Well-posedness for a two-dimensional dispersive model arising from capillary-gravity flows
Abstract
This paper is aimed to establish well-posedness in several settings for the Cauchy problem associated to a model arising in the study of capillary-gravity flows. More precisely, we determinate local well-posedness conclusions in classical Sobolev spaces and some spaces adapted to the energy of the equation. A key ingredient is a commutator estimate involving the Hilbert transform and fractional derivatives. We also study local well-posedness for the associated periodic initial value problem. Additionally, by determining well-posedness in anisotropic weighted Sobolev spaces as well as some unique continuation principles, we characterize the spatial behavior of solutions of this model. As a further consequence of our results, we derive new conclusions for the Shrira equation which appears in the context of waves in shear flows.
Keywords: Two-dimensional Benjamin-Ono equation; Cauchy problem; Local well-posedness; Weighted Sobolev spaces.
1 Introduction
This work concerns the initial value problem (IVP) for the equation:
(1.1) |
where denotes the Hilbert transform in the -direction defined via the Fourier transform as for , and its periodic equivalent for . This model was derived in [2] as an approximation to the equations for deep water gravity-capillary waves. Numerical results confirming existence of line solitary waves (solutions of the form , and real valued with suitable decay at infinity) as well as wave packet lump solitary waves were also presented in [2].
We are also interested in studying the IVP associated to the Shrira equation:
(1.2) |
This equation was deduced as a simplified model to describe a two-dimensional weakly nonlinear long-wave perturbation on the background of a boundary-layer type plane-parallel shear flow (see [38]). Existence and asymptotic behavior of solitary-wave solutions were studied in [10].
The models in (1.1) and (1.2) can be regarded, at least from a mathematical point of view, as two-dimensional versions of the Benjamin-Ono equation (see, [1, 14, 21, 34, 35, 39, 46] and the references therein):
(1.3) |
Alternatively, the equation in (1.1) can be considered as a two-dimensional extension of the so called Burgers-Hilbert equation (see, [3, 19]):
(1.4) |
This manuscript is intended to analyze well-posedness issues for the IVP (1.1) and (1.2). Here we adopt Kato’s notion of well-posedness, which consists of existence, uniqueness, persistence property (i.e., if the data a function space, then the corresponding solution describes a continuous curve in , ), and continuous dependence of the map data-solution. In this regard, referring to the IVP (1.1), by implementing a parabolic regularization argument (see [23]) local well-posedness (LWP) in and , were inferred in [8]. It was also showed in [8] that the IVP (1.1) is LWP in weighted Sobolev spaces , and .
Concerning the IVP (1.2), by adapting the short-time linear Strichartz estimate approach employed in [27, 31], LWP in was deduced in [5]. In [4], inspired by the works in [20, 30], LWP was established in assuming that the initial data satisfy for almost every . Recently, in [44], by employing short-time bilinear Strichartz estimates the conclusion on the periodic setting was improved to regularity without any assumption on the initial data. Now, with respect to weighted spaces, in [33] LWP was deduced in and , where for arbitrary initial data, and assuming that for almost every .
It is worth pointing out that the equation in (1.1) does not enjoy scale-invariance. In contrast, if solves the equation in (1.2), solves (1.2) whenever , and so this equation is -critical. On the other hand, real solutions of the IVP (1.1) formally satisfy the following conserved quantities (time invariant):
(1.5) | ||||
(1.6) |
and real solutions of (1.2) preserve the quantity and
(1.7) |
where is the fractional derivative operator in the variable defined by its Fourier transform as .
The aim of this paper is to obtain new well-posedness conclusions for both models (1.1) and (1.2) in the spaces , and some spaces adapted to (1.6) and (1.7). Furthermore, by establishing well-posedness in anisotropic spaces and some unique continuation principles, we will study the spatial behavior of solutions, determining that in general arbitrary polynomial type decay in the -spatial variable is not preserved by the flow of these equations.
Let us now state our results. We will mainly work on equation (1.1) without distinguishing between the signs of the term . Firstly, to justify the quantity (1.6), we consider the spaces defined by
(1.8) |
Our first conclusion establishes local well-posedness in the spaces and .
Theorem 1.1.
Let and let be any (fixed) of the spaces and . Then for any , there exist a time and a unique solution to the equation in (1.1) that belongs to
(1.9) |
if , or it belongs to
(1.10) |
if . Moreover, the flow map is continuous from to .
The Sobolev space is defined as usual with norm , and by . The proof of Theorem 1.1 is adapted from the ideas of Kenig [27] and Linares, Pilod and Saut [31]. A novelty in the present work is the study of the operators and which yields additional difficulties in contrast with the operator considered in the previous references. Among them, we required to deduce the following commutator relation:
Proposition 1.1.
Let , , with , then
(1.11) |
This estimate can be regarded as a nonlocal version of Calderon’s first commutator estimate deduced in [6, Lemma 3.1] and [29, Proposition 3.8] (see Proposition 2.1 in the present document). Proposition 1.1 is proved in the appendix, and it is useful to perform energy estimates involving the operator and the nonlinearity in the equation in (1.1).
We remark that Theorem 1.1 improves the conclusion in [8], lowering the regularity in the Sobolev scale to and obtaining well-posedness conclusion in spaces well-adapted to (1.6). Furthermore, we believe that these results could certainly be used to study existence and stability of solitary wave solutions, where one employs the quantity (see for instance [10]).
Next, we present our result in the periodic setting.
Theorem 1.2.
Let . Then for any , there exist and a unique solution of the IVP (1.1) that belongs to
Moreover, for any , there exists a neighborhood of in such that the flow map data-solution,
is continuous.
The function spaces and are defined in the Section 4 below. Theorem 1.2 is proved by means of the short-time Fourier restriction norm method developed by Ionescu, Kenig and Tataru [22], see also [40, 48]. Mainly, this technique combines energy estimates with linear and nonlinear estimates in short-time Bourgain’s spaces and their dual (see Section 4), where the former spaces enjoy the structure with localization in small time intervals whose length is of order , . We emphasize that up to our knowledge, Theorem 1.2 seems to be the first non-standard result dealing with the periodic equation (1.1).
Regarding the periodic quantity , we consider the Sobolev spaces
equipped with the norm . Then, since is a closed subspace of , replacing the spaces by in Section 4 below, the same proof of Theorem 1.2 yields:
Corollary 1.1.
Let . Then the IVP (1.1) is locally well-posed in .
Remarks.
-
(i)
Our local theory is still not sufficient to reach the energy spaces , determined by (1.6).
-
(ii)
For the one dimensional Benjamin-Ono equation (1.3), many authors, see [21, 35, 46] for instance, have applied the gauge transformation to establish local and global results. Unfortunately, we do not know if there exists such gauge transformation for (1.1). Additionally, we do not know if there is a maximal norm estimate available for solutions of (1.1), which would allow us to argue as in [28] to improve the results in Theorem 1.1.
-
(iii)
Concerning solutions of (1.1), we do not have a standard approach to derive bilinear estimates in the spaces and . As a consequence, the short-time Fourier restriction norm method applied to this case leads the same regularity attained in Theorem 1.1. For this reason, we have proved Theorem 1.1 by employing the short-time linear Strichartz approach instead, which also provides solutions in the class . The advantage of using this consequence lies in its application to methods based on energy estimates as the one we employ here to deduce well-posedness in weighted spaces.
Next, we study LWP issues in anisotropic weighted Sobolev spaces:
(1.12) |
and
(1.13) |
To motivate our results, we observe that for a function sufficiently regular with enough decay, requires the condition for almost every . Thus, formally transferring this idea to the equation in (1.1), we do not expect that in general solutions of this model propagate weights of arbitrary order in the -variable. Additionally, for arbitrary initial data, we contemplate to propagate weights of order for some . In this regard, we have:
Theorem 1.3.
In particular, Theorem 1.3 shows that the IVP (1.1) admits weights of arbitrary order in the -variable. The proof of these results follows the ideas of Fonseca, Linares and Ponce [12, 13, 14]. We emphasize that our conclusions involve further difficulties, since here we deal with anisotropic spaces in two spatial variables, and the -spatial decay allowed by (1.1) for arbitrary initial data does not even reach an integer number (cf. [14, Theorem 1] for the BO equation). Finally, we remark that Theorem 1.3 improves the range of weights determined in the work of [8], and we do not require the assumption .
Next, we state some unique continuation principles for solutions of the IVP (1.1).
Theorem 1.4.
Let , and . Let be a solution of the IVP (1.1) such that . If there exist two different times in for which
then for all and almost every .
