This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Waves of maximal height for a class of nonlocal equations with inhomogeneous symbols

Hung Le Department of Mathematical Sciences, Norwegian University of Science and Technology, 7491 Trondheim, Norway [email protected]
Abstract.

In this paper, we consider a class of nonlocal equations where the convolution kernel is given by a Bessel potential symbol of order α\alpha for α>1\alpha>1. Based on the properties of the convolution operator, we apply a global bifurcation technique to show the existence of a highest, even, 2π2\pi-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 waves
2010 Mathematics Subject Classification:
76B15, 76B03, 35S30, 35A20
HL was supported by the ERCIM ‘Alain Bensoussan’ Fellowship Programme.

1. Introduction

Our object of study is the following nonlocal equation

ut+Lux+uux=0,(t,x)×,u_{t}+Lu_{x}+uu_{x}=0,\qquad(t,x)\in\mathbb{R}\times\mathbb{R}, (1.1)

where LL denotes the Fourier multiplier operator with symbol m(ξ)=(1+ξ2)α/2m(\xi)=(1+\xi^{2})^{-\alpha/2}, α>1\alpha>1. Equation (1.1) is known as the fractional Korteweg–de Vries equation. We are looking for a 2π2\pi-periodic traveling-wave solution u(t,x)=ϕ(xμt)u(t,x)=\phi(x-\mu t), where μ>0\mu>0 denotes the speed of the right-propagating wave. In this moving frame, equation (1.1) takes the form

μϕ+Lϕ+12ϕ2=0,-\mu\phi+L\phi+\frac{1}{2}\phi^{2}=0, (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:

ϕϕ+γ,μμ+γ,BB+γ(1μ12γ)\phi\mapsto\phi+\gamma,\quad\mu\mapsto\mu+\gamma,\quad B\mapsto B+\gamma\left(1-\mu-\frac{1}{2}\gamma\right)

for any γ\gamma\in\mathbb{R}, which maps solutions of μϕ+Lϕ+12ϕ2=B-\mu\phi+L\phi+\frac{1}{2}\phi^{2}=B to solutions of a new equation of the same form. In this paper, the above multiplier m(ξ)m(\xi) has order of α-\alpha, which is less than 1-1. 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 1-1, 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 2π/32\pi/3. 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 ϕ\phi to the Whitham equation reached its critical height when ϕ=μ\phi=\mu (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 C1/2C^{1/2}-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 m(ξ)=tanh(ξ)/ξm(\xi)=\tanh(\xi)/\xi. They showed that the highest cusped wave has a singularity at its crest of the form |xlog(|x|)||x\log(|x|)|. 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 m(ξ)=(1+ξ2)1m(\xi)=(1+\xi^{2})^{-1}. 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 KK, we prove the existence and regularity properties of a highest, 2π2\pi-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

ut+2uux+Lux=0,u_{t}+2uu_{x}+Lu_{x}=0,

where LL is the Fourier multiplier operator given by the symbol

m(ξ)=tanhξξ.m(\xi)=\sqrt{\frac{\tanh\xi}{\xi}}.

This multiplier has order greater than 1-1, is inhomogeneous and completely monotone. The main result is the existence of a highest, cusped, PP-periodic traveling-wave solution, which is even, strictly decreasing, smooth on each half-period, and belongs to the Hölder space C12()C^{\frac{1}{2}}(\mathbb{R}). The proof is based on the regularity and monotonicity properties of the convolution kernel induced by mm. On the other hand, Bruell and Dhara [5] considered the fractional KdV equation as in equation (1.1), where LL is the Fourier multiplier operator given by the homogeneous symbol

m(ξ)=|ξ|α,α>1.m(\xi)=|\xi|^{-\alpha},\qquad\alpha>1.

This multiplier has order less than 1-1, is homogeneous, and completely monotone. The authors showed the existence of a highest, 2π2\pi-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 BB in equation (1.2) now becomes ϕ^(0)\widehat{\phi}(0).

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 mm for the fractional KdV equation (1.1) is Bessel potential function, which is inhomogeneous and of order strictly smaller than 1-1. 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 B=0B=0. Moreover, even though our kernel KK is even and smooth for all xx\in\mathbb{R} with rapid decay derivatives, we do not use a closed formula for KK in the physical space, as well as for its periodization as in [5, 12]. It is also worth to mention that the kernel KK 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 KK 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 μ\mu. We further show that the only possibility along the main bifurcation curve is that maxϕ\max\phi must approach μ\mu. Moreover, we investigate the solution ϕ\phi satisfying maxϕ=μ\max\phi=\mu. Using the properties of the convolution operator KK, we obtain Lipschitz continuous at its crest.

Let us formulate our main existence theorem which constructs solutions bifurcating from nontrivial, even, smooth, and 2π2\pi-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 k1k\geq 1, there exists a wave speed μk>0\mu^{*}_{k}>0 and a global bifurcation branch:

s(ϕk(s),μk(s)),s>0,s\mapsto(\phi_{k}(s),\mu_{k}(s)),\qquad s>0,

of nontrivial, 2πk\frac{2\pi}{k}-periodic, smooth, even solutions to the steady equation (1.2) for α>1\alpha>1, emerging from the bifurcation point (0,μk)(0,\mu_{k}^{*}). Moreover, given any unbounded sequence (sn)n(s_{n})_{n\in\mathbb{N}} of positive numbers sns_{n}, there exists a subsequence of (ϕk(sn))n(\phi_{k}(s_{n}))_{n\in\mathbb{N}}, which converges uniformly to a limiting traveling-wave solution (ϕ¯k,μ¯k)(\bar{\phi}_{k},\bar{\mu}_{k}) that solves (1.2) and satisfies

ϕ¯k(0)=μ¯k.\bar{\phi}_{k}(0)=\bar{\mu}_{k}.

The limiting wave is strictly increasing on (πk,0)\left(-\frac{\pi}{k},0\right) and exactly Lipschitz at x2πkx\in\frac{2\pi}{k}\mathbb{Z}.

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 KK corresponding to the symbol m(ξ)m(\xi) and show that the action of the operator KK on periodic functions on the real line is in fact similar to the periodized kernel KPK_{P} acts on these functions in one period. In Section 3, we discuss some interesting analytic features of the periodized kernel KPK_{P}. In fact, both KK and KPK_{P} 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 KPK_{P}, 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 μ\mu, 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 maxϕμ\max\phi\to\mu, and an analysis of solutions at their maximum maxϕ=μ\max\phi=\mu.

2. Functional-analytic setting and general conventions

Setting 𝕋:=[π,π]\mathbb{T}:=[-\pi,\pi], we denote by 𝒟(𝕋)\mathcal{D}(\mathbb{T}) the space of test functions on 𝕋\mathbb{T}, whose dual space, the space of distributions on 𝕋\mathbb{T}, is 𝒟(𝕋)\mathcal{D}^{\prime}(\mathbb{T}). If 𝒮()\mathcal{S}(\mathbb{Z}) is the space of rapidly decaying functions from \mathbb{Z} to \mathbb{C} and 𝒮()\mathcal{S}^{\prime}(\mathbb{Z}) denotes its dual space, let :𝒟(𝕋)𝒮()\mathcal{F}:\mathcal{D}^{\prime}(\mathbb{T})\to\mathcal{S}^{\prime}(\mathbb{Z}) be the Fourier transformation on the torus defined by the duality on 𝒟(𝕋)\mathcal{D}(\mathbb{T}) via

f(ξ)=f^(ξ):=𝕋f(x)exp(ixξ)dx,f𝒟(𝕋).\mathcal{F}f(\xi)=\widehat{f}(\xi):=\int_{\mathbb{T}}f(x)\exp(-ix\xi)\,\mathrm{d}x,\qquad f\in\mathcal{D}(\mathbb{T}).

To simplify our notations, throughout this paper, if ff and gg are elements in an ordered Banach space, we write fgf\lesssim g (fg)(f\gtrsim g) if there exists a constant c>0c>0 such that fcgf\leq cg (fcg)(f\geq cg). Moreover, the notation fgf\eqsim g is used whenever fgf\lesssim g and fgf\gtrsim g. We also denote +:=[0,]\mathbb{R}_{+}:=[0,\infty] and by 0\mathbb{N}_{0} the set of natural numbers including zero. The space (X;Y)\mathcal{L}(X;Y) denotes the set of all bounded linear operators from XX to YY.

Next, let us investigate how the convolution KK* acts on periodic functions. Suppose that fL()f\in L^{\infty}(\mathbb{R}) is periodic and even. Since KK is in L1()L^{1}(\mathbb{R}), we can write:

K(xy)\displaystyle\int_{-\infty}^{\infty}K(x-y) f(y)dy=n=ππK(xy+2nπ)f(y)dy\displaystyle f(y)\,\mathrm{d}y=\sum_{n=-\infty}^{\infty}\int_{-\pi}^{\pi}K(x-y+2n\pi)f(y)\,\mathrm{d}y
=ππ(n=K(xy+2nπ))f(y)dy=:ππKP(xy)f(y)dy.\displaystyle=\int_{-\pi}^{\pi}\left(\sum_{n=-\infty}^{\infty}K(x-y+2n\pi)\right)f(y)\,\mathrm{d}y=:\int_{-\pi}^{\pi}K_{P}(x-y)f(y)\,\mathrm{d}y.

It is clear from the definition of KP(x)K_{P}(x) that it is 2π2\pi-periodic, even, and continuous on 𝕋\mathbb{T}. Moreover, Minkowski’s inequality shows that KP(x)K_{P}(x) belongs to Lp(π,π)L^{p}(-\pi,\pi), for 1p<21\leq p<2. Therefore, Carleson–Hunt Theorem [17] implies that KP(x)K_{P}(x) can be approximated pointwise by its Fourier series:

KP(x)=12πnA^nexp(inx),a.e.,K_{P}(x)=\frac{1}{2\pi}\sum_{n\in\mathbb{Z}}\widehat{A}_{n}\exp(inx),\quad a.e.,

and the Fourier coefficients of KPK_{P} are given by

A^n=ππj=K(x+2jπ)exp(ixn)dx=j=ππK(x+2jπ)exp(i(x+2jπ)n)dx=K(x)exp(ixn)dx=K^(n).\displaystyle\begin{split}\widehat{A}_{n}&=\int_{-\pi}^{\pi}\sum_{j=-\infty}^{\infty}K(x+2j\pi)\exp(-ixn)\,\mathrm{d}x\\ &=\sum_{j=-\infty}^{\infty}\int_{-\pi}^{\pi}K(x+2j\pi)\exp(-i(x+2j\pi)n)\,\mathrm{d}x\\ &=\int_{-\infty}^{\infty}K(x)\exp(-ixn)\,\mathrm{d}x=\widehat{K}(n).\end{split} (2.1)

Thus, the periodic problem is given by the same multiplier as the problem on the line, so we have the representation:

Kf(x)=12πnf^(n)K^(n)exp(inx)=12πnf^(n)m(n)exp(inx).K*f(x)=\frac{1}{2\pi}\sum_{n\in\mathbb{Z}}\widehat{f}(n)\widehat{K}(n)\exp(inx)=\frac{1}{2\pi}\sum_{n\in\mathbb{Z}}\widehat{f}(n)\,m(n)\exp(inx).

