Waves of maximal height for a class of nonlocal equations with inhomogeneous symbols
Abstract.
In this paper, we consider a class of nonlocal equations where the convolution kernel is given by a Bessel potential symbol of order for . Based on the properties of the convolution operator, we apply a global bifurcation technique to show the existence of a highest, even, -periodic traveling-wave solution. The regularity of this wave is proved to be exactly Lipschitz.
Key words and phrases:
Whitham type,inhomogeneous,nonlocal,maximal height,water waves2010 Mathematics Subject Classification:
76B15, 76B03, 35S30, 35A201. Introduction
Our object of study is the following nonlocal equation
(1.1) |
where denotes the Fourier multiplier operator with symbol , . Equation (1.1) is known as the fractional Korteweg–de Vries equation. We are looking for a -periodic traveling-wave solution , where denotes the speed of the right-propagating wave. In this moving frame, equation (1.1) takes the form
(1.2) |
where the equation has been integrated once, and the constant of integration set to zero. In fact, there is no loss of generality in doing so due to the Galilean change of variables:
for any , which maps solutions of to solutions of a new equation of the same form. In this paper, the above multiplier has order of , which is less than . This type of dispersive terms tends to behave smoothly at far field and worse at the origin. Solutions to the type of equation (1.1) with negative orders feature wave breaking and peaked periodic waves; see discussions in [13, 15]. It should be mentioned that for a linear operator of order strictly smaller than , the regularity of the highest, periodic, traveling-wave solution does not depend on the order operator; see Theorem 4.5. Instead, the dispersive term may affect the angle at the wave crest.
Investigation on highest waves for various types of water-waves equations has a long history. One of the earliest works to study peaked waves was Stokes, who [19] conjectured that for a fixed wavelength and the gravitational acceleration, the Euler equation has a highest, periodic traveling-wave with a sharp crest of angle . However, attempts to prove the Stokes conjecture had failed until the work by Amick, Fraekel, and Toland [1]. Whitham [21] also made an assumption that the solution to the Whitham equation reached its critical height when (the notation has been changed to match equation (1.2)). Recently, Ehrnström and Wahlén [12] confirmed the existence and regularity of a highest and periodic traveling-wave solution for the Whitham equation, and hence successfully proved Whitham conjecture. This solution features a sharp crest and is exactly -Hölder continuous at the peak and smooth between its crest and trough on the half-period. After this breakthrough work, there have been an increasing number of studies on nonlocal equations using their approach. One of the earliest papers in this direction is to justify the convexity of the wave profile between consecutive crests using a computer-assisted proof [14]. A similar result as in [12] but for solitary waves was also obtained by Truong, Wahlén, and Wheeler [20] using a center manifold theorem approach. Moreover, Ehrnström, Johnson, and Claasen [9] proved the existence and regularity of a highest wave for the bidirectional Whitham equation with cubic nonlinearity and a Fourier multiplier symbol . They showed that the highest cusped wave has a singularity at its crest of the form . Utilizing the same approach as in [12], Arnesen [3] considered the Degasperis–Procesi equation. This equation has a nonlocal form with a quadratic nonlinearity and a Fourier multiplier symbol . Using the method in [12], Arnesen constructed a family of highest, peaked, periodic traveling waves and showed that any highest, even, periodic wave of the Degasperis–Procesi equation is exactly Lipschitz at its crest. The author, therefore, excluded the existence of even, periodic, cusped traveling-wave solutions.
Since the unidirectional Whitham equation is nonlocal, the equation (1.1) is obtained, up to a scaling factor, by replacing the Whitham dispersion relation by a more regular dispersive term. Building on the integral kernel , we prove the existence and regularity properties of a highest, -periodic traveling-wave solution of equation (1.2). Our most direct influence are the works by Ehrnström and Wahlén [12] and by Bruell and Dhara [5]. In the former work, the investigation started with the Whitham equation
where is the Fourier multiplier operator given by the symbol
This multiplier has order greater than , is inhomogeneous and completely monotone. The main result is the existence of a highest, cusped, -periodic traveling-wave solution, which is even, strictly decreasing, smooth on each half-period, and belongs to the Hölder space . The proof is based on the regularity and monotonicity properties of the convolution kernel induced by . On the other hand, Bruell and Dhara [5] considered the fractional KdV equation as in equation (1.1), where is the Fourier multiplier operator given by the homogeneous symbol
This multiplier has order less than , is homogeneous, and completely monotone. The authors showed the existence of a highest, -periodic traveling waves solution, which is Lipschitz continuous. Their method is based on the nonlocal approach in the paper [12] but adapt to treat homogeneous symbols. The convolution kernel associated with this homogeneous symbol cannot be identified with a positive, decaying function on the real line. To compensate for this lack of positivity, solutions are assumed to have the zero mean, and hence the constant of integration in equation (1.2) now becomes .
