Equation 2.21).

Step 4. Further regularity. We show by induction on that (Equation 2.24) can be generalized for as

where .

The procedure is absolutely similar to Step 3. It is strictly unchanged until Eq. (Equation 2.26) where we now use Moser’s tame estimates on products (e.g. Reference 1):

Lemma 2.11.

Let and . Then one has

This yields

Estimate (Equation 2.27) is modified along the same lines and it follows from the Sobolev embedding that is bounded from above by the right-hand side of (Equation 2.28).

An estimate on in is then provided as before using (Equation 2.21), which concludes the induction.

Step 5. Endgame. From the variational formulation of the problem, one easily gets the following lemma, whose proof we omit.

Lemma 2.12.

Let , and . If solves the boundary value problem Equation 2.13, then

and

Iterating estimates (Equation 2.24) and (Equation 2.28) and using the lemma gives the theorem.

2.3. Regularizing diffeomorphisms

If solves the boundary value problem (Equation 2.3), then one can give precise estimates on , owing to Prop. 2.7 and 2.8, and using Theorem 2.9. However, these estimates depend strongly on the diffeomorphism chosen to straighten the fluid domain. The trivial diffeomorphism given in Remark 2.4 is not the best choice possible: in order to control the -norm of its Sobolev component, one needs to control the -norm of the surface parameterization . The next proposition shows that there exist “regularizing” diffeomorphisms for which a linear control of the -norm suffices.

Proposition 2.13.

Let , , and let , . If there exists such that on , then there exists a diffeomorphism of the form Equation 2.6 such that

is -regular (with );

one has ;

one has and

Remark 2.14.
i.

The diffeomorphism provided by this lemma is a perturbation of the trivial diffeomorphism given in Remark 2.4. The -component remains unchanged, and the behavior at the surface is exactly the same. However, the Sobolev component is half a derivative smoother here than for the trivial diffeomorphism (where it has the smoothness of ). This is why we say that the diffeomorphism is “regularizing”.

ii.

Note that if for some , and if the condition is satisfied uniformly in , then one has and . This will be used in the proof of Prop. 3.19.

iii.

If with is such that on , then one can find a neighbourhood of in such that for all , . To each of these , one can associate a regularizing diffeomorphism by Prop. 2.13. The proof shows that if is small enough, then the mapping is affine. This mapping is therefore smooth and (using the notation of the proof) one can check that for all , one has for some . Hence,

Proof.

Note that the Jacobian of the mapping is equal to . Therefore, if satisfies the properties stated in the lemma, is indeed a diffeomorphism between and .

Let be given by

we look for such that satisfies

We construct such a mapping using a Poisson kernel extension of . Let be a smooth, compactly supported, function defined on and such that and . For any , and , we define as . From this definition it follows also that for all , one has

since .

Define now . It is obvious that satisfies the last three conditions of (Equation 2.29). For the first one, remark that

and that

Taking small enough, one can complete the proof from (Equation 2.30), (Equation 2.31), (Equation 2.32) and the assumption .

3. The Dirichlet-Neumann operator

The aim of this section is to investigate the properties of the Dirichlet-Neumann operator associated to a class of boundary value problems included in the general framework studied in Section 2. It is known that such operators depend analytically on the parameterization of the surface. Coifman and Meyer Reference 9 considered small Lipschitz perturbations of a line or plane, and Craig et al. Reference 12Reference 13 perturbations of hyperplanes in any dimension. These studies rely on subtle estimates of singular integral operators. More recently, Nicholls and Reitich Reference 28 addressed the analyticity of the Dirichlet-Neumann operator using a simple method based on a change of variables (see also Reference 2). Here, we are also interested in the dependence of the Dirichlet-Neumann operator on the fluid domain, but from a Sobolev rather than an analytical viewpoint. The sharp elliptic estimate of the previous section allows us to give “tame” estimates on the action of the Dirichlet-Neumann operator on Sobolev spaces. In this section, we also compute the principal symbol of the DN operator, give tame estimates of its commutators with spatial or time derivatives, and also study carefully its shape derivatives.

3.1. Definition and basic properties

As in Section 2, we consider a fluid domain of the form

where and satisfy

We also consider a constant coefficients elliptic operator , satisfying the coercivity condition (Equation 2.2). The boundary value problems we consider in this section are a particular case of the boundary value problems (Equation 2.3) since we only consider the case of a homogeneous source term and the Neumann boundary condition at the bottom. More precisely, let solve

where we recall that, as defined in (Equation 2.4), denotes the conormal derivative associated to .

For all and , and provided that and are smooth enough, we know by Theorem 2.9 that exists and is unique. Therefore, the following definition makes sense.

Definition 3.1.

Let , and assume that satisfy condition (Equation 3.1). We define the Dirichlet-Neumann operator to be the operator given by

where denotes the solution of (Equation 3.2).

Remark 3.2.
i.

Thus defined, is not exactly the Dirichlet-Neumann operator because of the scaling factor ; yet, we use this terminology for the sake of simplicity.

ii.

Thanks to the minus sign in the definition, maps the Dirichlet data to the (rescaled) outward normal derivative when .

As in Section 2, we can associate to (Equation 3.2) an elliptic boundary value problem on the flat strip : denoting by the “regularizing” diffeomorphism between and (and by its inverse) given in Prop. 2.13, and , one has

where is as given in Lemma 2.5.

Notation 3.3.

We denote by the solution of the b.v.p. (Equation 3.3).

Proceeding as in the proof of Lemma 2.6, one can define the Dirichlet-Neumann operator in terms of .

Proposition 3.4.

Under the same assumptions as in Def. 3.1, one has

where is as defined in Notation 3.3.

Before stating our main estimates on the DN operator, let us state some notation.

Notation 3.5.
i.

When a bottom parameterization () is given, we generically write .

ii.

For all , we denote generically by (resp. ) constants which depend on and (resp. , and ).

The next theorem shows that the DN operator is of order one and gives precise estimates on its operator norm.

Theorem 3.6.

Let and be two continuous functions satisfying Equation 3.1. Then:

i.

For all , if , then for all such that , one has

ii.

For all , if and if , then

for all such that , and where we used Notation 3.5.

Remark 3.7.

Note that the DN operator is defined for functions whose gradient is in some Sobolev space, but which are not necessarily in a Sobolev space themselves.

Proof.

We just prove the second part of the theorem; the proof of the first part is very similar. Owing to Prop. 3.4, we have

By the trace theorem, this yields

where the notation is as in Notation 2.1.

Using the decomposition of Lemma 2.8 and the tame product estimate of Lemma 2.11, one obtains

Now, remark that when the diffeomorphism between the flat strip and the fluid domain is the regularizing diffeomorphism of Prop. 2.13, the estimates of Lemma 2.8, together with the Sobolev embedding , give

Similarly, the constant which appears in Theorem 2.9 when one takes can be bounded from above by and the result follows therefore from (Equation 3.4), Theorem 2.9 and Remark 2.10.

Some important properties of the DN operator are listed in the next proposition.

Proposition 3.8.

Let satisfy Equation 3.1. Then:

i.

The operator is self-adjoint:

ii.

The operator is positive:

iii.

We also have the estimates

and for all , where is given in Lemma 2.5, one has

Remark 3.9.

Using the self-adjointness of , one could extend this operator to all Sobolev spaces , .

Proof.
i.

According to Prop. 3.4, and using Notation 3.3, one has , and Green’s identity yields . Using Prop. 3.4 once again yields the result.

ii.

Writing and integrating by parts, one obtains

where the last inequality uses the coercivity of proved in Lemma 2.5.

iii.

Proceeding as in ii, one has

and the first estimate follows from Lemma 2.12.

To prove the second estimate, remark first that by Poincaré’s inequality, we have , and therefore . As a consequence, we obtain . Using (Equation 3.6) and the estimate , we deduce . The end of the proof is then straightforward.

3.2. Symbol of the Dirichlet-Neumann operator

In order to compute the commutator of with differential operators, which is a crucial step to obtaining energy estimates for the water-waves equations, we need to know its principal symbol. Since this result is interesting in itself, we state it as a theorem (we use the classical notation to denote the pseudo-differential operator associated to the symbol ).

Theorem 3.10.

There exists an integer , depending only on , such that if and satisfy Equation 3.1, then, for , one has

where is as defined in Notation 3.5 and the symbol is given by

with .

Remark 3.11.
i.

