1. Introduction
Studies of the uniqueness, existence, and qualitative properties of solutions to dispersive partial differential equations constitute a substantial body of modern research. In this paper, we concentrate on analysing a specific problem: the nonlinear Benjamin equation
where
are real parameters and
is the standard Hilbert transform. The quasilinear PDE (
1) was originally introduced by T.B. Benjamin in his studies of the propagation of internal waves in a two-fluid system [
1,
2]. We note that setting
in (
1) provides the classical Korteweg–de Vries equation (KdV), while letting
recovers the well-known Benjamin–Ono (BO) equation as a special case.
The analysis of (
1), particularly the study of its low-regularity solutions, has garnered significant attention in the specialized literature. A variety of methods have been employed, ranging from bilinear estimates to the
I-method (see for example, [
3,
4,
5,
6]). Historically, the first low-regularity results were obtained in [
6], within the
settings. These results were subsequently extended to the
,
settings in [
4]. Our work extends these findings by exploring well-posedness in the newly introduced
spaces, offering enhanced control over asymptotic behavior. The well-posedness analysis in [
4,
6] is local in time.
The first global well-posedness results in
, for
, were obtained in [
5]. The local analysis follows a similar approach to that in [
4,
6], but employs a refined bilinear estimate for a frequency cut-off operator
I applied to the quadratic nonlinearity
. In addition to demonstrating local solvability, this technique provides explicit estimates for the growth rate of mild solutions and their intervals of existence. When combined with a continuation method, this approach establishes global well-posedness.
Furthermore, studies on variable dispersion, such as those examining special cases of the KdV and BO equations (see, for example, [
7,
8,
9,
10,
11] and references therein), have shown how dispersion affects wave propagation and stability. Our research endeavours to consider the existing understanding of the impact of variable dispersion on the well-posedness and asymptotic properties of solutions to the nonlinear Benjamin equation within the framework of weighted Sobolev spaces. The study of low dispersion properties of a fractional variant of the Benjamin–Bona–Mahony (BBM) equation within
-spaces for
has been carried out in [
12], highlighting the importance of these spaces in controlling the asymptotic behaviour of solutions.
The local and global analyses in [
4,
5,
6] do not extend to the critical case of
. The main obstacle is that the bilinear estimate used in [
4,
6], as well as its
I analogue from [
5], fails for
. The critical case was recently addressed in [
3], where the main tool used was the Besov–Bourgain spaces.
A distinctive feature of Equation (
1) is its non-locality, primarily due to the presence of the global in space Hilbert transform
. This non-locality presents significant challenges in predicting the long-term behaviour of internal waves in fluid dynamics. Furthermore, the Fourier symbol of the group generated by the linear part of (
1) has finite regularity at the origin, complicating the analysis of wave propagation. By addressing these challenges, our study contributes to the accurate modeling of wave propagation, which is critical in both natural and engineered systems. As a consequence of the non-locality, one cannot expect super-algebraic spatial decay of solutions, even for the rapidly decreasing initial data
from the Schwartz class
. In context of the closely related Benjamin–Ono equation, this situation and its implications were thoroughly investigated in [
13,
14,
15,
16,
17].
A systematic study of the interplay between asymptotic behavior, regularity, and the existence of global solutions to the Benjamin equation (
1) was initiated in [
18]. In particular, building on the ideas from [
13,
14,
15,
16,
17], it is demonstrated in [
18] that for initial data
, with
,
and
, the Equation (
1) is globally well-posed in
. The result indicates that, due to the jump discontinuity in the Fourier symbol of the operator
, one cannot expect solutions to decay algebraically at infinity faster than
. The situation improves slightly in the scale of weighted homogeneous Sobolev spaces
, where the propagation of the discontinuity in the Fourier image of the solution can be properly controlled for
,
.