In the present paper, we will combine the approach in [12] to treat inhomogeneous symbols, and the derivation for an estimate of a priori traveling-wave solution to nonlocal equations as in [5]. We recall that the symbol for the fractional KdV equation (1.1) is Bessel potential function, which is inhomogeneous and of order strictly smaller than . We also do not make the assumption of zero mean solution as in [5]. Since the kernel operator is positive, we choose the constant of integration in equation (1.2) to be . Moreover, even though our kernel is even and smooth for all with rapid decay derivatives, we do not use a closed formula for in the physical space, as well as for its periodization as in [5, 12]. It is also worth to mention that the kernel does not exhibit the completely monotone feature, which plays a main role for the analysis in [5, 12]. Instead, our investigation only relies on its qualitative behaviors such as positivity of and the signs of its derivatives; see Section 2 and 3. The properties of the convolution kernel are proved in the ongoing work by Bruell, Ehrnström, Johnson, and Wahlén [6].
We now briefly discuss the strategies we used to derive the results. We will be using a global bifurcation theory as in the paper [7] to build a global, locally analytic curve of periodic smooth waves where the solution maximum approaches . We further show that the only possibility along the main bifurcation curve is that must approach . Moreover, we investigate the solution satisfying . Using the properties of the convolution operator , we obtain Lipschitz continuous at its crest.
Let us formulate our main existence theorem which constructs solutions bifurcating from nontrivial, even, smooth, and -periodic traveling-wave solutions of equation (1.1). At the end of the bifurcation curve, the limiting solution reaches its highest and is precisely Lipschitz continuous.
Theorem 1.1 (Main theorem).
For each integer , there exists a wave speed and a global bifurcation branch:
of nontrivial, -periodic, smooth, even solutions to the steady equation (1.2) for , emerging from the bifurcation point . Moreover, given any unbounded sequence of positive numbers , there exists a subsequence of , which converges uniformly to a limiting traveling-wave solution that solves (1.2) and satisfies
The limiting wave is strictly increasing on and exactly Lipschitz at .
The outline of our study is as follows. In Section 2 we introduce notations and functional-analytic settings that we encounter in the analysis. We also inspect the convolution kernel corresponding to the symbol and show that the action of the operator on periodic functions on the real line is in fact similar to the periodized kernel acts on these functions in one period. In Section 3, we discuss some interesting analytic features of the periodized kernel . In fact, both and share many properties.
Section 4 is the heart of this paper. We provide some general estimates of an even solution to equation (1.2). In particular, using the regularity and monotonicity properties of the convolution kernel , we prove that if a traveling-wave solution, which is even, periodic, and monotone on half the period, attains its maximum equal to the wave speed , then it is Lipschitz continuous. Finally, Section 5 proved our main theorem, Theorem 1.1. It consists of the bifurcation analysis from [5, 11, 10], in which we rule out certain possibilities for the bifurcation curve. We, therefore, construct a global curve of sinusoidal, periodic smooth waves along which , and an analysis of solutions at their maximum .
2. Functional-analytic setting and general conventions
Setting , we denote by the space of test functions on , whose dual space, the space of distributions on , is . If is the space of rapidly decaying functions from to and denotes its dual space, let be the Fourier transformation on the torus defined by the duality on via
To simplify our notations, throughout this paper, if and are elements in an ordered Banach space, we write if there exists a constant such that . Moreover, the notation is used whenever and . We also denote and by the set of natural numbers including zero. The space denotes the set of all bounded linear operators from to .
Next, let us investigate how the convolution acts on periodic functions. Suppose that is periodic and even. Since is in , we can write:
It is clear from the definition of that it is -periodic, even, and continuous on . Moreover, Minkowski’s inequality shows that belongs to , for . Therefore, Carleson–Hunt Theorem [17] implies that can be approximated pointwise by its Fourier series:
and the Fourier coefficients of are given by
(2.1) |
Thus, the periodic problem is given by the same multiplier as the problem on the line, so we have the representation:
The periodized operator is introduced to aid the analysis of periodic solutions satisfying certain sign conditions in a half-period. From the above analysis, can be expressed as the Fourier series
(2.2) |
and the relationship between and is
(2.3) |
Note that this sum is absolutely convergent since has rapid decay.