The estimate of the theorem can be extended to higher-order Sobolev spaces, but we do not need such a result here.

ii.

The parameterization of the bottom does not appear in the principal symbol of . This is not surprising since the contribution to the surface of the bottom is “smoothed” by the elliptic equation.

iii.

For the water-waves equations, one has and takes the simple form

There is therefore an interesting phenomenological difference between the 1D and the 2D cases. In the latter, the principal symbol of the Dirichlet-Neumann operator is a pseudo-differential operator, while in the former, it is simply a Fourier multiplier: , which does not depend on the fluid domain.

Proof.

The proof of the theorem relies strongly on the factorization procedure of elliptic operators, as set forth in Reference 35 (see also Reference 30). Recall that, according to Prop. 3.4, one has , where denotes the solution of (Equation 3.3). The idea is to deduce an approximation of from an approximation of for which the conormal derivative at the surface can be explicitly computed. In order to find such an approximation, we first approximate the elliptic operator ; this is where we need the factorization procedure mentioned above.

Writing , one can check that the operator can be written in the form

We now look for an approximation of of the form

where, for all , denotes the pseudo-differential operator of the symbol . Obviously, if one wants the highest-order terms of to match those of , must be the roots of the second-order polynomial ; namely,

We now take the function we are looking for as an approximate solution of the equation ; from (Equation 3.8), it suffices to take an approximate solution of the backward evolution equation

we therefore take

where . Since the real part of is always positive, is smoothing for all . As a consequence, one has:

Lemma 3.12.

Let . Let and satisfy condition Equation 3.1. Then

where is as defined in Notation 3.5.

Proof.

From the explicit expression of given in (Equation 3.9), one deduces

where is a positive constant which depends on , , and . Let us define as

it is clear that is a symbol of order zero (uniformly in ). The operator acts therefore continuously on . Moreover, its operator norm can be bounded in terms of a finite number of -norms of space-frequency derivatives of the symbol (-derivatives with respect to and derivatives with respect to are enough; see Reference 23 and also Reference 25). Using the Sobolev embedding for , it follows that

Thus,

The gain of half a derivative claimed in the first estimate of the lemma is deduced from this expression by a classical computation (see e.g. Prop. 12.4 of Reference 34).

One can estimate the first-order derivatives of in the same way, which yields the second estimate of the lemma.

We now prove that is indeed an approximate solution of (Equation 3.10) and hence of the equation .

Lemma 3.13.

There exists an integer , depending only on , such that if and satisfy condition Equation 3.1, then

Proof.

Simple computations yield

where and are as in the proof of Lemma 3.12.

It is easy to check that is of order one, so that it acts continuously on with values in . As in the proof of Lemma 3.12 above, we can bound its norm in terms of a finite number of derivatives of the symbol. For large enough, we therefore have

where is as defined in Notation 3.5.

Similarly, the operator is of order , so that (taking a larger if necessary), one has

and one can conclude the proof as for Lemma 3.12.

We now proceed to estimate the difference :

Lemma 3.14.

There exists an integer , depending only on , such that if and satisfy Equation 3.1, then

Proof.

Since by definition , one gets

together with the boundary conditions and . Using Theorem 2.9, we therefore find

Using (Equation 3.7) and the definition of , one checks easily that is a first-order operator, so that with the help of Lemma 3.12, one gets the bound . Owing to Lemma 3.13, the same bound also holds on . Finally, since is a smoothing operator, such an estimate also holds for , and the proof of the lemma is complete.

We are now ready to finish the proof of the theorem. First remark that so that

by Lemma 3.14, we therefore have, for some ,

To prove that (Equation 3.15) coincides with the estimate of the theorem in the case , we must show that , which we do now.

Thanks to Lemmas 2.5 and 2.13, we know the explicit expression of ; from the definition of the conormal derivative, one can then compute easily

where .

The explicit expression of given in (Equation 3.9) yields also

Plugging this expression into (Equation 3.16) yields , which concludes the -estimate of the theorem. Recalling that the DN operator is self-adjoint (see Prop. 3.8), one deduces the -estimate by a standard duality argument; finally, the -estimate is obtained by interpolation.

3.3. Commutator estimates

This section is devoted to the proof of tame estimates of the commutator of the Dirichlet-Neumann operator with spatial derivatives and time derivative. The next proposition deals with the case of spatial derivatives.

Proposition 3.15.

There exists an integer , depending only on , such that if and satisfy Equation 3.1, then for all and , , one has

where is as defined in Notation 3.5.

Remark 3.16.

The interest of this commutator estimate is that it is “tame”: even though we have a loss of derivative (in the sense that one needs to control the -norm of and not only its -norm), this loss is linear, and the multiplicative constant which appears in front of it involves only Sobolev norms of independent of . This point is crucial to obtaining tame energy estimates later.

Proof.

First remark that the following identity holds for all , :

so that

Estimate of . The idea is to replace by its principal symbol computed in the previous section. One has, denoting ,

The operator is of order , and one can bound its operator norm , as in the proofs of Lemmas 3.12 and 3.13, in terms of the derivatives of the symbol given in Theorem 3.10. Thus, for some ,

Both and can be bounded using Theorem 3.10:

From (Equation 3.18)-(Equation 3.20), we deduce

Estimate of . Using Notation 3.3, it is easy to check that

with .

The first term of the r.h.s. of (Equation 3.22) is estimated as follows:

In order to estimate the second term of (Equation 3.22), first remark that solves the b.v.p.

by the trace theorem and Theorem 2.9, one therefore gets

The term which appears in both (Equation 3.23) and (Equation 3.25) is estimated in the following lemma:

Lemma 3.17.

Under the same assumptions as in the proposition, one has

Proof.

By Lemma 2.8, we can decompose into , so that .

One has , which is itself smaller than the r.h.s. of the estimate of the lemma, thanks to the estimates (Equation 3.5) and Theorem 2.9. In order to bound from above, remark that

and that for all , one has . It is then easy to obtain, using Lemma 2.11 and the estimates (Equation 3.5), that

Owing to Theorem 2.9, the r.h.s. of this latter estimate is smaller than the r.h.s. of the estimate given in the lemma, so that the proof is complete.

From (Equation 3.22), (Equation 3.23), (Equation 3.25) and the lemma one obtains

The proposition is therefore a consequence of (Equation 3.17), (Equation 3.21) and (Equation 3.26).

We end this section with two propositions concerning the commutator properties of the Dirichlet-Neumann operator with a general scalar-valued differential operator of order one, and with the time derivative (when the surface depends on time).

Proposition 3.18.

Let and suppose that and satisfy Equation 3.1. Let be a first-order differential operator on with coefficients in . Then, for all , one has

where is as in Notation 3.5, and denotes the sum of the -norm of all the coefficients of .

Proof.

With the same techniques as in the proof of the previous proposition, one can show that

where is the solution of the boundary value problem

Green’s identity asserts that

We also know that

Integrating by parts the first term of the r.h.s., one finds (recall that ),

From (Equation 3.27)-(Equation 3.30) we deduce

from which one obtains easily

Multiplying (Equation 3.28) by , integrating by parts and using Poincaré’s inequality, one obtains

and (Equation 3.31) yields therefore

and one concludes the proof with the help of Lemma 2.12.

With only minor modifications, and using Remark 2.14, the same proof gives:

Proposition 3.19.

Let and . Let and satisfy Equation 3.1 uniformly for . Then, for all , one has

where is as in Notation 3.5.

3.4. Shape derivative of the Dirichlet-Neumann operator

The Dirichlet-Neumann operator is linear but depends nonlinearly on the parameterization of the surface. It is known that this dependence is smooth, and even analytical Reference 9Reference 13Reference 28. The next theorem gives an explicit expression of its shape derivative, that is, of its derivative with respect to the surface parameterization. In Prop. 3.25 below we give tame estimates on the first and second shape derivatives.

Theorem 3.20.

Let and , . Suppose that and satisfy Equation 3.1. Then there exists a neighborhood of in such that for all given , the mapping

is well defined and differentiable. Moreover, for all , one has

where

with

Remark 3.21.

For the water-waves equations, one has , and is simply given by .

Proof.