The periodized operator KPK_{P} is introduced to aid the analysis of periodic solutions satisfying certain sign conditions in a half-period. From the above analysis, KPK_{P} can be expressed as the Fourier series

KP(x)=12πnm(n)exp(inx).K_{P}(x)=\frac{1}{2\pi}\sum_{n\in\mathbb{Z}}m(n)\exp(inx). (2.2)

and the relationship between KK and KPK_{P} is

KP(x)=nK(x+2nπ).K_{P}(x)=\sum_{n\in\mathbb{Z}}K(x+2n\pi). (2.3)

Note that this sum is absolutely convergent since KK has rapid decay.

The operator LL

Now let LL be the operator

L:fKf,L:f\mapsto K*f,

defined via duality on the space 𝒮()\mathcal{S}^{\prime}(\mathbb{R}) of tempered distributions. Then one can see that for a continuous periodic function ff, the operator LL is given by

ππKP(xy)f(y)dy,\int_{-\pi}^{\pi}K_{P}(x-y)f(y)\,\mathrm{d}y,

and more generally by

K(xy)f(y)dy\int_{\mathbb{R}}K(x-y)f(y)\,\mathrm{d}y

if ff is bounded and continuous. Thus, we have the following formula for LL

Lf(x)=12πnm(n)f^(n)exp(inx).Lf(x)=\frac{1}{2\pi}\sum_{n\in\mathbb{Z}}m(n)\widehat{f}(n)\exp(inx). (2.4)

Next, we shall say that a function φ:\varphi:\mathbb{R}\to\mathbb{R} is Hölder continuous of regularity β(0,1)\beta\in(0,1) at a point xx\in\mathbb{R} if

|φ|Cxβ:=suph0|φ(x+h)φ(x)|h|β<,|\varphi|_{C^{\beta}_{x}}:=\sup_{h\neq 0}\frac{|\varphi(x+h)-\varphi(x)}{|h|^{\beta}}<\infty,

and let

Cβ()\displaystyle C^{\beta}(\mathbb{R}) ={φC():supx|φ|Cxβ<}\displaystyle=\{\varphi\in C(\mathbb{R}):\sup_{x}|\varphi|_{C^{\beta}_{x}}<\infty\}

be the space of β\beta-Hölder continuous functions on \mathbb{R}. If kk\in\mathbb{N}, then

Ck,β()={φCk():φ(k)Cβ()}C^{k,\beta}(\mathbb{R})=\{\varphi\in C^{k}(\mathbb{R}):\varphi^{(k)}\in C^{\beta}(\mathbb{R})\}

denotes the space of kk-times continuously differentiable functions whose kk-th derivative is β\beta-Hölder continuous on \mathbb{R}.

Before we proceed to discuss the space of the operator LL, let us recall the definition of Besov spaces Bp,qs()B^{s}_{p,q}(\mathbb{R}) using the Littlewood–Paley decomposition. Let {γj}j\{\gamma_{j}\}_{j\in\mathbb{N}} be a family of smooth and compactly supported functions, where

suppγ0[2,2],suppγj{ξ:2j1|ξ|2j+1}forj1,\displaystyle\operatorname{supp}\gamma_{0}\subset[-2,2],\qquad\operatorname{supp}\gamma_{j}\subset\{\xi\in\mathbb{R}:2^{j-1}\leq|\xi|\leq 2^{j+1}\}\qquad\text{for}\quad j\geq 1,
j=0γj(ξ)=1\displaystyle\sum_{j=0}^{\infty}\gamma_{j}(\xi)=1

for all ξ\xi\in\mathbb{R}. For a tempered distribution f𝒮()f\in\mathcal{S}^{\prime}(\mathbb{R}), we let γj(D)f=1(γj(ξ)f^(ξ))\gamma_{j}(D)f=\mathcal{F}^{-1}(\gamma_{j}(\xi)\widehat{f}(\xi)), so that

f=j=0γj(D)f.f=\sum_{j=0}^{\infty}\gamma_{j}(D)f.

Then the Besov spaces Bp,qs()B^{s}_{p,q}(\mathbb{R}), ss\in\mathbb{R}, 1p1\leq p\leq\infty, 1q<1\leq q<\infty are defined by

{f𝒮():fBp,qs():=[j=0(2sjγj(D)fLp())q]1/2<},\left\{f\in\mathcal{S}^{\prime}(\mathbb{R}):\|f\|_{B^{s}_{p,q}(\mathbb{R})}:=\left[\sum_{j=0}^{\infty}(2^{sj}\|\gamma_{j}(D)f\|_{L^{p}(\mathbb{R})})^{q}\right]^{1/2}<\infty\right\},

and for 1p1\leq p\leq\infty and q=q=\infty, we instead define

Bp,qs()={f𝒮():fBp,s():=supj02sjγj(D)fLp()<}.B^{s}_{p,q}(\mathbb{R})=\left\{f\in\mathcal{S}^{\prime}(\mathbb{R}):\|f\|_{B^{s}_{p,\infty}(\mathbb{R})}:=\sup_{j\geq 0}2^{sj}\|\gamma_{j}(D)f\|_{L^{p}(\mathbb{R})}<\infty\right\}.

For a 2π2\pi-periodic tempered distribution f=12πkf^kexp(ik)f=\frac{1}{2\pi}\sum_{k\in\mathbb{Z}}\widehat{f}_{k}\exp(ik\cdot), we have the identity

γj(D)f=1πkγj(k)f^kexp(ikx),\gamma_{j}(D)f=\frac{1}{\pi}\sum_{k\in\mathbb{Z}}\gamma_{j}(k)\widehat{f}_{k}\exp(ikx),

so that γj(D)f\gamma_{j}(D)f is a trigonometric polynomial.

Moreover, for ss\in\mathbb{R}, we define the Zygmund spaces 𝒞s\mathcal{C}^{s} of order ss by

𝒞s(𝕋)=B,s(𝕋),\mathcal{C}^{s}(\mathbb{T})=B^{s}_{\infty,\infty}(\mathbb{T}),

and recall that when ss is a positive noninteger, we have 𝒞s=Cs,ss\mathcal{C}^{s}=C^{\lfloor s\rfloor,s-\lfloor s\rfloor}; and when ss is a nonnegative integer, we have Ws,𝒞sW^{s,\infty}\subsetneq\mathcal{C}^{s}. As a consequence of Littlewood–Paley theory, we have the relation 𝒞s(𝕋)=Cs(𝕋)\mathcal{C}^{s}(\mathbb{T})=C^{s}(\mathbb{T}) for any positive noninteger ss. Thus, the Hölder spaces on the torus are completely characterized by Fourier series. It ss\in\mathbb{N}, then Cs(𝕋)C^{s}(\mathbb{T}) is a proper subset of 𝒞s(𝕋)\mathcal{C}^{s}(\mathbb{T}) and

C1(𝕋)C1(𝕋)𝒞1(𝕋),C^{1}(\mathbb{T})\subsetneq C^{1-}(\mathbb{T})\subsetneq\mathcal{C}^{1}(\mathbb{T}),

where C1(𝕋)C^{1-}(\mathbb{T}) denotes the space of Lipschitz continuous functions on 𝕋\mathbb{T}.

To finish this section, we discuss the space of the operator LL. It follows from the well-known Bessel potential estimate

|Dξnm(ξ)|(1+|ξ|)αn,n0,|D_{\xi}^{n}m(\xi)|\lesssim(1+|\xi|)^{-\alpha-n},\quad n\geq 0,

that LL defines a bounded operator

L:Bp,qs(𝕋)Bp,qs+α(𝕋),L:B^{s}_{p,q}(\mathbb{T})\to B^{s+\alpha}_{p,q}(\mathbb{T}),

see, for example, [2, 4]. In particular, the operators

L:𝒞s(𝕋)𝒞s+α(𝕋)L:\mathcal{C}^{s}(\mathbb{T})\to\mathcal{C}^{s+\alpha}(\mathbb{T})

is bounded on 𝕋\mathbb{T}. Thus, LL is a smoothing operator of order α<1-\alpha<-1.

3. Properties of the periodized kernel KPK_{P}

This section is devoted to summarize some properties of the kernel KPK_{P}. We recall the Fourier representation of KPK_{P} (2.2) and the relationship (2.3) between KK and KPK_{P}. Recalling that K=1(m)K=\mathcal{F}^{-1}(m) and the Fourier multiplier is given by the Bessel potential m(ξ)=(1+ξ2)α/2m(\xi)=(1+\xi^{2})^{-\alpha/2} for α>1\alpha>1, we record some properties of the periodized operator KPK_{P}.

Theorem 3.1.

Let α>1\alpha>1. The kernel KPK_{P} has the following properties:

  1. (a)

    KPK_{P} is even, positive, 2π2\pi-periodic, and continuous. KPK_{P} is smooth on 𝕋\{0}\mathbb{T}\backslash\{0\}.

  2. (b)

    KPK_{P} is in W1,1(𝕋)W^{1,1}(\mathbb{T}). In particular, KPK_{P}^{\prime} is integrable. Moreover, KPK_{P} is Hölder continuous with β(0,α1)\beta\in(0,\alpha-1) if α(1,2)\alpha\in(1,2), Lipschitz continuous when α=2\alpha=2, and continuously differentiable if α>2\alpha>2.

  3. (c)

    KPK_{P} is decreasing on (0,π)(0,\pi).

Proof.

We first observe that this theorem holds when we replace KPK_{P} by KK; 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 KPK_{P} are inherited from KK. By [16, Proposition 1.2.5], we know that KK is smooth on 𝕋\{0}\mathbb{T}\backslash\{0\}, and so is KPK_{P} by the relation (2.3). Moreover, the regularity of KPK_{P} in part (b) follows from the regularity of KK 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 KPK_{P} is even and decreasing on (0,π)(0,\pi). The next lemma provides monotonicity property for the operator L=KL=K* based on KPK_{P}.

Lemma 3.2.

Let α>1\alpha>1. The operator LL is parity preserving on L(𝕋)L^{\infty}(\mathbb{T}). Moreover, if f,gL(𝕋)f,g\in L^{\infty}(\mathbb{T}) are odd functions satisfying f(x)g(x)f(x)\geq g(x) on [0,π][0,\pi], then either

Lf(x)>Lg(x)for all x(0,π),Lf(x)>Lg(x)\qquad\text{for all }\quad x\in(0,\pi),

or f=gf=g on 𝕋\mathbb{T}.

Proof.

To see that LL is parity-preserving, let fL(𝕋)f\in L^{\infty}(\mathbb{T}) be an odd function. Then since KK is even, we have

Lf(x)+Lf(x)\displaystyle Lf(x)+Lf(-x) =ππKP(xy)f(y)dy+ππKP(xy)f(y)dy\displaystyle=\int_{-\pi}^{\pi}K_{P}(x-y)f(y)\,\mathrm{d}y+\int_{-\pi}^{\pi}K_{P}(-x-y)f(y)\,\mathrm{d}y
=ππKP(xy)(f(y)+f(y))dy=0,\displaystyle=\int_{-\pi}^{\pi}K_{P}(x-y)(f(y)+f(-y))\,\mathrm{d}y=0,

which shows that LfLf is odd. Similarly, when ff is even, LfLf is also even.