Theorem 1.5.
Let and . Let be a solution of the IVP (1.1) such that . If there exist two different times in for which
Then the following identity holds true
(1.14) |
for almost every . In particular, if it follows
(1.15) |
Remarks.
-
(i)
Since the weight does not satisfy the condition (see [7, 45]) the assumption in the space is necessary for our arguments. Notice that for the condition does not make sense in general, for this reason, we have distinguished between part (ii) and (iii) of Theorem 1.3. Besides, by inspecting our arguments in Lemma 5.1 below and employing [47, Theorem 4.3], the hypothesis can be replaced by the assumption that for a.e. , the map belongs to the -closure of the space of square integrable continuous odd functions.
-
(ii)
Theorem 1.4 establishes that for arbitrary initial data in with and , is the largest possible decay for solutions of the IVP (1.1) on the -spatial variable. Consequently, for this regimen of indexes , Theorem 1.3 (i) is sharp. However, it still remains an open problem to derive a similar conclusion for the cases . Moreover, Theorem 1.4 shows that if with , and for almost every , then the corresponding solution of the IVP (1.1) satisfies
Although, there does not exist a non-trivial solution corresponding to data with a.e. with
-
(iii)
The condition in Theorem 1.5 can be relaxed assuming for instance that and . In addition, (1.15) provides some unique continuation principles for solutions of the equation in (1.1). Indeed, if for some positive odd integer number , then it must be the case that . Besides, if there exists three times such that , , and
then . Accordingly, Theorem 1.5 establishes that for any initial data , , (or with ), the decay is the largest possible in the -spatial decay. More precisely, if , , , then the corresponding solution of the IVP (1.1) satisfies
and there does not exist a non-trivial solution with initial data such that
All of the previous well-posedness conclusions were addressed by compactness method. As a matter of fact, we have that the local Cauchy problem for the equation in (1.1) cannot be solved for initial data in any isotropic or anisotropic spaces by a direct contraction principle based on its integral formulation.
Proposition 1.2.
Let (resp. ). Then there does not exist a time such that the Cauchy problem (1.1) admits a unique solution on the interval and such that the flow-map data-solution is -differentiable from to (resp. from to ).
We remark that a similar conclusion was derived before for (1.2) in [9]. Following these arguments or the ideas in [18, Theorem 1.4] for instance, it is not difficult to deduce Proposition 1.2. For the sake of brevity, we omit its proof.
Finally, we present our conclusions on the Shrira equation:
Theorem 1.6.
As a result of Theorem 1.6, we derive new well-posedness conclusions in the spaces where the energy (1.7) makes sense. Besides, in the periodic setting, we obtain the same well-posedness result stated for the two-dimensional case in the work of R. Schippa [44, Theorem 1.2], that is, we deduced that (1.2) is LWP in , . We remark that our results are provided by rather different considerations than those given in [44], where the author employed the setting of the periodic -/-spaces ([17]) combined with key short-time bilinear Strichartz estimates (see Section 3 of the aforementioned reference). Certainly, we believe that these considerations can be adapted to (1.1).
Regarding weighted spaces, our conclusions extend the results in [33], since here we deal with less regular solutions, and we improve the -spatial decay allowed by (1.6) to the interval . Actually, by increasing the required regularity, it is not difficult to adapt our result to solutions in anisotropic spaces . We remark that our proof of well-posedness in is applied directly to solutions in the space , in contrast, in [33] the author first derive well-posedness in weighted spaces for solutions with the additional property .
We will begin by introducing some notation and preliminaries. Sections 3 and 4 are devoted to prove Theorem 1.1 and Theorem 1.2 respectively. Theorems 1.3, 1.4 and 1.5 will be deduced in Section 5. Section 6 is aimed to prove Theorem 1.6. We conclude the paper with an appendix where we show Proposition 1.1.
2 Notation and preliminaries
Given two positive quantities and , means that there exists a positive constant such that . We write to symbolize that and . The Fourier variables of are denoted and in the periodic case as .
denotes the commutator between the operators and , that is
Given and integer, we define the Lebesgue spaces , by its norm as with the usual modification when . To emphasize the dependence on the variables when , we will denote by . We denote by the spaces of smooth functions of compact support and the space of Schwarz functions. The Fourier transform is defined by
For a given number , the operators , and are defined via the Fourier transform according to , and , respectively, where . The Sobolev spaces consist of all tempered distributions such that .
Since we will deal with the periodic and real equation in different sections, we will employ the same notation for the norm of the Sobolev spaces which consists of the periodic distributions such that . Recalling the spaces (1.8), we will denote by for or , and .
Now, if denotes a functional space (for instance those introduced above), we define the spaces and according to the norms
respectively, for all .
We define the unitary group of solutions of the linear problem determined by (1.1) by
(2.1) |
where
(2.2) |
The resonant function is given by
(2.3) |
The variable is assumed to be dyadic, i.e., . We will mostly use the dyadic numbers , then we set . Let even function, with and in . For each , we let and . We define the projector operators in by the relations
(2.4) | ||||
With a slightly abuse of notation, we will employ the same notation for the operators and defined in . We also require the following projections in our estimates
(2.5) | ||||
To obtain estimates for the nonlinear term the following Leibniz rules for fractional derivatives will be implemented in our arguments.
Lemma 2.1.
If and , then
Lemma 2.1 was proved by Kato and Ponce in [26]. We also need the following lemma whose proof can be consulted in [15].
Lemma 2.2.
Given and , it holds that
(2.6) | ||||
(2.7) |
with , .
Lemma 2.3.
Let , then
where , , .
Lemma 2.3 was deduced by Muscalu, Pipher, Tao and Thiele in [36]. The following commutator estimate will be useful in our considerations.
Proposition 2.1.
Let and , then
(2.8) |
The estimate (2.8) was established in [6, Lemma 3.1] and it was extended to the BMO spaces in [29, Proposition 3.8].
To deduce the LWP result in Theorem 1.1 on the space , we require the following set of inequalities, which were deduced in the proof of [27, Lemma 2.1] (see equations (2.5), (2.6) and (2.7) in this reference). See also [31, Lemma 4.6].
Lemma 2.4.
-
(i)
Let , then
(2.9) -
(ii)
If is a positive constant chosen small enough, then the following holds true. There exist
and such that
(2.10) (2.11) and
(2.12) for all .
3 Well-posedness for real solutions
This section is devoted to establish LWP for (1.1) in the spaces and . Since this conclusion in the former space can be deduced by the same reasoning applied to , or by following the ideas in [18, Theorem 1.3], we will restrict our considerations to prove Theorem 1.1 in .
However, given that the LWP result in , will be employed to deduce Theorem 1.3, we will present some remarks on this conclusion in the Subsection 3.3.3 below.
3.1 Linear Estimates
This section summarized some space-time estimates for the unitary group defined by (2.1).
Lemma 3.1.
The following estimate holds
whenever , and .
Notice that the endpoint Strichartz estimate corresponding to is not stated in the preceding lemma, as a consequence we need to lose a little bit of regularity to control this norm.
Corollary 3.1.
For each and , there exists such that
where the implicit constant depends on .
Proof.
Also, we require the following refined Strichartz estimate, which has been proved in different contexts (see [5, 27, 31]).
Lemma 3.2.
Let and . Then there exist and such that
(3.1) | ||||
whenever solves
(3.2) |
3.2 Energy Estimates
Lemma 3.3.
Let . Consider and be a solution of the IVP (1.1), then there exists a constant such that
(3.3) |
Proof.
The estimates of the norm in the space is deduced by applying standard energy estimates and Lemma 2.1. For a more detailed discussion, we refer to [27, Lemma 1.3] .
To deal with the component in the -norm, we apply to the equation in (1.1), we multiply then by and integrate in space to deduce
where we have used that the operator is skew-symmetric and is self-adjoint. Hence, by writing and using that defines a bounded operator in , we get
To control the norm , we apply to the equation in (1.1), multiplying the resulting expression by and integrating in space it is seen that
Once again, decomposing and using that is skew-symmetric, we get
(3.4) | ||||
Then the Cauchy-Schwarz inequality and Proposition 1.1 yield
(3.5) | ||||
and so we arrive at
Integrating in time the previous estimates yield the desired conclusion. ∎
Next we derive a priori estimates for the norms and in , whenever .
Lemma 3.4.
Let fixed. Consider solution of the IVP (1.1). Then, there exist and such that
(3.6) |
Proof.