In this paper, we continue the investigation of the nonlinear and non-local dispersive model (
1) in weighted Sobolev-like spaces, building on the work initiated in [
18]. In contrast with the previously cited works, our analysis relies on variable-weight Sobolev spaces
(For the definition of the scale
, see [
19] and
Section 2 below), which were introduced recently in [
19]. This approach enables a more precise description of how solutions behave at large distances, which is crucial for understanding wave propagation in non-local dispersive systems, thereby revealing the intricate mechanisms underlying their behaviour. In the context of the Benjamin equation (and in similar dispersive models, such as the Benjamin–Ono equation), these spaces arise naturally. Unlike the weighted spaces
and
from [
13,
14,
15,
16,
17,
18], they provide direct control over the large-
x asymptotic behavior of classical solutions and their weak derivatives up to and including order
s. This idea allows for a more refined analysis of the large-
x asymptotic behavior of solutions.
Thanks to this property, we can demonstrate that for sufficiently regular input data, the jump discontinuities in the frequency space—caused by the finite regularity of the Fourier symbol associated with the linear part of (
1)—cannot propagate beyond the origin. As a direct consequence, it follows that the large-
x asymptotic behavior of classical solutions
is determined entirely by the behaviour of their Fourier images
near the origin. Consequently, this asymptotic behavior can be computed explicitly in terms of quantities associated with the input data. Once the asymptotic identity is established, a range of qualitative results, similar to those in [
18] for the Benjamin equation and [
13,
14,
15,
16,
17] for the Benjamin–Ono equation, automatically follows. Although our analysis focuses solely on the Benjamin equation (
1), the approach is sufficiently general to be adapted for analyzing classical solutions to non-local quasi-linear dispersive models of Benjamin-type, where linear wave propagation operators exhibit finitely many jump discontinuities in the frequency domain.
The paper is organized as follows:
Section 2 provides a brief review of the basic properties of the scale
. In
Section 3, we derive several technical estimates related to both the linear and nonlinear parts of (
1).
Section 4 discusses the model’s well-posedness in weighted settings and its large-
x asymptotic behavior.
Section 6 concludes the paper.
Notation
- (i)
x is the physical variable
- (ii)
is the frequency variable
- (iii)
* denotes Fourier convolution
- (iv)
is the Fourier transform
- (v)
is the inverse Fourier transform
- (vi)
represent the weighted Lebesgue spaces.
- (vii)
denotes the space of real valued tempered distributions
- (viii)
c is a generic constant.
2. Function-Theoretic Framework
To begin, we define the normalized Fourier transform and its inverse
respectively.
In the forthcoming analysis, we employ the scale of the variable-weighted Sobolev spaces, introduced recently in [
19]. For real valued functions, these spaces are defined as
where
is the space of real valued tempered distributions,
are Fourier multipliers (projectors) associated with the Heaviside functions
,
, and
is the standard homogeneous Sobolev space of order
s (see e.g., [
20]). Note that the Fourier images of real valued distributions are Hermitian and, hence, the concrete choice of the sign (+ or −) in (
3) is irrelevant and the symmetric, positive definite bilinear form
defines a real Hilbert structure in
. An interpolation argument, employed in
Section 3, requires a complex version of
. In a usual way,
is complexified by letting
Directly from (
2)–(
5), it follows that for
,
, and
are equal as Banach spaces. However, when
, the former class carries information about the large
x asymptotic of all weak derivatives up to and including order
s, and, hence, is more suitable for the analysis of the fine interplay between regularity and the large
x behaviour of solutions of (
1).
Basic properties of
are established in [
19]. In particular, we have
where
is the standard complex interpolation functor (see [
20]). We note that the embeddings (
6) are dense, as the Schwartz class
of smooth rapidly decreasing test functions is dense in each of
,
, and
.
In addition to (
6)–(
8), we need the following technical
Lemma 1. The embeddingis dense and compact. Proof. As
, it is sufficient to verify the real valued version of (
9) only (for details see [
21]).
(a) The denseness and continuity of (
9) follows from (
6). To show compactness, consider
, with
. In view of the assumption
, we have
uniformly for
.
(b) As
u is assumed to be real valued, its Fourier image
is Hermitian (i.e.,
,
). Therefore, for
, we have
with
. To bound
, as in [
19], we let
By the definition of
, for
, we have
, with
. This identity, combined with the explicit formula
the assumption
and the Cauchy–Schwarz inequality, indicates that
when
. Hence,
, uniformly for
.
