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

Subharmonic Dynamics of Wave Trains in the Korteweg-de Vries / Kuramoto-Sivashinsky Equation

Mathew A. Johnson & Wesley R. Perkins Department of Mathematics, University of Kansas, 1460 Jayhawk Boulevard, Lawrence, KS 66045, USA; matjohn@ku.eduDepartment of Mathematics, Lehigh University, 17 Memorial Drive East, Bethlehem, PA 18015, USA; wesley.perkins@lehigh.edu
(August 14, 2025)
Abstract

We study the stability and nonlinear local dynamics of spectrally stable periodic wave trains of the Korteweg-de Vries / Kuramoto-Sivashinsky equation when subjected to classes of periodic perturbations. It is known that for each NN\in{\mathbb{N}}, such a TT-periodic wave train is asymptotically stable to NTNT-periodic, i.e., subharmonic, perturbations, in the sense that initially nearby data will converge asymptotically to a small Galilean boost of the underlying wave, with exponential rates of decay. However, both the allowable size of initial perturbations and the exponential rates of decay depend on NN and, in fact, tend to zero as NN\to\infty, leading to a lack of uniformity in such subharmonic stability results. Our goal here is to build upon a recent methodology introduced by the authors in the reaction-diffusion setting and achieve a subharmonic stability result which is uniform in NN. This work is motivated by the dynamics of such wave trains when subjected to perturbations which are localized (i.e., integrable on the line).

1 Introduction

In this work, we consider the local dynamics of wave trains, i.e., periodic traveling wave solutions, of the Korteweg-de Vries / Kuramoto-Sivashinsky (KdV/KS) equation

(1.1) ut+εuxxx+δ(uxx+uxxxx)+uux=0u_{t}+\varepsilon u_{xxx}+\delta\left(u_{xx}+u_{xxxx}\right)+uu_{x}=0

where x,tx,t\in{\mathbb{R}}. Here, ε0\varepsilon\geq 0 and δ>0\delta>0 are modeling parameters which may, without loss of generality, be chosen such that ε2+δ2=1\varepsilon^{2}+\delta^{2}=1: see Remark 1.1 below. The equation (1.1) is known to be a canonical model for pattern formation that has been used to describe many applications including plasma instabilities, turbulence in reaction diffusion equations, flame front propagation, and nonlinear wave dynamics in fluid mechanics [20, 21, 7, 15, 16]. In the case ε=0\varepsilon=0 and δ=1\delta=1, equation (1.1) becomes the classical Kuramoto-Sivashinsky equation, which is known to be a generic equation for chaotic dynamics, and there is a very large literature on these solutions, their bifurcations and period doubling cascades, and their stability: see, for example, [9] and reference therein. In this ε=0\varepsilon=0 case, (1.1) is also known to model thin film dynamics down a completely vertical wall [6]. When the angle of the wall is decreased from vertical, however, additional dispersive effects are present [9] and modeled by ε>0\varepsilon>0, and in the “flat” limit where the inclined wall becomes horizontal one recovers the completely integrable Korteweg-de Vries equation, corresponding here to ε=1\varepsilon=1 and δ=0\delta=0. Thus, in the context of inclined thin film flow one can consider the general model (1.1) as interpolating between the vertical wall (ε,δ)=(0,1)(\varepsilon,\delta)=(0,1) and the “flat” limit (ε,δ)(1,0+)(\varepsilon,\delta)\to(1^{-},0^{+}). For more information on the connection to thin film dynamics, as well as its derivation in this context from the viscous shallow water equations or the full Navier Stokes equations, see [24, 25].

Remark 1.1.

In the literature, one may encounter more general looking systems of the form

ut+εuxxx+δuxx+γuxxxx+Λuux=0u_{t}+\varepsilon u_{xxx}+\delta u_{xx}+\gamma u_{xxxx}+\Lambda uu_{x}=0

where here ε\varepsilon\in{\mathbb{R}} and δ,γ,Λ>0\delta,\gamma,\Lambda>0 are arbitrary constants. However, we note that, through a rescaling argument, such systems can always be put in the form (1.1), i.e., one can always take Λ=1\Lambda=1 and δ=γ\delta=\gamma with ε2+δ2=1\varepsilon^{2}+\delta^{2}=1. See [3, Section 2]. Note the particular scaling here is sometimes referred to as the “thin film” scaling: see [9] for example.

In this work, we are interested in understanding the stability and long-time dynamics of periodic traveling wave solutions of (1.1) to specific classes of perturbations. To begin our discussion, we briefly discuss the existence theory for periodic solutions of (1.1). First, note that traveling wave solutions of (1.1) correspond to solutions of the form u(x,t)=u¯(xct)u(x,t)=\bar{u}(x-ct), where u¯\bar{u} necessarily satisfies the profile ODE

(1.2) cu¯+εu¯′′′+δ(u¯′′+u¯′′′′)+u¯u¯=0.-c\bar{u}^{\prime}+\varepsilon\bar{u}^{\prime\prime\prime}+\delta\left(\bar{u}^{\prime\prime}+\bar{u}^{\prime\prime\prime\prime}\right)+\bar{u}\bar{u}^{\prime}=0.

The existence and local structure of periodic solutions of (1.2) has been studied extensively by several other authors. For general values of (ε,δ)(\varepsilon,\delta), an elementary Hopf bifurcation analysis [3] shows the existence of a 33-parameter family of asymptotically small amplitude periodic traveling wave solutions of (1.2) which, up to translation, can be parametrized by the wave speed cc and the period TT. In the “classical KS” limit |ε|1|\varepsilon|\ll 1, one can likewise use normal form analysis to establish a similar existence result [9], while the full bifurcation picture for the KS equation (ε=0\varepsilon=0) is known to be extremely complicated: see, for example, [14], where the authors prove the existence of a Shi’lnikov bifurcation which leads to cascades of period doubling, period multiplying kk-bifurcations and oscillatory homoclinic orbits as the period is increased, as well as the numerical bifurcation study in [4]. For a summary and more details, see [3, 11].

Using the above existence studies as motivation, and closely following the work in [3], we make the following assumption regarding the existence of periodic solutions of (1.2), as well as the structure of the local manifold of periodic solutions.

Assumption 1.2.

Suppose u¯()\bar{u}(\cdot) is a TT-periodic solution of (1.2) with speed cc. Then the set of periodic solutions near u¯\bar{u} forms a 33-dimensional manifold

{(x,t)u(xγ(c+δc)t;X):γ,(δc,X)Ω}\left\{(x,t)\mapsto u(x-\gamma-(c+\delta c)t;X):\gamma\in{\mathbb{R}},~(\delta c,X)\in\Omega\right\}

where Ω2\Omega\subset{\mathbb{R}}^{2} is some open set containing the point (δc,X)=(0,T)(\delta c,X)=(0,T) and each u(;X)u(\cdot;X) is an XX-periodic solution to (1.2) with wave speed c+δcc+\delta c.

Remark 1.3.

It is natural to assume (1.2) admits a 33-dimensional manifold of periodic solutions. Indeed, note that integrating the profile equation (1.2) once yields

cu¯+εu¯′′+δ(u¯+u¯′′′)+12u¯2=q-c\bar{u}+\varepsilon\bar{u}^{\prime\prime}+\delta\left(\bar{u}^{\prime}+\bar{u}^{\prime\prime\prime}\right)+\frac{1}{2}\bar{u}^{2}=q

for some constant of integration qq\in{\mathbb{R}}. Periodic solutions of (1.2) thus correspond to values

(T,c,q,u¯(0),u¯(0),u¯′′(0))6,(T,c,q,\bar{u}(0),\bar{u}^{\prime}(0),\bar{u}^{\prime\prime}(0))\in{\mathbb{R}}^{6},

where TT, cc, and qq denote the period, wave speed and constant of integration, subjected to the periodicity condition

(u¯(T),u¯(T),u¯′′(T))=(u¯(0),u¯(0),u¯′′(0)).\left(\bar{u}(T),\bar{u}^{\prime}(T),\bar{u}^{\prime\prime}(T)\right)=\left(\bar{u}(0),\bar{u}^{\prime}(0),\bar{u}^{\prime\prime}(0)\right).

leading, generically, to a 33-parameter family of periodic solutions parametrized, up to translation, by the period TT, the wave speed cc.

As stated above, our main goal is to study the stability and dynamics of periodic traveling wave solutions of (1.1) to specific classes of perturbations. Previously, there has been much work regarding both the spectral and nonlinear stability of such periodic solutions when subjected to localized perturbations, i.e., perturbations that are integrable on the line [3, 2, 11], as well as their dynamics under slow-modulations [17]. In these works, it is found that for each admissible pair (ε,δ)(\varepsilon,\delta) in (1.1) there exist periodic traveling wave solutions which are spectrally and nonlinearly stable to localized perturbations. Here, however, we study the stability and dynamics of TT-periodic traveling wave solutions of (1.1) when subject to so-called subharmonic perturbations, i.e., NTNT-periodic perturbations for some NN\in{\mathbb{N}}. Before we continue, we note that an extremely important feature of (1.1), which we utilize heavily in our forthcoming analysis, is the presence of a Galilean symmetry. In particular, if u(x,t)u(x,t) is a solution of (1.1), then so is the function

(1.3) u(xct,t)+cu(x-ct,t)+c

for any cc\in{\mathbb{R}}. Thanks to this Galilean invariance, as well as the translational invariance of (1.1), it follows that the stability of a particular wave only depends on one parameter: namely, the period TT. Thus, when discussing the stability of periodic traveling wave solutions of the KdV-KS equation (1.1), we identify all waves of a particular period. Furthermore, note we can view (1.3) as coupling between the wave speed and the mass of the wave, defined via

M:=0Tu(x,t)𝑑x,M:=\int_{0}^{T}u(x,t)dx,

which is readily seen to be a conserved quantity of (1.1) due to the conservative structure.

Now, suppose that u¯(x)\bar{u}(x) is a TT-periodic solution of (1.2) with wave speed cc, and note the linearization of (1.1) (in the appropriate co-moving frame) about u¯\bar{u} is given by the operator

:=x(cu¯x2)δ(x2+x4).\mathcal{L}:=\partial_{x}\left(c-\bar{u}-\partial_{x}^{2}\right)-\delta\left(\partial_{x}^{2}+\partial_{x}^{4}\right).