To show the second part of the lemma, let f,gL(𝕋)f,g\in L^{\infty}(\mathbb{T}) be odd functions such that f(x)g(x)f(x)\geq g(x) on [0,π][0,\pi]. By contradiction, assume that there exists x0(0,π)x_{0}\in(0,\pi) such that Lf(x0)=Lg(x0)Lf(x_{0})=Lg(x_{0}). Then we compute

Lf(x0)Lg(x0)=ππKP(x0y)(f(y)g(y))dy=0π(KP(x0y)KP(x0+y))(f(y)g(y))dy.\begin{split}Lf(x_{0})-Lg(x_{0})&=\int_{-\pi}^{\pi}K_{P}(x_{0}-y)(f(y)-g(y))\,\mathrm{d}y\\ &=\int_{0}^{\pi}(K_{P}(x_{0}-y)-K_{P}(x_{0}+y))(f(y)-g(y))\,\mathrm{d}y.\end{split}

Since 0<x0,y<π0<x_{0},y<\pi, we have

π<x0y<x0+y<2π,π<x0y<π,0<x0+y<2π.-\pi<x_{0}-y<x_{0}+y<2\pi,\quad-\pi<x_{0}-y<\pi,\quad 0<x_{0}+y<2\pi.

Then the evenness, periodicity, and monotonicity of KPK_{P} yield

KP(x0y)KP(x0+y)>0K_{P}(x_{0}-y)-K_{P}(x_{0}+y)>0

for all x0,yx_{0},y in the interval (0,π)(0,\pi), and therefore, Lf(x0)>Lg(x0)Lf(x_{0})>Lg(x_{0}), which is a contradiction to our assumption Lf(x0)=Lg(x0)Lf(x_{0})=Lg(x_{0}), unless f=gf=g on 𝕋\mathbb{T}. ∎

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 ϕ\phi is a solution, ϕ\phi must be real-valued, bounded, and satisfy equation (1.2) pointwise. Fixing α>1\alpha>1, to aid our analysis, we re-write equation (1.2) as

12(μϕ)2=12μ2Lϕ.\frac{1}{2}(\mu-\phi)^{2}=\frac{1}{2}\mu^{2}-L\phi. (4.1)

Our first result is a rough bound for a solution according to the wave speed μ\mu.

Lemma 4.1.

Let ϕC(𝕋)\phi\in C(\mathbb{T}) be solution of (4.1). Then if μ>1\mu>1, then

0ϕ(xm)2(μ1)ϕ(xM)0\leq\phi(x_{m})\leq 2(\mu-1)\leq\phi(x_{M})

and if μ1\mu\leq 1, then

2(μ1)ϕ(xm)0ϕ(xM),2(\mu-1)\leq\phi(x_{m})\leq 0\leq\phi(x_{M}),

where ϕ(xm):=minx𝕋ϕ(x)\phi(x_{m}):=\min_{x\in\mathbb{T}}\phi(x) and ϕ(xM):=maxx𝕋ϕ(x)\phi(x_{M}):=\max_{x\in\mathbb{T}}\phi(x).

Proof.

By Lemma 3.2, LL is a strictly monotone operator. Since furthermore Lc=cLc=c for constants cc, we therefore obtain that

12(μϕ)2=12μ2Lϕ12μ2ϕ(xm).\frac{1}{2}\left(\mu-\phi\right)^{2}=\frac{1}{2}\mu^{2}-L\phi\leq\frac{1}{2}\mu^{2}-\phi(x_{m}).

In particular, 12(μϕ(xm))212μ2ϕ(xm)\frac{1}{2}(\mu-\phi(x_{m}))^{2}\leq\frac{1}{2}\mu^{2}-\phi(x_{m}), or equivalently, we have the inequality

ϕ(xm)(12ϕ(xm)(μ1))0.\phi(x_{m})\left(\frac{1}{2}\phi(x_{m})-(\mu-1)\right)\leq 0.

Similar arguments can be made to obtain

12(μϕ)2=12μ2Lϕ12μ2ϕ(xM),\frac{1}{2}\left(\mu-\phi\right)^{2}=\frac{1}{2}\mu^{2}-L\phi\geq\frac{1}{2}\mu^{2}-\phi(x_{M}),

which yields that

ϕ(xM)(12ϕ(xM)(μ1))0.\phi(x_{M})\left(\frac{1}{2}\phi(x_{M})-(\mu-1)\right)\geq 0.

Combining both cases, we have the desired estimates. ∎

Remark 4.2.

If a solution satisfies ϕ(x)=0\phi(x)=0 for some xx, then at those xx, the equation (4.1) reduces to Lϕ=0L\phi=0. Since LL is a strictly monotone operator by Lemma 3.2, we must have either ϕ0\phi\equiv 0, or ϕ\phi is sign-changing.

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 C1C^{1} solution, which is nonincreasing on (0,π)(0,\pi), must be strictly decreasing and less than μ\mu on (0,π)(0,\pi). Thus, we conclude that maxϕ=ϕ(0)=μ\max\phi=\phi(0)=\mu.

Lemma 4.3.

Any nontrivial, 2π2\pi-periodic, even solution ϕC1(𝕋)\phi\in C^{1}(\mathbb{T}) of (1.2) which is nonincreasing on (0,π)(0,\pi) satisfies

ϕ(x)<0andϕ(x)<μon (0,π).\phi^{\prime}(x)<0\qquad\text{and}\qquad\phi(x)<\mu\qquad\text{on }(0,\pi).

For such a solution, one necessarily has μ>0\mu>0. Moreover, if ϕC2(𝕋)\phi\in C^{2}(\mathbb{T}), then ϕ′′(0)<0\phi^{\prime\prime}(0)<0.

Proof.

Since ϕC1(𝕋)\phi\in C^{1}(\mathbb{T}), we can take the derivative of equation (4.1) to obain

(μϕ)ϕ(x)=Lϕ(x).(\mu-\phi)\phi^{\prime}(x)=L\phi^{\prime}(x).

Since ϕ\phi is nonincreasing on (0,π)(0,\pi), we have ϕ(x)0\phi^{\prime}(x)\leq 0 on (0,π)(0,\pi). We want to show that Lϕ(x)<0L\phi^{\prime}(x)<0 for x(0,π)x\in(0,\pi). Since ϕ\phi^{\prime} is odd, nontrivial, and nonpositive on (0,π)(0,\pi), Lemma 3.2 implies Lϕ(x)<0L\phi^{\prime}(x)<0 on (0,π)(0,\pi). Thus, we have

(μϕ)ϕ(x)<0,(\mu-\phi)\phi^{\prime}(x)<0,

and hence we have shown that ϕ(x)<0\phi^{\prime}(x)<0 and ϕ(x)<μ\phi(x)<\mu on (0,π)(0,\pi). On the other hand, by Lemma 4.1, we know 0ϕ(0)=ϕ(xM)0\leq\phi(0)=\phi(x_{M}), so that μ>0\mu>0.

To prove the second part, assume that ϕC2(𝕋)\phi\in C^{2}(\mathbb{T}). Differentiating twice equation (4.1) gives

(μϕ)ϕ′′=Lϕ′′+(ϕ)2.(\mu-\phi)\phi^{\prime\prime}=L\phi^{\prime\prime}+(\phi^{\prime})^{2}.

Then evaluating this equality by x=0x=0 and using the evenness of KPK_{P} and ϕ′′\phi^{\prime\prime}, we compute

(μϕ)ϕ′′(0)\displaystyle(\mu-\phi)\phi^{\prime\prime}(0) =ππKP(y)ϕ′′(y)dy\displaystyle=\int_{-\pi}^{\pi}K_{P}(y)\phi^{\prime\prime}(y)\,\mathrm{d}y
=20πKP(y)ϕ′′(y)dy\displaystyle=2\int_{0}^{\pi}K_{P}(y)\phi^{\prime\prime}(y)\,\mathrm{d}y
=20ϵKP(y)ϕ′′(y)dy+2ϵπKP(y)ϕ′′(y)dy\displaystyle=2\int_{0}^{\epsilon}K_{P}(y)\phi^{\prime\prime}(y)\,\mathrm{d}y+2\int_{\epsilon}^{\pi}K_{P}(y)\phi^{\prime\prime}(y)\,\mathrm{d}y
=20ϵKP(y)ϕ′′(y)dy+2[KP(y)ϕ(y)]|y=ϵy=π2ϵπKP(y)ϕ(y)dy.\displaystyle=2\int_{0}^{\epsilon}K_{P}(y)\phi^{\prime\prime}(y)\,\mathrm{d}y+2[K_{P}(y)\phi^{\prime}(y)]|_{y=\epsilon}^{y=\pi}-2\int_{\epsilon}^{\pi}K_{P}^{\prime}(y)\phi^{\prime}(y)\,\mathrm{d}y.

Since ϕ\phi is in C2C^{2} and KK is integrable, the first integral vanishes as ϵ0\epsilon\to 0, so does the boundary term KP(ϵ)ϕ(ϵ)K_{P}(\epsilon)\phi^{\prime}(\epsilon). The term KP(π)ϕ(π)K_{P}(\pi)\phi^{\prime}(\pi) also vanishes since ϕ(π)=0\phi^{\prime}(\pi)=0. By Theorem 3.1 and what we just proved, both KPK_{P}^{\prime} and ϕ\phi^{\prime} are strictly negative on (0,π)(0,\pi), and hence we have

(μϕ)ϕ′′(0)=2limϵ0+ϵπKP(y)ϕ(y)dy<0,(\mu-\phi)\phi^{\prime\prime}(0)=-2\lim_{\epsilon\to 0^{+}}\int_{\epsilon}^{\pi}K_{P}^{\prime}(y)\phi^{\prime}(y)\,\mathrm{d}y<0,

which shows that ϕ′′(0)<0\phi^{\prime\prime}(0)<0. ∎

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 ϕ(0)=μ\phi(0)=\mu from below. Here we only rely on the boundedness of the solution.

Theorem 4.4.

Let ϕμ\phi\leq\mu be a bounded solution of (4.1). Then:

  1. (i)

    If ϕ<μ\phi<\mu uniformly on 𝕋\mathbb{T}, then ϕC()\phi\in C^{\infty}(\mathbb{R}).

  2. (ii)

    ϕ\phi is smooth on any open set where ϕ<μ\phi<\mu.

Proof.