To estimate
, we let
,
,
. The analysis presented in [
22] indicates that
Using these facts, the inclusion
, and the standard Minkowski inequality, we obtain initially
then, using the elementary inequality
(which holds uniformly for
, with some constant
) and Formulas (
3), (
10) and (
11),
provided
and
. We note that the right hand side of (
12) tends to zero, as
, as
, for all
, provided
.
(c) From parts (a) and (b) of the proof, it follows that the the unit ball of
is equibounded, equitight, and equicontinuous in
, provided
and
,
, for some
. Hence, on account of the Kolmogorov–Riesz theorem, the embedding (
9) is indeed compact if
. The proof is complete. □
4. Large- Asymptotic of Classical Solutions
Technical estimates of
Section 2 and
Section 3 allow us to deduce the global weighted solvability of (
1) directly from the well-known unweighted well-posedness results, see, e.g., the discussion in [
4,
5,
6,
18], the classical papers [
25,
26], and note that (
1) is the low order perturbation of the Korteweg–de Vries equation.
Theorem 1. Assume that , with and . Then, for any finite value of , the unique global classical solution u to (
1)
satisfies, Proof. (a) It is well known, that for the initial data
,
, the Benjamin equation is classically globally well-posed (see the remark preceding Theorem 1). In particular, for such input data and any finite value of
, we have
. Furthermore, for any
Hence, to complete the proof of Theorem 1, we need show that under the assumption
, the inclusion (
21) holds.
(b) For the sake of brevity, for
,
, we let
On account of Lemma 1, the embedding
is dense and compact. Hence, for
,
can be realized as a completion of
with respect to the inner product
, with some positive definite bounded symmetric map
. It is not difficult to verify that
extends uniquely to a symmetric positive definite bounded linear map
and that
. By virtue of (
24), the last inclusion indicates that
, with
, is compact.
(c) From part (a) of the proof, the spectrum of is discrete, while the collection of associated eigenfunctions is a complete orthogonal basis in and in , .
We denote
and let
be the associated orthogonal projector. From the definition of
, for
, we have
. Hence, in view of Lemma 2 and part (a) of the proof, the sequence of linear Galerkin approximations
is well defined.
The special choice of the orthogonal basis indicates that
Letting
in (
27) and using Lemmas 2 and 3, for
, we obtain
Bound (
27), combined with the standard Gronwall’s inequality, indicates that
uniformly in
, for every finite fixed value of
. In turn, for
, using (
14), (
16), (
24) and (
29), integrating over the interval
and using the Cauchy–Schwarz inequality, we obtain
where the generic constant
depends on the coefficients
,
,
, and
of (
1) only. From the last bound, it follows that
uniformly in
.
(d) The rest of the proof is standard. Using the uniform inclusions (
29) and (
30) and passing to subsequences, we conclude that: (i)
and
converge weak-* to some
w and
in
and
, respectively; (ii)
converges weakly to
(and by construction, strongly to
) in
; and (iii) in view of (
24),
, converges strongly in
. Passing to the weak limit in (
25) and (), we see that for
,
Subtracting (
22), (
23), (
31) and (
32) and letting
, in view of (
13) and (
16), we obtain
As
, Gronwall’s inequality gives
, and as
, (
21) follows. The proof is complete. □
By virtue of Theorem 1, Fourier images of classical solutions with input data
from
,
,
are regular away from the origin. Hence, the large-
x asymptotic of
is controlled solely by the small
behavior of its Fourier image
. Passing to the frequency domain in (
1), we have
where
is the Fourier symbol of operator
and
denotes the Fourier convolution square. The large-
x asymptotic of
is described by
Lemma 4. Assume , . If , for some positive integer p, thenuniformly in bounded time intervals. Proof. (a) We begin with some basic estimates. As
, we have
, with
. Consequently,
where
. The Cauchy–Schwarz inequality, (
37) and (
38) and assumption
, combined together, indicate that
with
, depending on
s,
r and
only. Further, on the account of (
6), we have
Consequently, for any
, with
, from the interpolation identity (