Since we are considering subharmonic perturbations of u¯\bar{u}, i.e., perturbations with period NTNT for some NN\in{\mathbb{N}}, here we consider \mathcal{L} as a closed, densely defined linear operator acting on Lper2(0,NT)L^{2}_{\rm per}(0,NT) with TT-periodic coefficients. To describe the spectrum of \mathcal{L} acting on Lper2(0,NT)L^{2}_{\rm per}(0,NT), first observe that one always has that

(1.4) u¯x=0and(1)=u¯x\mathcal{L}\bar{u}_{x}=0\quad{\rm and}\quad\mathcal{L}(1)=-\bar{u}_{x}

so that λ=0\lambda=0 is an eigenvalue of \mathcal{L} with algebraic multiplicity at least two and a Jordan chain of height at least one. With this in mind, and following the works [19, 18, 10, 13], we introduce the notion of spectral stability used throughout this work.

Definition 1.4.

A TT-periodic traveling wave solution u¯Hloc1()\bar{u}\in H^{1}_{\rm loc}({\mathbb{R}}) of (1.2) is said to be diffusively spectrally stable provided the following conditions hold:

  • (D1)

    The spectrum of the linear operator \mathcal{L} acting on L2()L^{2}({\mathbb{R}}) satisfies

    σL2()(){λ:(λ)<0}{0};\sigma_{L^{2}({\mathbb{R}})}\left(\mathcal{L}\right)\subset\left\{\lambda\in{\mathbb{R}}:\Re(\lambda)<0\right\}\cap\{0\};
  • (D2)

    There exists a θ>0\theta>0 such that for any ξ[π,π)\xi\in[-\pi,\pi) the real part of the spectrum of the Bloch operator ξ:=eiξxeiξx\mathcal{L}_{\xi}:=e^{-i\xi x}\mathcal{L}e^{i\xi x} acting on Lper2(0,1)L^{2}_{\rm per}(0,1) satisfies

    (σLper2(0,1)(ξ))θξ2;\Re\left(\sigma_{L^{2}_{\rm per}(0,1)}\left(\mathcal{L}_{\xi}\right)\right)\leq-\theta\xi^{2};
  • (D3)

    λ=0\lambda=0 is a TT-periodic eigenvalue of 0[ϕ]\mathcal{L}_{0}[\phi] with algebraic multiplicity two and geometric multiplicity one.

Refer to caption
Figure 1: Plot of the numerically obtained stability boundaries, computed in [3, Section 2], in the period (denoted here as XX) versus the parameter ε\varepsilon in (1.1). Note that δ=1ε2\delta=\sqrt{1-\varepsilon^{2}} is fixed by the choice of ε\varepsilon and that, by Galilean invariance, the stability of such a wave depends only on its period. In this figure, the shaded regions correspond to diffusively spectrally stable periodic traveling wave solutions of (1.1).

We note that in [3] it was shown that for every admissible pair of modeling parameters (ε,δ)(\varepsilon,\delta) there exists a range of periods for which all periodic traveling wave solutions of (1.1) with those period are diffusively spectrally stable: see Figure 1 and also [1, 5]. In the context of the KdV/KS equation (1.1), it is known that diffusively spectrally stable periodic traveling wave solutions (coupled with an additional non-degeneracy hypothesis, see Assumption 1.5 below) are nonlinearly stable to localized perturbations, in the sense that localized perturbations of such a periodic waves u¯\bar{u} converges to spatially localized phase modulations of u¯\bar{u}. Specifically, given such a periodic traveling wave u¯\bar{u} and initial data

u(x,0)=u¯(x)+v(x,0),v(,0)||L1()H5()1u(x,0)=\bar{u}(x)+v(x,0),\quad\|v(\cdot,0)||_{L^{1}({\mathbb{R}})\cap H^{5}({\mathbb{R}})}\ll 1

then for large time the associated solution u(x,t)u(x,t) satisfies

u(x,t)u¯(x+Ψ(x,t)),t1u(x,t)\approx\bar{u}(x+\Psi(x,t)),~~~t\gg 1

for some function Ψ(,t)W5,()\Psi(\cdot,t)\in W^{5,\infty}({\mathbb{R}}) which behaves essentially like a finite sum of (small) error functions: see [3]. As we will see below, the implication of diffusive spectral stability will be that such a TT-periodic traveling wave is necessarily spectrally stable to perturbations in Lper2(0,NT)L^{2}_{\rm per}(0,NT) for every NN\in{\mathbb{N}}.

Before continuing to state our main result, we introduce an additional non-degeneracy hypothesis. As we will see in Lemma 2.1 below, Assumption 1.2, along with the diffusive spectral stability assumption, implies that for |ξ||\xi| small the Bloch operators ξ\mathcal{L}_{\xi} have two eigenvalues near the origin which expand as

(1.5) λj(ξ)=iajξ+o(ξ),aj,j=1,2\lambda_{j}(\xi)=-ia_{j}\xi+o\left(\xi\right),~~a_{j}\in{\mathbb{R}},~~j=1,2

for |ξ|1|\xi|\ll 1. Following the previous work [3, 10], we make the following additional non-degeneracy hypothesis:

Assumption 1.5.

The coefficients aja_{j}\in{\mathbb{R}} in (1.5) are distinct.

By standard spectral perturbation theory, Assumption 1.5 ensures the analyticity of the functions λj()\lambda_{j}(\cdot) in (1.5): again, see Lemma 2.1 below. Furthermore, it is known that the coefficients aja_{j} are the characteristics of an associated Whitham averaged system, formally governing slowly modulated periodic solutions of (1.1). Consequently, Assumption 1.5 corresponds to strict hyperbolicity of the Whitham averaged system. See [17] for more details. Note also that all the diffusively spectrally stable periodic traveling wave solutions in Figure 1 were seen numerically to satisfy Assumption 1.5.

We now begin our discussion of our main results concerning the dynamics of TT-periodic, diffusively spectrally stable traveling wave solutions of (1.1) when subjected to subharmonic perturbations. First, in Section 2.1 we will see that the spectrum of \mathcal{L} acting on Lper2(0,NT)L^{2}_{\rm per}(0,NT) is given by the union of the (necessarily discrete) spectrum of the corresponding Bloch operators ξ\mathcal{L}_{\xi} acting in Lper2(0,T)L^{2}_{\rm per}(0,T) for the discrete (finite) subset of ξ[π/T,π/T)\xi\in[-\pi/T,\pi/T) such that eiξNT=1e^{i\xi NT}=1. Thus, such a traveling wave u¯\bar{u} is necessarily spectrally stable to all subharmonic perturbations. In particular, for each NN\in{\mathbb{N}} there exists a constant δN>0\delta_{N}>0 such that

(σLN2(){0})δN.\Re\left(\sigma_{L^{2}_{N}}\left(\mathcal{L}\right)\setminus\{0\}\right)\leq-\delta_{N}.

Since \mathcal{L} is clearly sectorial, it is easy to show that for NN\in{\mathbb{N}} fixed and δ(0,δN)\delta\in(0,\delta_{N}) there exists a constant Cδ>0C_{\delta}>0 such that

(1.6) et(1𝒫N)fLN2CeδtfLN2\left\|e^{\mathcal{L}t}\left(1-\mathcal{P}_{N}\right)f\right\|_{L^{2}_{N}}\leq Ce^{-\delta t}\|f\|_{L^{2}_{N}}

for all fLN2f\in L^{2}_{N}, where here 𝒫N\mathcal{P}_{N} denotes the projection of LN2L^{2}_{N} onto the NTNT-periodic generalized kernel of \mathcal{L}. Equipped with this linear estimate and exploiting the Galilean invariance (1.3), the following nonlinear stability result was established in [22, 23] for the case N=1N=1 and is easily extended to general NN\in{\mathbb{N}}.

Proposition 1.6 (Nonlinear Subharmonic Stability [22, 23]).

Let u¯Hloc1\bar{u}\in H^{1}_{\rm loc} be a TT-periodic traveling wave solution of (1.1) with wave speed cc. Assume that u¯\bar{u} is diffusively spectrally stable and additionally satisfies Assumption 1.5. Fix NN\in{\mathbb{N}} and take δN>0\delta_{N}>0 such that

max((σLper2(0,NT)()){0})=δN\max\left(\Re\left(\sigma_{L^{2}_{\rm per}(0,NT)}\left(\mathcal{L}\right)\right)\setminus\{0\}\right)=-\delta_{N}

holds. Then for each NN\in{\mathbb{N}} and for every δ(0,δN)\delta\in(0,\delta_{N}), there exists an ε=εδ>0\varepsilon=\varepsilon_{\delta}>0 and a constant C=Cδ>0C=C_{\delta}>0 such that whenever u0Hper1(0,NT)u_{0}\in H^{1}_{\rm per}(0,NT) and u0u¯H1(0,NT)<ε\|u_{0}-\bar{u}\|_{H^{1}(0,NT)}<\varepsilon, then the solution uu of (1.1) with initial data u(0)=u0u(0)=u_{0} exists globally in time and satisfies

u(+γ,t)u¯((δM)t)δMH1(0,NT)Ceδtu0u¯H1(0,NT)\left\|u(\cdot+\gamma_{\infty},t)-\bar{u}\left(\cdot-(\delta M)t\right)-\delta M\right\|_{H^{1}(0,NT)}\leq Ce^{-\delta t}\left\|u_{0}-\bar{u}\right\|_{H^{1}(0,NT)}

for all t>0t>0, where here γ=γ(N)\gamma_{\infty}=\gamma_{\infty}(N) is some constant and

δM:=1NT0NT(u0(z)u¯(z))𝑑z.\delta M:=\frac{1}{NT}\int_{0}^{NT}\left(u_{0}(z)-\bar{u}(z)\right)dz.