We can choose a neighborhood of such that for all , condition (Equation 3.1) is satisfied (taking smaller if necessary). To each it is therefore possible to associate a regularizing diffeomorphism as in Prop. 2.13. Taking smaller if necessary, and using Remark 2.14, we can assume that the mapping is affine. We denote by its derivative at . Since the matrix , given by Lemma 2.5 with , has coefficients in , it follows that the mapping

is smooth. We denote by its derivative at . Let us also denote by the solution of the boundary value problem

By Theorem 2.9, we know that . It is quite easy to prove that the mapping defined as

is continuous. Differentiating (Equation 3.32) with respect to , it is easy to show that is differentiable at and that for all , solves

The following is a key lemma. It gives an explicit function solving (Equation 3.33) except for the Dirichlet condition at the surface.

Lemma 3.22.

For all , the function solves

Remark 3.23.

The expression of given in the above lemma might not seem obvious. We sketch here a way to find it in the case where and for 1D surfaces. Denote by the solution of the Laplace equation (Equation 1.7) in with Dirichlet condition at the surface and homogeneous Neumann condition at the bottom. First write in variational form that solves this boundary value problem and then differentiate this variational equality with respect to using the classical work of Hadamard on shape functionals Reference 20 (see also Lemma 5.1 of Reference 15). This yields an expression of the derivative of the mapping . Pulling this expression back by the regularizing diffeomorphism yields an expression of the derivative of and hence of . The expression given in Lemma 3.22 is just a generalization of this expression found formally in the case of multi-dimensional surface waves.

Proof.

Let us compute (writing instead of ),

Using the fact that , we obtain

Still using the identity , one can remark that

and therefore, one can write

where the symmetric matrix is equal to

We now prove that . In order to do so, let us write the matrix in the form , where is a symmetric matrix, and . The matrix given by Lemma 2.5 can therefore be written

and it follows that for any , the matrix is given by

It is then easy, though tedious, to check that . From (Equation 3.34) we obtain therefore , and it remains only to check that satisfies the boundary conditions to conclude the proof of the lemma.

From Prop. 2.13 and Remark 2.14, one has and so that satisfies the Dirichlet boundary condition stated in the lemma on the upper boundary of the strip . To check that the Neumann condition of the lower boundary is also satisfied, recall that by definition

Now, recall that owing to Remark 2.14, one has , so that

One can check that this latter expression equals , which concludes the proof.

From (Equation 3.33) and Lemma 3.22, solves

by definition of the DN operator , it follows that or equivalently

To finish the proof, we write in terms of .

One has ; hence, using the fact that denotes the derivative of the mapping at applied to ,

Together with (Equation 3.35), and using the identity , with as defined in the statement of the theorem, this yields

Lemma 3.24.

Under the assumptions and with the notation of the theorem, one has

Proof.

Recall that owing to Prop. 2.13 and Remark 2.14, one has , , and . Using the same notation as in the proof of Lemma 3.22, one obtains

Using the expression of given in the proof of Lemma 3.22, we also compute

and the lemma follows.

The theorem is then a simple consequence of (Equation 3.37) and Lemma 3.24.

Theorem 3.20 is crucial in the symbolic analysis of the linearized water-wave equations. However, one can notice that the explicit expression it gives is not very useful at the time of giving estimates of the shape derivatives. Indeed, both terms of this expression are in , while the derivative of the DN operator belongs to . This means that there is a cancellation of the most singular components of both terms. Estimates of the shape derivatives have therefore to be done at an upper level.

Proposition 3.25.

Let and , . Suppose that , and satisfy Equation 3.1. Then the mapping

is and the successive derivatives are “tame”:

i.

For all , one has

ii.

For all ,

iii.

Similar estimates hold for , .

Proof.

Recall that if the diffeomorphism is the regularizing diffeomorphism constructed in Prop. 2.13, one has for some and where is the same compactly supported function as in the proof of Lemma 2.13. Therefore, for all , . From the explicit expression of given in Lemma 2.5, and with the same computations as for Lemma 2.8, one obtains therefore

recall also that owing to Theorem 2.9 and Remark 2.10 (with ), the solution to (Equation 3.32) satisfies for all the tame estimate

Now, recall that we saw in (Equation 3.36) that

where solves (Equation 3.33). From (Equation 3.38) and (Equation 3.39), together with Lemma 2.11, it is easy to see that the first term of the r.h.s. satisfies the estimate of the proposition. The estimate on the second term of the r.h.s. is deduced from Theorem 2.9 applied to the boundary value problem (Equation 3.33).

Since the methods for obtaining the estimates on higher derivatives of are absolutely similar, we omit the proof.

4. The linearized water-waves equations

4.1. Trigonalization of the linearized system

As seen in the introduction, the water-waves equations are

where, for the sake of simplicity, we wrote instead of , being the parameterization of the bottom.

We can write this system in condensed form as

with and

This section is devoted to the study of the linearized water-waves equations around an admissible reference state, in the following sense:

Definition 4.1.

Let . We say that is an admissible reference state if and , and if moreover

where we recall that is a parameterization of the bottom.

By definition, the linearized operator associated to (Equation 4.2) is given by ; from the explicit expression of given above, one computes

with , and, for all smooth enough,

and

According to Theorem 3.20, we have, for all ,

so that becomes

One can check that the principal part of the above operator admits as an eigenvalue of multiplicity two and a nontrivial Jordan block. Taking as a new unknown makes this Jordan block appear under its canonical form. Unexpectedly enough, this change of unknowns not only makes trigonal the principal symbol of but also gives an explicit and extremely simple expression of the lower-order terms:

Proposition 4.2.

Let , be an admissible reference state, and .

The following two assertions are equivalent:

i.

the pair solves on ;

ii.

the pair solves on , with

where .

Notation 4.3.

For all smooth enough, we write

where and are as defined in (Equation 4.5)-(Equation 4.6), so that .

The coefficient appearing in the trigonal operator obviously plays an important role. It is therefore interesting to give it a physical meaning. The pair being given as in Prop. 4.2, we can define a velocity potential by solving the Laplace equation (Equation 1.7) in the fluid domain with Dirichlet condition at the surface and homogeneous Neumann condition at the bottom. In accordance with (Equation 1.10), we introduce the pressure as

The following proposition shows that if solves the water-waves equations (Equation 4.1) at some time , then the pressure defined in (Equation 4.8) vanishes at the surface and the normal derivative of the pressure at the surface coincides with . The condition we shall impose later (see (Equation 4.10)) coincides therefore with the traditional Taylor criterion Reference 33Reference 3Reference 22Reference 37 that the interface is not accelerating into the fluid region more rapidly than the normal component of the gravity.

Proposition 4.4.

Let and be an admissible reference state. If for some , solves the water-waves equations Equation 4.1, then , defined in Equation 4.8, satisfies

Proof.

Let us remark that

where is defined in (Equation 4.5). It follows therefore from (Equation 4.8) that

From this expression, one deduces easily that if solves (Equation 4.1) at time .

We now prove the second statement of the proposition. One has by definition

At time , we just saw that , from which one deduces easily that . Now, from the definition (Equation 4.8) of , one computes

Remarking that

one obtains finally , which concludes the proof.

4.2. Study of the trigonal operator .

Because the principal part of has a Jordan block, the Cauchy problem

could be either ill- or well-posed. Such situations have been extensively studied for differential systems (see Reference 16 and the references therein for the study of general non-strictly-hyperbolic problems, and Reference 11 for a more related situation), and seem inherent to the water-waves problem Reference 10Reference 36Reference 37: in order to be well-posed, a Lévy condition is needed on the sub-principal symbol of . Since the operator is positive, the Lévy condition on becomes

where is defined in terms of as in Prop. 4.2. The next proposition shows that under this condition, the Cauchy problem associated to the trigonal operator is well-posed, and that one gets tame estimates on the solution.

Proposition 4.5.

Let , and let be an admissible reference state. Also let and . Then there is a unique solution to Equation 4.9 and for all , there exist , such that

The constants , depend on and through

where is an integer depending only on , and is as in Notation 3.5.

4.2.1. Proof of Prop. 4.5

As is often the case for equations similar to (Equation 4.9) (see e.g. Reference 36Reference 37), we first consider a parabolic regularization of (Equation 4.9):

Even for (Equation 4.12), well-posedness is not straightforward. As in Reference 36Reference 37, we choose to use an iterative scheme to prove it. Let us first introduce the notation

and

so that .

We seek a solution of (Equation 4.12) as a limit of the sequence defined for all as

Well-posedness of Cauchy problems of type (Equation 4.14) is ensured by the next lemma.

Lemma 4.6.