To show part (i)(i), let ϕ<μ\phi<\mu uniformly on 𝕋\mathbb{T} be a bounded solution to equation (4.1). We know from Section 2 that LL maps 𝒞s(𝕋)\mathcal{C}^{s}(\mathbb{T}) into 𝒞s+α(𝕋)\mathcal{C}^{s+\alpha}(\mathbb{T}) for any ss\in\mathbb{R}. In particular, LL maps L(𝕋)𝒞0(𝕋)L^{\infty}(\mathbb{T})\subset\mathcal{C}^{0}(\mathbb{T}) into 𝒞α(𝕋)\mathcal{C}^{\alpha}(\mathbb{T}). Moreover, if s>0s>0, the Nemytskii operator

fμ12μ2ff\mapsto\mu-\sqrt{\frac{1}{2}\mu^{2}-f}

maps L(𝕋)B,s(𝕋)L^{\infty}(\mathbb{T})\cap B^{s}_{\infty,\infty}(\mathbb{T}) into itself for f<12μ2f<\frac{1}{2}\mu^{2} and s>0s>0; see [4, Theorem 2.87]. Since ϕ<μ\phi<\mu, equation (4.1) gives

Lϕ<12μ2,L\phi<\frac{1}{2}\mu^{2},

and hence, we obtain the mapping

[Lϕ12μ2Lϕ][ϕLϕ]:L(𝕋)B,s(𝕋)B,s+α(𝕋)\left[L\phi\mapsto\sqrt{\frac{1}{2}\mu^{2}-L\phi}\right]\circ\left[\phi\mapsto L\phi\right]:L^{\infty}(\mathbb{T})\cap B_{\infty,\infty}^{s}(\mathbb{T})\to B_{\infty,\infty}^{s+\alpha}(\mathbb{T}) (4.2)

for all s0s\geq 0. Finally, equation (4.1) yields

ϕ=μμ22Lϕ\phi=\mu-\sqrt{\mu^{2}-2L\phi}

by the assumption ϕ<μ\phi<\mu. We also note that the mapping xxx\mapsto\sqrt{x} is real analytic for x>0x>0, and hence an iterative argument in ss shows that ϕC(𝕋)\phi\in C^{\infty}(\mathbb{T}). To show that ϕC()\phi\in C^{\infty}(\mathbb{R}), we note that if ϕ\phi is a periodic solution of (4.1), its translation ϕh:=ϕ(+h)\phi_{h}:=\phi(\cdot+h) for any hh\in\mathbb{R} is also periodic. However, the previous analysis implies that ϕhC(𝕋)\phi_{h}\in C^{\infty}(\mathbb{T}) for any hh\in\mathbb{R}, so we conclude that ϕC()\phi\in C^{\infty}(\mathbb{R}).

To show part (ii)(ii), let ϕL()\phi\in L^{\infty}(\mathbb{R}), and assume that ϕ𝒞locs(U)\phi\in\mathcal{C}_{\mathrm{loc}}^{s}(U) for some open set UU and any ss\in\mathbb{R}. Let φC0(U)\varphi\in C^{\infty}_{0}(U) be a smooth function compactly supported in UU, and φ~C0(U)\widetilde{\varphi}\in C^{\infty}_{0}(U) be a cut-off function with φ~=1\widetilde{\varphi}=1 in a neighborhood VUV\Subset U of suppφ\operatorname{supp}\varphi. Then

φLϕ=φL(φ~ϕ)+φL((1φ~)ϕ).\varphi L\phi=\varphi L(\widetilde{\varphi}\phi)+\varphi L((1-\widetilde{\varphi})\phi).

It is easy to see that φL(φ~ϕ)𝒞locs+α(U)\varphi L(\widetilde{\varphi}\phi)\in\mathcal{C}_{\mathrm{loc}}^{s+\alpha}(U) and φL((1φ~)ϕ)\varphi L((1-\widetilde{\varphi})\phi) vanishes by construction. Thus, Lϕ𝒞locs+α(U)L\phi\in\mathcal{C}_{\mathrm{loc}}^{s+\alpha}(U). Using the same iteration argument in ss as in the proof of part (i)(i) shows that ϕ\phi is smooth on UU. ∎

As a motivation, we shall see that when the order of the dispersive term is strictly smaller than 1-1, any decrease of the order does not affect the regularity of the wave solution.

Theorem 4.5.

Let ϕμ\phi\leq\mu be an even solution of (4.1), which is nonincreasing on [0,π][0,\pi], and attains its maximum at ϕ(0)=μ\phi(0)=\mu. Then ϕ\phi cannot belong to the class C1(𝕋)C^{1}(\mathbb{T}).

Proof.

By contradiction, suppose that ϕ\phi is in C1(𝕋)C^{1}(\mathbb{T}). Since KPK^{\prime}_{P} is integrable by Theorem 3.1, equation (4.1) gives

ddx(μϕ)2L2KL1ϕL.\left\|{\frac{\,\mathrm{d}}{\,\mathrm{d}x}(\mu-\phi)^{2}}\right\|_{L^{\infty}}\leq 2\left\|{K^{\prime}}\right\|_{L^{1}}\left\|{\phi}\right\|_{L^{\infty}}.

Thus, (μϕ)2(\mu-\phi)^{2} is twice continuously differentiable. Then fixing |x|1|x|\ll 1 and applying the Taylor theorem, we have

(μϕ)2(x)=(μϕ)2(0)+[(μϕ)2](0)x+12[(μϕ)2]′′(ξ)x2(\mu-\phi)^{2}(x)=(\mu-\phi)^{2}(0)+[(\mu-\phi)^{2}]^{\prime}(0)x+\frac{1}{2}[(\mu-\phi)^{2}]^{\prime\prime}(\xi)x^{2} (4.3)

for some ξ\xi between 0 and xx. Differentiating equation (4.1) twice gives

12[(μϕ)2]′′(ξ)=Kϕ(ξ),\frac{1}{2}[(\mu-\phi)^{2}]^{\prime\prime}(\xi)=-K^{\prime}*\phi^{\prime}(\xi),

which substitutes into the expression (4.3) yields

(μϕ)2(x)=Kϕ(ξ)x2.(\mu-\phi)^{2}(x)=-K^{\prime}*\phi^{\prime}(\xi)x^{2}.

On the other hand, since KPK_{P} is even, nonincreasing on (0,π)(0,\pi) by Theorem 3.1, and ϕ\phi is nonincreasing on (0,π)(0,\pi), we obtain

Kϕ(0)=ππKP(y)ϕ(y)dy=20πKP(y)ϕ(y)dy=C>0-K^{\prime}*\phi^{\prime}(0)=-\int_{-\pi}^{\pi}K_{P}^{\prime}(-y)\phi^{\prime}(y)\,\mathrm{d}y=2\int_{0}^{\pi}K_{P}^{\prime}(y)\phi^{\prime}(y)\,\mathrm{d}y=C>0

for some constant C>0C>0. Both KK and ϕ\phi are continuous, so there exists a constant C0>0C_{0}>0 such that

Kϕ(ξ)C0-K^{\prime}*\phi^{\prime}(\xi)\geq C_{0}

for any ξ\xi between 0 and xx. Using this bound and equation (4.3), we have the estimate

(μϕ)2(x)C0x2,(\mu-\phi)^{2}(x)\geq C_{0}x^{2},

which is equivalent to

μϕ(x)|x|1.\frac{\mu-\phi(x)}{|x|}\gtrsim 1.

for all |x|1|x|\ll 1. Taking the limit when x0x\to 0 leads to contradiction to the fact ϕ(0)=0\phi^{\prime}(0)=0. ∎

In the next theorem, we investigate the regularity of a solution when it touches μ\mu from below. It was shown that the solution has 12\frac{1}{2}-Hölder regularity at x=0x=0 for an inhomogeneous multiplier of order greater than 1-1 [12], and Lipschitz continuous at x=0x=0 for a homogeneous multiplier of order less than 1-1 [5]. For our case with inhomogeneous symbol of order less than 1-1, we, indeed, obtain a Lipschitz regularity at x=0x=0.

Theorem 4.6.

Let ϕμ\phi\leq\mu be a nontrivial, 2π2\pi-periodic, even solution of (4.1), which is nonincreasing on [0,π][0,\pi]. If ϕ\phi attains its maximum at ϕ(0)=μ\phi(0)=\mu, then the following holds:

  1. (i)

    ϕC(𝕋\{0})\phi\in C^{\infty}(\mathbb{T}\backslash\{0\}) and ϕ\phi is strictly decreasing on (0,π)(0,\pi).

  2. (ii)

    ϕC1(𝕋)\phi\in C^{1-}(\mathbb{T}), that is ϕ\phi is Lipschitz continuous.

  3. (iii)

    ϕ\phi is precisely Lipschitz continuous at x=0x=0, that is, there exist constants c1,c2>0c_{1},c_{2}>0 such that

    c1|x|μϕ(x)c2|x|c_{1}|x|\leq\mu-\phi(x)\leq c_{2}|x|

    for |x|1|x|\ll 1.

Proof.

Assume that ϕμ\phi\leq\mu is an even solution of equation (4.1), which is nonincreasing on (0,π)(0,\pi) and attains its maximum at ϕ(0)=μ\phi(0)=\mu.

  1. (i)(i)

    Let x(0,π)x\in(0,\pi) and h(π,0)h\in(-\pi,0). Then we rewrite equation (4.1) as

    12(2μϕ(x+h)ϕ(xh))\displaystyle\frac{1}{2}(2\mu-\phi(x+h)-\phi(x-h)) (ϕ(x+h)ϕ(xh))\displaystyle(\phi(x+h)-\phi(x-h))
    =Kϕ(x+h)Kϕ(xh)\displaystyle=K*\phi(x+h)-K*\phi(x-h)

    We want to show that the right-hand side is strictly negative. In fact, for any x(0,π)x\in(0,\pi) and h(π,0)h\in(-\pi,0), since ϕ\phi and KPK_{P} are even and periodic, we obtain

    Kϕ(x+\displaystyle K*\phi(x+ h)Kϕ(xh)\displaystyle h)-K*\phi(x-h)
    =0π(KP(x+y)KP(xy))(ϕ(yh)ϕ(y+h))dy.\displaystyle=\int_{0}^{\pi}(K_{P}(x+y)-K_{P}(x-y))(\phi(y-h)-\phi(y+h))\,\mathrm{d}y.

    By similar arguments as in the proof of Lemma 3.2, we know that KP(x+y)KP(xy)<0K_{P}(x+y)-K_{P}(x-y)<0 for x,y(0,π)x,y\in(0,\pi), We further have ϕ(yh)ϕ(y+h)0\phi(y-h)-\phi(y+h)\leq 0 for y(0,π)y\in(0,\pi) and h(π,0)h\in(-\pi,0) by the assumption that ϕ\phi is even and nonincreasing on (0,π)(0,\pi). Therefore, the integrand is nonpositive. Since ϕ\phi is a nontrivial solution and KPK_{P} is not a constant, we conclude that

    Kϕ(x+h)Kϕ(xh)>0K*\phi(x+h)-K*\phi(x-h)>0 (4.4)

    for any h(π,0)h\in(-\pi,0). Hence, Kϕ(x+h)=Kϕ(xh)K*\phi(x+h)=K*\phi(x-h) if and only if ϕ(x+h)=ϕ(xh)\phi(x+h)=\phi(x-h). Then inequality (4.4) implies

    ϕ(x+h)>ϕ(xh)for anyh(π,0).\phi(x+h)>\phi(x-h)\qquad\text{for any}\quad h\in(-\pi,0).

    Thus, ϕ\phi is strictly decreasing on (0,π)(0,\pi). In view of Therem 4.4, ϕ\phi is smooth on 𝕋\{0}\mathbb{T}\backslash\{0\}.