The above result establishes the nonlinear asymptotic stability of the Galilean family associated to u¯\bar{u}, showing that nearby NTNT-periodic solutions will, up to a spatial translation, asymptotically converge to a member of the Galilean family of u¯\bar{u}. However, Proposition 1.6 lacks uniformity in NN in two important ways. Specifically, both the exponential rate of decay δ\delta and the allowable size of initial perturbations ε=εδ\varepsilon=\varepsilon_{\delta} are controlled by the size of the spectral gap δN>0\delta_{N}>0. Since it is known that δN0\delta_{N}\to 0 as NN\to\infty, it follows that both δ\delta and ε\varepsilon in Proposition 1.6 must necessarily tend to zero as NN\to\infty. Of course, from a practical level it would be preferable to develop a theory which, for a given background wave u¯\bar{u}, yielded a fixed size for the theoretically prescribed domain of attraction as well as a uniform (in NN) rate of decay of initial perturbations. The fact that this is possible is precisely our main result.

Theorem 1.7 (Uniform Subharmonic Asymptotic Stability).

Let u¯Hloc1\bar{u}\in H^{1}_{\rm loc} be a TT-periodic traveling wave solution of (1.1) with wave speed cc. Assume that u¯\bar{u} is diffusively spectrally stable and, additionally, satisfies Assumption 1.5. There exists an ε>0\varepsilon>0 and a constant C>0C>0 such that, for every NN\in{\mathbb{N}}, whenever u0Lper1(0,NT)Hper5(0,NT)u_{0}\in L^{1}_{\rm per}(0,NT)\cap H^{5}_{\rm per}(0,NT) and

E0:=u0u¯Lper1(0,NT)Hper5(0,NT)<ε,E_{0}:=\left\|u_{0}-\bar{u}\right\|_{L^{1}_{\rm per}(0,NT)\cap H^{5}_{\rm per}(0,NT)}<\varepsilon,

there exists a function ψ~(x,t)\widetilde{\psi}(x,t) satisfying ψ~(,0)0\widetilde{\psi}(\cdot,0)\equiv 0 such that the solution uu of (1.1) with initial data u(0)=u0u(0)=u_{0} exists globally in time and satisfies

(1.7) u(ψ~(,t),t)u¯((δM)t)δMHper5(0,NT),x,tψ~(,t)Hper5(0,NT)CE0(1+t)1/4\left\|u\left(\cdot-\widetilde{\psi}(\cdot,t),t\right)-\bar{u}\left(\cdot-(\delta M)t\right)-\delta M\right\|_{H^{5}_{\rm per}(0,NT)},~~\left\|\nabla_{x,t}\widetilde{\psi}(\cdot,t)\right\|_{H^{5}_{\rm per}(0,NT)}\leq CE_{0}(1+t)^{-1/4}

for all t0t\geq 0.

The main difficulty in establishing the proof of Theorem 1.7 appears at the linear level. Specifically, one must develop a strategy to handle the accumulation of NTNT-periodic eigenvalues near λ=0\lambda=0 for N1N\gg 1 in a uniform way. In the proof of Proposition 1.6, this accumulation is handled by enclosing the origin in the spectral plane in an small ball B(0,rN)B(0,r_{N}) where the NN-dependent radius is chosen so that

(1.8) σLper2(0,NT)()B(0,rN)={0}\sigma_{L^{2}_{\rm per}(0,NT)}\left(\mathcal{L}\right)\cap B(0,r_{N})=\{0\}

and defining the associate Riesz spectral projection

(1.9) 𝒫N=12πiB(0,rN)dzz\mathcal{P}_{N}=\frac{1}{2\pi i}\int_{\partial B(0,r_{N})}\frac{dz}{z-\mathcal{L}}

onto the NTNT-periodic generalized kernel of \mathcal{L}. One then decomposes the linear solution operator (semigroup) as

et=et𝒫N+et(1𝒫N)e^{\mathcal{L}t}=e^{\mathcal{L}t}\mathcal{P}_{N}+e^{\mathcal{L}t}\left(1-\mathcal{P}_{N}\right)

and uses the exponential bound (1.6) to establish the result. The lack of uniformity in Proposition 1.6, however, stems from the fact that the radius rNr_{N} of the ball used in the Riesz projection (1.9) must necessarily tend to zero as NN\to\infty in order to maintain the spectral decomposition (1.8).

To establish the uniformity in Theorem 1.7, however, we cannot allow the radius of the ball about the origin to shrink. Instead, we define a ball B(0,r)B(0,r) with an NN-independent radius and define the associated Riesz spectral projection associated to the total generalized eigenspace corresponding to eigenvalues interior to the ball. Naturally, the dimension of this total eigenspace is tending to infinity as NN\to\infty, and we must work to establish uniform in NN decay estimates associated with the induced decomposition of the semigroup. This methodology was first introduced, at the linear level, by the authors and collaborators in [8] and was further extended, in the context of reaction-diffusion systems, to the full nonlinear level in [12], and is closely modeled off of the known stability theory for localized perturbations [10]. In the present work, the additional complication (compared to these previous subharmonic works) is the presence of a non-trivial Jordan block (1.4) which, in turn, yields slower uniform decay rates of the associated linear semigroups as compared to those rates in the reaction-diffusion context.

Finally, we make a few remarks concerning the connection between Proposition 1.6 and Theorem 1.7. The analysis in Section 4 shows that the modulation function ψ~\widetilde{\psi} in Theorem 1.7 can be decomposed as

ψ~(x,t)=1Nγ(t)+ψ(x,t)\widetilde{\psi}(x,t)=\frac{1}{N}\gamma(t)+\psi(x,t)

where here ψ(,t)WN5,\psi(\cdot,t)\in W^{5,\infty}_{N} satisfies

(1.10) ψ(,t)LNCE0\left\|\psi(\cdot,t)\right\|_{L^{\infty}_{N}}\leq CE_{0}

for some constant C>0C>0 independent of NN. Further, while our methods fail to directly yield convergence111This is due to the presence of a non-trivial Jordan block associated with the co-periodic Bloch opertaor 0\mathcal{L}_{0}, leading to slower linear decay rates in Section 3 when compared to, for example, the reaction-diffusion context where no such Jordan block is present (see [12] for details). of γ(t)\gamma(t) as tt\to\infty, one immediate observation is that for each fixed NN\in{\mathbb{N}} and NTNT-periodic solution u(x,t)u(x,t) of (1.1) with initial data sufficiently close (in LN1HN5L^{1}_{N}\cap H^{5}_{N}) to u¯\bar{u}, then using the notation from both Proposition 1.6 and Theorem 1.7 we have

u(+γ,t)u(xψ~(,t),t)LN2CN(1+t)1/4\left\|u(\cdot+\gamma_{\infty},t)-u(x-\widetilde{\psi}(\cdot,t),t)\right\|_{L^{2}_{N}}\leq C_{N}(1+t)^{-1/4}

for some (NN-dependent) constant CN>0C_{N}>0. This leads one to expect that, under relatively generic circumstances, for each fixed NN one should have

1Nγ(t)+ψ(x,t)γ\frac{1}{N}\gamma(t)+\psi(x,t)\to\gamma_{\infty}

as tt\to\infty. Understanding this convergence rigorously is an interesting remaining problem from our analysis.

The outline of this paper is as follows. In Section 2 we briefly review the application of Floquet-Bloch theory to subharmonic perturbations and collect several properties of the Bloch operators and their associated semigroups. We further establish basic “high-frequency” decay properties of the Bloch semigroups that arise as consequences of the diffusive spectral stability assumption. In Section 3 we provide a delicate decomposition of the semigroup ete^{\mathcal{L}t} acting on Lper2(0,NT)L^{2}_{\rm per}(0,NT) which will yield polynomial decay rates on the semigroup which are uniform in NN: see Proposition 3.1. This linear decomposition then motivates in Section 4 a nonlinear decomposition of a small Lper2(0,NT)L^{2}_{\rm per}(0,NT) neighborhood of the underlying diffusively stable TT-periodic background wave u¯\bar{u}. Equipped with this nonlinear decomposition, along with the linear estimates from Section 3, we apply a nonlinear iteration scheme to corresponding system of perturbation equations, yielding a proof of our main result Theorem 1.7 above.

Acknowledgments: The work of MAJ was partially funded by the NSF under grant DMS-2108749, as well as the Simons Foundation Collaboration grant number 714021.

2 Preliminaries

Here, we review several preliminary analytical results. We begin with a brief review of the application of Floquet-Bloch theory to subharmonic perturbations, and then use this to establish some elementary semigroup estimates for the associated Bloch operators. For notational convenience, for each NN\in{\mathbb{N}}, p1p\geq 1, and fixed period T>0T>0 we introduce the notation

LNp:=Lperp(0,NT)L^{p}_{N}:=L^{p}_{\rm per}(0,NT)

and similarly with all associated Sobolev spaces.

2.1 Floquet-Bloch Theory for Subharmonic Perturbations

We begin by reviewing the results of Floquet-Bloch theory when applied to subharmonic perturbations. For more details, see the recent works [8, 12].

Suppose that u¯\bar{u} is a TT-periodic traveling wave solution of (1.1) with wave speed cc\in{\mathbb{R}}, and consider the associated linearized operator \mathcal{L}. For each fixed NN\in{\mathbb{N}}, define222Observe that ΩN\Omega_{N} is always a finite set with |ΩN|=N|\Omega_{N}|=N and with the distance between any two closest elements being 2π/NT2\pi/NT.

ΩN:={ξ[π/T,π/T):eiξNT=1},\Omega_{N}:=\left\{\xi\in[-\pi/T,\pi/T):e^{i\xi NT}=1\right\},

and note that, since the coefficients of \mathcal{L} are TT-periodic, basic results in Floquet-theory implies that any NTNT-periodic solution of the ordinary differential equation

v=λv\mathcal{L}v=\lambda v

must be of the form

v(x)=eiξxTw(x)v(x)=e^{i\xi xT}w(x)

for some ξΩN\xi\in\Omega_{N} and wLper2(0,T)w\in L^{2}_{\rm per}(0,T). More specifically, one can show that λ\lambda\in{\mathbb{C}} is an NTNT-periodic eigenvalue of \mathcal{L} if and only if there exists a ξΩN\xi\in\Omega_{N} such that the problem

λw=eiξxeiξxw=:ξw,\lambda w=e^{-i\xi x}\mathcal{L}e^{i\xi x}w=:\mathcal{L}_{\xi}w,

admits a non-trivial solution in Lper2(0,T)L^{2}_{\rm per}(0,T), where here the operators ξ\mathcal{L}_{\xi} are known as the Bloch operators associated to \mathcal{L} and the parameter ξ\xi is referred to as the Bloch frequency: see also Definition 1.4. In fact, we have the spectral decomposition

σLN2()=ξΩNσLper2(0,T)(ξ),\sigma_{L^{2}_{N}}\left(\mathcal{L}\right)=\bigcup_{\xi\in\Omega_{N}}\sigma_{L^{2}_{\rm per}(0,T)}\left(\mathcal{L}_{\xi}\right),

which characterizes the NN-periodic spectrum of \mathcal{L} in terms of the union of a TT-periodic eigenvalues for the 11-parameter family of Bloch operators {ξ}ξΩN\{\mathcal{L}_{\xi}\}_{\xi\in\Omega_{N}}.

