Damping for fractional wave equations and applications to water waves
Abstract.
Motivated by numerically modeling surface waves for inviscid Euler equations, we analyze linear models for damped water waves and establish decay properties for the energy for sufficiently regular initial configurations. Our findings give the explicit decay rates for the energy, but do not address reflection/transmission of waves at the interface of the damping. Still for a subset of the models considered, this represents the first result proving the decay of the energy of the surface wave models.
1. Introduction
Motivated by highly successful numerical methods for damping the surface water wave equations proposed in the work [CFGK05], we wish to establish a theory of absorbing boundary conditions/perfectly matched layers as an approach to a damped linear water wave models. Such methods are essential to ensure that one can numerically simulate long-time behaviors of wave-trains without boundary interference. While nonlinear damping mechanisms have been proposed using nonlinear properties of the water wave models in the work [Ala17, Ala18, ABHK18], implementation of such methods can be numerically very stiff since the nonlinear damping mechanisms involve many spatial derivatives of the underlying models. However, the proposed methods in the work [CFGK05] are extremely non-stiff, which we argue is strongly related to them arising due to linear mechanisms for damping. Such connections should be explored further to fully understand the efficacy of these existing methods. The connection between damped waves and absorbing boundary conditions has long been understood for models with local differential operators, see for instance [Joh21, Nat13] and references therein, so we endeavor here to extend our understanding of damping effects the inherently non-local models that arise the water wave problem.
Let us very briefly recall a Hamiltonian formulation of the evolution of a fluid interface in the gravity-capillary water wave system subject to an external pressure. This equations can be written in terms of the surface height, denoted , and the velocity potential of the fluid restricted to the surface, denoted . In [ACM+22], the authors derive a robust method for numerically solving the Euler equations in very general geometric setting using the coordinate equations
Here is the un-normalized Dirichlet-to-Neumann map, is the tangent angle of the surface, is an arc-length parameter, are the normal and tangential derivatives at the surface and is an external pressure term in which we can introduce damping or forcing on the equations. We will consider especially a form of damping introduced by Clamond-Fructus-Frue-Kristiansen in [CFGK05].
In the case of non-zero surface tension () one can use a prescribed to stabilize small waves similar to the work of Alazard et al, see [ABHK18]. When , we can think of these conditions as numerical boundary conditions that absorb energy and allow for as little reflection as possible. Generically, one numerically solves the water waves problem on a periodic domain of length and take a connected interval on which we will damp the fluid with corresponding indicator function . We will consider here the damping properties of the numerically effective damping term,
as proposed in [CFGK05], which we will denote as Linear Damping of the water wave problem related to the order of regularity required to establish the model equation (1.1) from a full paradifferential diagonalization of the water wave equations, see [ABZ11].
Using the paradifferential formulation of the water waves developed in for instance [ABZ11], one observes the following leading order linear model for damped gravity water waves
(1.1) |
This model can be studied from the classical point of view of scattering theory and perfectly matched layers, though the non-local nature of the operator means that many known techniques fail and more refined tools are required. To that end, we study (1.1) here using propagation estimates in the study of semiclassical scattering operators, which have been developed quite thoroughly in the recent book [DZ19] for operators of the form . However, the non-locality of operators of the form results in some important modifications that we illuminate here. Much of our analysis should be extendable to other non-local wave equation models with appropriate modifications. The well-posedness of a nonlinear model related to (1.1) in the setting of the water waves with surface-tension was established in the recent work of [Moo22], but the strength and speed of damping that arises from such a method is not clear. Here, we are able to prove the polynomial decay of the energy for the linear model.
As discussed in [Ala17], there is a long-standing connection between damped wave equations, absorbing boundary conditions in numerical analysis and the notions of so-called control and observability estimates for a given equation on the support of the damping function.111See for instance [Ala17] or [BZ19] for careful definitions of control/observability estimates if the reader is unfamiliar. While our approach here does not use such an estimates directly, some important surveys and results in this direction for a variety of models that have similar proof strategies include [ABBG+12, BLR92, BZ19, BZ04, Mac21, Phu07, RT75, Zua05, Zua07]. We also highlight a specific version of absorbing boundary condition was introduced in [JKR14], though we point out that the model we consider here can be easily generalized to higher dimensional water wave models.
A related damping model is of the form
(1.2) |
This results from a similar paradifferential diagonalization of a damping that is guaranteed to lead to nonlinear damping by consideration of the Hamiltonian energy for the water wave equations, see [Ala18]. The techniques we apply here can likely be applied to study damping of this form with appropriate modifications, in particular with respect to the required regularity of the initial data. However, for the sake of smoothness of exposition, we focus only on equation (1.1) in our analysis below.