7), it follows that
for some
. This estimate, combined with the definition (
2)–(
5) of
-norm, implies that
From our assumption,
. Hence,
(b) We write
The inclusion
indicates that
. Therefore, (
35) and (
36), combined with the Fubini–Tonelli theorem, provides
Partial integration gives
Hence, from the last two formulas, we infer
It is not difficult to verify that
is a linear combination of the powers
,
. In view of (
39), this observation allows us to conclude that each inner product appearing in (
41) is finite. Differentiating and using the Faa di Bruno formula, we have
where
are polynomials of degree
in
. Hence, the inner integral in (
42) can be written as the sum of terms appearing in (
40) and, consequently, is absolutely integrable in
. On the account of classical Lebesgue–Riemann lemma, we conclude that
uniformly in bounded time intervals. Finally, using (
41)–(
43) and the elementary identities
after some simplifications, we arrive at (
35) and (
36). □
As
and
, with
,
are Banach algebras (see [
19]), it follows that
. Further, on the account of (
2)–(
5) and (
6)–(
8), we have
. Hence, for
, repeating verbatim arguments of Lemma 4, we obtain
uniformly in bounded time intervals.
For the sake of brevity, we let
The bulk asymptotic of the classical solutions to (
1) is given by the following
Theorem 2. Assume , with . Then,uniformly in bounded time intervals. Proof. (a) Directly from (
35), (
36) and (
44), it follows that
for large values of
x and uniformly in bounded time intervals. We still need to compute moments, associated with
, and the Hilbert transform.
(b) Elementary calculations and conservation of
-norm along classical trajectories of (
1) indicate that
From the remark preceding formula (
44),
while by our assumption
. Hence, the restrictions
as well as their derivatives of orders
extend continuously to
. Using this fact, it is not difficult to verify that
Direct calculations yield
In connection with the last formula, we note that the quantity
is preserved along the classical trajectories of (
1) and, hence,
. Applying the functional
to the both sides of (
1), using partial integration and the identity
, we arrive at
so that
. Combining all the formulas together and integrating
with respect to
s over
, after some simplifications, (
45) follows. □
The proof of Theorem 2 relies on Lemma 4 with
. The cases of
can be treated using analogy, and the resulting asymptotic formulas are obtained by truncating (
45) at the
p-th power of
.
As an immediate consequence of Theorem 2, we have the following analogue of qualitative results obtained in [
18] for the Benjamin equation (see also [
13,
14,
15,
16,
17] for the closely related Benjamin–Ono equation).
Corollary 1. Assume , with .
- (i)
If , then , for each .
- (ii)
If , and , then , for each .
- (iii)
If , and , then , for each .
- (iv)
If , , and , then , for each .
- (v)
If , , at three distinct time points , and , then , for all .
- (vi)
If , and the associated nontrivial classical solution satisfies , , at time , then , at . In particular, there exists a classical solution to (1) with . - (vii)
There exists two distinct and nontrivial classical solutions and of (1), satisfying and , for each and any finite value of .
Proof. (a) Items (i)–(iv) follow from a direct evaluation of the coefficients appearing in (
45). In particular, in (i)
, from assumptions (ii), (iii), and (iv), we have
,
and
, respectively.
(b) If (v) holds, then the quadratic polynomial has three distinct real roots and, hence, vanishes identically. In particular, this implies that , and as the last quantity is invariant under the Benjamin flow, for all .
(c) Under the assumptions of (vi), at time
, the principal asymptotic terms in (
45) vanish identically. This implies, in particular, that
Hence,
, at
. Further, for
, with
, we have
and
. As a consequence, the associated classical solution satisfies (vi) with
.
(d) Finally, choosing the initial data
,
, so that
,
,
, and
, it is not difficult to verify that
Further, for the data from the Schwartz class
, the inclusion
holds for any
. Hence, repeating the calculations of Lemma 4 and Theorem 2 verbatim, but with
, we obtain
The last identity yields (vii). □