From above, it is natural when studying such NTNT-periodic spectral problems to desire to decompose arbitrary functions as superpositions of functions of the form eiξw()e^{i\xi\cdot}w(\cdot) with wLper2(0,T)w\in L^{2}_{\rm per}(0,T). To this end, given a function gLN2g\in L^{2}_{N} we define the TT-periodic Bloch transform of gg as

T(g)(ξ,x):=e2πix/Tg^(ξ+2π/T),ξΩN,x\mathcal{B}_{T}(g)(\xi,x):=\sum_{\ell\in{\mathbb{Z}}}e^{2\pi i\ell x/T}\hat{g}\left(\xi+2\pi\ell/T\right),~~\xi\in\Omega_{N},~~x\in{\mathbb{R}}

where now g^\hat{g} denotes the Fourier transform of gg on the torus given by

(2.1) g^(z):=NT/2NT/2eizyg(y)𝑑y.\hat{g}(z):=\int_{-NT/2}^{NT/2}e^{-izy}g(y)dy.

In particular, observe that for any ξΩN\xi\in\Omega_{N} and gLN2g\in L^{2}_{N}, the function T(g)(ξ,)\mathcal{B}_{T}(g)(\xi,\cdot) is TT-periodic and, furthermore, the function gg can be recovered via the inverse Bloch representation formula

g(x)=1NTξΩNeiξxT(g)(ξ,x).g(x)=\frac{1}{NT}\sum_{\xi\in\Omega_{N}}e^{i\xi x}\mathcal{B}_{T}(g)(\xi,x).

One can check that the TT-periodic Bloch transform

T:LN22(ΩN,Lper2(0,T))\mathcal{B}_{T}:L^{2}_{N}\mapsto\ell^{2}\left(\Omega_{N},L^{2}_{\rm per}(0,T)\right)

as defined above satisfies the subharmonic Parseval identify

(2.2) f,gLN2=1NT2ξΩNT(f)(ξ,),T(g)(ξ,)L2(0,T)\left<f,g\right>_{L^{2}_{N}}=\frac{1}{NT^{2}}\sum_{\xi\in\Omega_{N}}\left<\mathcal{B}_{T}(f)(\xi,\cdot),\mathcal{B}_{T}(g)(\xi,\cdot)\right>_{L^{2}(0,T)}

valid for all f,gLN2f,g\in L^{2}_{N}. Finally, we note that if fLper2(0,T)f\in L^{2}_{\rm per}(0,T) and gLper2(0,NT)g\in L^{2}_{\rm per}(0,NT) then

T(fg)(ξ,x)=f(x)T(g)(ξ,x)\mathcal{B}_{T}(fg)(\xi,x)=f(x)\mathcal{B}_{T}(g)(\xi,x)

and, additionally, we have the identity

(2.3) f,gL2(0,NT)=1Tf,T(g)(0,)L2(0,T).\left<f,g\right>_{L^{2}(0,NT)}=\frac{1}{T}\left<f,\mathcal{B}_{T}(g)(0,\cdot)\right>_{L^{2}(0,T)}.

For proofs of (2.2) and (2.3), see [8, Section 2].

With the above functional analytic tools in hand, we readily see that for our given linearized operator \mathcal{L} and vLN2v\in L^{2}_{N} we have

T(v)(ξ,x)=ξT(v)(ξ,x)\mathcal{B}_{T}\left(\mathcal{L}v\right)(\xi,x)=\mathcal{L}_{\xi}\mathcal{B}_{T}(v)(\xi,x)

so that the operators ξ\mathcal{L}_{\xi} may be viewed as operator-valued symbols under T\mathcal{B}_{T} acting on LN2L^{2}_{N}. Similarly, since the operators \mathcal{L} and ξ\mathcal{L}_{\xi} are clearly sectorial on their respective domains, they clearly generate analytic semigroups on LN2L^{2}_{N} and Lper2(0,T)L^{2}_{\rm per}(0,T), respectively, and further their associated semigroups satisfy

(2.4) T(etv)(ξ,x)=(eξtT(v)(ξ,))(x)andetv(x)=1NTξΩNeiξxeξt(v)(ξ,x).\mathcal{B}_{T}\left(e^{\mathcal{L}t}v\right)(\xi,x)=\left(e^{\mathcal{L}_{\xi}t}\mathcal{B}_{T}(v)(\xi,\cdot)\right)(x)~~{\rm and}~~e^{\mathcal{L}t}v(x)=\frac{1}{NT}\sum_{\xi\in\Omega_{N}}e^{i\xi x}e^{\mathcal{L}_{\xi}t}\mathcal{B}(v)(\xi,x).

In particular, we see that the TT-periodic Bloch transform T\mathcal{B}_{T} diagonalizes the periodic coefficient operator differential \mathcal{L} acting on LN2L^{2}_{N} in the same way that the Fourier transform diagonalizes constant-coefficient differential operators acting on L2()L^{2}({\mathbb{R}}). This decomposition formula will be used heavily in our forthcoming linear stability estimates.

2.2 Diffusive Spectral Stability & Basic Properties of Bloch Semigroups

With the above preliminaries, we now establish some important immediate consequences of the diffusive spectral stability assumption. As a first result, we describe the unfolding of the Jordan block (1.4) for the Bloch operators ξ\mathcal{L}_{\xi} for |ξ|1|\xi|\ll 1. This result for the KdV/KS system (1.1) was established in [3, Section 3.1]. See also [10] for a more general version of the same result.

Lemma 2.1 (Spectral Preparation).

Suppose that u¯\bar{u} is a TT-periodic traveling wave solution of (1.1) which is diffusively spectrally stable and, additionally, satisfies Assumption 1.5. Then the following properties hold.

  • (i)

    For any fixed ξ0(0,π/T)\xi_{0}\in(0,\pi/T) there exists a constant δ0>0\delta_{0}>0 such that

    (σ(ξ))<δ0\Re\left(\sigma\left(\mathcal{L}_{\xi}\right)\right)<-\delta_{0}

    for all ξ[π/T,π/T)\xi\in[-\pi/T,\pi/T) with |ξ|>ξ0|\xi|>\xi_{0}.

  • (ii)

    There exist positive constants ξ1\xi_{1} and δ1\delta_{1} such that for all |ξ|<ξ1|\xi|<\xi_{1}, the spectrum of ξ\mathcal{L}_{\xi} decomposes into two disjoint subsets

    σ(ξ)=σ(ξ)σ0(ξ)\sigma\left(\mathcal{L}_{\xi}\right)=\sigma_{-}\left(\mathcal{L}_{\xi}\right)\bigcup\sigma_{0}\left(\mathcal{L}_{\xi}\right)

    with the following properties:

    • (a)

      (σ(ξ))<δ1\Re\left(\sigma_{-}\left(\mathcal{L}_{\xi}\right)\right)<-\delta_{1} and (σc(ξ))>δ1\Re\left(\sigma_{c}\left(\mathcal{L}_{\xi}\right)\right)>-\delta_{1}.

    • (b)

      The set σ0(ξ)\sigma_{0}\left(\mathcal{L}_{\xi}\right) consists of precisely two eigenvalues, which are analytic in ξ\xi and expand as

      λj(ξ)=iξajdjξ2+𝒪(ξ3)j=1,2\lambda_{j}(\xi)=-i\xi a_{j}-d_{j}\xi^{2}+\mathcal{O}(\xi^{3})~~~j=1,2

      for |ξ|1|\xi|\ll 1 and some constants aja_{j}\in{\mathbb{R}} (distinct) and dj>0d_{j}>0.

    • (c)

      The left and right TT-periodic eigenfunctions ϕj(ξ)\phi_{j}(\xi) and ϕ~(ξ)\tilde{\phi}(\xi) of ξ\mathcal{L}_{\xi} associated with λj(ξ)\lambda_{j}(\xi) above, normalized so that

      ϕ~j(ξ),ϕk(ξ)L2(0,1)=iξδkj,1j,k2,\left<\tilde{\phi}_{j}(\xi),\phi_{k}(\xi)\right>_{L^{2}(0,1)}=i\xi\delta^{j}_{k},~~1\leq j,k\leq 2,

      are given as

      ϕj(ξ)\displaystyle\phi_{j}(\xi) =(iξ)β1(j)(ξ)q1(ξ)+β2(j)(ξ)q2(ξ)\displaystyle=(i\xi)\beta_{1}^{(j)}(\xi)q_{1}(\xi)+\beta_{2}^{(j)}(\xi)q_{2}(\xi)
      ϕ~j(ξ)\displaystyle\tilde{\phi}_{j}(\xi) =β~1(j)(ξ)q~1(ξ)+(iξ)β~2(j)(ξ)q~2(ξ)\displaystyle=\tilde{\beta}_{1}^{(j)}(\xi)\tilde{q}_{1}(\xi)+(i\xi)\tilde{\beta}_{2}^{(j)}(\xi)\tilde{q}_{2}(\xi)

      where the functions qj,q~j:[ξ0,ξ0]Lper2(0,T)q_{j},\tilde{q}_{j}:[-\xi_{0},\xi_{0}]\to L^{2}_{\rm per}(0,T) are analytic functions such that {qj(ξ,)}j=1,2\{q_{j}(\xi,\cdot)\}_{j=1,2} and {q~j(ξ,)}j=1,2\{\tilde{q}_{j}(\xi,\cdot)\}_{j=1,2} are dual bases of the total eigenspace of ξ\mathcal{L}_{\xi} associated with the spectrum σ0(ξ)\sigma_{0}(\mathcal{L}_{\xi}) chosen to satisfy

      (q1(0,)q2(0,))=(1u¯)and(q~1(0,)q~2(0,))=(1ψadj)\left(\begin{array}[]{c}q_{1}(0,\cdot)\\ q_{2}(0,\cdot)\end{array}\right)=\left(\begin{array}[]{c}1\\ \bar{u}^{\prime}\end{array}\right)~~{\rm and}~~\left(\begin{array}[]{c}\tilde{q}_{1}(0,\cdot)\\ \tilde{q}_{2}(0,\cdot)\end{array}\right)=\left(\begin{array}[]{c}1\\ \psi^{\rm adj}\end{array}\right)

      where ψadjLper2(0,T)\psi^{adj}\in L^{2}_{\rm per}(0,T) is a generalized left eigenfunction satisfying ψadj=1\mathcal{L}^{\dagger}\psi^{adj}=1 as well as

      ψadj,1L12=0andψadj,u¯L12=1,\left<\psi^{adj},1\right>_{L^{2}_{1}}=0{~~\rm and}~~\left<\psi^{adj},\bar{u}^{\prime}\right>_{L^{2}_{1}}=1,

      and where the βj,β~j:[ξ0,ξ0]\beta_{j},\tilde{\beta}_{j}:[-\xi_{0},\xi_{0}]\to{\mathbb{C}} are analytic functions.