Let , be an admissible reference state, and also let and . For all , the Cauchy problem

admits a unique solution . Moreover, for all there exist and such that

Proof.

In order to perform energy estimates on the equation, we seek a change of unknowns which symmetrizes the operator . Let ; one has (note the importance here of the Lévy condition (Equation 4.10)). The operator is a symmetrizer of in the sense that , with

that is, the principal part of is an anti-adjoint operator of order one. The natural energy associated to the equation is therefore defined as As usual, one computes

Estimate of . By Cauchy-Schwartz and then Hölder’s inequality, one obtains easily

Estimate of . One has

and since the principal symbols of and are the same, we deduce from the decomposition above that the operator is of order one with skew-symmetric principal symbol. Classical results of pseudo-differential calculus yield therefore

Estimate of . Since , one obtains easily and thus

Endgame. Using (Equation 4.15), (Equation 4.16), (Equation 4.17) and (Equation 4.18) one obtains

For large enough (in order for the prefactor of to be negative in the r.h.s. of the inequality above),we have therefore

Now, remark that for some constant depending on , , and . Equation (Equation 4.19) gives therefore the desired energy estimate in the -norm, and it is routine to conclude the proof by classical duality arguments.

Owing to this lemma, we have the following estimate for (Equation 4.14):

From the definition of , one obtains easily, for all ,

so that one has finally, for all ,

Proving the convergence of the iterative scheme (Equation 4.14) is then classical. We have therefore:

Lemma 4.7.

Let and be an admissible reference state satisfying Equation 4.10. Also let and . Then, for all , there exists a unique solution to Equation 4.12.

We now turn to give precise energy estimates on the solution to (Equation 4.12) given by Lemma 4.7.

Let us denote by the spatial part of the operator , so that , and decompose it as with

where and is some real positive constant (which we add here because we will need the operator to control the -norm as in Prop. 3.8).

As in the proof of Lemma 4.6, the strategy consists of symmetrizing the principal part of the operator, namely, . The operator which symmetrizes is given here by

where denotes the square root of the operator . The natural energy to consider here is therefore

In fact, we do not work directly with all : the estimates of Theorem 2.9 show that it is convenient to work with Sobolev spaces , . Instead of taking in the definition above, we change it slightly as

when , we write simply instead of . The link between spaces of finite energy for (Equation 4.21) and Sobolev spaces is made in the next lemma.

Lemma 4.8.

Let and be a reference state satisfying Equation 4.10. Then there exists such that for all and ,

where is as in the statement of Prop. 4.5.

Notation 4.9.

From now on, we always take and write simply instead of .

Proof.

For all , , write with

Upper and lower bounds for are easy to find:

Remark now that , so that using Prop. 3.8 (and assuming that is large enough), one obtains

where , and is as in Notation 3.5.

The lemma follows therefore from (Equation 4.21) and (Equation 4.22)-(Equation 4.24).

Before addressing the heart of the proof, let us recall some useful nonlinear estimates.

Lemma 4.10.

Let and such that . Let and and define as in Equation 4.13. Then:

i.

For all and , one has

ii.

For all , and , one has

Proof.

The first point of the lemma is the classical Kato-Ponce estimate Reference 24. The second one is a consequence of this estimate since one has

Lemma 4.11.

Let and be an admissible reference state satisfying Equation 4.10. Then, for all , the solution to Equation 4.12 satisfies

where the constant is as in the statement of Prop. 4.5.

Proof.

Throughout this proof, we write . We proceed as in the proof of Lemma 4.6. One computes

where the sums are taken over all , .

Estimate of . From the definitions of and , one computes easily

Estimate of . Using the fact that the operator is anti-adjoint, one finds

Using Lemma 4.10 one can control the first term of the r.h.s. and remarking that , one can control also the second one:

Using Hölder’s inequality and Lemma 4.8, one obtains therefore

where, throughout this proof, is a positive constant which depends on the same parameters as in the statement of Prop. 4.5.

Estimate of . Using the fact that the operators and are respectively anti- and self-adjoint, one computes

By Prop. 3.8 we have

we then use Lemmas 4.8 and 4.10, as well as Hölder’s inequality to find

To control , one uses successively Prop. 3.18 and Lemma 4.8 to find

From (Equation 4.28), (Equation 4.29) and (Equation 4.30), we obtain finally

Estimate of . One has

Using the Cauchy-Schwartz inequality and Prop. 3.15, we obtain

where is the same as in Prop. 3.15.

It is then easy to deduce that

For , we proceed as for and find

From (Equation 4.32), (Equation 4.33) and (Equation 4.34), we have therefore

Finally, from (Equation 4.26), (Equation 4.27), (Equation 4.31) and (Equation 4.35) one obtains the estimate:

Estimate of . Without any particular difficulty, this term is bounded from above by

Estimate of . Remark that this term can be decomposed into ; the first term of this decomposition is easy to bound; for the second, we use Prop. 3.19, so that finally

End of the proof. From (Equation 4.25), (Equation 4.36), (Equation 4.38) and (Equation 4.38), we obtain, as in the proof of Lemma 4.6,

When is negative, the estimate of the lemma follows easily from this expression.

We can now prove the well-posedness of (Equation 4.9). In order to do this, we show that the sequence , where denotes the solution to (Equation 4.12), converges to a solution of (Equation 4.9) when .

Let us first prove that is a Cauchy sequence. Let and write . One has

Remark now that, as a first consequence of Lemma 4.11, for all , there exists such that , for all . Applying Lemma 4.11 to yields therefore

From a Gronwall-type argument, we deduce

and it follows therefore from Lemma 4.8 that is a Cauchy sequence in . The sequence is therefore convergent in this space, and the limit solves (Equation 4.9). The estimate given in the proposition is simply obtained by taking in Lemmas 4.8 and 4.11.

4.3. Tame estimates for the water-waves equations

In this section, we give our main result concerning the linearized water-waves equations: the Cauchy problem

is well-posed, and the solution satisfies tame estimates. We first need to introduce two scales of Banach spaces, namely and , in which the estimates can be written simply, and in which a Nash-Moser scheme can be constructed.

Definition 4.12.

Let and . Define the Banach spaces and as

and endow them with the norms

Notation 4.13.

An admissible reference state does not necessarily belong to the Banach scale because is not necessarily in a Sobolev space (though its gradient is). However, we abusively use the notation to denote the quantity

Proposition 4.14.

Let , and be an admissible reference state satisfying Equation 4.10. Also let and . Then there is a unique solution to Equation 4.39. Moreover, for all , , the following estimate holds:

for some depending only on .

Proof.

Denote and let and . Prop. 4.5 asserts that there exists a unique solution to the Cauchy problem (Equation 4.9). Owing to Prop. 4.2, we know that solves the Cauchy problem (Equation 4.39). We now proceed to derive tame estimates on from the energy estimate (Equation 4.11).

Taking in (Equation 4.11), one obtains by a simple Gronwall argument that

for some depending only on . Plugging this expression into (Equation 4.11), and estimating the quantities and which appear in (Equation 4.11) in terms of , and by standard tame estimates, one obtains (taking a larger if necessary),

from which it is easy to deduce (using the formula ),

In order to obtain a control of in we still need to control and in and respectively.

Since one has ; from the expression of given in (Equation 4.4) and the tame estimates of Prop. 3.25, one deduces

which, together with (Equation 4.41), yields

Finally, one has . One can compute from the expression of given in (Equation 4.4) and prove that it is a tame bilinear mapping using Prop. 3.25. Using (Equation 4.41) and (Equation 4.42) we can then obtain a tame estimate on (we do not detail the proof since it does not raise any particular difficulty). Namely,

The proposition is then a consequence of (Equation 4.41), (Equation 4.42) and (Equation 4.43) for all , , . By interpolation, we deduce it for all , .

4.4. On the Lévy condition

As seen in Prop. 4.4, the Lévy condition (Equation 4.10), namely , is equivalent to the traditional Taylor criterion. Early works Reference 27Reference 10Reference 38 assume smallness conditions on , which implies that this criterion holds. One of Wu’s key results Reference 36Reference 37 is that, both for 1D or 2D surface waves, one has indeed as soon as the reference state solves the water-wave equations (Equation 4.1). We investigate in this section if this result extends to the present case of finite depth. We first set some notation.

Let be the lower boundary of the fluid domain. One can define the mapping on as

so that is the inward unit normal vector to at . This mapping is regular and its derivative at is a linear map from into . Since by construction, is an endomorphism of . By definition, the second fundamental form of is defined as

where denotes the usual scalar product of .