1.1. Main results
Here, we study a linear model for the damped water wave equation explicitly framed on a periodic domain, where we are able to give quantitative estimates on the damping rates of (1.1). To be precise, let be the circle and satisfy . For , we define the fractional Laplacian operator as follows
(1.3) |
For , we consider the damped fractional wave equation
(1.4) |
The energy of the solution to (1.4) is defined by
(1.5) |
For a localized damping function (meaning that ), the geometric control condition fails. In this case, we show that for any , the energy of the solution decays as . By constructing quasi-modes, we show that this polynomial rate is sharp for localized damping.
In order to explore the possibility of larger decay rates, we now consider damping with finitely many zeros. In this case, the energy decay rates depend on the “switching on” behavior of the damping function near zeros of . To quantify this connection, we introduce the following definition:
Definition 1.1.
We say , has finite degeneracy, if has finitely many zeros , , and for each , there exists , such that
(1.6) |
Given , we denote by the Hölder space of those continuous periodic functions such that
We are now ready to state the main results on the energy decay rates.
Theorem 1.
Remarks. 1. The energy decay rate for localized damping (see (1.7)) is sharp. This follows from the sharpness of the resolvent bound (1.12) (see Remark after Theorem 2) and the equivalence between resolvent bounds and polynomial decay rates proved by Anantharaman–Léautaud (see [AL14, Proposition 2.4], which is also stated in §5 in the current paper).
2. Suppose for . If , then the polynomial rates in (1.7), (1.8) can be improved to and , respectively. See Remark in §5.
3. The first result in Theorem 1 can be generalized to tori of higher dimensions . Indeed, suppose , , , and
(1.9) |
where is the cosphere bundle of . Then there exists such that for any , we have
The proof is the same as in the 1D case presented in this paper.
4. For comparison, we recall the usual damped wave equation on
(1.10) |
It is known that under the same dynamical condition (1.9), the energy of the solution to (1.10) decays exponentially (see for instance [Zwo12, Theorem 5.10]). A result [Phu07] for a damped wave equation on a bounded domain results in polynomial rates for damped wave equations without geometric control conditions provided an observability estimate holds and with very minimal regularity requirements on the damping function.
The energy decay rates for the damped equation (1.4) are closely related to the resolvent estimates (via semi-group theory or Fourier transform, see for instance [AL14, Zwo12]) of the stationary operator
(1.11) |
Theorem 1 follows from the following resolvent bounds.
Theorem 2.
Remark. Let us prove that (1.12) is optimal for cut-off function which does not have full support on that is, is a localized damping function. To see this, let be such that and . For all integers , define . Since , we have
We will establish in §2.1 that the commutator . This means that is a bounded operator and hence there is such that
Therefore, there is no such and , such that , implies
This shows that (1.12) is optimal for localized damping functions.
As we show in Lemma 5.2,
is a meromorphic family of operators with finite rank poles. The poles of are called resonances for . We denote the set of resonances by . Using Theorem 2 and Grushin problems, we give the following description of the distribution of resonances:
Theorem 3.
Suppose with , , . Then
-
1.
There exists , such that
(1.14) Moreover, if has finite degeneracy as in Definition 1.1, then for any , there exists , such that
(1.15) -
2.
For , we denote and the set of resonances of . Then for each , , there exists , an open neighborhood of , and , such that
(1.16)
Remarks. 1. Despite the asymptotic expansion of the resonances in (1.16), Theorem 3 does not imply the existence of a resonance-free strip with constant width for , because the expansions are not uniform in .
2. For the damped wave equation framed on a compact manifold,
where is the usual Laplace-Beltrami operator, the distribution of resonances and corresponding energy decay rates have been studied in the works [MM82], [Sjö00] and [Ana10]. In [MM82], Markus–Matsaev established a Weyl law for the resonances in terms of counting how many resonances can exist at a given energy. In [Sjö00], Sjöstrand further proved that the imaginary parts of the resonances “concentrate” (in a suitable sense) on the half average (with respect to the Liouville measure on the cosphere bundle of the manifold) of the damping function. In [Ana10], Anantharaman proved a fractal Weyl law for the resonances and studied several inverse problems. It is an important topic of future work to see if analogous results hold for damped fractional wave equations of the form studied here.
1.2. Outline of Paper
In §2, we recall some properties of the semiclassical calculus for operators on the torus that we will require for our analysis. The propagation estimates required to prove the resolvent estimate and the resulting resolvent mapping properties are proven in §3. We prove a stronger resolvent estimate for damping functions that vanish to a given order at a finite number of points on in §4. In §5, we give an overview of the proof from [AL14] (simplified in our particular setting) that the resolvent bounds proved are equivalent to energy decay bounds for the damped fractional wave equation. To give insight into the properties of the resolvent, in §6 we prove that the low energy resonances can be approximated well by a finite approximation that can be constructed explicitly using a Grushin problem. Finally, in §7 we provide some numerical simulations demonstrating various aspects of our theorems in practice, both for the exact linear fractional wave model, as well as for water wave models with Clamond Damping. This includes a means of approximating the low-energy resonances and comparing to the asymptotics in the previous section.