The operator
Now let be the operator
defined via duality on the space of tempered distributions. Then one can see that for a continuous periodic function , the operator is given by
and more generally by
if is bounded and continuous. Thus, we have the following formula for
(2.4) |
Next, we shall say that a function is Hölder continuous of regularity at a point if
and let
be the space of -Hölder continuous functions on . If , then
denotes the space of -times continuously differentiable functions whose -th derivative is -Hölder continuous on .
Before we proceed to discuss the space of the operator , let us recall the definition of Besov spaces using the Littlewood–Paley decomposition. Let be a family of smooth and compactly supported functions, where
for all . For a tempered distribution , we let , so that
Then the Besov spaces , , , are defined by
and for and , we instead define
For a -periodic tempered distribution , we have the identity
so that is a trigonometric polynomial.
Moreover, for , we define the Zygmund spaces of order by
and recall that when is a positive noninteger, we have ; and when is a nonnegative integer, we have . As a consequence of Littlewood–Paley theory, we have the relation for any positive noninteger . Thus, the Hölder spaces on the torus are completely characterized by Fourier series. It , then is a proper subset of and
where denotes the space of Lipschitz continuous functions on .
3. Properties of the periodized kernel
This section is devoted to summarize some properties of the kernel . We recall the Fourier representation of (2.2) and the relationship (2.3) between and . Recalling that and the Fourier multiplier is given by the Bessel potential for , we record some properties of the periodized operator .
Theorem 3.1.
Let . The kernel has the following properties:
-
(a)
is even, positive, -periodic, and continuous. is smooth on .
-
(b)
is in . In particular, is integrable. Moreover, is Hölder continuous with if , Lipschitz continuous when , and continuously differentiable if .
-
(c)
is decreasing on .
Proof.
We first observe that this theorem holds when we replace by ; see, for example, by Grafakos [16]. Due to the Fourier representation (2.2) and the relation (2.3), the evenness, positivity, periodicity, and regularity properties of are inherited from . By [16, Proposition 1.2.5], we know that is smooth on , and so is by the relation (2.3). Moreover, the regularity of in part (b) follows from the regularity of in [16, Section 1.3]. Part (c), however, is proved in the ongoing work by Bruell, Ehrnström, Johnson, and Wahlén [6]. ∎
We know from the above theorem that is even and decreasing on . The next lemma provides monotonicity property for the operator based on .
Lemma 3.2.
Let . The operator is parity preserving on . Moreover, if are odd functions satisfying on , then either
or on .
Proof.
To see that is parity-preserving, let be an odd function. Then since is even, we have
which shows that is odd. Similarly, when is even, is also even.
To show the second part of the lemma, let be odd functions such that on . By contradiction, assume that there exists such that . Then we compute
Since , we have
Then the evenness, periodicity, and monotonicity of yield
for all in the interval , and therefore, , which is a contradiction to our assumption , unless on . ∎
4. A priori properties of periodic traveling-wave solutions
In this section, we provide some basic properties of solutions to the main equation (1.2) such as its a priori estimate, monotonicity, and regularity. When we say is a solution, must be real-valued, bounded, and satisfy equation (1.2) pointwise. Fixing , to aid our analysis, we re-write equation (1.2) as
(4.1) |
Our first result is a rough bound for a solution according to the wave speed .
Lemma 4.1.
Proof.
By Lemma 3.2, is a strictly monotone operator. Since furthermore for constants , we therefore obtain that
In particular, , or equivalently, we have the inequality
Similar arguments can be made to obtain
which yields that
Combining both cases, we have the desired estimates. ∎
Remark 4.2.
We come to the first set of results in this section, which we need to rule out the closed loop possibility of the solution in the global bifurcation analysis. It shows that any nontrivial, even, periodic solution, which is nonincreasing on , must be strictly decreasing and less than on . Thus, we conclude that .
Lemma 4.3.
Any nontrivial, -periodic, even solution of (1.2) which is nonincreasing on satisfies
For such a solution, one necessarily has . Moreover, if , then .