In the next proposition, we show that the Lévy condition (Equation 4.10) is satisfied provided that a certain smallness condition holds on the second fundamental form evaluated at the bottom values of the velocity field.

Proposition 4.15.

Let and be an admissible reference state, and denote by the velocity potential associated to . Assume that for some , solves the water-waves equations Equation 4.1 and that

Then there exists such that on .

Remark 4.16.
i.

The velocity potential associated to is found by solving the Laplace equation (Equation 1.7) in the fluid domain, with Dirichlet condition at the surface and homogeneous Neumann boundary condition at the bottom. This latter condition ensures that for all , lives in , so that the expression makes sense.

ii.

If the bottom is flat, then everywhere, and criterion (Equation 4.45) is always satisfied. Thus, in the case of flat bottoms, Wu’s result remains true: the generalized Taylor’s sign condition holds provided that the reference state solves the water-waves equations (Equation 1.11) at time .

iii.

By continuity arguments, Wu’s result can also be extended to “nearly flat” bottoms: no smallness condition on the reference state is required for the generalized Taylor’s sign condition to hold, provided that the bottom parameterization is flat enough (how flat depending on ).

iv.

In 1D, the criterion given in the proposition reads simply

and is therefore always satisfied in the regions where the bottom surface is concave.

v.

As we will see later, Taylor’s sign criterion is a sufficient condition for the well-posedness of the water-waves equations for small times. This condition is almost necessary, but the criterion given in Lemma 4.15 gives only a sufficient condition for Taylor’s sign condition to be satisfied. Its interest lies in its simple geometric form. It is for instance obvious that this sufficient condition is fulfilled for flat or nearly flat bottoms, which is far from transparent if one works directly with Taylor’s sign condition.

Proof.

Recall that where and are given by (Equation 4.5) and (Equation 4.6). Since (and its derivatives involved in and ) vanishes at infinity, so do and ; the acceleration of gravity being strictly positive, one deduces that there exist and such that whenever , which is precisely the property we want to prove. The remainder of the proof consists therefore in showing that there exists such that on the ball .

We know by Prop. 4.4 that , where . Prop. 4.4 also asserts that on the surface; it follows that solves the boundary value problem

The next lemma makes the link between the Neumann condition at the bottom and the second fundamental form (recall that by assumption, belongs to ).

Lemma 4.17.

The velocity potential being defined as above, one has

Proof.

Step 1. Geometric tools. The first step consists in reparameterizing the fluid domain in the neighborhood of . For small enough, one can define the mapping

if is small enough, is a -parameterization of its range . We now want to define the gradient in these new coordinates. Let us denote by the gradient on the submanifold and introduce defined as

One can prove (Reference 5, see also Reference 15 for the 1D case) that for any function defined on one has

where , denotes the orthogonal projection of on (which is unique if is small enough) and . From (Equation 4.47), it follows in particular that

and that the tangential component of is exactly .

Step 2. We now use the tools introduced above to prove the result. According to (Equation 4.48) and with the same notation as in the first step, one finds . By definition, one also has , so that using (Equation 4.47), one obtains

Using (Equation 4.46), this yields

Since by (Equation 4.48) we have and because by assumption is tangent to , it follows from (Equation 4.49) that , which is the result claimed in the lemma.

Remarking that , the assumption made in the statement of the proposition ensures that . Now, remark that is subharmonic because ; whenever reaches its minimum, it is therefore necessarily on the boundary of the fluid domain and at such a point the outward normal derivative is strictly negative. From the observation made above, cannot reach its minimum on . Its minimum is therefore reached on the surface, where vanishes identically. Hence, is positive in the fluid domain. Moreover, any point of the surface being a minimum for the subharmonic function , one has everywhere on the surface.

As said above, one has . It follows that one has everywhere on . By a continuity argument, there exists such that for all in the ball . Taking concludes the proof of the proposition.

5. The nonlinear equations

In this section, we construct a solution to the water-waves equations. The crucial step is the tame estimate on the linearized equation proved in the previous section. The iterative scheme we use here is of Nash-Moser type. We first state a Nash-Moser implicit function theorem in Section 5.1 and then use it to solve the water-waves equations in 5.2.

5.1. A simple Nash-Moser implicit function theorem

For the sake of simplicity, we do not use an optimal form of the Nash-Moser theorem. A very simple version of this result can be found in Reference 31; for the sake of completeness, we reproduce here this result.

Let and , be two scales of Banach spaces and denote , . Assume also that there exist some smoothing operators satisfying for every , and and ,

We also assume that whenever .

Theorem 5.1.

Let and assume that there exist , an integer , a real number and constants and such that for any ,

Moreover, one assumes that for every such that , there exists an operator satisfying for any , and

Then if is sufficiently small (with respect to some upper bound of , and where depends only on ), there exists a function such that .

Remark 5.2.

The proof of Reference 31 shows in fact that and that for all , assuming that instead of ensures the existence of a solution instead of .

5.2. Resolution of the water-waves equations

We are now ready to state the main theorem of this paper (recall that denotes the second fundamental form of the bottom, as defined in (Equation 4.44)):

Theorem 5.3.

Let , and be such that , with ( depending only on ). Assume moreover that

and

where is the velocity field associated to . Then there exists and a unique solution to the water-waves equations Equation 1.11 with initial conditions and such that .

Remark 5.4.
i.

The initial velocity field associated to is given by the expression , where is the velocity potential found by solving the Laplace equation (Equation 1.7) in the fluid domain with Dirichlet condition at the surface and homogeneous boundary condition at the bottom.

ii.

In the case of flat bottoms, everywhere and the assumption on made in the theorem is always satisfied. For uneven bottoms, the smallness assumption made on is weaker than the smallness assumptions made, in the case of 1D surface waves, by Yosihara Reference 38.

iii.

One can replace the assumption on by the (sharper) assumption that on , where is defined in (Equation 4.7) and , with and defined as in (Equation 4.3).

iv.

It is physically reasonable to assume that the velocity decays at infinity, but it would be too restrictive to suppose that the velocity potential also does. This is why we take such that , and not simply .

Proof.

The result is obtained as a consequence of the Nash-Moser Theorem 5.1. We work here with the scale of Banach spaces and given in Def. 4.12. It is classical that is equipped with a family of smoothing operators satisfying (Equation 5.50). Direct use of Nash-Moser’s theorem would restrict us to the case of small initial data . To avoid this, we proceed as in Reference 21 (p. 195), exploiting the fact that the water-waves equations are solvable at . Given any initial condition such that , one can find such that

We then define as and introduce the mapping :

so that . Clearly, if , then furnishes a solution to the Cauchy problem (Equation 1.11) with initial condition .

Let us check that the assumptions of Theorem 5.1 are satisfied. One has, for all ,

From the explicit expression of given by (Equation 4.3) and the tame estimates on the Dirichlet-Neumann operator and its derivatives given in Theorem 3.6 and Prop. 3.25, it is easy to deduce that for all ,

(note the above estimate only involves the gradient of , which is made possible by Theorem 3.6; see Remark 3.7).

Taking and some , the condition implies that and hence remains bounded. Defining as the supremum of all the constants which appear in (Equation 5.53) when remains in the ball gives therefore the first condition of (Equation 5.51).

For all , one has

and

checking that the last two conditions of (Equation 5.51) are satisfied is thus obtained in the same way as for the first one, using Prop. 3.25.

We now turn to check condition (Equation 5.52). From the expression of given in (Equation 5.54), it is obvious that the right inverse must be defined as

In order to deduce the estimate (Equation 5.52) from Prop. 4.14, we must show that for all in the ball , is an admissible reference state satisfying (Equation 4.10) uniformly, i.e. that there exists and such that

and

where is as defined in (Equation 4.7).

Lemma 5.5.

Under the assumptions of the theorem, there exists such that if , then Equation 5.55 and Equation 5.56 are satisfied (for a possibly smaller ).

Proof.

To prove (Equation 5.55), write , so that using the assumption made on the initial data, , where we used the fact that . Sobolev embeddings then yield , from which the conclusion is easy.

To prove (Equation 5.56), remark that . It follows that . Since by construction, solves the water-waves equations (Equation 1.11) at time , we deduce from Props. 4.4 and 4.15 that there exists such that . The end of the proof is then straightforward.