Acknowledgements. We would like to thank Jared Wunsch for many helpful discussions, and Ruoyu P. T. Wang for showing us useful references. J.L.M was supported in part by NSF Applied Math Grant DMS-1909035 and NSF Applied Math Grant DMS-2307384.
2. Semiclassical analysis on the circle
2.1. Semiclassical pseudodifferential operators
We consider the following symbol class
here is the cotangent bundle of , , and . With the best constants as semi-norms, the class is a Fréchet space. For , which could depend on with semi-norms uniform in , we define its semiclassical quantization by
(2.1) |
Let , be the Fourier coefficients of , as in (1.3). Then we have
(2.2) |
Similarly, we define its microlocal quantization by
We also use the notations , for the microlocal or semiclassical quantization of . Let be the set of pseudodifferential operators that consists of semiclassical quantizations of all symbols in . We define the semiclassical symbol map
Later we will identify with if no ambiguity.
Remark. is a microlocal operator with symbol such that , . However, is not a semiclassical operator, as its “semiclassical symbol” is not a smooth function.
Consider a family of symbols which is bounded in and is in . Let be the semiclassical quantization of . The semiclassical wavefront set of is defined to be the essential support of , that is, if and only if there exists a neighborhood of in , such that for any ,
We also define the semiclassical elliptic set of by
We record the following formula for the symbol calculus of operator compositions: if , , then
In particular, we can compute the commutator of two pseudodifferential operators
(2.3) |
where .
2.2. Semiclassical Fourier multipliers
When the symbol in (2.2) does not depend on , we say is a semiclassical Fourier multiplier. In this section, we generalize the definition of semiclassical Fourier multipliers to bounded symbols and prove a commutator estimate for semiclassical Fourier multipliers and functions with Hölder continuity.
Given a bounded function , we define the semiclassical Fourier multiplier by
for . We call the symbol of the semi-classical Fourier multiplier .
Proposition 2.1.
Let be a smooth function, bounded together with all its derivatives, which in addition vanishes on a neighborhood of the origin. Consider two real numbers . There exists a constant such that, for all and for all ,
(2.4) |
Proof.
We introduce the symbols defined by . The key point is to observe that is a bounded family in . The wanted estimate (2.4) will then be a direct consequence of the following lemma.
Lemma 2.2.
Consider two real numbers and such that . For any bounded subset of , there exists a constant such that for all symbol , all , and all ,
(2.5) |
where is the Fourier multiplier with symbol .
Proof.
To prove this result, it is convenient to use the paradifferential calculus of Bony [Bon81] and the Littlewood-Paley decomposition. We start by introducing some notations. Fix a function with support in the interval and equal to when . Then set which is supported in . Then, for all , one has , which one can use to decompose tempered distribution (this setting includes in particular periodic function ). For , we set
We also use the notation for (so that ).
Given a function , denote by denotes the multiplication operator and denote by the operator of paramultiplication by , defined by
Now rewrite the commutator as
The claim then follows from the bounds
(2.6) | |||
(2.7) | |||
(2.8) |
where maps bounded sets to bounded sets. The estimate (2.8) is a direct result of Plancherel Theorem. We refer the reader to [Hör97, Proposition 10.2.2] and [Hör97, Theorem 9.6.4′] for the proof of (2.6). Let us now prove (2.7). To do so, observe that
We then use the Bernstein’s inequality and the characterization of Sobolev and Hölder spaces in terms of Littlewood-Paley decomposition, to write
where we used the assumption to insure that the series converges. ∎
This concludes the proof of Proposition 2.1. ∎
3. Resolvent estimates for localized damping
This section is devoted to proving the resolvent bound for localized damping, that is, the first part of Theorem 2. This resolvent bound gives energy decay for solutions to (1.4) when using [AL14], and the proof of the energy decay is streamlined in §5.
To take advantage of semiclassical analysis, we introduce the semiclassical rescaling
(3.1) |
and define
(3.2) |
We notice that
We start by stating an equivalent version of the first part of Theorem 2 in the semiclassical scale.
Proposition 3.1.
Suppose with . Let be as in (3.2). Then there exist , and such that, for all , all complex number and for all , there holds
(3.3) |
Proof of Proposition 3.1.
We denote by various constants independent of and whose value may change from line to line. We write to say that for such a constant .
We first assume that is a real number with for some sufficiently small.