We will follow the arguments in [27] and [31]. By applying Lemma 3.2 with we find
(3.7) | ||||
Taking small such that , Young’s inequality yields
(3.8) |
Hence the previous display and Plancherel’s identity show
(3.9) |
This completes the estimate for the first two terms on the right-hand side (r.h.s) of (3.7). Next, we deal with the third factor on the r.h.s of (3.7). An application of (2.7) allows us to deduce
(3.10) | ||||
which holds for . Since , the previous inequality completes the study of third term in (3.7). Next, we decompose the remaining factor in (3.7) as follows
To deal with , we employ the point-wise inequality
(3.11) |
valid for and small satisfying . Hence the fractional Leibniz’s rule (2.6), Plancherel’s identity and (3.11) show
This completes the analyze of . On the other hand, employing Lemma 2.3, we further decompose
(3.12) | ||||
where . Since the norms , we use (3.8) with and Plancherel’s identity to infer .
To deal with , we let small satisfying , then we employ (2.9) to control the norm . The estimate for is a consequence of Plancherel’s identity and (3.11) with . This yields the desired bound for .
Next, by employing (2.11), (2.12) in Lemma 2.4, it is seen
for some and fixed and where . Given that
Plancherel’s identity yields
(3.13) |
From this we get . According to (3.12), this completes the estimate of . Collecting the bounds derived for and , we obtain
On the other hand, the estimate concerning is obtained by applying Lemma 3.2 with , estimate (2.7), (3.9) and the inequality
(3.14) |
valid for . To avoid repetition we shall omit its proof. However, we emphasized that this estimate does not require to implement Lemma 2.4. The proof is complete. ∎
Additionally, we require to control the norm . This estimate will be useful to close the argument leading to the proof of Theorem 1.1 in the space .
Lemma 3.5.
3.3 Proof of Theorem 1.1
Our results relay on existence of smooth solutions for the IVP (1.1). To achieve this conclusion in the spaces , we require the following lemma.
Lemma 3.6.
Let . Then it holds
(3.15) | |||
(3.16) |
for every .
Proof.
Whenever , local well-posedness in for the IVP (1.1) follows from a parabolic regularization argument. Roughly speaking, an additional term is added to the equation, after which the limit is taken. This technique was applied in [8] for the IVP (1.1) establishing LWP in for all .
Furthermore, employing Lemma 3.6, it is possible to apply a parabolic regularization argument adapting the ideas in [8] or [25, Section 6.2] (see also [32, Theorem 9.2]), to obtain local well-posedness for the IVP (1.1) in , . Summarizing the preceding discussion we have:
Lemma 3.7.
Let and be any (fixed) of the spaces and . Then for any , there exist and a unique solution of the IVP (1.1). In addition, the flow-map is continuous in the -norm.
The proof of Lemma 3.7 also provides existence of smooth solutions and a blow-up criterion. More precisely, let , where is any (fixed) of the spaces and , then there exists a solution to the IVP (1.1), where is the maximal time of existence of satisfying and the following blow-up alternative holds true
(3.17) |
if .
We require of some additional a priori estimates.
Lemma 3.8.
Let . then there exists such that for all there is a solution of the IVP (1.1) where . Moreover, there exists a constant such that
and
(3.18) |
whenever .
Proof.
Now we can prove the existence of solutions.
3.3.1 Existence of solution
Lemma 3.9.
Let and such that . Assume that , then
(3.19) |
for each and .
Proof.
By support considerations we observe
Integrating the above expression, we use Plancherel’s identity and Lebesgue dominated convergence theorem to verify that the first norm on the left-hand side (l.h.s) of (3.19) satisfy the desired limit. A similar argument provides the required limit for the second norm on the l.h.s of (3.19). ∎
Now, we gather the previous result to derive some conclusion for the smooth solutions generated by some approximations of the initial data.
Let , . For each dyadic number , Lemma 3.8 assures the existence of a time (for some constant ) independent of and smooth solutions of the IVP (1.1) with initial data such that
(3.20) |
and
(3.21) |
Additionally, we combine Lemma 3.5, (3.20) and (3.21) to infer
(3.22) |
provided that is chosen large enough. Now, let , , and , so satisfies the equation
(3.23) |
with initial condition . Thus, by employing similar energy estimates leading to (3.3), together with (3.19), we deduce
(3.24) |
whenever .
Accordingly, we shall prove that is a Cauchy sequence in . Let us first estimate the sequence in .
Lemma 3.10.
Let , . If , , then
(3.25) |
and
(3.26) |
provided that with large enough.
Proof.
Since (3.25) and (3.26) are inferred as in the proof of Lemma 3.4, we will only deduce (3.25). Recalling that satisfies (3.23), we apply Lemma 3.2 with to get
(3.27) | ||||
for some with to be determined along the proof. Now, we proceed to estimate each of the factors .
In view of (3.24), it follows that whenever . To study , we employ Young’s inequality to derive
(3.28) |
Plancherel’s identity shows
(3.29) |
Therefore, choosing , where is small satisfying , we have from (3.24) and (3.29) that
Next, by employing (2.7), we follow the arguments in (3.10) to deduce
for all , with , where we have used (3.20), (3.21) and (3.24). Now, we divide the remaining term as follows
By employing the fractional Leibniz’s rule (2.6) in the -variable, (3.20) and a similar argument to (3.28), it is seen
for all such that . On the other hand, from Lemma 2.3 it is deduced
where and satisfy the conditions in Lemma 2.4 (ii). An application of (3.28) shows
for each , where is small satisfying . Now, we combine estimate (3.9) and(3.20) to derive
Additionally, employing (2.9) and identity (3.29), it is not difficult to see
for all and . Finally, gathering together estimates (2.10), (2.12), (3.20) and (3.21), we deduce
so that Young’s inequality and (3.24) yield
for all and () given by Lemma 2.4. Collecting all the previous estimates
Plugging the bounds obtained for the terms , in (3.27), we obtain
This completes the deduction of (3.25) provided that and is chosen sufficiently large. ∎
Next, we shall prove that is a Cauchy sequences in the space .
Proposition 3.1.
Let , . If , then
(3.30) |
Proof.
We apply to (3.23) and then multiplying by and integrating the resulting expression in space, we deduce
(3.31) | ||||
Integrating by parts,
which together with Lemma 2.1 yield
(3.32) | ||||
On the other hand,
then Lemma 2.1 gives
(3.33) | ||||
To control , we employ the fact that solves the equation in (1.1) to apply energy estimates with Lemma 2.1 to find
(3.34) |
where we have also used Gronwall’s inequality, (3.20) and (3.21). Therefore, gathering (3.32)-(3.34), and (3.20) and (3.21) in (3.31), after Gronwall’s inequality we get
which holds in virtue of Lemma 3.10 and (3.19). Once we have established that as the estimate for the norm of will be completed.
Now, applying to (3.23), and then multiplying by and integrating in space, we have
Now, given that determines a skew-symmetric operator, it is seen
(3.35) | ||||
so that Proposition 1.1 applied to the -variable gives
Therefore, the preceding differential inequality, Gronwall’s lemma, (3.21) and (3.19) imply
Finally, we proceed to estimate the norm. Since solves (3.23), we apply to this equation, multiplying by , then integrating in space, we deduce
(3.36) | ||||
where we have employed the decomposition . Arguing as in (3.35), Proposition 1.1 shows
(3.37) | ||||
On the other hand, we use Hölder’s inequality to find
(3.38) |
According to the above estimate, we are led to bound the norms and . Thus, given that satisfies the equation in (1.1), integrating by parts it follows that
Hence, Gronwall’s inequality and (3.21) yield
(3.39) |
Now, from the fact that solves (3.23) and integrating by parts we find
(3.40) | ||||
To estimate , we employ that solves the equation in (1.1) to get
From this estimate and (3.39), it is seen
Then, in view of (3.20)-(3.22), (3.34), (3.39) and Gronwall’s inequality
where we have used that . Consequently, the previous estimate allows us to deduce
Now, by using (3.39) and Hölder’s inequality,
Thus, inserting the above estimates in (3.40), applying Gronwall’s inequality together with (3.20), (3.21) and (3.25) reveal
(3.41) | ||||
Going back to , we plug (3.39) and (3.41) into (3.38) to obtain
(3.42) |
Now, collecting (3.37), (3.42) in (3.36),
Then, applying Gronwall’s inequality to the preceding inequality, together with (3.20), (3.21) yield
where we have used (3.25) with and large enough. This completes the proof of (3.30). ∎
3.3.2 Uniqueness and Continuous Dependence
Uniqueness of solution in the classes
can be easily obtained by applying energy estimates at the -level following the same ideas in Lemma 3.3. For the sake of brevity, we omit these computations. Continuous dependence on the spaces , follows from the continuity of the flow-map data solution in Lemma 3.7 and the ideas in [31].