Noting again that the Bloch operators ξ\mathcal{L}_{\xi} are sectorial on Lper2(0,T)L^{2}_{\rm per}(0,T), the spectral results in Lemma 2.1 immediately imply the following elementary estimates for the Bloch semigroups eξte^{\mathcal{L}_{\xi}t}.

Proposition 2.2.

Suppose that u¯\bar{u} is a TT-periodic traveling wave solution of (1.1) which satisfies the hypothesis of Lemma 2.1. Then the following properties hold.

  • (i)

    For any fixed ξ0(0,π/T)\xi_{0}\in(0,\pi/T), there exist positive constants C0C_{0} and d0d_{0} such that

    eξtfB(Lper2(0,T))C0ed0t\left\|e^{\mathcal{L}_{\xi}t}f\right\|_{B\left(L^{2}_{\rm per}(0,T)\right)}\leq C_{0}e^{-d_{0}t}

    is valid for all t0t\geq 0 and all ξ[π/T,π/T)\xi\in[-\pi/T,\pi/T) with |ξ|>ξ0|\xi|>\xi_{0}.

  • (ii)

    With ξ1\xi_{1} chosen as in Lemma 2.1(ii), there exist positive constants C1C_{1} and d1d_{1} such that for any |ξ|<ξ1|\xi|<\xi_{1}, if Π(ξ)\Pi(\xi) denotes the (rank-two) spectral projection onto the generalized eigenspaces associated to the critical eigenvalues {λj(ξ)}j=1,2\{\lambda_{j}(\xi)\}_{j=1,2}, then

    eξt(1Π(ξ))B(Lper2(0,T))C1ed1t\left\|e^{\mathcal{L}_{\xi}t}\left(1-\Pi(\xi)\right)\right\|_{B\left(L^{2}_{\rm per}(0,T)\right)}\leq C_{1}e^{-d_{1}t}

    is valid for all t0t\geq 0.

Note that Proposition 2.2(ii) is a natural extension to |ξ|1|\xi|\ll 1 of the linear bound (1.6), established for ξ=0\xi=0 which was key in establishing the fixed-NN subharmonic stability result Proposition 1.6.

3 Subharmonic Linear Estimates

The goal of this section is to obtain decay estimates on the semigroup ete^{\mathcal{L}t} acting on LN2L^{2}_{N} which are uniform in NN. To this end, we use (2.4) to study the action of ete^{\mathcal{L}t} on LN2L^{2}_{N} in terms of the associated Bloch operators. Naturally, by Lemma 2.1 we expect for each fixed NN\in{\mathbb{N}} that the long-time behavior of ete^{\mathcal{L}t} is dominated by the generalized kernel of \mathcal{L}. As discussed in the introduction, the main difficulty in obtaining such a result while maintaining uniformity in NN is the accumulation of NTNT-periodic eigenvalues near the origin as NN\to\infty. This is overcome through a delicate decomposition of ete^{\mathcal{L}t}, separating the action into appropriate critical (i.e., corresponding to spectrum accumulating near the origin) and non-critical frequency components. This methodology is heavily influenced by the corresponding decomposition used in the case of localized perturbations of periodic waves: see, for example, [10]. Note, further, that while our basic strategy is similar to the recent subharmonic analysis of reaction diffusion equations conducted in [12], the new challenge here is the fact that the linearized operator \mathcal{L} now has two spectral curves which pass through the origin, a reflection of the non-trivial Jordan structure in (1.4).

To begin, let ξ1(0,π/T)\xi_{1}\in(0,\pi/T) be defined as in Lemma 2.1 and let ρ\rho be a smooth cutoff function satisfying ρ(ξ)=1\rho(\xi)=1 for |ξ|<ξ1/2|\xi|<\xi_{1}/2 and ρ(ξ)=0\rho(\xi)=0 for |ξ|>ξ1|\xi|>\xi_{1}. Given a function vLN2v\in L^{2}_{N}, we then use (2.4) to decompose the action of ete^{\mathcal{L}t} on vv into high-frequency and low-frequency components via

(3.1) {eLtv(x)=1NTξΩNρ(ξ)eiξxeLξtT(v)(ξ,x)+1NTξΩN(1ρ(ξ))eiξxeLξtT(v)(ξ,x)=:Slf,N(t)v(x)+Shf,N(t)v(x).\left\{\begin{aligned} e^{Lt}v(x)&=\frac{1}{NT}\sum_{\xi\in\Omega_{N}}\rho(\xi)e^{i\xi x}e^{L_{\xi}t}\mathcal{B}_{T}(v)(\xi,x)+\frac{1}{NT}\sum_{\xi\in\Omega_{N}}(1-\rho(\xi))e^{i\xi x}e^{L_{\xi}t}\mathcal{B}_{T}(v)(\xi,x)\\ &=:S_{lf,N}(t)v(x)+S_{hf,N}(t)v(x).\end{aligned}\right.

Using Proposition 2.2(i), it follows that there exist constants C,η>0C,\eta>0, both independent of NN, such that

maxξΩN(1ρ(ξ))eLξtB(L12)Ceηt\max_{\xi\in\Omega_{N}}(1-\rho(\xi))\|e^{L_{\xi}t}\|_{B\left(L^{2}_{1}\right)}\leq Ce^{-\eta t}

which, by the subharmonic Parseval identity (2.2), implies the exponential estimate

Shf,N(t)vLN22\displaystyle\|S_{hf,N}(t)v\|_{L^{2}_{N}}^{2} =1NT2ξΩN(1ρ(ξ))eLξtT(v)(ξ,)Lx2(0,1)2\displaystyle=\frac{1}{NT^{2}}\sum_{\xi\in\Omega_{N}}\left\|(1-\rho(\xi))e^{L_{\xi}t}\mathcal{B}_{T}(v)(\xi,\cdot)\right\|_{L^{2}_{x}(0,1)}^{2}
1NT2ξΩN(1ρ(ξ))2eLξtB(L12)2T(v)(ξ,)Lx2(0,1)2\displaystyle\leq\frac{1}{NT^{2}}\sum_{\xi\in\Omega_{N}}(1-\rho(\xi))^{2}\|e^{L_{\xi}t}\|_{B\left(L^{2}_{1}\right)}^{2}\left\|\mathcal{B}_{T}(v)(\xi,\cdot)\right\|_{L^{2}_{x}(0,1)}^{2}
Ce2ηt(1NT2ξΩNT(v)(ξ,)Lx2(0,1)2)\displaystyle\leq Ce^{-2\eta t}\left(\frac{1}{NT^{2}}\sum_{\xi\in\Omega_{N}}\left\|\mathcal{B}_{T}(v)(\xi,\cdot)\right\|_{L^{2}_{x}(0,1)}^{2}\right)
=Ce2ηtvLN22,\displaystyle=Ce^{-2\eta t}\|v\|_{L^{2}_{N}}^{2},

on the high-frequency component of the solution operator.

For the low-frequency component, define for each |ξ|<ξ0|\xi|<\xi_{0} the rank-two spectral projection onto the critical modes of ξ\mathcal{L}_{\xi} by

{Π(ξ):L12j=12ker(Lξλj(ξ)I)Π(ξ)f=j=12qj(ξ)q~j(ξ,),fL2(0,1).\left\{\begin{aligned} &\Pi(\xi):L^{2}_{1}\to\bigoplus_{j=1}^{2}{\rm ker}\left(L_{\xi}-\lambda_{j}(\xi)I\right)\\ &\Pi(\xi)f=\sum_{j=1}^{2}q_{j}(\xi)\left<\tilde{q}_{j}(\xi,\cdot),f\right>_{L^{2}(0,1)}.\end{aligned}\right.

Note we also have the alternate representation formula

Π(ξ)f=1iξj=12ϕj(ξ)ϕ~j(ξ,),fL2(0,1),\Pi(\xi)f=\frac{1}{i\xi}\sum_{j=1}^{2}\phi_{j}(\xi)\left<\tilde{\phi}_{j}(\xi,\cdot),f\right>_{L^{2}(0,1)},

which more explicitly demonstrates the singularity as ξ0\xi\to 0 of the eigenprojection. Using this, the low-frequency operator Slf,N(t)S_{lf,N}(t) can be further decomposed into the contribution from the critical modes near (λ,ξ)=(0,0)(\lambda,\xi)=(0,0) and the contribution from the low-frequency spectrum bounded away from ξ=0\xi=0 via

(3.2) {Slf,N(t)v(x)=1NTξΩNρ(ξ)eiξxeLξtΠ(ξ)T(v)(ξ,x)+1NTξΩNρ(ξ)eiξxeLξt(1Π(ξ))T(v)(ξ,x)=:Sc,N(t)v(x)+S~lf,N(t)v(x).\left\{\begin{aligned} S_{lf,N}(t)v(x)&=\frac{1}{NT}\sum_{\xi\in\Omega_{N}}\rho(\xi)e^{i\xi x}e^{L_{\xi}t}\Pi(\xi)\mathcal{B}_{T}(v)(\xi,x)\\ &\quad\qquad+\frac{1}{NT}\sum_{\xi\in\Omega_{N}}\rho(\xi)e^{i\xi x}e^{L_{\xi}t}\left(1-\Pi(\xi)\right)\mathcal{B}_{T}(v)(\xi,x)\\ &=:S_{c,N}(t)v(x)+\widetilde{S}_{lf,N}(t)v(x).\end{aligned}\right.