  2. (ii)(ii)

    To prove the Lipschitz regularity at the maximum point, we make use of a bootstrap argument. By contradiction, assume that the solution ϕ<μ\phi<\mu is not Lipschitz continuous. Suppose that ϕ\phi is only a bounded function. Then recalling that LL maps L𝒞0(𝕋)L^{\infty}\subset\mathcal{C}^{0}(\mathbb{T}) into 𝒞α(𝕋)\mathcal{C}^{\alpha}(\mathbb{T}) for α>1\alpha>1 and the expression

    12(2μϕ(x)ϕ(y))(ϕ(x)ϕ(y))=Lϕ(x)Lϕ(y),\frac{1}{2}(2\mu-\phi(x)-\phi(y))(\phi(x)-\phi(y))=L\phi(x)-L\phi(y),

    we have the estimate

    12(ϕ(x)ϕ(y))2|Lϕ(x)Lϕ(y)||xy|.\frac{1}{2}(\phi(x)-\phi(y))^{2}\leq|L\phi(x)-L\phi(y)|\simeq|x-y|.

    Thus, it is straightforward to see that ϕ\phi is 12\frac{1}{2}-Hölder continuous.

    Next, evaluating equation (4.1) at x=0x=0 gives

    12μ2+Kϕ(0)=0,-\frac{1}{2}\mu^{2}+K*\phi(0)=0,

    and then subtracting from equation (4.1), we obtain

    12(μϕ)2(x)=Kϕ(0)Kϕ(x)\frac{1}{2}(\mu-\phi)^{2}(x)=K*\phi(0)-K*\phi(x)

    for x(0,π)x\in(0,\pi). Since ϕ\phi is smooth on 𝕋\{0}\mathbb{T}\backslash\{0\} by Theorem 4.4, differentiating the above equality yields

    (μϕ)ϕ(x)=(Kϕ)(x)(Kϕ)(0),(\mu-\phi)\phi^{\prime}(x)=(K*\phi)^{\prime}(x)-(K*\phi)^{\prime}(0),

    where we are using the fact that (Kϕ)(0)=0(K*\phi)^{\prime}(0)=0 and (Kϕ(0))=0(K*\phi(0))^{\prime}=0. If ϕ\phi is 12\frac{1}{2}-Hölder continuous, then Kϕ𝒞12+α(𝕋)K*\phi\in\mathcal{C}^{\frac{1}{2}+\alpha}(\mathbb{T}). Since α>1\alpha>1, we gain at least some Hölder regularity for (Kϕ)(K*\phi)^{\prime}. Thus, the right-hand side of expression ((ii)(ii)) can be estimated by a constant multiple of |x|β|x|^{\beta} for some β(12,1]\beta\in\left(\frac{1}{2},1\right], and hence,

    (μϕ)ϕ(x)|x|β.(\mu-\phi)\phi^{\prime}(x)\lesssim|x|^{\beta}. (4.5)

    By assumption that ϕ\phi is not Lipschitz continuous at x=0x=0, the above estimate guarantees that ϕ\phi is at least β\beta-Hölder continuous, where β>12\beta>\frac{1}{2}. We aim to bootstrap this argument to obtain Lipschitz regularity of ϕ\phi at x=0x=0. If 12+α>2\frac{1}{2}+\alpha>2, we use that Kϕ𝒞12+α(𝕋)C2(𝕋)K*\phi\in\mathcal{C}^{\frac{1}{2}+\alpha}(\mathbb{T})\subset C^{2}(\mathbb{T}), which guarantees that its derivative is at least Lipschitz continuous (β=1\beta=1 in expression (4.5)), and hence, ϕ\phi is Lipschitz continuous. On the other hand, if 12+α2\frac{1}{2}+\alpha\leq 2, then ϕ\phi is γ\gamma-Hölder continuous for some γ>β\gamma>\beta. We then repeat the argument finitely many times to yield that ϕ\phi is indeed Lipschitz continuous at x=0x=0, that is

    μϕ(x)|x|,for|x|1.\mu-\phi(x)\lesssim|x|,\qquad\mathrm{for}\quad|x|\ll 1. (4.6)
  3. (iii)(iii)

    From the upper bound (4.6), it remains to show that

    (μϕ)ϕ(x)|x|(\mu-\phi)\phi^{\prime}(x)\gtrsim|x| (4.7)

    for |x|1|x|\ll 1. Taking the derivative of equation (4.1) and evaluating at any ξ(0,π)\xi\in(0,\pi) we have

    (μϕ)ϕ(ξ)=(Kϕ)(ξ)=0π(KP(ξy)KP(ξ+y))ϕ(y)dy.(\mu-\phi)\phi^{\prime}(\xi)=(K*\phi)^{\prime}(\xi)=\int_{0}^{\pi}(K_{P}(\xi-y)-K_{P}(\xi+y))\phi^{\prime}(y)\,\mathrm{d}y.

    On the other hand, evaluating the upper bound established in (4.6) at x=ξ(0,π)x=\xi\in(0,\pi) gives μϕ(ξ)|ξ|\mu-\phi(\xi)\lesssim|\xi|. Therefore, dividing the above equation by (μϕ)(ξ)>0(\mu-\phi)(\xi)>0 to obtain the inequality

    ϕ(ξ)0πKP(ξy)KP(ξ+y)|ξ|ϕ(y)dy.\phi^{\prime}(\xi)\gtrsim\int_{0}^{\pi}\frac{K_{P}(\xi-y)-K_{P}(\xi+y)}{|\xi|}\phi^{\prime}(y)\,\mathrm{d}y. (4.8)

    Our aim is to show that lim infξ0+ϕ(ξ)\liminf_{\xi\to 0^{+}}\phi^{\prime}(\xi) is strictly bounded away from 0. We compute

    limξ0+KP(ξy)KP(ξ+y)|ξ|\displaystyle\lim_{\xi\to 0^{+}}\frac{K_{P}(\xi-y)-K_{P}(\xi+y)}{|\xi|}
    =limξ0+(KP(yξ)KP(y)ξ+KP(y)KP(y+ξ)ξ)ξ|ξ|=2KP(y)\displaystyle=\lim_{\xi\to 0^{+}}\left(\frac{K_{P}(y-\xi)-K_{P}(y)}{\xi}+\frac{K_{P}(y)-K_{P}(y+\xi)}{\xi}\right)\frac{\xi}{|\xi|}=2K_{P}^{\prime}(y)

    for any y(0,π)y\in(0,\pi). Then since KPK^{\prime}_{P} is integrable, taking the limit on both sides of the inequality (4.8) yields

    lim infξ0ϕ(ξ)20πKP(y)ϕ(y)dy=c\liminf_{\xi\to 0}\phi^{\prime}(\xi)\gtrsim 2\int_{0}^{\pi}K_{P}^{\prime}(y)\phi^{\prime}(y)\,\mathrm{d}y=c (4.9)

    for some constant c>0c>0, since ϕ\phi and KPK_{P} are strictly decreasing on (0,π)(0,\pi). Then for any 0<x10<x\ll 1, applying the Mean Value Theorem, we have

    ϕ(0)ϕ(x)x=ϕ(z)\frac{\phi(0)-\phi(x)}{x}=\phi^{\prime}(z)

    for some z(0,x)z\in(0,x). Combing with the expression (4.9) gives our desired estimate (4.7), and hence, we conclude that ϕ\phi is exactly Lipschitz continuous at x=0x=0. ∎

From Theorem 4.4 and 4.6, we conclude that any even, periodic solution ϕμ\phi\leq\mu 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 ϕμ\phi\leq\mu be an even solution of equation (4.1), which is nonincreasing on (0,π)(0,\pi). Then there exists a constant λ=λ(α)>0\lambda=\lambda(\alpha)>0, depending only on the kernel KK, such that

μϕ(π)λπ.\mu-\phi(\pi)\geq\lambda\pi.
Proof.

Choose any x[π4,3π4]x\in\left[\frac{\pi}{4},\frac{3\pi}{4}\right]. Then by the evenness of KK and the fact that ϕ\phi is nonincreasing on (0,π)(0,\pi), we have the estimate

(μϕ(π))ϕ(x)(μϕ(x))ϕ(x)=0π(KP(xy)KP(x+y))ϕ(y)dyπ/43π/4(KP(xy)KP(x+y))ϕ(y)dy,\begin{split}(\mu-\phi(\pi))\phi^{\prime}(x)&\leq(\mu-\phi(x))\phi^{\prime}(x)\\ &=\int_{0}^{\pi}(K_{P}(x-y)-K_{P}(x+y))\phi^{\prime}(y)\,\mathrm{d}y\\ &\leq\int_{\pi/4}^{3\pi/4}(K_{P}(x-y)-K_{P}(x+y))\phi^{\prime}(y)\,\mathrm{d}y,\end{split} (4.10)

since the integrand is nonpositive. Notice that KP(xy)KP(x+y)>0K_{P}(x-y)-K_{P}(x+y)>0 for x,y(0,π)x,y\in(0,\pi), and there exists a constant λ=λ(α)>0\lambda=\lambda(\alpha)>0, depending only on the kernel KK, such that

K(xy)K(x+y)2λfor allx,y(π4,3π4).K(x-y)-K(x+y)\geq 2\lambda\qquad\text{for all}\quad x,y\in\left(\frac{\pi}{4},\frac{3\pi}{4}\right).

Thus, integrating the inequality (4.10) with respect to xx over (π4,3π4)\left(\frac{\pi}{4},\frac{3\pi}{4}\right) yields

(μϕ(π))\displaystyle(\mu-\phi(\pi)) (ϕ(3π4)ϕ(π4))\displaystyle\left(\phi\left(\frac{3\pi}{4}\right)-\phi\left(\frac{\pi}{4}\right)\right)
π/43π/4(π/43π/4K(xy)K(x+y)dx)ϕ(y)dy\displaystyle\leq\int_{\pi/4}^{3\pi/4}\left(\int_{\pi/4}^{3\pi/4}K(x-y)-K(x+y)\,\mathrm{d}x\right)\phi^{\prime}(y)\,\mathrm{d}y
λπ(ϕ(3π4)ϕ(π4)).\displaystyle\leq\lambda\pi\left(\phi\left(\frac{3\pi}{4}\right)-\phi\left(\frac{\pi}{4}\right)\right).

From Theorem 4.6, we know that ϕ\phi is strictly decreasing on (π4,3π4)\left(\frac{\pi}{4},\frac{3\pi}{4}\right), and hence we can divide the above inequality by the quantity ϕ(3π4)ϕ(π4)<0\phi\left(\frac{3\pi}{4}\right)-\phi\left(\frac{\pi}{4}\right)<0 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, 2π2\pi-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 ϕ\phi, which is even, strictly decreasing on (0,π)(0,\pi) and attains its maximum ϕ(0)=μ\phi(0)=\mu. By Theorem 4.6, the highest wave is a peaked traveling-wave solution of

ut+Lux+uux=0for α>1.u_{t}+Lu_{x}+uu_{x}=0\qquad\text{for }\quad\alpha>1.