3.3.3 Solutions in
With the aim of Lemmas 3.2, 3.7 and the blow-up alternative (3.17), the proof of local well-posedness in , follows the same ideas in [18, Theorem 1.3].
Similarly, the proof of local well-posedness in can also be deduced from the arguments employed above to estimate the -norm of the space . However, when replacing by in our estimates, we require to employ Lemmas 2.1 and 2.2 in the full spatial variables, as a consequence the estimate of for solutions of the IVP (1.1) is essential in this part.
4 Periodic Solutions
4.1 Function spaces and additional notation
We will follow the notation in [22] (see also, [40, 42, 43, 48]). We recall that . The variable is presumed to be dyadic. Given , we define and . For each , we set and . Let , we set
(4.1) |
where is defined as in (2.2).
Now, we introduce some family of projectors required for our arguments. To simplify notation, we will employ the same symbols used in (2.5) restricted to this section. We define the projector operators in by the relation
for all and , here stands for the indicator function on the set . Given a dyadic number , we define the operator by the Fourier multiplier , where with dyadic. We also set .
For a time , let be the greatest dyadic number such that . Let and , we define the dyadic -type normed spaces
where the functions and are defined as in Section 2. We will denote by the space . Next we introduce the spaces according to uniformly on time intervals of size :
and
Let and be any of the spaces or , we set
equipped with the norm:
Then for a given , we define the spaces and from their frequency localized version and by using the Littlewood-Paley decomposition as follows
(4.2) |
and
Next, we define the associated energy spaces endowed with the norm
In the subsequent considerations and will denote the spaces above with parameter .
4.1.1 Basic Properties
Now we collect some basic properties of the spaces and . These results have been deduced in different contexts in [16, 22, 42, 41, 48] for instance.
Lemma 4.1.
Let , , and satisfying
Then for any , and ,
(4.3) |
and
(4.4) |
The implicit constants above are independent of , and in consequence of the definition of the spaces .
Additionally, we require the next conclusion:
Lemma 4.2.
Let , and a bounded interval. Then
for all whose Fourier transform is in and the implicit constant is independent of .
The following lemma will be useful to obtain time factors in the energy estimates.
Lemma 4.3.
Let and . Then for any ,
where the implicit constant is independent of and , and in consequence of the definition of the spaces .
Proof.
The proof follows the same arguments in [16, Lemma 3.4]. ∎
Lemma 4.4.
Let , then
whenever and the implicit constant is independent of .
We also need the following linear estimate which is deduced in much the same way as in [22, Proposition 3.2] (see also [48, Proposition 6.2]).
Proposition 4.1.
Assume that , and with
Then
(4.5) |
where the implicit constant is independent of , and in consequence of the definition of the spaces , and .
To obtain a priori estimates for smooth solutions we need the following lemma.
Lemma 4.5.
Let , . Then the mapping is increasing and continuous on and
(4.6) |
Proof.
The proof follows the same line of arguments in [48, Lemma 6.3]. ∎
4.2 Bilinear estimates
Next, we obtain the crucial bilinear estimates, which will be applied in both the short time estimates and energy estimates in the subsequent subsections. Recalling the notation introduced in (4.1), we have:
Proposition 4.2.
Assume that and functions supported in for .
-
(i)
It holds that
(4.7) -
(ii)
Suppose that . If for some , then
(4.8) otherwise
(4.9) -
(iii)
If ,
(4.10)
Before proving Proposition 4.2, we require the following elementary result.
Lemma 4.6.
Let be two intervals in , and a function with . Suppose that . Then
Proof of Proposition 4.2.
We notice that
(4.11) |
where . Let us first establish (i). In view of the above display, we can assume that . Let , then is supported in
and , , so that
(4.12) |
where is defined as in (2.3). Thus, by applying the Cauchy-Schwarz inequality in and then in we get
(4.13) | ||||
In this manner, the same procedure applied above now to the spatial variables on the r.h.s of (4.13) yields (4.7).
Next, we deduce (ii). By (4.11), we shall assume that and , that is, . We consider the sets:
(4.14) | ||||
Accordingly, we divide given by (4.12) as
(4.15) |
where corresponds to the restriction of to the domain . Now, we proceed to estimate each of the factor , .
Estimate for . By support considerations, it must follow that , or equivalently, . Thus, recalling (2.3), the resonant function for this case satisfies
(4.16) | ||||
So, we divide , where consists of the elements in satisfying that and those for which . Thus, we find
(4.17) |
in each of the regions and . Now, since in the support of , we further divide the region of integration according to the cases where and , namely
(4.18) |
To estimate , we use that , (4.17) and Lemma 4.6, together with the Cauchy-Schwarz inequality in the variable to find
(4.19) | ||||
where we have employed the Cauchy-Schwarz inequality in , and the last line is obtained by the same inequality in . The estimate for is deduced changing the roles of by in the preceding argument. This completes the study of .
Estimate for . In this case and , then
We write , where and . Consequently, in each of the sets , it holds
(4.20) |
Now, since , with , (4.20) establishes that in each of the regions defined by restricted to , either or . In consequence, we can further divide restricted to each , as in (4.18) to apply a similar argument to (4.19), which ultimately leads to the desired estimate.
Estimate for and . In these cases both regions of integration can be bounded directly by means of the Cauchy-Schwarz inequality without any further consideration on the resonant function. Indeed, in the support of , we have that and so
(4.21) | ||||
where we have employed that , together with consecutive applications of the Cauchy-Schwarz inequality. On the other hand, to estimate , we split the region of integration in two parts for which at least one of the variables among and is not considered in the summation. This in turn allows us to perform some simple modifications to the previous argument dealing with to bound by the r.h.s of (4.21).
Collecting the estimates for , , we complete the deduction of (ii).
Next, we consider (iii). In virtue of (4.11), we shall assume that and . As before, we decompose , where corresponds to the restriction of (given by (4.12)) to the domain determined by (4.14).
Since , (4.17) allows us to estimate exactly as in (4.19). The estimate for is obtained without considering the resonant function as in the study of above for each . For the sake of brevity, we omit these estimates.
In the case of , we notice that (4.20) shows that and could vanish in the support of the integral. Instead, we split , where , , we have
(4.22) |
in each of the regions and . Thus, (4.22) and similar considerations in the deduction of (4.19) yield
This completes the deduction of (iii). The proof is complete. ∎
By duality and Proposition 4.2, we obtain the following bilinear estimates.
Corollary 4.1.
Let be dyadic numbers and supported in for .
-
(1)
It holds that
(4.23) -
(2)
Suppose that . If for some , then
(4.24) otherwise
(4.25) -
(3)
If ,
(4.26)
4.3 Short time bilinear estimates
In this section, we derive the crucial key bilinear estimates for the equation and the difference of solutions.
Proposition 4.3.
Let , , then
(4.27) | ||||
(4.28) |
for all and where the implicit constants are independent of , and the definition of the spaces involved.
We split the proof of Proposition 4.3 in the following technical lemmas.
Lemma 4.7 ().
Let satisfying . Then,
whenever and .
Proof.
We use the definition of the space to find
where
(4.29) | ||||
with . Now, we define
(4.30) | ||||
for , and we set similarly and . Therefore, from the definition of the spaces , (4.24) and (4.25), we find
(4.31) | ||||
since , in the first line above we have used that the sum over on the left-hand side of (4.31) can be controlled by the right-hand side of this inequality. Therefore, the above expression and Lemma 4.1 yield the deduction of the lemma. ∎
Lemma 4.8 ().
Let satisfying . Then,
whenever and .
Proof.
Following the same arguments and notation as in the proof of Lemma 4.7, we write
(4.32) | ||||
To estimate the first term on the right-hand side of (4.32), we employ (4.26) and the restrictions to find
(4.33) |
Thus, we add the above expression over with , then we apply Lemma 4.1 to the resulting inequality to obtain the desired bound. Next, we deal with the second sum on the right-hand side of (4.32). Interpolating (4.23) and (4.26), it is seen
(4.34) | ||||
for all and . Under these considerations, either or , which implies
Then, plugging the previous estimate in (4.34) and recalling that , we get
(4.35) | ||||
Therefore, taking sufficiently close to , we sum (4.35) over with and then we apply Lemma 4.1 to derive the desired estimate for the second term on the r.h.s of (4.32). ∎
Lemma 4.9 ().