This lemma shows that the estimate (Equation 5.52) assumed in Theorem 5.1 is a consequence of Prop. 4.14 (taking a larger if necessary). We can therefore use Theorem 5.1, which asserts that one can solve the equation provided that for some . Now, recall that and that, by construction, . One has therefore

which, taking a smaller if necessary, is smaller than .

We have therefore proved the existence of a solution to , i.e. a solution to the water-waves equations (Equation 1.11); the case of finitely regular solutions is handled as in Remark 5.2.

We now turn to prove uniqueness. Let and be two solutions in , for some , being as above. The difference solves therefore

Using Prop. 3.25, it is easy to obtain that for all , one has , where the constant depends on the norm of and in . Proceeding as in the proof of Prop. 4.14, one obtains the estimate

for some integer . Bounding from above by and using a classical Gronwall argument yields , whence the uniqueness.

Appendix A. Proof of Lemma 2.8

Owing to Lemma 2.5, the nonconstant coefficients of are of the form (up to a multiplicative constant)

It is clear that one can write , with and , so that

Similarly, one can write and with

and

It follows easily that

which achieves the proof of the first estimate of the lemma. We now turn to estimate the Sobolev norms of and . Remark that they are both of the form , with , and

Let us denote . For all , , one can show by induction that is a sum of terms of the form

where , satisfy the relation

Decomposing into , one obtains the following estimate:

where the are such that .

Let be defined as

so that by (Equation A.3), one has .

If , then necessarily

and therefore

If , then remark that

Denoting by the -norm which appears in (Equation A.4) and using Young’s inequality, one has therefore

Recalling that for all , one has

and using (Equation A.5), it follows that

Plugging the estimates (Equation A.1) into this inequality and using (Equation A.4) and (Equation A.6), one obtains the second estimate of the lemma.

Acknowledgements

I want to address my warmest thanks to Guy Métivier who made precious comments about this work. I also acknowledge stimulating discussions with J. Bona, F. Boyer, G. Carbou, T. Colin and J.-L. Joly.

Table of Contents

  1. Abstract
  2. 1. Introduction
    1. 1.1. Presentation of the problem
    2. 1.2. Presentation of the results
    3. Theorem 1.1.
    4. 1.3. Notation
  3. 2. Elliptic boundary value problems on a strip
    1. 2.1. Reduction to an elliptic equation on a flat strip
    2. Assumption 2.2.
    3. Definition 2.3.
    4. Lemma 2.5.
    5. Lemma 2.6.
    6. Proposition 2.7.
    7. Lemma 2.8.
    8. 2.2. Variable coefficients elliptic equations on a flat strip
    9. Theorem 2.9.
    10. Lemma 2.11.
    11. Lemma 2.12.
    12. 2.3. Regularizing diffeomorphisms
    13. Proposition 2.13.
  4. 3. The Dirichlet-Neumann operator
    1. 3.1. Definition and basic properties
    2. Definition 3.1.
    3. Proposition 3.4.
    4. Theorem 3.6.
    5. Proposition 3.8.
    6. 3.2. Symbol of the Dirichlet-Neumann operator
    7. Theorem 3.10.
    8. Lemma 3.12.
    9. Lemma 3.13.
    10. Lemma 3.14.
    11. 3.3. Commutator estimates
    12. Proposition 3.15.
    13. Lemma 3.17.
    14. Proposition 3.18.
    15. Proposition 3.19.
    16. 3.4. Shape derivative of the Dirichlet-Neumann operator
    17. Theorem 3.20.
    18. Lemma 3.22.
    19. Lemma 3.24.
    20. Proposition 3.25.
  5. 4. The linearized water-waves equations
    1. 4.1. Trigonalization of the linearized system
    2. Definition 4.1.
    3. Proposition 4.2.
    4. Proposition 4.4.
    5. 4.2. Study of the trigonal operator .
    6. Proposition 4.5.
    7. Lemma 4.6.
    8. Lemma 4.7.
    9. Lemma 4.8.
    10. Lemma 4.10.
    11. Lemma 4.11.
    12. 4.3. Tame estimates for the water-waves equations
    13. Definition 4.12.
    14. Proposition 4.14.
    15. 4.4. On the Lévy condition
    16. Proposition 4.15.
    17. Lemma 4.17.
  6. 5. The nonlinear equations
    1. 5.1. A simple Nash-Moser implicit function theorem
    2. Theorem 5.1.
    3. 5.2. Resolution of the water-waves equations
    4. Theorem 5.3.
    5. Lemma 5.5.
  7. Appendix A. Proof of Lemma 2.8
  8. Acknowledgements

Mathematical Fragments

Equation (1.1)
Equation (1.2)
Equation (1.3)
Equation (1.4)
Equation (1.6)
Equation (1.7)
Equation (1.10)
Equation (1.11)
Equation (1.12)
Equation (1.13)
Equation (1.14)
Equation (2.2)
Equation (2.3)
Equation (2.4)
Notation 2.1.

For all open sets , we denote by , and the canonical norms of , and respectively. When no confusion is possible on the domain , we write simply , and .

Equation (2.5)
Equation (2.6)
Assumption 2.2.

One has with and . Moreover, there exists such that on .

Remark 2.4.

The most simple diffeomorphism between and is given by

and hence . If and , it is clear that is -regular, with , , and .

Equation (2.7)
Lemma 2.5.

Suppose that the mapping , given by Equation 2.6 satisfies Assumption 2.2. Let with satisfying Equation 2.2. Then the equation holds in if and only if the equation holds in , where and are deduced from and via formula Equation 2.7, and , with

Moreover, one has, for all ,

Equation (2.9)
Equation (2.10)
Equation (2.11)
Lemma 2.6.

Suppose that the mapping , given by Equation 2.6, satisfies Assumption 2.2. For all , one has

Proposition 2.7.

Suppose that the mapping , given by Equation 2.6, satisfies Assumption 2.2. Then is a (variational, classical) solution of Equation 2.3 if and only if given by Equation 2.7 is a (variational, classical) solution of

where is as given in Lemma 2.5.

Lemma 2.8.

Let and assume that the mapping , given by Equation 2.6, is -regular. Then one can write with , and

Equation (2.13)
Equation (2.15)
Theorem 2.9.

Let , . Let , and .

i.

If satisfies Equation 2.15, then there exists a unique solution to Equation 2.13. Moreover,

ii.

If and are such that satisfies Equation 2.15, then there exists a unique solution to Equation 2.13. Moreover, when ,

where .

Remark 2.10.
i.

The proof below shows that the quantity can be estimated more precisely than . Namely, one can replace the quantities and in both estimates of the theorem by and respectively. This remark is very useful when giving estimates on the Dirichlet-Neumann operator.

ii.

The second estimate of the theorem remains of course valid when , but in that case, the first estimate of the theorem is more precise.

Equation (2.16)
Equation (2.18)
Equation (2.19)
Equation (2.20)
Equation (2.21)
Equation (2.22)
Equation (2.24)
Equation (2.25)
Equation (2.26)
Equation (2.27)
Equation (2.28)
Lemma 2.11.

Let and . Then one has

Lemma 2.12.

Let , and . If solves the boundary value problem Equation 2.13, then

and

Proposition 2.13.

Let , , and let , . If there exists such that on , then there exists a diffeomorphism of the form Equation 2.6 such that

is -regular (with );

one has ;

one has and

Remark 2.14.
i.

The diffeomorphism provided by this lemma is a perturbation of the trivial diffeomorphism given in Remark 2.4. The -component remains unchanged, and the behavior at the surface is exactly the same. However, the Sobolev component is half a derivative smoother here than for the trivial diffeomorphism (where it has the smoothness of ). This is why we say that the diffeomorphism is “regularizing”.

ii.

Note that if for some , and if the condition is satisfied uniformly in , then one has and . This will be used in the proof of Prop. 3.19.

iii.

If with is such that on , then one can find a neighbourhood of in such that for all , . To each of these , one can associate a regularizing diffeomorphism by Prop. 2.13. The proof shows that if is small enough, then the mapping is affine. This mapping is therefore smooth and (using the notation of the proof) one can check that for all , one has for some . Hence,

Equation (2.29)
Equation (2.30)
Equation (2.31)
Equation (2.32)
Equation (3.1)
Equation (3.2)
Definition 3.1.

Let , and assume that satisfy condition (Equation 3.1). We define the Dirichlet-Neumann operator to be the operator given by

where denotes the solution of (Equation 3.2).