We use the subscript XevenX_{\mathrm{even}} for the restriction of a Banach space XX to its subset of even functions. Let β(1,2)\beta\in(1,2) and set

F:Cevenβ(𝕋)×+Cevenβ(𝕋),F:C^{\beta}_{\mathrm{even}}(\mathbb{T})\times\mathbb{R}_{+}\to C^{\beta}_{\mathrm{even}}(\mathbb{T}),

where

F(ϕ,μ):=μϕLϕ12ϕ2,(ϕ,μ)Cevenβ(𝕋)×+.F(\phi,\mu):=\mu\phi-L\phi-\frac{1}{2}\phi^{2},\qquad(\phi,\mu)\in C^{\beta}_{\mathrm{even}}(\mathbb{T})\times\mathbb{R}_{+}. (5.1)

Then F(ϕ,μ)=0F(\phi,\mu)=0 if and only if ϕ\phi is an even Cβ(𝕋)C^{\beta}(\mathbb{T})-solution of (1.2) corresponding to the wave speed μ+\mu\in\mathbb{R}_{+}. We have the first local bifurcation result as follow.

Theorem 5.1 (Local bifurcation).

For each integer k1k\geq 1, the point (0,μk)(0,\mu_{k}^{*}), where μk=(1+k2)α/2\mu_{k}^{*}=(1+k^{2})^{-\alpha/2} is a bifurcation point. More precisely, there exits ϵ0>0\epsilon_{0}>0 and an analytic curve through (0,μk)(0,\mu_{k}^{*}),

{(ϕk(ϵ),μk(ϵ)):|ϵ|<ϵ0}Cevenβ(𝕋)×+,\{(\phi_{k}(\epsilon),\mu_{k}(\epsilon)):|\epsilon|<\epsilon_{0}\}\subset C^{\beta}_{\mathrm{even}}(\mathbb{T})\times\mathbb{R}_{+},

of nontrivial, 2πk\frac{2\pi}{k}-periodic, even solutions of (5.1) with μk(0)=μk\mu_{k}(0)=\mu_{k}^{*} and

Dϵϕk(0)=ϕk(x)=cos(xk).D_{\epsilon}\phi_{k}(0)=\phi_{k}^{*}(x)=\cos(xk).

In a neighborhood of the bifurcation point (0,μk)(0,\mu_{k}^{*}) these are all the nontrivial solutions of F(ϕ,μ)=0F(\phi,\mu)=0 in Cevenβ(𝕋)×+C^{\beta}_{\mathrm{even}}(\mathbb{T})\times\mathbb{R}_{+}.

Proof.

We will prove the result using the analytic version of the Crandall–Rabinowitz Theorem [7, Theorem 8.4.1]. It is clear that F(0,μ)=0F(0,\mu)=0 for any μ+\mu\in\mathbb{R}_{+}. We are looking for 2π2\pi-periodic, even, nontrivial solutions bifurcating from the line {(0,μ):μ}\{(0,\mu):\mu\in\mathbb{R}\} of trivial solutions. The wave speed μ>0\mu>0 shall be the bifurcation parameter. The linearization of FF around the trivial solution (ϕ=0,μ)(\phi=0,\mu) is given by

DϕF(0,μ):Cevenβ(𝕋)Cevenβ(𝕋),ϕ(μIdL)ϕ.D_{\phi}F(0,\mu):C^{\beta}_{\mathrm{even}}(\mathbb{T})\to C^{\beta}_{\mathrm{even}}(\mathbb{T}),\qquad\phi\mapsto(\mu\,\mathrm{Id}-L)\phi.

Then from Section 2, we know that L:Cevenβ(𝕋)Cevenβ+α(𝕋)L:C^{\beta}_{\mathrm{even}}(\mathbb{T})\to C^{\beta+\alpha}_{\mathrm{even}}(\mathbb{T}) is parity preserving and a smoothing operator, which implies that it is compact on Cevenβ(𝕋)C^{\beta}_{\mathrm{even}}(\mathbb{T}). Hence, DϕF(0,μ)D_{\phi}F(0,\mu) is a compact perturbation of an isomorphism, and therefore constitutes a Fredholm operator of index 0. The nontrivial kernel of DϕF(0,μ)D_{\phi}F(0,\mu) is spanned by functions ψCevenβ(𝕋)\psi\in C^{\beta}_{\mathrm{even}}(\mathbb{T}) satisfying

ψ^(k)(μ(1+k2)α/2)=0\widehat{\psi}(k)(\mu-(1+k^{2})^{-\alpha/2})=0

for all kk. For μ(0,1]\mu\in(0,1], we see that suppψ{±(μ2/α1)1/2}\operatorname{supp}\psi\subseteq\{\pm(\mu^{-2/\alpha}-1)^{1/2}\}. Therefore, the kernel of DϕF(0,μ)D_{\phi}F(0,\mu) is one-dimensional if and only if μ=μk:=(1+k2)α/2\mu=\mu_{k}^{*}:=(1+k^{2})^{-\alpha/2} for some kk\in\mathbb{Z}, in which case it is given by

kerDϕF(0,μ)=span{ϕk}withϕk(x):=cos(xk).\ker D_{\phi}F(0,\mu)=\mathrm{span}\{\phi_{k}^{*}\}\qquad\text{with}\quad\phi_{k}^{*}(x):=\cos(xk).

We also note that μk\mu_{k}^{*} are all simple eigenvalues of LL. Finally, we observe that

DϕμF(0,μk)ϕk=ϕkD_{\phi\mu}F(0,\mu_{k}^{*})\phi_{k}^{*}=\phi_{k}^{*}

is not in the range of DϕF(0,μk)D_{\phi}F(0,\mu_{k}^{*}), 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 F(ϕ,μ)=0F(\phi,\mu)=0. Recalling β(1,2)\beta\in(1,2), set

S:={(ϕ,μ)U:F(ϕ,μ)=0},S:=\{(\phi,\mu)\in U:F(\phi,\mu)=0\},

where the admissible set UU is given by

U:={(ϕ,μ)Cevenβ(𝕋)×+:ϕ<μ}.U:=\{(\phi,\mu)\in C^{\beta}_{\mathrm{even}}(\mathbb{T})\times\mathbb{R}_{+}:\phi<\mu\}.

Then all bounded solutions ϕ\phi of equation (4.1) satisfy F(ϕ,μ)=0F(\phi,\mu)=0 for all (ϕ,μ)S(\phi,\mu)\in S. We start the analysis by providing the LL^{\infty}-bound for a solution.

Lemma 5.2 (LL^{\infty} bound).

Let μ>0\mu>0. Then any bounded solution ϕ\phi to equation (1.2) satisfies:

ϕL2(μ+KL1).\left\|{\phi}\right\|_{L^{\infty}}\leq 2(\mu+\left\|{K}\right\|_{L^{1}}).
Proof.

From equation (4.1), we have the estimate

ϕL22(μ+KL1)ϕL,\left\|{\phi}\right\|_{L^{\infty}}^{2}\leq 2(\mu+\left\|{K}\right\|_{L^{1}})\left\|{\phi}\right\|_{L^{\infty}},

where we have used the fact that the kernel KK is integrable. Therefore, ϕ0\phi\equiv 0, which holds trivially, or the desired estimate follows by dividing by ϕL\left\|{\phi}\right\|_{L^{\infty}}. ∎

Before extending the local branches globally, we need two helping lemmas.

Lemma 5.3.

The Frechét derivative DϕF(ϕ,μ)D_{\phi}F(\phi,\mu) is a Fredholm operator of index 0 for all (ϕ,μ)U(\phi,\mu)\in U.

Proof.

We have

DϕF(ϕ,μ)=(μϕ)IdLD_{\phi}F(\phi,\mu)=(\mu-\phi)\mathrm{Id}-L

for any given (ϕ,μ)U(\phi,\mu)\in U. Since (μϕ)Id(\mu-\phi)\mathrm{Id} is an isomorphism on Cβ(𝕋)C^{\beta}(\mathbb{T}) and LL is compact on Cβ(𝕋)C^{\beta}(\mathbb{T}), the operator DϕF(ϕ,μ)D_{\phi}F(\phi,\mu) is Fredholm. From the proof of Theorem 5.1, we know that DϕF(0,μ)D_{\phi}F(0,\mu) has Fredholm index 0, and so is DϕF(ϕ,μ)D_{\phi}F(\phi,\mu) due to the fact that the index is continuous. ∎

Lemma 5.4.

For any (ϕ,μ)S(\phi,\mu)\in S, the function ϕ\phi is smooth, and any bounded and closed subset of SS is compact in Cevenβ(𝕋)×+C^{\beta}_{\mathrm{even}}(\mathbb{T})\times\mathbb{R}_{+}.

Proof.

If (ϕ,μ)S(\phi,\mu)\in S, then ϕ<μ\phi<\mu, and we write equation (1.2) in the form

ϕ=μμ22Lϕ=:F~(ϕ,μ).\phi=\mu-\sqrt{\mu^{2}-2L\phi}=:\widetilde{F}(\phi,\mu).

The function F~\widetilde{F} is a bounded and linear mapping from UU into Cevenβ+α(𝕋)C^{\beta+\alpha}_{\mathrm{even}}(\mathbb{T}). Moreover, by Theorem 4.4, we know that ϕ\phi is smooth.

Let ASUA\subset S\subset U be a bounded and closed set. Then F~(A)={ϕ:(ϕ,μ)A}\widetilde{F}(A)=\{\phi:(\phi,\mu)\in A\} is relatively compact in Cevenβ(𝕋)C^{\beta}_{\mathrm{even}}(\mathbb{T}). Since AA is closed, any sequence {(ϕn,μn)}n\{(\phi_{n},\mu_{n})\}_{n\in\mathbb{N}} has a convergent subsequence in AA by Arzela–Ascoli’s lemma. We conclude that AA is compact in Cevenβ(𝕋)×+C^{\beta}_{\mathrm{even}}(\mathbb{T})\times\mathbb{R}_{+}. ∎

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 μ(ϵ)\mu(\epsilon) is not identically zero for 0<ϵ10<\epsilon\ll 1. However, the latter claim is an immediate consequence of Theorem 5.7 below. This theorem is an adaptation of [11, Theorem 4.4] with UU and SS as above.

Theorem 5.5 (Global bifurcation).

The local bifurcation curve s(ϕk(s),μk(s))s\mapsto(\phi_{k}(s),\mu_{k}(s)) from Theorem 5.1 of solutions of equation (5.1) extends to a global continuous curve of solutions 𝔊:+S\mathfrak{G}:\mathbb{R}_{+}\to S, that allows a local real-analytic reparameterization around each s>0s>0. One of the following alternatives holds:

  1. (i)

    (ϕk(s),μk(s))Cβ(𝕋)×+\|(\phi_{k}(s),\mu_{k}(s))\|_{C^{\beta}(\mathbb{T})\times\mathbb{R}_{+}}\to\infty as ss\to\infty.

  2. (ii)

    There exists a subsequence (ϕk(sn),μk(sn))n(\phi_{k}(s_{n}),\mu_{k}(s_{n}))_{n\in\mathbb{N}} such that the pair (ϕk(sn),μk(sn))(\phi_{k}(s_{n}),\linebreak[1]\mu_{k}(s_{n})) approaches the boundary of SS as nn\to\infty.

  3. (iii)

    The function s(ϕk(s),μk(s))s\mapsto(\phi_{k}(s),\mu_{k}(s)) is (finitely) periodic.

We apply the Lyapunov–Schmidt reduction, in order to establish the bifurcation formulas. Let kk\in\mathbb{N} be a fixed number and set

M:=span{cos(xl):lk},N:=kerDϕF(0,μk)=span{ϕk}.M:=\mathrm{span}\{\cos(xl):l\neq k\},\qquad N:=\ker D_{\phi}F(0,\mu_{k}^{*})=\mathrm{span}\{\phi_{k}^{*}\}.