Let satisfying . Then,
whenever and .
Proof.
Following the same notation employed in the proof of Lemma 4.7, we have
(4.36) | ||||
To estimate the first term on the r.h.s of (4.36), we use (4.24) to deduce
(4.37) |
where . These restrictions imply, , then when , we have
(4.38) | ||||
Now, when , we use instead
(4.39) |
By support considerations, it must follow that , whenever . This implies that summing over in (4.39) yields a factor of order . This remark completes the estimates for the first sum in (4.36). The remaining sum in (4.36) is bounded directly by (4.25) and arguing as above. The proof of the lemma is now complete. ∎
Lemma 4.10 ().
Let satisfying . Then,
whenever and .
Proof.
We are in conditions to prove Proposition 4.3.
Proof of Proposition 4.3.
We will adapt the ideas in [40] for our considerations. We will only deduce (4.27), since (4.28) is obtained by a similar reasoning. For each , we choose extensions , of and satisfying, and . By the definition of the space and Minkowski inequality we have
where
To estimate , we use Lemma 4.7, the fact that for small enough and the definition of to derive
The estimate for is obtained symmetrically as above. Next, we use Lemma 4.8 and that to obtain
Let fixed, then Lemma 4.9 and the Cauchy-Schwarz inequality yield
which holds given that . The estimate for follows from Lemma 4.10 and similar considerations as above. This concludes the deduction of (4.27). ∎
4.4 Energy estimates.
This section is devoted to derive the estimates required to control the -norm of regular solutions and the difference of solutions.
Lemma 4.11.
Let , then there exists small enough such that for it holds that
(4.41) |
for each function , .
Proof.
In view of (4.11), we will assume that . Let be an extension of to such that for each . Additionally, let be a smooth function supported in such that
Then, we write
(4.42) | ||||
where
Let us estimate the sum over in (4.42). Recalling the dyadic defining the spaces , we denote by
for each , and . Now since there are at most integers in , we employ (4.8) and (4.9) when , or (4.10) if to deduce that
(4.43) | ||||
Next we deal with the sum over in (4.42). We consider fixed and let
for each , and . We treat first the case . Since , we have
From (4.8) and the fact that , we get
(4.44) | ||||
In the regions where , we use Lemmas 4.2 and 4.3, together with the fact that to deduce
(4.45) | ||||
Now, we deal with the case . Interpolating the right-hand side of (4.44) with the bound derived for using (4.7) instead of (4.8), we find for all that
(4.46) | ||||
Therefore, the estimate for is now a consequence of (4.45) and (4.46). On the other hand, we can implement (4.9) and the same ideas dealing with (4.45) to derive the following bound
This completes the analysis of in the region . Next we treat the case . Interpolating (4.7) and (4.10), we obtain for all that
(4.47) | ||||
Therefore, taking and employing a similar reasoning to (4.46), the estimate for when is a consequence of (4.47). Gathering all the previous results, by setting we obtain (4.41). ∎
Lemma 4.12.
Assume that , , then there exists such that for ,
whenever and .
Proof.
We divide the integral expression in the following manner
(4.48) | ||||
Integrating by parts and using (4.41), the first term on the right-hand side of the above expression satisfies
(4.49) |
The estimate for is deduced arguing as in [22, Lemma 6.1] (see equation (6.10)) and following the same ideas leading to (4.41). For the sake of brevity, we omit its proof. ∎
Proposition 4.4.
Let and . Then for any solution of the IVP (1.1) on ,
(4.50) |
where the implicit constant above is independent of the definition of the spaces involved.
Proof.
According to the definition of the spaces and the fact that solves the IVP (1.1), it is enough to derive a bound for the sum over of the following expression
(4.51) |
Now we split the estimate for the integral term above according to the iterations: ,
(4.52) |
,
(4.53) |
,
(4.54) |
and ,
(4.55) |
In view of Lemma 4.12, the iteration satisfies
(4.56) |
Summing the above expression over and , we can modify the power of by an arbitrary small factor to apply the Cauchy-Schwarz inequality in the sum over . Next, we apply the same inequality for the sum over , obtaining (4.50). Now, recalling (4.49) in the proof of Lemma 4.12, we notice that the iteration satisfies the same estimate in (4.56).
Next we apply (4.41) to control the iterations as follows
(4.57) |
Since , we can increase the power in by a small factor to apply the Cauchy-Schwarz inequality separately in each of the sums over to derive the desired result. The estimate for is obtained by (4.41) and a similar reasoning to the iteration . This completes the estimate for the r.h.s of (4.51) and in turn the deduction of (4.50). ∎
We also require the following result to deal with the difference of solutions.
Proposition 4.5.
Let , . Consider solutions of the IVP (1.1) with initial data respectively, then
(4.58) |
and
(4.59) | ||||
where the implicit constants are independent of and the spaces involved.
Proof.
We shall argue as in the proof of Proposition 4.4. Letting , we find that solves the equation:
(4.60) |
with initial condition . Let , the definition of the -norm and the fact that solves (4.60) yield
(4.61) | ||||
Then, we are reduced to estimate the integral term on the right-hand side of the last inequality. Arguing as in the proof of Proposition 4.4, applying Lemmas 4.11 and 4.12, we obtain
(4.62) |
and
(4.63) | ||||
where we emphasize that the last term on the right-hand side of (4.63) appears from the estimate dealing with the iteration and Lemma 4.11, since in this case
with . It remains to control the term in the integral in (4.61). We divide our considerations as in the proof of Proposition 4.4 according to the iterations: , , and . Notice that in this case we cannot apply Lemma 4.12 to control the iteration. We use instead Lemma 4.11 to find for that
(4.64) | ||||
Summing (4.64) over and , we use that for to apply the Cauchy-Schwarz inequality on the sum over and then on to control the resulting expression by the r.h.s of (4.58) if , or by the last term on the r.h.s of (4.59) if .
The remaining iterations are treated as in the proof of Proposition 4.4, and their resulting bounds are the same displayed on the right-hand sides of (4.62) if and (4.63) if respectively. The proof of the proposition is now complete.
∎
4.5 Proof of Theorem 1.2
We follow similar considerations as in [22, 48] to prove Theorem 1.2. We begin by recalling the local well-posedness result for smooth initial data, which can be deduced as in [24, Theorem 2.1].
Theorem 4.1.
Let . Then there exist and a unique solution of the IVP (1.1). Moreover, the existence time is a non-increasing function of and the flow-map is continuous.
We divide the proof of Theorem 1.2 in the following main parts.
4.5.1 A priori estimates for smooth solutions
Proposition 4.6.
Proof.
We consider fixed and as in the statement of the proposition. In virtue of Theorem 4.1, there exist and solution of the IVP (1.1) with initial data . Then for a given to be chosen later, we collect the estimates (4.5), (4.27) and (4.50) to find for each that
(4.66) |
where . We emphasize that our arguments indicate that the implicit constants in (4.66) and are independent of and in consequence of the definition of the spaces involved (which depend on ). Letting and , (4.66) yields
(4.67) |
Considering now , in (4.66), we also find
(4.68) |
Since the mapping is decreasing and continuous with , from (4.6) it follows that
(4.69) |
where the implicit constant is independent of and the definition of the spaces involved. Thus, we can choose sufficiently small, such that according to the constants in (4.67) and (4.69). Then, for this time and the associated spaces , we can apply a bootstrap argument relaying on (4.67), (4.69) and the continuity of , to obtain , for any . Consequently, Lemma 4.4 reveals
Therefore, up to choosing smaller at the beginning of the argument, from (4.68) we infer
In this manner, the preceding result and Theorem 4.1 allow us to extend , if necessary, to the whole interval . This completes the proof of the proposition. ∎
4.5.2 -Lipschitz bounds and uniqueness
Let be two solutions of the IVP (1.1) defined on with initial data such that , where we denote by and the spaces defined at time and . Notice that this implies that , whenever . We collect (4.5), (4.28) and (4.58) to get
(4.70) |
where the implicit constants above are independent of the definition of the spaces. Let , satisfying . Following a similar reasoning as in the proof of Proposition 4.6, there exists a time sufficiently small, for which with respect to the constants in (4.70) and . Consequently, (4.70) and Lemma 4.4 yield
for any . Thus, if , the last equation reveals that on . Since depends on , we can employ the same spaces to repeat this procedure a finite number of times obtaining uniqueness in the whole interval .
4.5.3 Existence and continuity of the flow-map
Let and fixed. For a given with , we consider a sequence converging to in , such that . We denote by the solution of the IVP (1.1) with initial data determined by Theorem 4.1. Therefore, according to Proposition 4.6, there exists , such that and (4.65) holds. We shall prove that defines a Cauchy sequence in for some . To this aim, we will proceed as in [22, 48].