As above, by possibly choosing η>0\eta>0 smaller, Proposition 2.2(ii) implies that there exists a constant C>0C>0 such that for each NN\in{\mathbb{N}} we have

eLξt(1Π(ξ))fL12CeηtfL12\left\|e^{L_{\xi}t}(1-\Pi(\xi))f\right\|_{L^{2}_{1}}\leq Ce^{-\eta t}\|f\|_{L^{2}_{1}}

for each fL12f\in L^{2}_{1}, and hence another application of Parseval’s identity (2.2) yields

S~lf,N(t)vLN2CeηtvLN2.\left\|\widetilde{S}_{lf,N}(t)v\right\|_{L^{2}_{N}}\leq Ce^{-\eta t}\|v\|_{L^{2}_{N}}.

For the critical component, we decompose Sc,N(t)S_{c,N}(t) further as

Sc,N(t)v(x)\displaystyle S_{c,N}(t)v(x) =1NTeL0tΠ(0)T(v)(0,x)\displaystyle=\frac{1}{NT}e^{L_{0}t}\Pi(0)\mathcal{B}_{T}(v)(0,x)
+1NTξΩN{0}ρ(ξ)eiξxj=121iξeλj(ξ)tϕj(ξ,x)ϕ~j(ξ,),T(v)(ξ,)L12\displaystyle\qquad+\frac{1}{NT}\sum_{\xi\in\Omega_{N}\setminus\{0\}}\rho(\xi)e^{i\xi x}\sum_{j=1}^{2}\frac{1}{i\xi}e^{\lambda_{j}(\xi)t}\phi_{j}(\xi,x)\left<\tilde{\phi}_{j}(\xi,\cdot),\mathcal{B}_{T}(v)(\xi,\cdot)\right>_{L^{2}_{1}}

For the non-zero frequencies, noting that Lemma 2.1 implies that the eigenfunctions {ϕj(ξ)}j=12\left\{\phi_{j}(\xi)\right\}_{j=1}^{2} of LξL_{\xi} expand as

ϕj(ξ,x)\displaystyle\phi_{j}(\xi,x) =β2(j)(ξ)[u¯(x)+(q2(ξ,x)u¯(x))]+(iξ)β1(j)(ξ)q1(ξ,x)\displaystyle=\beta_{2}^{(j)}(\xi)\left[\bar{u}^{\prime}(x)+\left(q_{2}(\xi,x)-\bar{u}^{\prime}(x)\right)\right]+(i\xi)\beta_{1}^{(j)}(\xi)q_{1}(\xi,x)
=β2(j)(ξ)u¯(x)+𝒪(ξ),\displaystyle=\beta_{2}^{(j)}(\xi)\bar{u}^{\prime}(x)+\mathcal{O}(\xi),

for 0<|ξ|10<|\xi|\ll 1 suggests the decomposition

(3.3) {Sc,N(t)v(x)=1NTeL0tΠ(0)T(v)(0,x)+u¯(x)(1NTξΩN{0}ρ(ξ)eiξxj=121iξeλj(ξ)tβ2(j)(ξ)ϕ~j(ξ,),T(v)(ξ,)L12)+1NTξΩN{0}ρ(ξ)eiξxj=12eλj(ξ)tϕj(ξ,x)β2(j)(ξ)u¯(x)iξϕ~j(ξ,),T(v)(ξ,)L12=:1NTeL0tΠ(0)T(v)(0,x)+u¯(x)sp,N(t)v(x)+S~c,N(t)v(x).\left\{\begin{aligned} S_{c,N}(t)v(x)&=\frac{1}{NT}e^{L_{0}t}\Pi(0)\mathcal{B}_{T}(v)(0,x)\\ &\quad+\bar{u}^{\prime}(x)\left(\frac{1}{NT}\sum_{\xi\in\Omega_{N}\setminus\{0\}}\rho(\xi)e^{i\xi x}\sum_{j=1}^{2}\frac{1}{i\xi}e^{\lambda_{j}(\xi)t}\beta_{2}^{(j)}(\xi)\left<\tilde{\phi}_{j}(\xi,\cdot),\mathcal{B}_{T}(v)(\xi,\cdot)\right>_{L^{2}_{1}}\right)\\ &\qquad+\frac{1}{NT}\sum_{\xi\in\Omega_{N}\setminus\{0\}}\rho(\xi)e^{i\xi x}\sum_{j=1}^{2}e^{\lambda_{j}(\xi)t}\frac{\phi_{j}(\xi,x)-\beta_{2}^{(j)}(\xi)\bar{u}^{\prime}(x)}{i\xi}\left<\tilde{\phi}_{j}(\xi,\cdot),\mathcal{B}_{T}(v)(\xi,\cdot)\right>_{L^{2}_{1}}\\ &=:\frac{1}{NT}e^{L_{0}t}\Pi(0)\mathcal{B}_{T}(v)(0,x)+\bar{u}^{\prime}(x)s_{p,N}(t)v(x)+\widetilde{S}_{c,N}(t)v(x).\end{aligned}\right.

Additionally, note the ξ=0\xi=0 term above can be expressed as

e0tΠ(0)T(v)(0,x)=e0t1,B1(v)(0,)L12+u¯(x)ψadj,B1(v)(0,)L12.e^{\mathcal{L}_{0}t}\Pi(0)\mathcal{B}_{T}(v)(0,x)=e^{\mathcal{L}_{0}t}\left<1,B_{1}(v)(0,\cdot)\right>_{L^{2}_{1}}+\bar{u}^{\prime}(x)\left<\psi^{adj},B_{1}(v)(0,\cdot)\right>_{L^{2}_{1}}.

Observing that333Indeed, observe that ddt(1tu¯)=u¯=0(1)=0(1tu¯)\frac{d}{dt}\left(1-t\bar{u}^{\prime}\right)=-\bar{u}^{\prime}=\mathcal{L}_{0}(1)=\mathcal{L}_{0}(1-t\bar{u}^{\prime}) and clearly the associated initial condition is satisfied.

e0t1=1tu¯e^{\mathcal{L}_{0}t}1=1-t\bar{u}^{\prime}

and noting that (2.3) implies the identities

0TT(v)(0,x)𝑑x=T0NTv(x)𝑑xandψadj,T(v)(0,)L12=Tψadj,vLN2\int_{0}^{T}\mathcal{B}_{T}(v)(0,x)dx=T\int_{0}^{NT}v(x)dx~~{\rm and}~~\left<\psi^{adj},\mathcal{B}_{T}(v)(0,\cdot)\right>_{L^{2}_{1}}=T\left<\psi^{adj},v\right>_{L^{2}_{N}}

we find the representation

1NTe0tΠ(0)T(v)(0,x)\displaystyle\frac{1}{NT}e^{\mathcal{L}_{0}t}\Pi(0)\mathcal{B}_{T}(v)(0,x) =1N0NTv(z)𝑑z\displaystyle=\frac{1}{N}\int_{0}^{NT}v(z)dz
+1Nu¯(x)(ψadj,vLN2t0NTv(z)𝑑z).\displaystyle\quad+\frac{1}{N}\bar{u}^{\prime}(x)\left(\left<\psi^{adj},v\right>_{L^{2}_{N}}-t\int_{0}^{NT}v(z)dz\right).

Observe that the above representation of the ξ=0\xi=0 contribution of the solution operator clearly demonstrates the expected linear instability444This is naturally expected due to the Jordan block at the λ=0\lambda=0 eigenvalue of 0\mathcal{L}_{0}. of the underlying wave u¯\bar{u}. In the forthcoming analysis nonlinear analysis, this linear instability will be compensated by allowing for Galilean boosts of the underlying wave.

Taken together, it follows that the linear solution operator ete^{\mathcal{L}t} acting on LN2L^{2}_{N} can be decomposed as

(3.4) {etv(x)=u¯(x)(1Nψadj,vLN2tN0NTv(z)𝑑z+sp,N(t)v(x))+1N0NTv(z)𝑑z+S~N(t)v(x),\left\{\begin{aligned} e^{\mathcal{L}t}v(x)&=\bar{u}^{\prime}(x)\left(\frac{1}{N}\left<\psi^{adj},v\right>_{L^{2}_{N}}-\frac{t}{N}\int_{0}^{NT}v(z)dz+s_{p,N}(t)v(x)\right)\\ &\qquad+\frac{1}{N}\int_{0}^{NT}v(z)dz+\widetilde{S}_{N}(t)v(x),\end{aligned}\right.

where here

S~N(t):=Shf,N(t)+S~lf,N(t)+S~c,N(t)\widetilde{S}_{N}(t):=S_{hf,N}(t)+\widetilde{S}_{lf,N}(t)+\widetilde{S}_{c,N}(t)

and the operators Shf,N(t)S_{hf,N}(t) and S~lf,N(t)\widetilde{S}_{lf,N}(t) are defined in (3.1) and (3.2), respectively, and S~c,N(t)\widetilde{S}_{c,N}(t) and sp,N(t)s_{p,N}(t) are defined in (3.3). With this decomposition in hand, we now establish temporal estimates on the above which are uniform in NN\in{\mathbb{N}}.

Proposition 3.1 (Linear Estimates).