1. Estimate of .
We claim that
(3.4) |
To see this, observe that, by definition of we have
Consequently, by taking the -scalar product with , we get
Therefore, it follows from the Cauchy-Schwarz inequality that
Since , this immediately implies the wanted estimate (3.4).
2. Propagation estimates.
The estimate of the remaining component is divided into two steps. We begin with the most delicate part, which consists in estimate the microlocal component of where the operator is not elliptic. The analysis will therefore rely on a propagation argument.
More precisely, consider a cut-off function satisfying
We want to estimate the -norm of . We claim that, for some exponent ,
(3.5) |
To prove this claim we use two different kinds of localization.
Localization in frequency. We further decompose the problem into waves traveling to the left and waves traveling to the right. To do so, consider two cut-off functions with and such that
To obtain the wanted estimate (3.5), it is sufficient to prove that and are bounded by the right-hand side of (3.5).
These two terms will be treated similarly (see the explanations at the end of this step) and for notational simplicity we focus on the estimate of . Our aim is thus to prove that, for some exponent , we have
(3.6) |
Localization in space. In addition to the previous localization in frequency, we see that to prove (3.6), by using a suitable partition of unity, it is sufficient to prove that, for any point , there exist and a function with such that
(3.7) |
We will use suitable cut-off functions , as given by the following
Lemma 3.2.
Assume that . Then there exist two functions such that
and moreover is such that its derivative satisfies for some functions satisfying and

Proof of Lemma 3.2.
Introduce a -periodic function which is even and such that
Given three parameters and to be determined, define
Now pick such that for all and set
Now let be the unique function with mean value and such that . We then set and choose and small enough. ∎
Commutator argument. Given the functions and as introduced above, we consider the operator defined by
(3.8) |
The idea is to exploit the fact that is self-adjoint to write under the form of a commutator:
(3.9) |
We then notice that by the assumption on the support of , we have which in turn implies that
We thus end up with
Once this formula is established, one can compute this commutator using only the Leibniz formula. Indeed, directly from the definitions of and (see (3.8)), we have (noticing that )
By combining the previous identities and using again the fact that , we conclude that
Now we use the special form of the function , that is the fact that . Introduce the semiclassical operators and defined by
Then, we have
(3.10) |
Having analyzed the first term in the right-side of (3.9), we now estimate the second one. To do so, we begin by writing the latter under the form
(3.11) |
Since , one has the obvious inequality
To estimate the commutator we shall make use of the assumption that belongs to the Hölder space for some exponent . Let . It follows from Proposition 2.1 applied with that
(3.12) |
Since is bounded uniformly in , it follows that
(3.13) |
Now, by combining (3.9) together with (3.10) and (3.13), we find
By using again the fact that is bounded uniformly in , by dividing each side of the previous inequality by and using the Cauchy-Schwarz inequality, this yields
It remains to bound (resp. ) from below (resp. above). To do so, set . Since belongs to , it follows from Proposition 2.1 applied with and any arbitrary real number in , that
Since , we deduce that
Now consider the function as given by Lemma 3.2. Since can be written under the form
we have
Using again Proposition 2.1 to estimate the commutator, we get that
On the other hand, by applying Proposition 2.1 with , we deduce that
This immediately implies that
Consequently, we end up with
(3.14) |
Lastly, to prove similar estimates for , it suffices to construct auxiliary functions , as in Lemma 3.2 with mere changes (see Figure 2)

3. Elliptic estimates.
Having estimated the main contribution of the frequencies of size near in the semiclassical scale, we now turn to the estimation of the low and high frequency components. More precisely, we want to estimate the -norms of and where and are semiclassical Fourier multipliers with symbols and in , such that
(3.15) |
High frequency estimate. We begin by estimating . To do so, we write that, by the definition of ,
Since , we have
On the other hand, since on the support of , directly from the Plancherel theorem, we have
Therefore, for and small enough, we get
Low frequency estimates. The low frequency component is estimated in a similar way. Namely, we write
and then observe that , to obtain
Since on the support of , directly from the Plancherel theorem, we have
Therefore, for , small enough, we get
4. End of the proof when is a real number.
The Plancherel theorem implies that
Therefore, it follows from the previous -estimates for and that
and hence we have
Now, it follows from (3.4) and (3.5) that
(3.16) |
We conclude that
Consequently, taking small enough, we obtain
This concludes the proof of the desired result (3.3) when .
5. End of the proof when is a complex number. Lastly, if is a complex number, we notice that
Thus for any complex number such that and , we have
Consequently, for all and small enough, the estimate (3.3) still holds. ∎
4. resolvent bounds for finitely degenerate damping
The goal of this section is to prove the second part of Theorem 2. We start by rewriting the resolvent bounds in the semiclassical scale.