Then Cevenβ(𝕋)=MNC^{\beta}_{\mathrm{even}}(\mathbb{T})=M\oplus N and a continuous projection onto the one-dimensional space NN is given by

Πϕ=ϕ,ϕkL2ϕk,\Pi\phi=\langle\phi,\phi_{k}^{*}\rangle_{L^{2}}\phi_{k}^{*},

where ,L2\langle\cdot,\cdot\rangle_{L^{2}} denotes the inner product in L2(𝕋)L^{2}(\mathbb{T}). Let us recall the Lyapunov–Schmidt reduction theorem from [18, Theorem I.2.3].

Theorem 5.6 (Lyapunov–Schmidt reduction).

There exists a neighborhood 𝒪×YU\mathcal{O}\times Y\subset U of (0,μk)(0,\mu_{k}^{*}) such that the problem

F(ϕ,μ)=0for(ϕ,μ)𝒪×YF(\phi,\mu)=0\qquad\mathrm{for}\quad(\phi,\mu)\in\mathcal{O}\times Y (5.2)

is equivalent to the finite-dimensional problem

Φ(ϵϕk,μ):=ΠF(ϵϕk+ψ(ϵϕk,μ),μ)=0\Phi(\epsilon\phi_{k}^{*},\mu):=\Pi F(\epsilon\phi_{k}^{*}+\psi(\epsilon\phi_{k}^{*},\mu),\mu)=0 (5.3)

for functions ψC(𝒪N×Y,M)\psi\in C^{\infty}(\mathcal{O}_{N}\times Y,M) and 𝒪NN\mathcal{O}_{N}\subset N an open neighborhood of the zero function in NN. One has that Φ(0,μk)=0\Phi(0,\mu_{k}^{*})=0, ψ(0,μk)=0\psi(0,\mu_{k}^{*})=0, Dϕψ(0,μk)=0D_{\phi}\psi(0,\mu_{k}^{*})=0, and solving problem (5.3) provides a solution

ϕ=ϵϕk+ψ(ϵϕk,μ)\phi=\epsilon\phi_{k}^{*}+\psi(\epsilon\phi_{k}^{*},\mu)

of the infinite-dimensional problem (5.2).

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).

The bifurcation curve found in Theorem 5.5 satisfies

ϕk(ϵ)=ϵcos(kx)+ϵ24(1m(k)m(0)+1m(k)m(2k)cos(2kx))+O(ϵ3)\phi_{k}(\epsilon)=\epsilon\cos(kx)+\frac{\epsilon^{2}}{4}\left(\frac{1}{m(k)-m(0)}+\frac{1}{m(k)-m(2k)}\cos(2kx)\right)+O(\epsilon^{3}) (5.4)

and

μk(ϵ)=m(k)+ϵ24(1m(k)m(0)+12(m(k)m(2k)))+O(ϵ3)\mu_{k}(\epsilon)=m(k)+\frac{\epsilon^{2}}{4}\left(\frac{1}{m(k)-m(0)}+\frac{1}{2(m(k)-m(2k))}\right)+O(\epsilon^{3}) (5.5)

in Cevenβ(𝕋)×+C^{\beta}_{\mathrm{even}}(\mathbb{T})\times\mathbb{R}_{+} as ϵ0\epsilon\to 0, where m(k)=(1+k2)α/2m(k)=(1+k^{2})^{-\alpha/2}, α>1\alpha>1. In particular, μ¨(0)>0\ddot{\mu}(0)>0 for any k1k\geq 1, that is, the local bifurcation in Theorem 5.1 describes a supercritical pitchfork bifurcation.

Proof.

For (ϕk(s),μk(s))k(\phi_{k}(s),\mu_{k}(s))_{k\in\mathbb{N}} along the analytic local bifurcation curve in Theorem 5.1, we fix kk\in\mathbb{N} and suppress the subscript kk to lighten our notation. We first show that

μ(s)=μ(s)\mu(s)=\mu(-s)

after a suitable choice of parameterization. Since ϕ(s,)\phi(s,\cdot) is even, it has a Fourier cosine representation

ϕ(s,)=12[ϕ(s)]0+j=1[ϕ(s)]jcos(j),\phi(s,\cdot)=\frac{1}{2}[\phi(s)]_{0}+\sum_{j=1}^{\infty}[\phi(s)]_{j}\cos(j\cdot),

where

[ϕ(s)]j=1πππϕ(s,y)cos(jy)dy[\phi(s)]_{j}=\frac{1}{\pi}\int_{-\pi}^{\pi}\phi(s,y)\cos(jy)\,\mathrm{d}y

for j=0,1,2,j=0,1,2,\dots 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 [ϕ(s)]1=s[\phi(s)]_{1}=s. We observe that if ϕ(s)\phi(s) is an even and 2π2\pi-periodic function, then ϕ(s,+π)\phi(s,\cdot+\pi) is also an even and 2π2\pi-periodic function, and its Fourier coefficient satisfies

[ϕ(s,+π)]1=1πππϕ(s,y+π)cos(y)dy=[ϕ(s)]1.[\phi(s,\cdot+\pi)]_{1}=\frac{1}{\pi}\int_{-\pi}^{\pi}\phi(s,y+\pi)\cos(y)\,\mathrm{d}y=-[\phi(s)]_{1}.

Since [ϕ(s,+π)]1=[ϕ(s)]1=s[\phi(s,\cdot+\pi)]_{1}=-[\phi(s)]_{1}=-s, uniqueness of the bifurcation curve yields

(ϕ(s,+π),μ(s))=(ϕ(s,),μ(s)),(\phi(s,\cdot+\pi),\mu(s))=(\phi(-s,\cdot),\mu(-s)),

which means μ(s)=μ(s)\mu(s)=\mu(-s). Therefore, by the analyticity of μ(s)\mu(s), we can write

μ(s)=n=0μ2ns2n,\mu(s)=\sum_{n=0}^{\infty}\mu_{2n}s^{2n},

where the sum is uniformly convergent in a neighborhood of 0. Similarly, ϕ(s)\phi(s) has an expansion

ϕ(s)=n=1ϕnsn\phi(s)=\sum_{n=1}^{\infty}\phi_{n}s^{n}

with convergence in Cevenβ(𝕋)C_{\mathrm{even}}^{\beta}(\mathbb{T}). Substituting these two formulae into the main equation (1.2) and equating coefficients of equal order in ss, we obtain

Lϕ1μ0ϕ1\displaystyle L\phi_{1}-\mu_{0}\phi_{1} =0,\displaystyle=0, (5.6)
Lϕ2μ0ϕ2\displaystyle L\phi_{2}-\mu_{0}\phi_{2} =12ϕ12,\displaystyle=-\frac{1}{2}\phi_{1}^{2}, (5.7)
Lϕ3μ0ϕ3\displaystyle L\phi_{3}-\mu_{0}\phi_{3} =μ2ϕ1ϕ1ϕ2.\displaystyle=\mu_{2}\phi_{1}-\phi_{1}\phi_{2}. (5.8)

By Theorem 5.1, ϕ1=ϕk(x)=cos(kx)\phi_{1}=\phi_{k}^{*}(x)=\cos(kx) and μ0=μk=m(k)\mu_{0}=\mu_{k}^{*}=m(k), 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 μn\mu_{n} and ϕn\phi_{n} can be determined by solving the corresponding equation.

Recalling the formula for LL given by (2.4) and expressing ϕ2(x)\phi_{2}(x) as a Fourier series, equation (5.7) gives

12πnm(n)ϕ^2(n)exp(inx)m(k)2πnϕ^2(n)exp(inx)=1414cos(2kx),\frac{1}{2\pi}\sum_{n\in\mathbb{Z}}m(n)\widehat{\phi}_{2}(n)\exp(inx)-\frac{m(k)}{2\pi}\sum_{n\in\mathbb{Z}}\widehat{\phi}_{2}(n)\exp(inx)=-\frac{1}{4}-\frac{1}{4}\cos(2kx),

and then comparing coefficients of cosine functions, we obtain the relation

{12πm(0)ϕ^2(0)m(k)2πϕ^2(0)=14,1πm(2k)ϕ^2(2k)m(k)πϕ^2(2k)=14,\begin{dcases}\frac{1}{2\pi}m(0)\widehat{\phi}_{2}(0)-\frac{m(k)}{2\pi}\widehat{\phi}_{2}(0)&=-\frac{1}{4},\\ \frac{1}{\pi}m(2k)\widehat{\phi}_{2}(2k)-\frac{m(k)}{\pi}\widehat{\phi}_{2}(2k)&=-\frac{1}{4},\end{dcases}

which yields

ϕ2(x)=14(m(k)m(0))+14(m(k)m(2k))cos(2kx).\phi_{2}(x)=\frac{1}{4(m(k)-m(0))}+\frac{1}{4(m(k)-m(2k))}\cos(2kx).

Next, the right-hand side of equation (5.8) becomes

(μ214(m(k)m(0))18(m(k)m(2k)))cos(kx)18(m(k)m(2k))cos(3kx).\left(\mu_{2}-\frac{1}{4(m(k)-m(0))}-\frac{1}{8(m(k)-m(2k))}\right)\cos(kx)-\frac{1}{8(m(k)-m(2k))}\cos(3kx).

Recalling the above parameterization [ϕ(s)]1=s[\phi(s)]_{1}=s, which implies that [ϕn]1=0[\phi_{n}]_{1}=0 for all n2n\geq 2, we find that

μ2=14(m(k)m(0))+18(m(k)m(2k))>0.\mu_{2}=\frac{1}{4(m(k)-m(0))}+\frac{1}{8(m(k)-m(2k))}>0.

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 (iii)(iii) in Theorem 5.5 is excluded, and both alternatives (i)(i) and (ii)(ii) occur simultaneously as ss\to\infty along the bifurcation branch 𝔊\mathfrak{G}. This implies that the highest wave is reached as a limit of the global bifurcation curve.

Lemma 5.8.

Any sequence of solutions (ϕk,μk)kS(\phi_{k},\mu_{k})_{k\in\mathbb{N}}\subset S to equation (4.1) with (μk)k(\mu_{k})_{k\in\mathbb{N}} bounded has a subsequence which converges uniformly to a solution ϕ\phi.

Proof.

Recalling the estimate of ϕ\phi in Lemma 5.2,

ϕL2(μ+KL1),\left\|{\phi}\right\|_{L^{\infty}}\leq 2(\mu+\left\|{K}\right\|_{L^{1}}),

we see that if (μk)k(\mu_{k})_{k\in\mathbb{N}} is bounded, then (ϕk)k(\phi_{k})_{k\in\mathbb{N}} is bounded. Since KK is integrable and continuous on \mathbb{R}, Dominated Convergence Theorem allows us to conclude that (Lϕk)k(L\phi_{k})_{k\in\mathbb{N}} is equicontinuous. Then Arzela–Ascoli’s lemma implies that there exists a subsequence which converges uniformly to a solution ϕ\phi of equation (4.1). ∎

Let

𝒦k:={ϕCevenβ(𝕋):ϕ is 2π/k-periodic and nonincreasing in (0,π/k)}\mathcal{K}_{k}:=\{\phi\in C^{\beta}_{\mathrm{even}}(\mathbb{T}):\phi\text{ is }2\pi/k\text{-periodic and nonincreasing in }(0,\pi/k)\}

be a closed cone in Cβ(𝕋)C^{\beta}(\mathbb{T}). The next proposition exclude alternative (iii)(iii) in Theorem 5.5.