Equation (3.3)
Notation 3.3.

We denote by the solution of the b.v.p. (Equation 3.3).

Proposition 3.4.

Under the same assumptions as in Def. 3.1, one has

where is as defined in Notation 3.3.

Notation 3.5.
i.

When a bottom parameterization () is given, we generically write .

ii.

For all , we denote generically by (resp. ) constants which depend on and (resp. , and ).

Theorem 3.6.

Let and be two continuous functions satisfying Equation 3.1. Then:

i.

For all , if , then for all such that , one has

ii.

For all , if and if , then

for all such that , and where we used Notation 3.5.

Remark 3.7.

Note that the DN operator is defined for functions whose gradient is in some Sobolev space, but which are not necessarily in a Sobolev space themselves.

Equation (3.4)
Equation (3.5)
Proposition 3.8.

Let satisfy Equation 3.1. Then:

i.

The operator is self-adjoint:

ii.

The operator is positive:

iii.

We also have the estimates

and for all , where is given in Lemma 2.5, one has

Equation (3.6)
Theorem 3.10.

There exists an integer , depending only on , such that if and satisfy Equation 3.1, then, for , one has

where is as defined in Notation 3.5 and the symbol is given by

with .

Equation (3.7)
Equation (3.8)
Equation (3.9)
Equation (3.10)
Lemma 3.12.

Let . Let and satisfy condition Equation 3.1. Then

where is as defined in Notation 3.5.

Lemma 3.13.

There exists an integer , depending only on , such that if and satisfy condition Equation 3.1, then

Lemma 3.14.

There exists an integer , depending only on , such that if and satisfy Equation 3.1, then

Equation (3.15)
Equation (3.16)
Proposition 3.15.

There exists an integer , depending only on , such that if and satisfy Equation 3.1, then for all and , , one has

where is as defined in Notation 3.5.

Equation (3.17)
Equation (3.18)
Equation (3.20)
Equation (3.21)
Equation (3.22)
Equation (3.23)
Equation (3.25)
Equation (3.26)
Proposition 3.18.

Let and suppose that and satisfy Equation 3.1. Let be a first-order differential operator on with coefficients in . Then, for all , one has

where is as in Notation 3.5, and denotes the sum of the -norm of all the coefficients of .

Equation (3.27)
Equation (3.28)
Equation (3.30)
Equation (3.31)
Proposition 3.19.

Let and . Let and satisfy Equation 3.1 uniformly for . Then, for all , one has

where is as in Notation 3.5.

Theorem 3.20.

Let and , . Suppose that and satisfy Equation 3.1. Then there exists a neighborhood of in such that for all given , the mapping

is well defined and differentiable. Moreover, for all , one has

where

with

Equation (3.32)
Equation (3.33)
Lemma 3.22.

For all , the function solves

Equation (3.34)
Equation (3.35)
Equation (3.36)
Equation (3.37)
Lemma 3.24.

Under the assumptions and with the notation of the theorem, one has

Proposition 3.25.

Let and , . Suppose that , and satisfy Equation 3.1. Then the mapping

is and the successive derivatives are “tame”:

i.

For all , one has

ii.

For all ,

iii.

Similar estimates hold for , .

Equation (3.38)
Equation (3.39)
Equation (4.1)
Equation (4.2)
Equation (4.3)
Equation (4.4)
Equation (4.5)
Equation (4.6)
Proposition 4.2.

Let , be an admissible reference state, and .

The following two assertions are equivalent:

i.

the pair solves on ;

ii.

the pair solves on , with

where .

Notation 4.3.

For all smooth enough, we write

where and are as defined in (Equation 4.5)-(Equation 4.6), so that .

Equation (4.8)
Proposition 4.4.

Let and be an admissible reference state. If for some , solves the water-waves equations Equation 4.1, then , defined in Equation 4.8, satisfies

Equation (4.9)
Equation (4.10)
Proposition 4.5.

Let , and let be an admissible reference state. Also let and . Then there is a unique solution to Equation 4.9 and for all , there exist , such that

The constants , depend on and through

where is an integer depending only on , and is as in Notation 3.5.

Equation (4.12)
Equation (4.13)
Equation (4.14)
Lemma 4.6.

Let , be an admissible reference state, and also let and . For all , the Cauchy problem

admits a unique solution . Moreover, for all there exist and such that

Equation (4.15)
Equation (4.16)
Equation (4.17)
Equation (4.18)
Equation (4.19)
Lemma 4.7.

Let and be an admissible reference state satisfying Equation 4.10. Also let and . Then, for all , there exists a unique solution to Equation 4.12.

Equation (4.21)
Lemma 4.8.

Let and be a reference state satisfying Equation 4.10. Then there exists such that for all and ,

where is as in the statement of Prop. 4.5.

Equation (4.22)
Equation (4.24)
Lemma 4.10.

Let and such that . Let and and define as in Equation 4.13. Then:

i.

For all and , one has

ii.

For all , and , one has

Lemma 4.11.

Let and be an admissible reference state satisfying Equation 4.10. Then, for all , the solution to Equation 4.12 satisfies

where the constant is as in the statement of Prop. 4.5.

Equation (4.25)
Equation (4.26)
Equation (4.27)
Equation (4.28)
Equation (4.29)
Equation (4.30)
Equation (4.31)
Equation (4.32)
Equation (4.33)
Equation (4.34)
Equation (4.35)
Equation (4.36)
Equation (4.38)
Equation (4.39)
Definition 4.12.

Let and . Define the Banach spaces and as

and endow them with the norms

Proposition 4.14.

Let , and be an admissible reference state satisfying Equation 4.10. Also let and . Then there is a unique solution to Equation 4.39. Moreover, for all , , the following estimate holds:

for some depending only on .

Equation (4.41)
Equation (4.42)
Equation (4.43)
Equation (4.44)
Proposition 4.15.

Let and be an admissible reference state, and denote by the velocity potential associated to . Assume that for some , solves the water-waves equations Equation 4.1 and that

Then there exists such that on .

Equation (4.46)
Equation (4.47)
Equation (4.48)
Equation (4.49)
Equation (5.50)
Theorem 5.1.

Let and assume that there exist , an integer , a real number and constants and such that for any ,

Moreover, one assumes that for every such that , there exists an operator satisfying for any , and

Then if is sufficiently small (with respect to some upper bound of , and where depends only on ), there exists a function such that .

Remark 5.2.

The proof of Reference 31 shows in fact that and that for all , assuming that instead of ensures the existence of a solution instead of .

Theorem 5.3.

Let , and be such that , with ( depending only on ). Assume moreover that

and

where is the velocity field associated to . Then there exists and a unique solution to the water-waves equations Equation 1.11 with initial conditions and such that .

Equation (5.53)
Equation (5.54)
Equation (5.55)
Equation (5.56)
Equation (A.1)
Equation (A.3)
Equation (A.4)
Equation (A.5)
Equation (A.6)