Proof.
Since , we can take the derivative of equation (4.1) to obain
Since is nonincreasing on , we have on . We want to show that for . Since is odd, nontrivial, and nonpositive on , Lemma 3.2 implies on . Thus, we have
and hence we have shown that and on . On the other hand, by Lemma 4.1, we know , so that .
To prove the second part, assume that . Differentiating twice equation (4.1) gives
Then evaluating this equality by and using the evenness of and , we compute
Since is in and is integrable, the first integral vanishes as , so does the boundary term . The term also vanishes since . By Theorem 3.1 and what we just proved, both and are strictly negative on , and hence we have
which shows that . ∎
In both papers [5, 12], a solution to the nonlocal equation is smooth when it is below its maximum value. In the next theorem, we show that the same conclusion is reached when the solution approaches the maximum value from below. Here we only rely on the boundedness of the solution.
Theorem 4.4.
Let be a bounded solution of (4.1). Then:
-
(i)
If uniformly on , then .
-
(ii)
is smooth on any open set where .
Proof.
To show part , let uniformly on be a bounded solution to equation (4.1). We know from Section 2 that maps into for any . In particular, maps into . Moreover, if , the Nemytskii operator
maps into itself for and ; see [4, Theorem 2.87]. Since , equation (4.1) gives
and hence, we obtain the mapping
(4.2) |
for all . Finally, equation (4.1) yields
by the assumption . We also note that the mapping is real analytic for , and hence an iterative argument in shows that . To show that , we note that if is a periodic solution of (4.1), its translation for any is also periodic. However, the previous analysis implies that for any , so we conclude that .
To show part , let , and assume that for some open set and any . Let be a smooth function compactly supported in , and be a cut-off function with in a neighborhood of . Then
It is easy to see that and vanishes by construction. Thus, . Using the same iteration argument in as in the proof of part shows that is smooth on . ∎
As a motivation, we shall see that when the order of the dispersive term is strictly smaller than , any decrease of the order does not affect the regularity of the wave solution.
Theorem 4.5.
Let be an even solution of (4.1), which is nonincreasing on , and attains its maximum at . Then cannot belong to the class .
Proof.
By contradiction, suppose that is in . Since is integrable by Theorem 3.1, equation (4.1) gives
Thus, is twice continuously differentiable. Then fixing and applying the Taylor theorem, we have
(4.3) |
for some between and . Differentiating equation (4.1) twice gives
which substitutes into the expression (4.3) yields
On the other hand, since is even, nonincreasing on by Theorem 3.1, and is nonincreasing on , we obtain
for some constant . Both and are continuous, so there exists a constant such that
for any between and . Using this bound and equation (4.3), we have the estimate
which is equivalent to
for all . Taking the limit when leads to contradiction to the fact . ∎
In the next theorem, we investigate the regularity of a solution when it touches from below. It was shown that the solution has -Hölder regularity at for an inhomogeneous multiplier of order greater than [12], and Lipschitz continuous at for a homogeneous multiplier of order less than [5]. For our case with inhomogeneous symbol of order less than , we, indeed, obtain a Lipschitz regularity at .
Theorem 4.6.
Let be a nontrivial, -periodic, even solution of (4.1), which is nonincreasing on . If attains its maximum at , then the following holds:
-
(i)
and is strictly decreasing on .
-
(ii)
, that is is Lipschitz continuous.
-
(iii)
is precisely Lipschitz continuous at , that is, there exist constants such that
for .
Proof.
Assume that is an even solution of equation (4.1), which is nonincreasing on and attains its maximum at .
-
Let and . Then we rewrite equation (4.1) as
We want to show that the right-hand side is strictly negative. In fact, for any and , since and are even and periodic, we obtain
By similar arguments as in the proof of Lemma 3.2, we know that for , We further have for and by the assumption that is even and nonincreasing on . Therefore, the integrand is nonpositive. Since is a nontrivial solution and is not a constant, we conclude that
(4.4) for any . Hence, if and only if . Then inequality (4.4) implies
Thus, is strictly decreasing on . In view of Therem 4.4, is smooth on .
-
To prove the Lipschitz regularity at the maximum point, we make use of a bootstrap argument. By contradiction, assume that the solution is not Lipschitz continuous. Suppose that is only a bounded function. Then recalling that maps into for and the expression
we have the estimate
Thus, it is straightforward to see that is -Hölder continuous.