Proposition 4.1.
Suppose has finite degeneracy as in Definition 1.1. Then for any , there exists , , , such that for , , , such that for any , we have
We reduce the proof of the resolvent bound for to the resolvent bounds for . The basic idea is that near the characteristic set of , the frequency is comparable to , hence we can “replace” by . Away from the characteristic set of , we have the ellipticity of . is easier to analyze as it is an ordinary differential operator and we can use integrating factors to simplify .

Proposition 4.2.
Proof of Proposition 4.1 using Proposition 4.2.
Let , , be as in the proof of Proposition 3.1. Then for , , we have
Thus for sufficiently small, using Plancherel theorem, we have
(4.2) |
A similar argument shows that
(4.3) |
Notice that we have the identity
Use Proposition 4.2 and we have
Since , we know . Therefore
(4.4) |
Gathering estimates (4.2), (4.3), (4.4), we find that
Thus when is sufficiently small, we have
This proves Proposition 4.1 when . The proof of Proposition 4.1 is completed by applying the same triangle inequality argument as in the last step of the proof of Proposition 3.1 when is complex. ∎
The rest of this section is devoted to proving Proposition 4.2, that is, the spectral gap and the resolvent bound for . The main idea is to consider a second microlocalization near zeros of : away from the zeros of , has lower bounds; near zeros of , we use the smallness of the second microlocalization and the explicit Green’s formula for .
We start by recording a property of when it has finite degeneracy.
Lemma 4.3.
Let be as in Definition 1.1, then there exists , , such that
-
1.
For any , , we have
-
2.
If is a local minimum of such that , then we have
Proof.
The first conclusion follows from Taylor expansion
and a similar expansion for .
For the second claim, we notice that is a non-vanishing continuous function on . Since is compact, there exists such that for . Thus the second claim can be achieved by shrinking the value of . ∎
Proof of Proposition 4.2.
We only prove for , the proof for is similar.
1. Resonances of .
To study the resonances of , we introduce the integrating factor
One can check that
Notice that has eigenvalues , and is a constant function. Therefore, for , the resonances for must satisfy
We can solve
Thus resonances of lie in .
2. Estimates for away from zeros of .
It suffices to prove the resolvent bound for . When is complex we apply the same triangle inequality argument as in the last step of the proof of Proposition 3.1.
Let , , be the zeros of . For , we denote
Let such that
For , we introduce cut-off functions
Functions , satisfies the following conditions

We first consider the estimates for . Notice that
Here we used the fact that
which follows from Lemma 4.3. Now we find that
Use Cauchy’s inequality and we find
Now notice that
and we conclude
(4.5) |
3. Estimates for near zeros of .
We now turn to estimate . We first solve
Since , we know , when , . Hence
Integrate over and notice that is supported in , and we find
Again notice
and we conclude that
It follows by the definition of now that
(4.6) |
4. End of the proof.
5. Polynomial energy decay
Here, for the benefit of the reader, we give a sketch proof of the equivalence between the resolvent estimates proved above and the corresponding energy and semigroup bounds for the damped wave equation. This gives an overview of the works of [AL14, BD08] taking and , which builds on ideas developed in [Leb94, LR97]. Since we are working in a very explicit setting with relatively simple operators, connecting the resolvent estimate to the damped wave decay can be done in a fairly explicit manner.
To start with, we rewrite the damped fractional wave equation as
(5.1) |
where , is an operator on with domain . The space is equipped with a seminorm .
The following proposition established by Anantharaman–Léautaud [AL14] connects the resolvent estimates and energy decay rates of the damped fractional wave equation.
Proposition 5.1 ([AL14, Proposition 2.4]).
Suppose . Let be as in (1.11). Then for , the following statements are equivalent
-
1.
There exists such that for any , there holds
Notice that .
-
2.
There exists , such that for any , , there holds
One would like to use the semi-group of to solve the damped wave system (5.1), hence it is important to understand the spectrum of . As we see below, the set of resonances of is the same set of resonances of . Hence it suffices to study the resonances of .
Lemma 5.2.
Suppose and is as in (1.11). Then the resolvent
is a meromorphic family of operators with finite rank poles. Moreover, is holomorphic in the region , where
Proof.
Take , . We have the following resolvent identity
(5.2) |
Notice that is invertible. Let
Then is a family of compact operators, hence is a family of Fredholm operators. For , we have
Then for , is a family of invertible Fredholm operators. The analytic Fredholm theory (see for instance [DZ19, Theorem C.8]) implies that
is a meromorphic family of operators with finite rank poles. This and (5.2) show that is a meromorphic family of operators with finite rank poles.
To see that is holomorphic in , we write , , then
If , or , then there exists such that for , we have
Therefore implies . Hence can not be a resonance in this case.