For a fixed and integers, we have
(4.71) | ||||
for all . Using Sobolev embedding and (4.65), we get
(4.72) | ||||
Then, the standard energy method and the above inequality show that the second term on the right-hand side of (4.71) is controlled as follows
(4.73) |
for each and some constant depending on . Therefore, it remains to estimate the first and last term in (4.71). By symmetry of the argument, we will restrict our considerations to study the former term. To simplify notation, let us denote by , and , then taking , we gather (4.5), (4.27) and (4.59) to find
(4.74) |
for all , and where is fixed. The above set of inequalities reveal
(4.75) | ||||
Repeating the arguments in the proof of Proposition 4.6, using (4.66) with and , we choose small so that
and such that, employing (4.70) and similar considerations as in the uniqueness proof above,
Furthermore, we can choose smaller, if necessary, to assure that with respect to the implicit constant in (4.75), and such that . Then gathering these estimates in (4.75), we get
where, given that , we have used that . From the inequality above and Lemma 4.4, we arrive at
(4.76) |
where . Therefore, according to our previous discussion, this completes the estimate for the first and third terms on the r.h.s of (4.71). Noticing that for large, , we can take large in (4.76), and then large in (4.73), to obtain that is a Cauchy sequence in for a fixed time .
Since each of the elements in the sequence solves the integral equation associated to (1.1) in , we find that the limit of this sequence is in fact a solution of the IVP (1.1) with initial data . This completes the existence part. Finally, it is not difficult to obtain the continuity of the flow-map from the same property for smooth solutions in Theorem 4.1 and the preceding arguments. We refer to [48] for a more detailed discussion.
5 Well-posedness results in weighted spaces
This section is aimed to establish Theorem 1.3. We will start introducing some preliminary results.
Given , we define the truncated weights according to
in such a way that is smooth and non-decreasing in with for all and there exists a constant independent of such that . To explicitly show the dependence on the spatial variables , we will denote by and .
Since we are interested in performing energy estimates with the weights and then taking the limit , we must assure that the computations involving the Hilbert transform and the aforementioned weights are independent of the parameter . In this direction we have:
Proposition 5.1.
For any and any , the Hilbert transform is bounded in with a constant depending on but independent of .
We recall the following characterization of the spaces .
Next, we proceed to show several consequences of Theorem 5.1. When and one can deduce
(5.2) |
and it holds
(5.3) |
Proposition 5.2.
Let . If such that there exists for which , are defined and , then for any , and consequently .
Proposition 5.3.
Let . For any
(5.4) |
and
(5.5) |
for all .
Proof.
The following result will be useful to study the behavior of solutions of (1.1) in , whenever .
Lemma 5.1.
Let and such that . Then, .
Proof.
Since the case can be easily verified, we will restrict our considerations to the case . We first notice that the same argument in the deduction of (5.5) establishes
Thus, an application of (5.2) and the previous result reduces our analysis to prove
(5.6) |
However, the preceding estimate is a consequence of [47, Proposition 3.2] and the assumption . ∎
We shall also employ the following interpolation inequality which is proved in much the same way as in [14, Lemma 1]:
Lemma 5.2.
Let . Assume that and , . Then for any ,
(5.7) |
Moreover, the inequality (5.7) is still valid with instead of with a constant independent of .
Now we are in the condition to prove Theorem 1.3
5.1 Proof of Theorem 1.3
In view of Theorem 1.1, for a given there exist and solution of the IVP (1.1). Let defined by
(5.8) |
In what follows, we will assume that is sufficiently regular to perform all the computations required in this section. Indeed, recalling the comments in the Subsection 3.3.3, we consider the sequence of smooth solutions with , then (3.46) holds and in the topology. Consequently, applying our arguments to and then taking the limit , we can impose the required assumptions on .
5.1.1 Proof of Theorem 1.3 (i)
Let us first prove the persistence property . We begin by deriving some estimates in the spaces and .
Estimate for the -norm. Here, fixed. We apply to the equation in (1.1) to find
(5.9) |
multiplying then by and integrating in space, we infer
(5.10) | ||||
Multiplying the equation in (1.1) by and then integrating in space, it is seen that
(5.11) | ||||
Adding the differential equations (5.10) and (5.11), after integrating by parts in the variable we deduce
(5.12) | ||||
Now, since , with implicit constant independent of , integrating by parts and using the Cauchy-Schwarz inequality we find
Notice that the norm is controlled by (5.8). Next, since , Proposition 5.1 shows
Hence, we employ Hölder’s inequality to get
Thus, gathering the previous estimates,
(5.13) | ||||
Estimate for the -norm. In this case, is arbitrary. Multiplying the equation in (1.1) by and integrating in space yield
(5.14) | ||||
Since the weight function does not depend on , writing and using that determines a skew-symmetric operator, we have that . Similarly, integrating by parts on the variable and writing , it follows that .
Now, integrating by parts and using that is skew-symmetric, it is not difficult to see
(5.15) |
From the fact that , with a constant independent of and (5.8), it follows
(5.16) |
whenever . Now, if , we have
(5.17) |
where we have employed the identity . To estimate the last expression in the preceding inequality, we choose , in (5.7) and applying Young’s inequality it is seen that
(5.18) | ||||
Thus, choosing , (5.16)-(5.18) and (5.8) imply
(5.19) |
Finally,
(5.20) |
Plugging the estimates for , in (5.14) yields
(5.21) |
This completes the desired estimate for the -norm;
Now, we collect the estimates derived for the norms and to conclude Theorem 1.3 (i). Letting
the inequalities (5.13) and (5.21) assure that there exists some constant independent of such that
(5.22) |
Then, Gronwall’s inequality implies
(5.23) | ||||
Thus, taking in the previous inequality shows
(5.24) | ||||
This shows that . Now, we shall prove that . Firstly, since , it is not difficult to see that is weakly continuous. The same is true for the map on . On the other hand, (5.24) implies
(5.25) | ||||
Clearly, weak continuity implies that the right-hand side of (5.25) goes to zero as . This shows right continuity at the origin of the map . Taking any and using that the equation in (1.1) is invariant under the transformations: and , right continuity at the origin yields continuity to the whole interval , in other words, .
The continuous dependence on the initial data follows from this property in and the same reasoning above applied to the difference of two solutions. This completes the proof of Theorem 1.3 (i).
5.1.2 Proof of Theorem 1.3 (ii) and (iii)
Let be the solution of the IVP (1.1) with for Theorem 1.3 (ii), or satisfying , for Theorem 1.3 (iii). Since we have already established that solutions of the IVP (1.1) preserve arbitrary polynomial decay in the -variable, we will restrict our considerations to deduce , . Once this has been done, following the arguments in (5.25), we will have that .
Moreover, the continuous dependence on the spaces and , follows by the same energy estimate leading to applied to the difference of two solutions.
Now, to assure the persistence property in for Theorem 1.3 (iii), we require the following claim:
Claim 1.
Let , fixed and
be a solution of the IVP (1.1). Assume that for a.e . Then, for every and almost every .
Proof.
Since solves the integral equation associated to (1.1), taking its Fourier transform we find
(5.26) |
where is defined by (2.2). Now, the assumptions imposed on the solution show
Hence, the above conclusion, Fubini’s theorem and Sobolev’s embedding on the -variable determines and are continuous on for every and almost every . From this, (5.26) yields the desired result. ∎
We begin by considering the case . We employ the differential equation (5.12) with the present restrictions on . Thus, we will derive bounds for and defined as in (5.12) for this case. Integrating by parts we get
(5.27) | ||||
where, given that , we have used . On the other hand,
(5.28) | ||||
Hence, Proposition 2.1 and Hölder’s inequality allow us to deduce
(5.29) | ||||
Since, with an implicit constant independent of , we combine the estimates for and to obtain the same differential inequality (5.13) adapted for this case. Consequently, this estimate, Gronwall’s inequality and the assumption imply . The proof of Theorem 1.3 (ii) is complete.
On the other hand, under the hypothesis of Theorem 1.3 (iii), since a.e , Lemma 5.1 and Plancherel’s identity assure that for . Then Gronwall’s inequality and the differential inequality (5.13) for this case yield , whenever . This consequence and Claim 1 complete the LWP results in , .