Next, evaluating equation (4.1) at gives
and then subtracting from equation (4.1), we obtain
for . Since is smooth on by Theorem 4.4, differentiating the above equality yields
where we are using the fact that and . If is -Hölder continuous, then . Since , we gain at least some Hölder regularity for . Thus, the right-hand side of expression () can be estimated by a constant multiple of for some , and hence,
(4.5) By assumption that is not Lipschitz continuous at , the above estimate guarantees that is at least -Hölder continuous, where . We aim to bootstrap this argument to obtain Lipschitz regularity of at . If , we use that , which guarantees that its derivative is at least Lipschitz continuous ( in expression (4.5)), and hence, is Lipschitz continuous. On the other hand, if , then is -Hölder continuous for some . We then repeat the argument finitely many times to yield that is indeed Lipschitz continuous at , that is
(4.6) -
From the upper bound (4.6), it remains to show that
(4.7) for . Taking the derivative of equation (4.1) and evaluating at any we have
On the other hand, evaluating the upper bound established in (4.6) at gives . Therefore, dividing the above equation by to obtain the inequality
(4.8) Our aim is to show that is strictly bounded away from . We compute
for any . Then since is integrable, taking the limit on both sides of the inequality (4.8) yields
(4.9) for some constant , since and are strictly decreasing on . Then for any , applying the Mean Value Theorem, we have
for some . Combing with the expression (4.9) gives our desired estimate (4.7), and hence, we conclude that is exactly Lipschitz continuous at . ∎
From Theorem 4.4 and 4.6, we conclude that any even, periodic solution of equation (4.1), which is monotone on half the period, is Lipschitz continuous. Therefore, if such a solution occurs, it must be the highest, peaked wave. Before closing this section, we provide a lemma to prove that one of the alternatives in the global bifurcation arguments in Section 5 does not occur.
Lemma 4.7.
Let be an even solution of equation (4.1), which is nonincreasing on . Then there exists a constant , depending only on the kernel , such that
Proof.
Choose any . Then by the evenness of and the fact that is nonincreasing on , we have the estimate
(4.10) |
since the integrand is nonpositive. Notice that for , and there exists a constant , depending only on the kernel , such that
Thus, integrating the inequality (4.10) with respect to over yields
From Theorem 4.6, we know that is strictly decreasing on , and hence we can divide the above inequality by the quantity to obtain the claim. ∎
5. Global bifurcation and conclusion of the main theorem
In the last section, we proved the existence of nontrivial, highest, even, -periodic solutions of equation (1.2) using an analytic bifurcation technique. In fact, both local and global bifurcation studies on nonlocal equations have been investigated intensively. For instance, the existence of smooth, small-amplitude, periodic traveling-wave solutions to Whitham equation was established by Ehrnström and Kalisch in [10] using Crandall–Rabinowitz local bifurcation theorem. The authors also investigated numerically a global branch of solutions approaching a highest, cusped, traveling-wave solution. An analytic proof for the latter fact was provided in [8] using a variational approach. Recently, Truong, Wahlén, and Wheeler [20] attacked a similar problem using the center manifold theorem for the Whitham equation.
In this paper, we make use of the same arguments in [11] which gave a general functional-analytic framework for bifurcation theory to Whitham equation. Our aim is to extend the local bifurcation branch found by the analytic version of Crandall–Rabinowitz theorem to a global one, and then characterize the end of this bifurcation curve. We will show that the global bifurcation curve reaches a limiting highest wave , which is even, strictly decreasing on and attains its maximum . By Theorem 4.6, the highest wave is a peaked traveling-wave solution of
We use the subscript for the restriction of a Banach space to its subset of even functions. Let and set
where
(5.1) |
Then if and only if is an even -solution of (1.2) corresponding to the wave speed . We have the first local bifurcation result as follow.
Theorem 5.1 (Local bifurcation).
For each integer , the point , where is a bifurcation point. More precisely, there exits and an analytic curve through ,
of nontrivial, -periodic, even solutions of (5.1) with and
In a neighborhood of the bifurcation point these are all the nontrivial solutions of in .
Proof.