Proposition 5.9.

The solutions ϕk(s)\phi_{k}(s), s>0s>0 on the global bifurcation curve \mathfrak{R} belong to 𝒦k\{0}\mathcal{K}_{k}\backslash\{0\} and alternative (iii)(iii) in Theorem 5.5 does not occur. In particular, the bifurcation curve (ϕk(s),μk(s))(\phi_{k}(s),\mu_{k}(s)) has no intersection with the trivial solution line for any s>0s>0.

Proof.

Due to [7, Theorem 9.2.2] the statement holds true if the following conditions are satisfied

  1. (a)

    𝒦k\mathcal{K}_{k} is a cone in a real Banach space.

  2. (b)

    (ϕk(ϵ),μk(ϵ))𝒦k×(\phi_{k}(\epsilon),\mu_{k}(\epsilon))\subset\mathcal{K}_{k}\times\mathbb{R} provided ϵ\epsilon is small enough.

  3. (c)

    If μ\mu\in\mathbb{R} and ϕkerDϕF(0,μ)𝒦k\phi\in\ker D_{\phi}F(0,\mu)\cap\mathcal{K}_{k}, then ϕ=γϕ\phi=\gamma\phi^{*} for γ0\gamma\geq 0 and μ=μk\mu=\mu_{k}^{*}.

  4. (d)

    Each nontrivial point on the bifurcation curve which also belongs to 𝒦k×\mathcal{K}_{k}\times\mathbb{R} is an interior point of 𝒦k×\mathcal{K}_{k}\times\mathbb{R} in SS.

Conditions (a), (b), (c) are satisfied because of the local bifurcation result in Theorem 5.1, so it remains to verify condition (d). Let (ϕ,μ)𝒦k×(\phi,\mu)\in\mathcal{K}_{k}\times\mathbb{R} be a nontrivial solution on the bifurcation curve found in Theorem 5.5. By Theorem 4.4, ϕ\phi is smooth and together with Lemma 4.3, we have ϕ<0\phi^{\prime}<0 on (0,π)(0,\pi) and ϕ′′(0)<0\phi^{\prime\prime}(0)<0. Choose a solution φ\varphi lying within |δ|1|\delta|\ll 1 small enough neighborhood in Cβ(𝕋)C^{\beta}(\mathbb{T}) such that φ<μ\varphi<\mu and ϕφCβ<δ\|\phi-\varphi\|_{C^{\beta}}<\delta. In view of the mapping (4.2), an iteration process on the regularity index yields that ϕφC2<δ~\|\phi-\varphi\|_{C^{2}}<\widetilde{\delta}, where δ~=δ~(δ)>0\widetilde{\delta}=\widetilde{\delta}(\delta)>0 can be made arbitrarily small by choosing δ\delta small enough. It follows that for δ\delta small enough, φ<μ\varphi<\mu is a smooth, even, nonincreasing on (0,π)(0,\pi) solution, and hence (ϕ,μ)(\phi,\mu) belongs to the interior of 𝒦k×\mathcal{K}_{k}\times\mathbb{R} in SS, which concludes the proof. ∎

Remark 5.10.

From the proof of Theorem 5.1 and Lemma 4.3, we have the bound 0<μ10<\mu\leq 1. Moreover, integrating equation (4.1) on \mathbb{R} and using the fact that m(0)=1m(0)=1, we obtain

(μ1)ϕ(x)dx=12ϕ2(x)dx.(\mu-1)\int_{\mathbb{R}}\phi(x)\,\mathrm{d}x=\frac{1}{2}\int_{\mathbb{R}}\phi^{2}(x)\,\mathrm{d}x.

Thus, the only way to reach μ=1\mu=1 is by approaching ϕ=0\phi=0, and hence when s>0s>0 is small, we have 0<μ<10<\mu<1. By Proposition 5.9, the bifurcation curve does not intersect the trivial solution line for any s>0s>0, so μ(s)<1\mu(s)<1 for all ss.

In the next lemma, we show that along the bifurcation curve, the wave speed μ\mu is, in fact, bounded away from 0.

Lemma 5.11.

Along the bifurcation curve in Theorem 5.5 we have that

μ(s)1\mu(s)\gtrsim 1

uniformly for all s0s\geq 0.

Proof.

By contradiction, assume that there exists a sequence (sn)k+(s_{n})_{k\in\mathbb{N}}\in\mathbb{R}_{+} with limnsn=\lim_{n\to\infty}s_{n}=\infty such that μ(sn)0\mu(s_{n})\to 0 as nn\to\infty, while ϕ(sn)ϕ0\phi(s_{n})\to\phi_{0} as nn\to\infty along the bifurcation curve 𝔊\mathfrak{G} found in Theorem 5.5. In view of Lemma 5.4, there exists a subsequence of (sn)n(s_{n})_{n\in\mathbb{N}} such that ϕ(sn)\phi(s_{n}) converges to a solution ϕ0\phi_{0} of equation (5.1). Along the bifurcation curve, we have that ϕ(sn)<μ(sn)0\phi(s_{n})<\mu(s_{n})\to 0 as nn\to\infty, and it follows that ϕ00\phi_{0}\leq 0. In view of Lemma 4.1, we have maxxϕ0(x)=0\max_{x}\phi_{0}(x)=0, whence ϕ00\phi_{0}\equiv 0 by Remark 4.2. Finally, Lemma 4.7 leads to

0=limn(μ(sn)ϕ(sn)(π))λπ>0,0=\lim_{n\to\infty}(\mu(s_{n})-\phi(s_{n})(\pi))\geq\lambda\pi>0,

which is a contradiction. Thus, we have μ(s)1\mu(s)\gtrsim 1 uniformly for all s0s\geq 0. ∎

At this point, we know that the wave speed is bounded away from 0 and 11. Before proving the main result, we show that alternatives (i)(i) and (ii)(ii) in Theorem 5.5 occur simultaneously.

Theorem 5.12.

In Theorem 5.5, alternatives (i)(i) and (ii)(ii) both occur.

Proof.

Let (ϕk(s),μk(s))(\phi_{k}(s),\mu_{k}(s)), ss\in\mathbb{R}, be the bifurcation curve found in Theorem 5.5. By Proposition 5.9, we know that ϕk(s)\phi_{k}(s) is nontrivial, even, and nonincreasing on (0,π)(0,\pi), and alternative (iii)(iii) in Theorem 5.5 does not occur. Thus, it is either alternative (i)(i) or alternative (ii)(ii) in Theorem 5.5 occurs.

Suppose that alternative (i)(i) occurs. Then we have either ϕk(s)Cβ\|\phi_{k}(s)\|_{C^{\beta}}\to\infty for some β(1,2)\beta\in(1,2) or |μk(s)||\mu_{k}(s)|\to\infty as ss\to\infty. By Remark 5.10, μk(s)\mu_{k}(s) is bounded between 0 and 11, so the only possibility is ϕk(s)Cβ(𝕋)\|\phi_{k}(s)\|_{C^{\beta}(\mathbb{T})} is unbounded. In this case, alternative (ii)(ii) must occur, that is, the quantity

lim infsinfx(μk(s)ϕk(s)(x))=0\liminf_{s\to\infty}\inf_{x\in\mathbb{R}}(\mu_{k}(s)-\phi_{k}(s)(x))=0

holds. Indeed, by contradiction, suppose that

lim infsinfx(μk(s)ϕk(s)(x))δ\liminf_{s\to\infty}\inf_{x\in\mathbb{R}}(\mu_{k}(s)-\phi_{k}(s)(x))\geq\delta

for some δ>0\delta>0. For any such solution (ϕk,μk)(\phi_{k},\mu_{k}), equation (1.2) gives

|ϕk(x)ϕk(y)|=2|Lϕk(x)Lϕk(y)|2μkϕk(x)ϕk(y)|Lϕk(x)Lϕk(y)|δ.|\phi_{k}(x)-\phi_{k}(y)|=\frac{2|L\phi_{k}(x)-L\phi_{k}(y)|}{2\mu_{k}-\phi_{k}(x)-\phi_{k}(y)}\leq\frac{|L\phi_{k}(x)-L\phi_{k}(y)|}{\delta}.

By Lemma 5.2, ϕk\phi_{k} is uniformly bounded in C(𝕋)C(\mathbb{T}). Then since LL maps Cβ(𝕋)C^{\beta}(\mathbb{T}) to Cβ+αC^{\beta+\alpha}, there exists some constant C>0C>0 depending on δ\delta and LL such that

ϕkCβ(𝕋)C,\|\phi_{k}\|_{C^{\beta}(\mathbb{T})}\leq C,

which is a contradiction. Therefore, alternative (ii)(ii) must occur.

On the other hand, suppose that alternative (ii)(ii), but not alternative (i)(i), occurs. Then fixing kk, there exists a sequence (ϕk(sn),μk(sn))n(\phi_{k}(s_{n}),\mu_{k}(s_{n}))_{n\in\mathbb{N}} in Theorem 5.5 solving equation (1.2) satisfying ϕk(sn)0\phi_{k}^{\prime}(s_{n})\leq 0 on (0,π)(0,\pi), ϕk(sn)<μk(sn)\phi_{k}(s_{n})<\mu_{k}(s_{n}), and

lim infn|μk(sn)ϕk(sn)(0)|=0,\liminf_{n\to\infty}|\mu_{k}(s_{n})-\phi_{k}(s_{n})(0)|=0,

while ϕk(sn)\phi_{k}(s_{n}) remains uniformly bounded in Cβ(𝕋)C^{\beta}(\mathbb{T}) for β(1,2)\beta\in(1,2). Taking a limit along a subsequence in CβC^{\beta^{\prime}} for some β(1,β)\beta^{\prime}\in(1,\beta) yields a contradiction to Theorem 4.6. Therefore, both alternatives (i)(i) and (ii)(ii) 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 (ϕk(s),μk(s))(\phi_{k}(s),\mu_{k}(s)) be the global bifurcation curve 𝔊\mathfrak{G} found in Theorem 5.5 and let (sn)n(s_{n})_{n\in\mathbb{N}} be a sequence in +\mathbb{R}_{+} tending to infinity. By Lemma 5.11 and Remark 5.10, we know that (μk(sn))n(\mu_{k}(s_{n}))_{n\in\mathbb{N}} is bounded between and away from 0 and 11. Moreover, Lemma 5.8 gives the existence of a subsequence (ϕk(snl),μk(snl))(\phi_{k}(s_{n_{l}}),\mu_{k}(s_{n_{l}})) converging uniformly to a solution (ϕ0,μ0)(\phi_{0},\mu_{0}) as ll\to\infty. Finally, from Theorem 5.12 and Theorem 4.6, we conclude that ϕ0(0)=μ0\phi_{0}(0)=\mu_{0} with ϕ¯\bar{\phi} 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.