Now, we assume that . We write with . We first notice that Claim 1, identity (5.1) and the preceding well-posedness conclusion yield . Thus, we multiply the equation in (1.1) by and (5.9) by , then integrating in space and adding the resulting expressions reveal
(5.30) | ||||
Integrating by parts on the -variable,
(5.31) | ||||
The Cauchy-Schwarz inequality and Proposition 5.1 determine
(5.32) | ||||
Where we have used the identity . By complex interpolation (5.7) with and , we argue as in (5.18), using that to deduce . This previous estimate, the fact that , and (5.32) complete the study of .
On the other hand, since with implicit constant independent of , the estimate for follows the same ideas employed to estimate
5.2 Proof of Theorem 1.4
Without loss of generality we shall assume that , i.e., and . So that where , and . The solution of the IVP (1.1) can be represented by Duhamel’s formula
(5.34) |
Since our arguments require localizing near the origin, we consider a function such that when . Then taking the Fourier transform to the integral equation (5.34), we have
(5.35) |
where, recalling (2.2), .
Claim 2.
Let Then it holds
(5.36) |
Let us assume for the moment that Claim 2 holds, then
(5.37) |
We first notice that since , Fubini’s theorem and Sobolev embedding on the -variable determines that is continuous in for almost every . Therefore, given that (5.37) holds at , Fubini’s theorem shows that for almost every , then an application of Proposition 5.2 imposes that for almost every . From this fact, the integral equation (5.35) and Claim 2, we deduce Theorem 1.4, that is, for all and almost every .
Proof of Claim 2.
In virtue of Theorem 5.1,
(5.38) | ||||
To estimate the r.h.s of the last inequality, we decompose where . Then, writing and using (5.2) and Proposition 5.3,
(5.39) | ||||
where the last line is obtained by (2.7). We employ Sobolev’s embedding and complex interpolation (5.7) to deduce
(5.40) | ||||
where . Hence, (5.38), (5.39) and (5.40) yield
This completes the proof of Claim 2. ∎
5.3 Proof of Theorem 1.5
Here we assume that , , , where . Without loss of generality, we let , that is, and . Taking the Fourier transform in (5.34) and differentiating on the variable yield
(5.41) | ||||
where , we have used that and the identity
setting .
Claim 3.
It holds that
Proof.
We first deal with the term determined by the homogeneous part of the integral equation. We use Theorem 5.1, (5.2) and Proposition 5.3 to find
(5.42) | ||||
To estimate the last term on the r.h.s of the above expression, we use (5.2), (5.3), Plancherel’s identity and Young’s inequality to get
(5.43) | ||||
where we have also used (5.7) with and . Gathering (5.42) and (5.43), we complete the analysis of . Next, we shall prove that
(5.44) |
where is defined according to the norm . Once this has been established, following the reasoning in (5.42) and (5.43), it will follow
Indeed, (2.7) and Sobolev’s embedding show , whenever . Now, complex interpolation (5.7), Young’s inequality and Sobolev’s embedding determine
(5.45) | ||||
Since is a Banach algebra, , so it remains to derive a bound for the second term on the right hand side of equation (5.45). Let , applying Sobolev’s embedding and complex interpolation we find
(5.46) | ||||
Notice that since , . Plugging (5.46) in (5.45), we complete the deduction of (5.44). To prove the remaining estimate, i.e.,
(5.47) |
we write , then according to the arguments in (5.42) and (5.43), to deduce (5.47), it is enough to show
(5.48) |
To this aim, after some computations applying Theorem 5.1 and property (5.2), we employ complex interpolation and Young’s inequality to show
Then, (5.46) allows us to conclude that . Finally, since , , there exists some such that , then we have
(5.49) | ||||
Now, taking such that , (5.49) shows that . This in turn confirms the validity of (5.48). ∎
Consequently, from (5.41) and Claim 3, it follows:
if and only if | (5.50) | |||
Now, since (5.49) establishes that , Sobolev’s embedding determines that can be regarded as a continuous function on the and variables. Additionally, since , Fubinni’s theorem and Sobolev’s embedding shows that is continuous in for almost every . Given that (5.50) holds at , according to the preceding discussions and Proposition 5.2, we deduce
so that
(5.51) |
for almost every . This completes the deduction of identity (1.14). Now, recalling that the quantity is invariant for solution of the equation in (1.1), and that determines a continuous map, we let in (5.51) to find
(5.52) | ||||
6 Proof of Theorem 1.6
This section is aimed to briefly indicate the modifications needed to prove Theorem 1.6. We first recall that the IVP (1.2) is LWP in the space , by the results established in [5]. To prove well-posedness in the space determined by the norm
the key ingredient is the refined Strichartz estimate deduced in [5]:
Once the above lemma has been established, the proof of LWP in follows the same line of arguments leading to the conclusion of Theorem 1.1. Actually, this case does not require to estimate the norm , which slightly simplifies our arguments. We emphasize that Lemma 3.6 assures the existence of solutions of the IVP (1.2) in the space . Consequently, it follows that (1.2) is LWP in , .
On the other hand, setting
the resonant function determined by the equation in (1.2) is given by
Then, it is not difficult to see:
Proposition 6.1.
This in turn allows us to follow the same reasoning leading to the deduction of Theorem 1.2 to derive that the IVP (1.2) is LWP in , .
Concerning well-posedness in weighted spaces, here we replace equation (5.9) by
Then, employing the above identity, we can adapt the arguments in the proof of Theorem 1.3 to obtain the same well-posedness conclusion in anisotropic spaces for the equation in (1.2). Besides, the arguments in Proposition 5.3 show
whenever fixed and for all . Thus, the previous estimate allows us to follow the same arguments in the proof of Theorems 1.4 and 1.5 to obtain the same conclusions for the IVP (1.2). However, instead of (1.14) we get
for almost . This encloses the discussion leading to the deduction of Theorem 1.6.
7 Appendix: proof of Proposition 1.1
We first require to further decompose the lower frequency operator introduced in (2.4). Thus, for all dyadic number , let , and we denote by the associated operator defined as in (2.4), i.e., the operator determined by the -multiplier by the function .
We shall use the following result.
Lemma 7.1.
Let such that for some . Consider the operator determined by . Then
(7.1) |
In the above, denotes the usual Hardy-Littlewood maximal function.
Additionally, we will apply the following particular case of the Fefferman-Stein inequality:
Lemma 7.2.
([11]) Let be a sequence of locally integrable functions in . Let . Then
(7.2) |
Now, we are in the condition to deduce Proposition 1.1.
Proof of Proposition 1.1.
When on the l.h.s of (1.11), by writing and using that determines a bounded operator in , we have that (1.11) follows from Proposition 2.1.
We will assume that with . We write
(7.3) |
then neglecting the null measure sets where or , we observe that the integral in (7.3) is not null only when , in order words, when . Thus, by Bony’s paraproduct decomposition we find
where and . Now, we proceed to estimate each of the factors , . Since , , and the Hilbert transform determines a bounded operator in , by the Littlewood-Paley inequality and support considerations we have
(7.4) | ||||
for some adapted projections supported in frequency on the set , and with dyadic. Now, by employing Lemma 7.1, we deduce
Inserting the above expression on the r.h.s of (7.4), applying (7.2) and Lemma 7.1, we get
(7.5) | ||||
To estimate the preceding inequality, we write , then employing Lemma 7.1, it follows
so that
(7.6) |
Hence, plugging (7.6) in (7.5), by the Fefferman-Stein inequality and the Littlewood-Paley inequality, we conclude
(7.7) |
Now, replacing by in the arguments above, we derive the same estimate in (7.7) for the term .
A similar reasoning yields the desired estimate for . Indeed, since , , by Littlewood-Paley inequality
Now, by Lemma 7.1 it follows
Thus, the preceding estimates and (7.2) reveal
The estimate for follows from the same arguments employed to analyze . The proof of Proposition 1.1 is complete.
∎
Acknowledgements
This work was supported by CNPq Brazil. The author wishes to express his gratitude to Prof. Felipe Linares for bringing this problem to his attention and for the valuable suggestions regarding the manuscript.
References
- [1] L. Abdelouhab, J. Bona, M. Felland, and J.-C. Saut. Nonlocal models for nonlinear, dispersive waves. Physica D: Nonlinear Phenomena, 40(3):360–392, 1989.
- [2] B. Akers and P. A. Milewski. A Model Equation for Wavepacket Solitary Waves Arising from Capillary-Gravity Flows. Studies in Applied Mathematics, 122(3):249–274, 2009.
- [3] J. Biello and J. K. Hunter. Nonlinear Hamiltonian waves with constant frequency and surface waves on vorticity discontinuities. Communications on Pure and Applied Mathematics, 63(3):303–336, 2010.