We will prove the result using the analytic version of the Crandall–Rabinowitz Theorem [7, Theorem 8.4.1]. It is clear that for any . We are looking for -periodic, even, nontrivial solutions bifurcating from the line of trivial solutions. The wave speed shall be the bifurcation parameter. The linearization of around the trivial solution is given by
Then from Section 2, we know that is parity preserving and a smoothing operator, which implies that it is compact on . Hence, is a compact perturbation of an isomorphism, and therefore constitutes a Fredholm operator of index . The nontrivial kernel of is spanned by functions satisfying
for all . For , we see that . Therefore, the kernel of is one-dimensional if and only if for some , in which case it is given by
We also note that are all simple eigenvalues of . Finally, we observe that
is not in the range of , which means the tranversality condition is satisfied. Thus, the assumptions of the Crandall–Rabinowitz theorem are fulfilled. ∎
Next, our aim is to extend the local bifurcation branch found in Theorem 5.1 to a global continuum of solutions of . Recalling , set
where the admissible set is given by
Then all bounded solutions of equation (4.1) satisfy for all . We start the analysis by providing the -bound for a solution.
Lemma 5.2 ( bound).
Let . Then any bounded solution to equation (1.2) satisfies:
Proof.
From equation (4.1), we have the estimate
where we have used the fact that the kernel is integrable. Therefore, , which holds trivially, or the desired estimate follows by dividing by . ∎
Before extending the local branches globally, we need two helping lemmas.
Lemma 5.3.
The Frechét derivative is a Fredholm operator of index for all .
Proof.
We have
for any given . Since is an isomorphism on and is compact on , the operator is Fredholm. From the proof of Theorem 5.1, we know that has Fredholm index 0, and so is due to the fact that the index is continuous. ∎
Lemma 5.4.
For any , the function is smooth, and any bounded and closed subset of is compact in .
Proof.
If , then , and we write equation (1.2) in the form
The function is a bounded and linear mapping from into . Moreover, by Theorem 4.4, we know that is smooth.
Let be a bounded and closed set. Then is relatively compact in . Since is closed, any sequence has a convergent subsequence in by Arzela–Ascoli’s lemma. We conclude that is compact in . ∎
According to [7, Theorem 9.1.1], Lemma 5.3 and 5.4 allow us to extend the local branches found in Theorem 5.1 to global curves once we establish that any of the derivatives is not identically zero for . However, the latter claim is an immediate consequence of Theorem 5.7 below. This theorem is an adaptation of [11, Theorem 4.4] with and as above.
Theorem 5.5 (Global bifurcation).
The local bifurcation curve from Theorem 5.1 of solutions of equation (5.1) extends to a global continuous curve of solutions , that allows a local real-analytic reparameterization around each . One of the following alternatives holds:
-
(i)
as .
-
(ii)
There exists a subsequence such that the pair approaches the boundary of as .
-
(iii)
The function is (finitely) periodic.
We apply the Lyapunov–Schmidt reduction, in order to establish the bifurcation formulas. Let be a fixed number and set
Then and a continuous projection onto the one-dimensional space is given by
where denotes the inner product in . Let us recall the Lyapunov–Schmidt reduction theorem from [18, Theorem I.2.3].
Theorem 5.6 (Lyapunov–Schmidt reduction).
The next theorem gives bifurcation formulas for the curve, which also justifies our arguments to extend the local curves globally.
Theorem 5.7 (Bifurcation formulas).
Proof.
For along the analytic local bifurcation curve in Theorem 5.1, we fix and suppress the subscript to lighten our notation. We first show that
after a suitable choice of parameterization. Since is even, it has a Fourier cosine representation
where
for Here we use a slightly different convention for the Fourier series compared to Section 2. Also, we choose the parameter in the local bifurcation curve so that . We observe that if is an even and -periodic function, then is also an even and -periodic function, and its Fourier coefficient satisfies
Since , uniqueness of the bifurcation curve yields
which means . Therefore, by the analyticity of , we can write
where the sum is uniformly convergent in a neighborhood of . Similarly, has an expansion
with convergence in . Substituting these two formulae into the main equation (1.2) and equating coefficients of equal order in , we obtain
(5.6) | ||||
(5.7) | ||||
(5.8) |
By Theorem 5.1, and , which implies expression (5.6) is satisfied. Under the assumption that the right-hand sides of above equations lie in the range of the linear operators from the left-hand sides, the coefficients and can be determined by solving the corresponding equation.