Suppose that u¯\bar{u} is a TT-periodic diffusively spectrally stable traveling wave solution of (1.1) with wave speed cc. Given any MM\in{\mathbb{N}} there exists a constant C>0C>0 such that for all NN\in{\mathbb{N}}, t0t\geq 0 and all 1,m,rM1\leq\ell,m,r\leq M we have

(3.5) xtmsp,N(t)vLN2C(1+t)1/4(+m)/2vLN2\left\|\partial_{x}^{\ell}\partial_{t}^{m}s_{p,N}(t)v\right\|_{L^{2}_{N}}\leq C(1+t)^{1/4-(\ell+m)/2}\|v\|_{L^{2}_{N}}

and

(3.6) xtmsp,N(t)xrvLN2C(1+t)1/4(+m)/2vLN2.\left\|\partial_{x}^{\ell}\partial_{t}^{m}s_{p,N}(t)\partial_{x}^{r}v\right\|_{L^{2}_{N}}\leq C(1+t)^{-1/4-(\ell+m)/2}\|v\|_{L^{2}_{N}}.

Further, there exist constants C,η>0C,\eta>0 such that for all t0t\geq 0 NN\in{\mathbb{N}}, and all 0,m,r1M0\leq\ell,m,r-1\leq M we have

(3.7) xtmS~N(t)vLN2C((1+t)1/4vLN1+eηtvHN+4m)\left\|\partial_{x}^{\ell}\partial_{t}^{m}\widetilde{S}_{N}(t)v\right\|_{L^{2}_{N}}\leq C\left((1+t)^{-1/4}\|v\|_{L^{1}_{N}}+e^{-\eta t}\|v\|_{H^{\ell+4m}_{N}}\right)

and

(3.8) xtmS~N(t)xrvLN2C((1+t)3/4vLN1+eηtvHNr++4m).\left\|\partial_{x}^{\ell}\partial_{t}^{m}\widetilde{S}_{N}(t)\partial_{x}^{r}v\right\|_{L^{2}_{N}}\leq C\left((1+t)^{-3/4}\|v\|_{L^{1}_{N}}+e^{-\eta t}\|v\|_{H^{r+\ell+4m}_{N}}\right).

Finally, there exists a constant C>0C>0, independent of NN, such that for all t0t\geq 0 we have

(3.9) sp,N(t)vLNCvLN1LN2andsp,N(t)xvLNC(1+t)1/2vLN1LN2.\left\|s_{p,N}(t)v\right\|_{L^{\infty}_{N}}\leq C\|v\|_{L^{1}_{N}\cap L^{2}_{N}}~~{\rm and}~~\left\|s_{p,N}(t)\partial_{x}v\right\|_{L^{\infty}_{N}}\leq C(1+t)^{-1/2}\|v\|_{L^{1}_{N}\cap L^{2}_{N}}.
Proof.

We begin by establishing the LN2L^{2}_{N} estimates above. To this end, first note that the definition of T\mathcal{B}_{T} implies that

ϕ~j(ξ,),T(v)(ξ,)L12\displaystyle\left<\tilde{\phi}_{j}(\xi,\cdot),\mathcal{B}_{T}(v)(\xi,\cdot)\right>_{L^{2}_{1}} =0Tϕ~j(ξ,x)¯e2πix/Tv^(ξ+2π/T)\displaystyle=\int_{0}^{T}\overline{\tilde{\phi}_{j}(\xi,x)}\sum_{\ell\in{\mathbb{Z}}}e^{2\pi i\ell x/T}\hat{v}(\xi+2\pi\ell/T)
=v^(ξ+2π/T)0Te2πix/Tϕ~j(ξ,x)¯𝑑x\displaystyle=\sum_{\ell\in{\mathbb{Z}}}\hat{v}(\xi+2\pi\ell/T)\int_{0}^{T}e^{2\pi i\ell x/T}\overline{\tilde{\phi}_{j}(\xi,x)}dx
=v^(ξ+2π/T)ϕ~j^(ξ,2π/T)¯.\displaystyle=\sum_{\ell\in{\mathbb{Z}}}\hat{v}(\xi+2\pi\ell/T)\overline{\widehat{\tilde{\phi}_{j}}(\xi,2\pi\ell/T)}.

Since g^L()gLN1\|\hat{g}\|_{L^{\infty}({\mathbb{R}})}\leq\|g\|_{L^{1}_{N}} by (2.1), an application of Cauchy-Schwarz implies the existence of a constant C>0C>0, independent of NN, such that

ρ(ξ)|ϕ~j(ξ,),T(v)(ξ,)L12|2\displaystyle\rho(\xi)\left|\left<\tilde{\phi}_{j}(\xi,\cdot),\mathcal{B}_{T}(v)(\xi,\cdot)\right>_{L^{2}_{1}}\right|^{2} ρ(ξ)vLN12((1+||2)1/2|ϕ~j^(ξ,2π/T)¯|(1+||2)1/2)\displaystyle\leq\rho(\xi)\|v\|_{L^{1}_{N}}^{2}\left(\sum_{\ell\in{\mathbb{Z}}}(1+|\ell|^{2})^{1/2}\left|\overline{\widehat{\tilde{\phi}_{j}}(\xi,2\pi\ell/T)}\right|(1+|\ell|^{2})^{-1/2}\right)
CvLN12supξ[π/T,π/T)(ρ(ξ)ϕ~j(ξ,)Hper1(0,T)2),\displaystyle\leq C\|v\|_{L^{1}_{N}}^{2}\sup_{\xi\in[-\pi/T,\pi/T)}\left(\rho(\xi)\|\tilde{\phi}_{j}(\xi,\cdot)\|_{H^{1}_{\rm per}(0,T)}^{2}\right),

valid for all ξΩN\xi\in\Omega_{N}. Using the subharmonic Parseval identity (2.2), along with Lemma 2.1, we find that

(3.10) xtmsp,N(t)vLN22\displaystyle\left\|\partial_{x}^{\ell}\partial_{t}^{m}s_{p,N}(t)v\right\|_{L^{2}_{N}}^{2}\leq 1NT2j=12ξΩN{0}ρ(ξ)(iξ)1(λj(ξ))meλj(ξ)tϕ~j(ξ,),T(v)(ξ,)L12L122\displaystyle\frac{1}{NT^{2}}\sum_{j=1}^{2}\sum_{\xi\in\Omega_{N}\setminus\{0\}}\left\|\rho(\xi)(i\xi)^{\ell-1}(\lambda_{j}(\xi))^{m}e^{\lambda_{j}(\xi)t}\left<\tilde{\phi}_{j}(\xi,\cdot),\mathcal{B}_{T}(v)(\xi,\cdot)\right>_{L^{2}_{1}}\right\|_{L^{2}_{1}}^{2}
CvLN12(1NTξΩN{0}|ξ|2(+m1)e2dξ2t),\displaystyle\leq C\|v\|_{L^{1}_{N}}^{2}\left(\frac{1}{NT}\sum_{\xi\in\Omega_{N}\setminus\{0\}}|\xi|^{2(\ell+m-1)}e^{2d\xi^{2}t}\right),

where here C,d>0C,d>0 are constants which are independent of NN. By Similar considerations, we find

(3.11) xtmS~N(t)vLN22Ce2ηtvLN22+CvLN12(1NTξΩN{0}|ξ|2(+m)e2dξ2t)\left\|\partial_{x}^{\ell}\partial_{t}^{m}\widetilde{S}_{N}(t)v\right\|_{L^{2}_{N}}^{2}\leq Ce^{-2\eta t}\|v\|_{L^{2}_{N}}^{2}+C\|v\|_{L^{1}_{N}}^{2}\left(\frac{1}{NT}\sum_{\xi\in\Omega_{N}\setminus\{0\}}|\xi|^{2(\ell+m)}e^{2d\xi^{2}t}\right)

where, again, the constants C,d>0C,d>0 are independent of NN.

Continuing, note that integration by parts yields the identity

T(xg)(ξ,x)=(x+iξ)T(g)(ξ,x).\mathcal{B}_{T}(\partial_{x}g)(\xi,x)=(\partial_{x}+i\xi)\mathcal{B}_{T}(g)(\xi,x).

It follows that for each 1j21\leq j\leq 2 and integer r1r\geq 1 we have

(3.12) ϕ~j(ξ,),T(xrg)(ξ,)L12=ϕ~j(ξ,),(x+iξ)rT(g)(ξ,)L12\displaystyle\left<\tilde{\phi}_{j}(\xi,\cdot),\mathcal{B}_{T}(\partial_{x}^{r}g)(\xi,\cdot)\right>_{L^{2}_{1}}=\left<\tilde{\phi}_{j}(\xi,\cdot),(\partial_{x}+i\xi)^{r}\mathcal{B}_{T}(g)(\xi,\cdot)\right>_{L^{2}_{1}}
=r=0r(rr)(iξ)rϕ~j(ξ,),xrrT(g)(ξ,)L12\displaystyle\qquad\qquad=\sum_{r^{\prime}=0}^{r}\left(\begin{array}[]{c}r\\ r^{\prime}\end{array}\right)(i\xi)^{r^{\prime}}\left<\tilde{\phi}_{j}(\xi,\cdot),\partial_{x}^{r-r^{\prime}}\mathcal{B}_{T}(g)(\xi,\cdot)\right>_{L^{2}_{1}}
=(1)rxrϕ~j(ξ,),T(g)(ξ,)L12+r=1r(rr)(iξ)rϕ~j(ξ,),xrrT(g)(ξ,)L12,\displaystyle\qquad\qquad=(-1)^{r}\left<\partial_{x}^{r}\tilde{\phi}_{j}(\xi,\cdot),\mathcal{B}_{T}(g)(\xi,\cdot)\right>_{L^{2}_{1}}+\sum_{r^{\prime}=1}^{r}\left(\begin{array}[]{c}r\\ r^{\prime}\end{array}\right)(i\xi)^{r^{\prime}}\left<\tilde{\phi}_{j}(\xi,\cdot),\partial_{x}^{r-r^{\prime}}\mathcal{B}_{T}(g)(\xi,\cdot)\right>_{L^{2}_{1}},

where the last equality follows by integrating by parts in the r=0r^{\prime}=0 term. Since Lemma 2.1 implies that ϕ~j(0,x)\tilde{\phi}_{j}(0,x) is a constant vector for each ξ\xi, we have

xϕ~j(ξ,x)=ξx(ϕ~j(ξ,x)ϕ~j(0,x)ξ),\partial_{x}\tilde{\phi}_{j}(\xi,x)=\xi\partial_{x}\left(\frac{\tilde{\phi}_{j}(\xi,x)-\tilde{\phi}_{j}(0,x)}{\xi}\right),

which is clearly 𝒪(ξ)\mathcal{O}(\xi) by the analytic dependence of ϕ~j(ξ,x)\tilde{\phi}_{j}(\xi,x) on ξ\xi. Taken together, it follows that for each r1r\geq 1 we have