- [4] E. Bustamante, J. Jiménez, and J. Mejía. Periodic Cauchy problem for one two-dimensional generalization of the Benjamin-Ono equation in Sobolev spaces of low regularity. Nonlinear Analysis, 188:50–69, 2019.
- [5] E. Bustamante, J. Jiménez, and J. Mejía. The Cauchy problem for a family of two-dimensional fractional Benjamin-Ono equations. Communications on Pure & Applied Analysis, 18(3):1177–1203, 2019.
- [6] L. Dawson, H. McGahagan, and G. Ponce. On the Decay Properties of Solutions to a Class of Schrödinger Equations. Proceedings of the American Mathematical Society, 136(6):2081–2090, 2008.
- [7] J. Duoandikoetxea. Fourier Analysis. Crm Proceedings & Lecture Notes. American Mathematical Soc., 2001.
- [8] O. Duque Gómez. Sobre una versión bidimensional de la ecuación Benjamin-Ono generalizada. PhD thesis, Universidad Nacional de Colombia-Bogotá, 2014.
- [9] A. Esfahani and A. Pastor. Ill-posedness results for the (generalized) Benjamin-Ono-Zakharov-Kuznetsov equation. Proceedings of the American Mathematical Society, 139(3):943–956, 2011.
- [10] A. Esfahani and A. Pastor. Two dimensional solitary waves in shear flows. Calculus of Variations and Partial Differential Equations, 57(4):102, 2018.
- [11] C. Fefferman and E. M. Stein. Some Maximal Inequalities. American Journal of Mathematics, 93(1):107–115, 1971.
- [12] G. Fonseca, F. Linares, and G. Ponce. The IVP for the Benjamin-Ono equation in weighted Sobolev spaces II. Journal of Functional Analysis, 262(5):2031 – 2049, 2012.
- [13] G. Fonseca, F. Linares, and G. Ponce. The IVP for the dispersion generalized Benjamin-Ono equation in weighted Sobolev spaces. Annales de l’Institut Henri Poincare (C) Non Linear Analysis, 30(5):763 – 790, 2013.
- [14] G. Fonseca and G. Ponce. The IVP for the Benjamin-Ono equation in weighted Sobolev spaces. Journal of Functional Analysis, 260(2):436–459, 2011.
- [15] L. Grafakos and S. Oh. The Kato-Ponce Inequality. Communications in Partial Differential Equations, 39(6):1128–1157, 2014.
- [16] Z. Guo and T. Oh. Non-Existence of Solutions for the Periodic Cubic NLS below . International Mathematics Research Notices, 2018(6):1656–1729, 2018.
- [17] M. Hadac, S. Herr, and H. Koch. Well-posedness and scattering for the KP-II equation in a critical space. Annales de l’Institut Henri Poincare (C) Non Linear Analysis, 26(3):917–941, 2009.
- [18] J. Hickman, F. Linares, O. Riaño, K. Rogers, and J. Wright. On a Higher Dimensional Version of the Benjamin–Ono Equation. SIAM Journal on Mathematical Analysis, 51(6):4544–4569, 2019.
- [19] J. K. Hunter, M. Ifrim, D. Tataru, and T. K. Wong. Long time solutions for a Burgers-Hilbert equation via a modified energy method. Proceedings of the American Mathematical Society, 143(8):3407–3412, 2015.
- [20] A. Ionescu and C. Kenig. Local and global wellposedness of periodic KP-I equations. Annals of Mathematics Studies, (163):181–211, 2007.
- [21] A. D. Ionescu and C. E. Kenig. Global Well-Posedness of the Benjamin-Ono Equation in Low-Regularity Spaces. Journal of the American Mathematical Society, 20(3):753–798, 2007.
- [22] A. D. Ionescu, C. E. Kenig, and D. Tataru. Global well-posedness of the KP-I initial-value problem in the energy space. Inventiones mathematicae, 173(2):265–304, 2008.
- [23] R. J. Iório. On the Cauchy problem for the Benjamin-Ono equation. Communications in Partial Differential Equations, 11(10):1031–1081, 1986.
- [24] R. J. Iório and W. V. L. Nunes. On equations of KP-type. Proceedings of the Royal Society of Edinburgh: Section A Mathematics, 128(4):725–743, 1998.
- [25] R. J. Iorio, Jr and V. d. M. a. Iorio. Fourier Analysis and Partial Differential Equations. Cambridge Studies in Advanced Mathematics. Cambridge University Press, 2001.
- [26] T. Kato and G. Ponce. Commutator estimates and the Euler and Navier-Stokes equations. Communications on Pure and Applied Mathematics, 41(7):891–907, 1988.
- [27] C. E. Kenig. On the local and global well-posedness theory for the KP-I equation. Annales de l’Institut Henri Poincare (C) Non Linear Analysis, 21(6):827 – 838, 2004.
- [28] C. E. Kenig and K. D. Koenig. On the local well-posedness of the Benjamin-Ono and modified Benjamin-Ono equations. Mathematical Research Letters, 10(6):879–895, 2003.
- [29] D. Li. On Kato-Ponce and Fractional Leibniz. Revista Matemática Iberoamericana, 35(1):23–100, 2019.
- [30] F. Linares, M. Panthee, T. Robert, and N. Tzvetkov. On the periodic Zakharov-Kuznetsov equation. Discrete & Continuous Dynamical Systems-A, 39(6):3521–3533, 2019.
- [31] F. Linares, D. Pilod, and J.-C. Saut. The Cauchy Problem for the Fractional Kadomtsev-Petviashvili Equations. SIAM J. Math. Analysis, 50:3172–3209, 2018.
- [32] F. Linares and G. Ponce. Introduction to Nonlinear Dispersive Equations. Universitext. Springer New York, 2015.
- [33] J. Lizarazo. El problema de Cauchy de la clase de ecuaciones de dispersión generalizada de Benjamin-Ono bidimensionales. PhD thesis, Universidad Nacional de Colombia-Bogotá, 2018.
- [34] L. Molinet. Global well-posedness in for the periodic Benjamin-Ono equation. American Journal of Mathematics, 130(3):635–683, 2008.
- [35] L. Molinet and D. Pilod. The Cauchy problem for the Benjamin-Ono equation in revisited. Anal. PDE, 5(2):365–395, 2012.
- [36] C. Muscalu, J. Pipher, T. Tao, and C. Thiele. Bi-parameter paraproducts. Acta Math., 193(2):269–296, 2004.
- [37] J. Nahas and G. Ponce. On the Persistent Properties of Solutions to Semi-Linear Schrödinger Equation. Communications in Partial Differential Equations, 34(10):1208–1227, 2009.
- [38] D. E. Pelinovsky and V. I. Shrira. Collapse transformation for self-focusing solitary waves in boundary-layer type shear flows. Physics Letters A, 206(3):195 – 202, 1995.
- [39] G. Ponce. On the global well-posedness of the Benjamin-Ono equation. Differential Integral Equations, 4(3):527–542, 1991.
- [40] F. Ribaud and S. Vento. Local and global well-posedness results for the Benjamin-Ono-Zakharov-Kuznetsov equation. Discrete & Continuous Dynamical Systems-A, 37(1):449–483, 2017.
- [41] T. Robert. Global well-posedness of partially periodic KP-I equation in the energy space and application. Annales de l’Institut Henri Poincare (C) Non Linear Analysis, 35(7):1773 – 1826, 2018.
- [42] T. Robert. On the Cauchy problem for the periodic fifth-order KP-I equation. Differential Integral Equations, 32(11/12):679–704, 2019.
- [43] R. Schippa. On the Cauchy problem for higher dimensional Benjamin-Ono and Zakharov-Kuznetsov equations. arXiv e-prints, page arXiv:1903.02027, 2019.
- [44] R. Schippa. On short-time bilinear Strichartz estimates and applications to the Shrira equation. Nonlinear Analysis, 198:111910, 2020.
- [45] E. M. Stein. The characterization of functions arising as potentials. Bull. Amer. Math. Soc., 67(1):102–104, 1961.
- [46] T. Tao. Global well-posedness of the Benjamin-Ono equation in . Journal of Hyperbolic Differential Equations, 01(01):27–49, 2004.
- [47] D. Yafaev. Sharp Constants in the Hardy-Rellich Inequalities. Journal of Functional Analysis, 168(1):121–144, 1999.
- [48] Y. Zhang. Local well-posedness of KP-I initial value problem on torus in the Besov space. Communications in Partial Differential Equations, 41(2):256–281, 2016.