Recalling the formula for given by (2.4) and expressing as a Fourier series, equation (5.7) gives
and then comparing coefficients of cosine functions, we obtain the relation
which yields
Next, the right-hand side of equation (5.8) becomes
Recalling the above parameterization , which implies that for all , we find that
Thus, the bifurcation formulas (5.4) and (5.5) hold, and the local branch in Theorem 5.1 is supercritical pitchfork bifurcation. ∎
The remaining of the section is to devote showing that alternative in Theorem 5.5 is excluded, and both alternatives and occur simultaneously as along the bifurcation branch . This implies that the highest wave is reached as a limit of the global bifurcation curve.
Lemma 5.8.
Any sequence of solutions to equation (4.1) with bounded has a subsequence which converges uniformly to a solution .
Proof.
Recalling the estimate of in Lemma 5.2,
we see that if is bounded, then is bounded. Since is integrable and continuous on , Dominated Convergence Theorem allows us to conclude that is equicontinuous. Then Arzela–Ascoli’s lemma implies that there exists a subsequence which converges uniformly to a solution of equation (4.1). ∎
Proposition 5.9.
The solutions , on the global bifurcation curve belong to and alternative in Theorem 5.5 does not occur. In particular, the bifurcation curve has no intersection with the trivial solution line for any .
Proof.
Due to [7, Theorem 9.2.2] the statement holds true if the following conditions are satisfied
-
(a)
is a cone in a real Banach space.
-
(b)
provided is small enough.
-
(c)
If and , then for and .
-
(d)
Each nontrivial point on the bifurcation curve which also belongs to is an interior point of in .
Conditions (a), (b), (c) are satisfied because of the local bifurcation result in Theorem 5.1, so it remains to verify condition (d). Let be a nontrivial solution on the bifurcation curve found in Theorem 5.5. By Theorem 4.4, is smooth and together with Lemma 4.3, we have on and . Choose a solution lying within small enough neighborhood in such that and . In view of the mapping (4.2), an iteration process on the regularity index yields that , where can be made arbitrarily small by choosing small enough. It follows that for small enough, is a smooth, even, nonincreasing on solution, and hence belongs to the interior of in , which concludes the proof. ∎
Remark 5.10.
From the proof of Theorem 5.1 and Lemma 4.3, we have the bound . Moreover, integrating equation (4.1) on and using the fact that , we obtain
Thus, the only way to reach is by approaching , and hence when is small, we have . By Proposition 5.9, the bifurcation curve does not intersect the trivial solution line for any , so for all .
In the next lemma, we show that along the bifurcation curve, the wave speed is, in fact, bounded away from .
Lemma 5.11.
Proof.
By contradiction, assume that there exists a sequence with such that as , while as along the bifurcation curve found in Theorem 5.5. In view of Lemma 5.4, there exists a subsequence of such that converges to a solution of equation (5.1). Along the bifurcation curve, we have that as , and it follows that . In view of Lemma 4.1, we have , whence by Remark 4.2. Finally, Lemma 4.7 leads to
which is a contradiction. Thus, we have uniformly for all . ∎
At this point, we know that the wave speed is bounded away from and . Before proving the main result, we show that alternatives and in Theorem 5.5 occur simultaneously.
Theorem 5.12.
In Theorem 5.5, alternatives and both occur.
Proof.
Let , , be the bifurcation curve found in Theorem 5.5. By Proposition 5.9, we know that is nontrivial, even, and nonincreasing on , and alternative in Theorem 5.5 does not occur. Thus, it is either alternative or alternative in Theorem 5.5 occurs.
Suppose that alternative occurs. Then we have either for some or as . By Remark 5.10, is bounded between and , so the only possibility is is unbounded. In this case, alternative must occur, that is, the quantity
holds. Indeed, by contradiction, suppose that
for some . For any such solution , equation (1.2) gives
By Lemma 5.2, is uniformly bounded in . Then since maps to , there exists some constant depending on and such that
which is a contradiction. Therefore, alternative must occur.
On the other hand, suppose that alternative , but not alternative , occurs. Then fixing , there exists a sequence in Theorem 5.5 solving equation (1.2) satisfying on , , and
while remains uniformly bounded in for . Taking a limit along a subsequence in for some yields a contradiction to Theorem 4.6. Therefore, both alternatives and occur simultaneously. ∎
We have all the ingredients to prove the main result in this paper. The proof of Theorem 1.1 will show that the limiting wave at the end of the bifurcation curve is even, periodic, highest, and exactly Lipschitz at its crest.
Proof of Theorem 1.1.