If , or , then for , we compute
Since either or for all , we again see that implies that . ∎
The same result holds for . In fact,
Lemma 5.3.
Let , be as in (5.1). Then the resolvent
is a meromorphic family of operators with finite rank poles and the poles (resonances) are exactly the resonances of . In particular, is a resonance of and .
Proof.
We now sketch the proof of Proposition 5.1 below. For further details of the proof, we refer to [AL14, §4].
Sketch proof of Proposition 5.1.
Notice that is the only resonance of with a nonnegative imaginary part – which correspond to the nondecaying part in the solution. Hence it is natural to split the eigenspace of . For that, we let be the orthogonal projection onto . We define
Then the solution to the damped wave system (5.1) can be expressed in terms of the semi-group of :
(5.4) |
Notice now that
Thus the first statement in Proposition 5.1 is equivalent to the following semi-group bound: there exists such that
(5.5) |
We now recall the result [BT10, Theorem 2.4] by Borichev–Tomilov and realize that (5.5) is equivalent to the following resolvent estimate for : there exists such that
(5.6) |
Use the following identity on for
and we find that
Thus (5.6) is equivalent to the bound for the resolvent of : there exists such that
(5.7) |
6. asymptotics of the resonances for small damping
This section is devoted to proving Theorem 3. Since the first statement is a direct result of the resolvent bounds in Theorem 2, we focus on the proof of the second statement. The main tools we use are Grushin problems – we refer to [SZ07] for an introduction and applications of Grushin problems.
Recall the notation
The operator has resonances , , . Notice that if is a resonance of , then is also a resonance of . Thus we only need to consider the resonances with .
Proof of the second statement of Theorem 3.
For , , we propose the following Grushin problem for in a neighborhood of :
where are given by
Here the inner product on is defined by
For , let be the operator defined by
where is the orthogonal projection onto the orthogonal complement of . We have the following explicit formula for :
(6.1) |
From (6.1) we know is in fact defined and holomorphic in a neighborhood of . Moreover, for near , is a bounded operator. Let be the orthogonal projection onto , then we have
A direct computation then shows that has an inverse
where and are given by
Notice that
Hence for in an open neighborhood of , there holds
Now by [DZ19, Lemma C.3], we know is invertible for in a neighborhood of . We denote the inverse of by
Then is holomorphic near and . Recall that the invertibility of is equivalent to the invertibility of . In fact, there holds the following Schur complement formula
Since is a matrix, its invertibility is much easier to characterize. Let , then we know is holomorphic near and if and only if . Since , we know . By the Weierstrass Preparation Theorem, we know that for in an open neighborhood of , there exist holomorphic functions , and such that
Thus in , the zeros of are
Since , are holomorphic functions, we have the Taylor expansion
(6.2) |
Thus one of the following statement is true:
-
1.
Either both are holomorphic functions of for near , that is,
(6.3) -
2.
Or both have power series expansion in terms of when is near and :
(6.4)
On the other hand, we have the following expansion of (see for instance [DZ19, Lemma C.3]):
(6.5) |
In particular, we know is holomorphic near . This indicates that is holomorphic. Therefore is holomorphic near and in the expansion (6.2), we must have . As a result, in either (6.3) or (6.4), we must have
(6.6) |
Inserting (6.6) in (6.5) gives
where is a matrix and each entry of is of . Use the fact that and we find
from which we solve
This completes the proof. ∎
7. Numerical Simulations
We run our numerical approximations on three models of damping to highlight the theoretical results above. The three damping functions we consider are:
where the function is a low-frequency damping, is a Gaussian damping with exponential decay at high frequency, and is a compactly supported damping function with slowly decaying Fourier modes.
The equations to (1.1) are rather straightforward to discretize as ODE systems in Fourier space, all the time dependent solvers are run by using Fourier pseudospectral methods in space with spatial grid points and integrating in time using the ode45 package in MATLAB with relative and absolute tolerance levels set at to ensure high accuracy of the solutions.
7.1. Demonstration of Polynomial Decay rates
We can observe that decay rates for the damped fractional wave dynamics appear to be numerically very close to the polynomial rates predicted by our estimates for well-chosen initial data. Indeed, in Figure 5, we observe relative decay properties comparing to the polynomial rate by plotting the evolution of for (1.1) using initial conditions that are high frequency and localized far from the damping. In particular, we take
and observe that the decay rates are quite remarkably close to our predicted polynomial rates.
To highlight this polynomial behavior, in Figure 6 we plot for both damping and damping with the same initial condition as above. For the low frequency damping, , we observe similar decay rates to those predicted in Theorem 1, namely close to .