References

Reference [1]
S. Alinhac, P. Gérard, Opérateurs pseudo-différentiels et théorème de Nash-Moser, Savoirs Actuels. InterEditions, Paris; Editions du Centre National de la Recherche Scientifique (CNRS), Meudon, 1991. 190 pp. MR 1172111 (93g:35001)
Reference [2]
J. Bona, T. Colin, D. Lannes, Long wave approximations for water-waves, Arch. Ration. Mech. Anal., to appear.
Reference [3]
T. Beale, T. Hou, J. Lowengrub, Growth rates for the linearized motion of fluid interfaces away from equilibrium, Comm. Pure Appl. Math. 46 (1993), no. 9, 1269–1301. MR 1231428 (95c:76016)
Reference [4]
G. Birkhoff, Helmholtz and Taylor instability, 1962 Proc. Sympos. Appl. Math., Vol. XIII pp. 55–76, American Mathematical Society, Providence, R.I. MR 0137423 (25:875)
Reference [5]
G. Carbou, Penalization method for viscous incompressible flow around a porous thin layer, Nonlinear Anal. Real World Appl. 5 (2004), no. 5, 815–855. MR 2085697 (2005d:76034)
Reference [6]
M. Christ, J.-L. Journé, Polynomial growth estimates for multilinear singular integral operators. Acta Math. 159 (1987), no. 1-2, 51–80. MR 0906525 (89a:42024)
Reference [7]
R. R. Coifman, G. David, Y. Meyer, La solution des conjecture de Calderón Adv. in Math. 48 (1983), no. 2, 144–148. MR 0700980 (84i:42025)
Reference [8]
R. R. Coifman, A. McIntosh, Y. Meyer, L’intégrale de Cauchy définit un opérateur borné sur pour les courbes lipschitziennes, Ann. of Math. (2) 116 (1982), no. 2, 361–387.MR 0672839 (84m:42027)
Reference [9]
R. Coifman, Y. Meyer, Nonlinear harmonic analysis and analytic dependence, Pseudodifferential operators and applications (Notre Dame, Indiana, 1984), 71–78, Proc. Sympos. Pure Math., 43, Amer. Math. Soc., Providence, RI, 1985. MR 0812284
Reference [10]
W. Craig, An existence theory for water waves and the Boussinesq and Korteweg-de Vries scaling limits, Comm. Partial Differential Equations 10 (1985), no. 8, 787–1003. MR 0795808 (87f:35210)
Reference [11]
W. Craig, Nonstrictly hyperbolic nonlinear systems, Math. Ann. 277 (1987), no. 2, 213–232. MR 0886420 (88d:35134)
Reference [12]
W. Craig, D. Nicholls, Traveling gravity water waves in two and three dimensions. Eur. J. Mech. B Fluids 21 (2002), no. 6, 615–641. MR 1947187 (2004b:76018)
Reference [13]
W. Craig, U. Schanz, C. Sulem, The modulational regime of three-dimensional water waves and the Davey-Stewartson system, Ann. Inst. H. Poincaré Anal. Non Linéaire 14 (1997), no. 5, 615–667.MR 1470784 (99i:35144)
[14]
W. Craig, C. Sulem, P.-L. Sulem, Nonlinear modulation of gravity waves: a rigorous approach, Nonlinearity 5 (1992), no. 2, 497–522. MR 1158383 (93k:76012)
Reference [15]
M. Dambrine, M. Pierre, About stability of equilibrium shapes, M2AN Math. Model. Numer. Anal. 34 (2000), no. 4, 811–834.MR 1784487 (2001h:49037)
Reference [16]
H. Flaschka, G. Strang, The correctness of the Cauchy problem, Advances in Math. 6 (1971), 347–379.MR 0279425 (43:5147)
Reference [17]
P. R. Garabedian, M. Schiffer, Convexity of domain functionals J. Analyse Math. 2, (1953), 281–368. MR 0060117 (15:627a)
Reference [18]
J. E. Gilbert, M. A. Murray, Clifford algebras and Dirac operators in harmonic analysis. Cambridge Studies in Advanced Mathematics, 26. Cambridge University Press, Cambridge, 1991. MR 1130821 (93e:42027)
Reference [19]
P. Grisvard, Elliptic problems in nonsmooth domains. Monographs and Studies in Mathematics, 24. Pitman (Advanced Publishing Program), Boston, MA, 1985. MR 0775683 (86m:35044)
Reference [20]
J. Hadamard, Mémoire sur le problème d’analyse relatif à l’équilibre des plaques élastiques encastrées, [B] Mém. Sav. étrang. (2) 33, Nr. 4, 128 S. (1908).
Reference [21]
R. Hamilton, The inverse function theorem of Nash and Moser, Bull. Amer. Math. Soc. (N.S.) 7 (1982), no. 1, 65–222. MR 0656198 (83j:58014)
Reference [22]
T. Hou, Z. Teng, P. Zhang, Well-posedness of linearized motion for -D water waves far from equilibrium, Comm. Partial Differential Equations 21 (1996), no. 9-10, 1551–1585.MR 1410841 (98c:76013)
Reference [23]
I. L. Hwang, The -boundedness of pseudodifferential operators, Trans. Amer. Math. Soc. 302 (1987), no. 1, 55–76.MR 0887496 (88e:47096)
Reference [24]
T. Kato, G. Ponce, Commutator estimates and the Euler and Navier-Stokes equations, Comm. Pure Appl. Math. 41 (1988), no. 7, 891–907.MR 0951744 (90f:35162)
Reference [25]
D. Lannes, Sharp estimates for pseudo-differential operators with symbols of limited smoothness and commutators, Preprint Université Bordeaux 1.
Reference [26]
J.L. Lions, E. Magenes, Problèmes aux limites non homogènes applications. Vol. 1. (French), Travaux et Recherches Mathématiques, No. 17 Dunod, Paris 1968. MR 0247243 (40:512)
Reference [27]
V. I. Nalimov, The Cauchy-Poisson problem. (Russian) Dinamika Splošn. Sredy Vyp. 18 Dinamika Zidkost. so Svobod. Granicami, (1974), 104–210, 254. MR 0609882 (58:29458)
Reference [28]
D. Nicholls, F. Reitich, A new approach to analyticity of Dirichlet-Neumann operators, Proc. Roy. Soc. Edinburgh Sect. A 131 (2001), no. 6, 1411–1433. MR 1869643 (2003b:35216)
Reference [29]
L. V. Ovsjannikov, Cauchy problem in a scale of Banach spaces and its application to the shallow water theory justification. Applications of methods of functional analysis to problems in mechanics (Joint Sympos., IUTAM/IMU, Marseille, 1975), pp. 426–437. Lecture Notes in Math., 503. Springer, Berlin, 1976. MR 0670760 (58:32358)
Reference [30]
M. Sablé-Tougeron, Régularité microlocale pour des problèmes aux limites non linéaires, Ann. Inst. Fourier 36, No.1, 39-82 (1986). MR 0840713 (88b:35021)
Reference [31]
X. Saint Raymond, A simple Nash-Moser implicit function theorem, Enseign. Math. (2) 35 (1989), no. 3-4, 217–226.MR 1039945 (91g:58018)
[32]
G. Schneider, C. E. Wayne, The long-wave limit for the water wave problem. I. The case of zero surface tension. Comm. Pure Appl. Math. 53 (2000), no. 12, 1475–1535. MR 1780702 (2002c:76025a)
Reference [33]
G. Taylor, The instability of liquid surfaces when accelerated in a direction perpendicular to their planes. I., Proc. Roy. Soc. London. Ser. A. 201, (1950). 192–196. MR 0036104 (12:58f)
Reference [34]
M. Taylor, Partial differential equations. II. Qualitative studies of linear equations, Applied Mathematical Sciences, 116. Springer-Verlag, New York, 1996. MR 1395149 (98b:35003)
Reference [35]
F. Trèves Introduction to pseudodifferential and Fourier integral operators. Vol. 1. Pseudodifferential operators, The University Series in Mathematics. Plenum Press, New York-London, 1980. MR 0597144 (82i:35173)
Reference [36]
S. Wu, Well-posedness in Sobolev spaces of the full water wave problem in -D, Invent. Math. 130 (1997), no. 1, 39–72. MR 1471885 (98m:35167)
Reference [37]
S. Wu, Well-posedness in Sobolev spaces of the full water wave problem in 3-D, J. Amer. Math. Soc. 12 (1999), no. 2, 445–495.MR 1641609 (2001m:76019)
Reference [38]
H. Yosihara, Gravity waves on the free surface of an incompressible perfect fluid of finite depth. Publ. Res. Inst. Math. Sci. 18 (1982), no. 1, 49–96. MR 0660822 (83k:76017)

Article Information

MSC 2000
Primary: 35Q35 (Other equations arising in fluid mechanics), 76B03 (Existence, uniqueness, and regularity theory), 76B15 (Water waves, gravity waves; dispersion and scattering, nonlinear interaction)
Secondary: 35J67 (Boundary values of solutions to elliptic PDE), 35L80 (Hyperbolic PDE of degenerate type)
Keywords
  • Water-waves
  • Dirichlet-Neumann operator
  • free surface
Author Information
David Lannes
MAB, Université Bordeaux 1 et CNRS UMR 5466, 351 Cours de la Libération, 33405 Talence Cedex, France
lannes@math.u-bordeaux1.fr
Additional Notes

This work was partly supported by the ‘ACI jeunes chercheurs du Ministère de la Recherche “solutions oscillantes d’EDP” et “Dispersion et non-linéarités ”, GDR 2103 EAPQ CNRS and the European network HYKE, funded by the EC as contract HPRN-CT-2002-00282.

Journal Information
Journal of the American Mathematical Society, Volume 18, Issue 3, ISSN 1088-6834, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on and published on .
Copyright Information
Copyright 2005 American Mathematical Society; reverts to public domain 28 years from publication
Article References

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.