Let be the global bifurcation curve found in Theorem 5.5 and let be a sequence in tending to infinity. By Lemma 5.11 and Remark 5.10, we know that is bounded between and away from and . Moreover, Lemma 5.8 gives the existence of a subsequence converging uniformly to a solution as . Finally, from Theorem 5.12 and Theorem 4.6, we conclude that with being precisely Lipschitz continuous at the maximum point. This finishes the proof of Theorem 1.1. ∎
Acknowledgements
H.L. would like to express his sincere gratitude to Mats Ehrnström for his continuous support and insightful comments.
References
- [1] C. J. Amick, L. E. Fraenkel, and J. F. Toland. On the Stokes conjecture for the wave of extreme form. Acta Math., 148:193–214, 1982.
- [2] W. Arendt and S. Bu. Operator-valued Fourier multipliers on periodic Besov spaces and applications. Proc. Edinb. Math. Soc. (2), 47(1):15–33, 2004.
- [3] M. N. Arnesen. A non-local approach to waves of maximal height for the Degasperis-Procesi equation. J. Math. Anal. Appl., 479(1):25–44, 2019.
- [4] H. Bahouri, J. Chemin, and R. Danchin. Fourier analysis and nonlinear partial differential equations, volume 343 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer, Heidelberg, 2011.
- [5] G. Bruell and R. N. Dhara. Waves of maximal height for a class of nonlocal equations with homogeneous symbols. 2018.
- [6] G. Bruell, M. Ehrnstöm, M. A. Johnson, and E. Wahlén. Waves of greatest height for nonlocal dispersive equations. Work in progress.
- [7] B. Buffoni and J. F. Toland. Analytic theory of global bifurcation. Princeton Series in Applied Mathematics. Princeton University Press, Princeton, NJ, 2003. An introduction.
- [8] M. Ehrnström, M. D. Groves, and E. Wahlén. On the existence and stability of solitary-wave solutions to a class of evolution equations of whitham type. Nonlinearity, 25(10):2903–2936, sep 2012.
- [9] M. Ehrnström, M. A. Johnson, and K. M. Claassen. Existence of a highest wave in a fully dispersive two-way shallow water model. Arch. Ration. Mech. Anal., 231(3):1635–1673, 2019.
- [10] M. Ehrnström and H. Kalisch. Traveling waves for the Whitham equation. Differential Integral Equations, 22(11-12):1193–1210, 2009.
- [11] M. Ehrnström and H. Kalisch. Global bifurcation for the Whitham equation. Math. Model. Nat. Phenom., 8(5):13–30, 2013.
- [12] M. Ehrnström and E. Wahlén. On Whitham’s conjecture of a highest cusped wave for a nonlocal dispersive equation. Ann. Inst. H. Poincaré Anal. Non Linéaire, 36(6):1603–1637, 2019.
- [13] M. Ehrnström and Y. Wang. Enhanced existence time of solutions to evolution equations of whitham type. arXiv preprint arXiv:2008.12722, 2020.
- [14] A. Enciso, J. Gómez-Serrano, and B. Vergara. Convexity of whitham’s highest cusped wave. arXiv preprint arXiv:1810.10935, 2018.
- [15] A. Geyer and D. E. Pelinovsky. Spectral instability of the peaked periodic wave in the reduced ostrovsky equation. 03 2019.
- [16] L. Grafakos. Modern Fourier analysis, volume 250 of Graduate Texts in Mathematics. Springer, New York, third edition, 2014.
- [17] O. G. Jørsboe and L. Mejlbro. The Carleson-Hunt theorem on Fourier series, volume 911 of Lecture Notes in Mathematics. Springer-Verlag, Berlin-New York, 1982.
- [18] H. Kielhöfer. Bifurcation theory, volume 156 of Applied Mathematical Sciences. Springer, New York, second edition, 2012. An introduction with applications to partial differential equations.
- [19] G. Stokes. On the theory of oscillatory waves. Mathematical and Physical Papers vol.1, pages 197–229, 1880.
- [20] T. Truong, E. Wahlén, and M. H. Wheeler. Global bifurcation of solitary waves for the whitham equation. arXiv preprint arXiv:2009.04713, 2020.
- [21] G. B. Whitham. Linear and nonlinear waves. Wiley-Interscience [John Wiley & Sons], New York-London-Sydney, 1974. Pure and Applied Mathematics.