7.2. Approximation of the resonances at low frequency
To begin, by implementing a low frequency approximation scheme, we can analyze the behavior of the resonances at low energy. For that we define
We introduce the discretization of using Fourier modes:
Since is a linear space of dimension , we know for each , is a matrix. Indeed, if we expand using Fourier series
then using the basis of , becomes a matrix
where is the upper-triangular Jordan block.
Because is a quadratic polynomial of with matrix coefficients, we know is polynomial of of order . Therefore is meromorphic function with poles (counting multiplicities). As a result, the inverse matrix
is a meromorphic family of matrices with poles (counting multiplicities). We denote the set of poles of , i.e., zeros of , by . We regard as a low frequency approximation to .
For a given damping function , we can thus numerically construct function whose form we can compute symbolically in MATLAB. We can then find the roots of this polynomial equation in .
We uniformly take as an approximation, and observe the following each of our three experiments. Depending upon the potential, we have a number of resonances concentrating around for as predicted by the asymptotics in §6. The observed resonances are computed using . The resonance with smallest imaginary part is for with respectively. We observed that these resonances were stable under refinement of the approximation parameter . Of course, there is resonance for every example corresponding to the constant solution.
We can compare our asymptotics from §6 to the computed approximate resonances with for a variety of damping functions and parameters . We consider again three experiments again motivated by low frequency damping (), Gaussian damping () and compactly supported damping (), but with varying amplitude depending upon the parameter :
In particular, our experiments above corresponded to for respectively. Figures 7-9 demonstrate quite well that our asymptotics remain quite robust for each of these potentials for small and still give a fair amount of insight especially into the real part of the resonances even for large . These resonances give insight into how states that are low-frequency and overlapping with the damping function decay in a significant fashion under the evolution of (1.1).















7.3. Improved decay for higher regularity initial data
Given our approximate resonance values, if we had a resonance free strip otherwise, we should have that where is defined as in (1.5). We have seen that this is not expected given our observed polynomial decay rates above. However, given highly regular data, we can prove that our decay rate will be better than any polynomial as in the Remark at the end of §5. However, we cannot provide such resonance free strips with our existing propagation estimates and based upon our estimates believe that no such strip exists at high energies. Even so, we can test the evolution of (1.1) for and compare the observed rate of convergence for to those suggested by our computed resonances with very regular initial data. These findings are displayed in Figure 10, where from left to right we plot the numerically observed decay rates of for each , . Specifically, we plot vs. and compare to the exponential decay rate we would expect from the low-energy resonances. To ensure we observe slow decay rates in particular, in each of these simulations, the initial data is taken to be
which by our approximation theory is the function that dominates the low energy resonances we can compute that are closest to the real axis. As suggested by our results, the decay rate does not fit an explicit asymptotic profile related to a specific resonance, though it does decay strongly and appears to decay at close to the expected exponential rate corresponding to the computed resonances at low energy.



7.4. The full water wave problem
To demonstrate how effective the damping we present here can be in the full model however, we also have included a model wave train from a forced-damped water wave model solved with very high precision using the techniques of [ACM+22]. We illustrate this in Figure 11. As can be seen clearly, the damping appears to have a very strong local effect and allows for nonlinear wave trains to exist stably far from the damping.

References
- [ABBG+12] Fatiha Alabau-Boussouira, Roger Brockett, Olivier Glass, Jérôme Le Rousseau, Enrique Zuazua, Sylvain Ervedoza, and Enrique Zuazua. The wave equation: Control and numerics. Control of Partial Differential Equations: Cetraro, Italy 2010, Editors: Piermarco Cannarsa, Jean-Michel Coron, pages 245–339, 2012.
- [ABHK18] Thomas Alazard, Pietro Baldi, and Daniel Han-Kwan. Control of water waves. Journal of the European Mathematical Society, 20(3):657–745, 2018.
- [ABZ11] Thomas Alazard, Nicolas Burq, and Claude Zuily. On the water-wave equations with surface tension. Duke Mathematical Journal, 158(3):413–499, 2011.
- [ACM+22] David M Ambrose, Roberto Camassa, Jeremy L Marzuola, Richard M McLaughlin, Quentin Robinson, and Jon Wilkening. Numerical algorithms for water waves with background flow over obstacles and topography. Advances in Computational Mathematics, 48(4):1–62, 2022.
- [AL14] Nalini Anantharaman and Matthieu Léautaud. Sharp polynomial decay rates for the damped wave equation on the torus. Analysis & PDE, 7(1):159–214, 2014.
- [Ala17] Thomas Alazard. Stabilization of the water-wave equations with surface tension. Annals of PDE, 3:1–41, 2017.
- [Ala18] Thomas Alazard. Stabilization of gravity water waves. Journal de Mathématiques Pures et Appliquées, 114:51–84, 2018.
- [Ana10] Nalini Anantharaman. Spectral deviations for the damped wave equation. Geom. Funct. Anal., 20:593–626, 2010.
- [BD08] Charles J. K. Batty and Thomas Duyckaerts. Non-uniform stability for bounded semi-groups on Banach spaces. J. Evol. Equ., 8(4):765–780, 2008.
- [BLR92] Claude Bardos, Gilles Lebeau, and Jeffrey Rauch. Sharp sufficient conditions for the observation, control, and stabilization of waves from the boundary. SIAM journal on control and optimization, 30(5):1024–1065, 1992.
- [Bon81] Jean-Michel Bony. Calcul symbolique et propagation des singularités pour les équations aux dérivées partielles non linéaires. Annales scientifiques de l’École normale supérieure, 14(2):209–246, 1981.
- [BT10] Alexander Borichev and Yuri Tomilov. Optimal polynomial decay of functions and operator semigroups. Mathematische Annalen, 347:455–478, 2010.
- [BZ04] Nicolas Burq and Maciej Zworski. Geometric control in the presence of a black box. Journal of the American Mathematical Society, 17(2):443–471, 2004.
- [BZ19] Nicolas Burq and Maciej Zworski. Rough controls for schrödinger operators on 2-tori. Annales Henri Lebesgue, 2:331–347, 2019.
- [CFGK05] Didier Clamond, Dorian Fructus, John Grue, and Øyvind Kristiansen. An efficient model for three-dimensional surface wave simulations. part ii: Generation and absorption. Journal of computational physics, 205(2):686–705, 2005.
- [DZ19] Semyon Dyatlov and Maciej Zworski. Mathematical theory of scattering resonances, volume 200. American Mathematical Soc., 2019.
- [Hör97] Lars Hörmander. Lectures on nonlinear hyperbolic differential equations, volume 26. Springer Science & Business Media, 1997.
- [JKR14] Geri I Jennings, Smadar Karni, and JB Rauch. Water wave propagation in unbounded domains. part i: nonreflecting boundaries. Journal of Computational Physics, 276:729–739, 2014.
- [Joh21] Steven G Johnson. Notes on perfectly matched layers (pmls). arXiv preprint arXiv:2108.05348, 2021.
- [Leb94] G. Lebeau. Équations des ondes amorties. In Séminaire sur les Équations aux Dérivées Partielles, 1993–1994, pages Exp. No. XV, 16. École Polytech., Palaiseau, 1994.
- [LR97] Gilles Lebeau and Luc Robbiano. Stabilisation de l’équation des ondes par le bord. Duke Math. J., 86(3):465–491, 1997.
- [Mac21] Fabricio Macià. Observability results related to fractional schrödinger operators. Vietnam Journal of Mathematics, 49(3):919–936, 2021.
- [MM82] A. S. Markus and V. I. Matsaev. Comparison theorems for spectra of linear operators and spectral asymptotics. Trudy Moskov. Mat. Obshch., 45:133–181, 1982.
- [Moo22] Gary Moon. A toy model for damped water waves. arXiv preprint arXiv:2203.16645, 2022.
- [Nat13] Frédéric Nataf. Absorbing boundary conditions and perfectly matched layers in wave propagation problems. Direct and inverse problems in wave propagation and applications, 14:219–231, 2013.
- [Phu07] Kim Dang Phung. Polynomial decay rate for the dissipative wave equation. Journal of Differential Equations, 240(1):92–124, 2007.
- [RT75] Jeffrey Rauch and Michael Taylor. Decay of solutions to nondissipative hyperbolic systems on compact manifolds. Comm. Pure Appl. Math., 28(4):501–523, 1975.
- [Sjö00] Johannes Sjöstrand. Asymptotic distribution of eigenfrequencies for damped wave equations. Publ. Res. Inst. Math. Sci., 30(5):573–611, 2000.
- [SZ07] Johannes Sjöstrand and Maciej Zworski. Elementary linear algebra for advanced spectral problems. Ann. I. Fourier, 57(7):2095–2141, 2007.
- [Wan22] Jian Wang. Dynamics of resonances for 0th order pseudodifferential operators. Commun. Math. Phys., 391:643–668, 2022.
- [Zua05] Enrique Zuazua. Propagation, observation, and control of waves approximated by finite difference methods. SIAM review, 47(2):197–243, 2005.
- [Zua07] Enrique Zuazua. Controllability and observability of partial differential equations: some results and open problems. In Handbook of differential equations: evolutionary equations, volume 3, pages 527–621. Elsevier, 2007.
- [Zwo12] Maciej Zworski. Semiclassical analysis, volume 138. American Mathematical Society, 2012.