(3.13) xtmsp,N(t)xrvLN22\displaystyle\left\|\partial_{x}^{\ell}\partial_{t}^{m}s_{p,N}(t)\partial_{x}^{r}v\right\|_{L^{2}_{N}}^{2}\leq CvLN12(1NTξΩN{0}|ξ|2(+m)e2dξ2t)\displaystyle C\|v\|_{L^{1}_{N}}^{2}\left(\frac{1}{NT}\sum_{\xi\in\Omega_{N}\setminus\{0\}}|\xi|^{2(\ell+m)}e^{2d\xi^{2}t}\right)

and, similarly,

(3.14) xtmS~N(t)xrvLN22Ce2ηtvHNr2+CvLN12(1NTξΩN{0}|ξ|2(+m+1)e2dξ2t)\left\|\partial_{x}^{\ell}\partial_{t}^{m}\widetilde{S}_{N}(t)\partial_{x}^{r}v\right\|_{L^{2}_{N}}^{2}\leq Ce^{-2\eta t}\|v\|_{H^{r}_{N}}^{2}+C\|v\|_{L^{1}_{N}}^{2}\left(\frac{1}{NT}\sum_{\xi\in\Omega_{N}\setminus\{0\}}|\xi|^{2(\ell+m+1)}e^{2d\xi^{2}t}\right)

To complete the proof of the LN2L^{2}_{N} bounds, it remains to provide bounds on the discrete, NN-dependent sums in (3.10)-(3.14). These uniform bounds come by directly applying Lemma A.1 in [12], which states that for any integer ω0\omega\geq 0, there exists a constant C>0C>0, independent of NN, such that555As motivation, observe that the sum 1NTξΩNξ2ωe2dξ2t\frac{1}{NT}\sum_{\xi\in\Omega_{N}}\xi^{2\omega}e^{-2d\xi^{2}t} can be considered as a Riemann sum approximation for the integral π/Tπ/Tξ2ωe2dξ2t𝑑ξ\int_{-\pi/T}^{\pi/T}\xi^{2\omega}e^{-2d\xi^{2}t}d\xi, which exhibits the stated temporal decay via a routine scaling argument.

1NξΩNξ2ωe2dξ2tC(1+t)ω1/2.\frac{1}{N}\sum_{\xi\in\Omega_{N}}\xi^{2\omega}e^{-2d\xi^{2}t}\leq C(1+t)^{-\omega-1/2}.

Applying this result to (3.10)-(3.14) with the appropriate values of ω\omega establishes the LN2L^{2}_{N} estimates stated in (3.5)-(3.8).

It remains to establish the LL^{\infty} estimates stated in (3.9). To this end, note that by similar estimates as above, the LL^{\infty} difference between sp,N(t)v(x)s_{p,N}(t)v(x) and the function

(3.15) j=121NTk{0}eiξkedjξk2tiξkβ2j(0)ϕ~j(0,),T(v)(ξk,)L12=:sp,N(1)(t)v(x)+sp,N(2)(t)v(x)\sum_{j=1}^{2}\frac{1}{NT}\sum_{k\in{\mathbb{Z}}\setminus\{0\}}e^{i\xi_{k}}\frac{e^{-d_{j}\xi_{k}^{2}t}}{i\xi_{k}}\beta_{2}^{j}(0)\left<\widetilde{\phi}_{j}(0,\cdot),\mathcal{B}_{T}(v)(\xi_{k},\cdot)\right>_{L^{2}_{1}}=:s^{(1)}_{p,N}(t)v(x)+s^{(2)}_{p,N}(t)v(x)

is controlled by (1+t)1/2vLN1(1+t)^{-1/2}\|v\|_{L^{1}_{N}}, where here ξk:=2πk/NT\xi_{k}:=2\pi k/NT for each kk\in{\mathbb{Z}}. Further, since for each j=1,2j=1,2 the function ϕ~j(0,x)\widetilde{\phi}_{j}(0,x) is actually a constant equal to ν~j:=(β~1(j)(0),β~2(j)(0))\tilde{\nu}_{j}:=\left(\tilde{\beta}_{1}^{(j)}(0),\tilde{\beta}_{2}^{(j)}(0)\right), we note that

ϕ~j(0,),T(v)(ξk,)L12=ν~jv^(ξk)\left<\widetilde{\phi}_{j}(0,\cdot),\mathcal{B}_{T}(v)(\xi_{k},\cdot)\right>_{L^{2}_{1}}=\tilde{\nu}_{j}\cdot\hat{v}(\xi_{k})

and hence, since (3.15) is in the form of a Fourier series itself, for each j=1,2j=1,2 the function sp,N(j)(t)v(x)s^{(j)}_{p,N}(t)v(x) can be recognized as the convolution of ν~jv\tilde{\nu}_{j}\cdot v with

(3.16) 1NTk{0}eiξkedjξk2tiξkβ2j(0).\frac{1}{NT}\sum_{k\in{\mathbb{Z}}\setminus\{0\}}e^{i\xi_{k}}\frac{e^{-d_{j}\xi_{k}^{2}t}}{i\xi_{k}}\beta_{2}^{j}(0).

Recognizing the above sums as TT-periodic Fourier series representations for Error functions, which are clearly bounded in LL^{\infty}, this establishes the first LL^{\infty} bound in (3.9). The second bound in (3.9) now follows by using precisely the same procedure as above, while noting that (3.12) implies the extra derivative on vv yields an extra 𝒪(ξ)\mathcal{O}(\xi) factor in the sum (3.16) which, in turn, yields the additional (1+t)1/2(1+t)^{-1/2} decay. ∎

Before continuing to our nonlinear analysis, we first provide an interpretation of the above decomposition of the linear solution operator. To this end, suppose u¯\bar{u} is a TT-periodic diffusively spectrally stable periodic traveling wave solution of (1.1), and suppose that u(x,t)u(x,t) is a solution (in same co-moving frame) with initial data u(x,0)=ϕ(x)+εv(x)u(x,0)=\phi(x)+\varepsilon v(x) with |ε|1|\varepsilon|\ll 1 and vLN1LN2v\in L^{1}_{N}\cap L^{2}_{N}. From the decomposition (3.4) and the linear estimates in Proposition 3.1, it is then natural to suspect that

u(x,t)\displaystyle u(x,t) u¯(x)+εetv(x)\displaystyle\approx\bar{u}(x)+\varepsilon e^{\mathcal{L}t}v(x)
u¯(x)+εu¯(x)(1Nψadj,vLN2+sp,N(t)v(x)tN0NTv(z)𝑑z)+εN0NTv(z)𝑑z\displaystyle\approx\bar{u}(x)+\varepsilon\bar{u}^{\prime}(x)\left(\frac{1}{N}\left<\psi^{adj},v\right>_{L^{2}_{N}}+s_{p,N}(t)v(x)-\frac{t}{N}\int_{0}^{NT}v(z)dz\right)+\frac{\varepsilon}{N}\int_{0}^{NT}v(z)dz
u¯(x+ε(1Nψadj,vLN2+sp,N(t)v(x))εtN0NTv(z)𝑑z)+εN0NTv(z)𝑑z,\displaystyle\approx\bar{u}\left(x+\varepsilon\left(\frac{1}{N}\left<\psi^{adj},v\right>_{L^{2}_{N}}+s_{p,N}(t)v(x)\right)-\frac{\varepsilon t}{N}\int_{0}^{NT}v(z)dz\right)+\frac{\varepsilon}{N}\int_{0}^{NT}v(z)dz,

which is a (small, via the LL^{\infty} bounds in (3.9)) spatio-temporal phase modulation of the background wave u¯\bar{u} together with an (identical) wave-speed and mass correction. In particular, recalling that solutions of the KdV/KS equation (1.1) are invariant under the Galilean transformation (1.3) the above suggests that the initially nearby wave u(x,t)u(x,t), up to spatio-temporal phase modulation, will asymptotically approach a member of the Galilean family associated to the background wave u¯\bar{u}. In the next section, we verify this intuition.

4 Uniform Nonlinear Asymptotic Stability

Our goal is to now use the linear decomposition (3.4) in order to establish the proof of Theorem 1.7. As discussed above, a small subharmonic perturbation of a TT-periodic, diffusively spectrally stable periodic traveling wave solution u¯\bar{u} of (1.1) will, for long time, behave like a coupled temporal mass modulation and spatio-temporal phase modulation of background wave u¯\bar{u}.

4.1 Nonlinear Decomposition, Perturbation Equations & Damping

Suppose that u¯\bar{u} is a TT-periodic diffusively spectrally stable periodic traveling wave solution of (1.1). Motivated by the decomposition (3.4), we develop a decomposition of nonlinear, subharmonic perturbations of the background wave u¯\bar{u} which accounts for the phase and mass modulations predicted by the linear theory.

To this end, suppose that u(x,t){u}(x,t) is a solution (in the same co-moving frame) with initial data u(x,0)LN1LN2{u}(x,0)\in L^{1}_{N}\cap L^{2}_{N} which is close (in LN2L^{2}_{N}) to u¯\bar{u}. By conservation of mass, we note that

0NTu(x,t)𝑑x=0NTu(x,0)𝑑x\int_{0}^{NT}{u}(x,t)dx=\int_{0}^{NT}{u}(x,0)dx

for all t0t\geq 0 for which it is defined. In particular, unless the initial data u(x,0){u}(x,0) has the same mass as the underlying wave u¯\bar{u}, we should not expect asymptotic convergence of u(x,t){u}(x,t) to u¯\bar{u}. Recalling, however, that solutions of (1.1) obey the Galilean invariance (1.3), it is natural to suspect that the added mass from the initial perturbation may induce a Galilean boost of the background wave u¯\bar{u}. Specifically, given such a u(x,0)LN1LN2{u}(x,0)\in L^{1}_{N}\cap L^{2}_{N} and defining

δM:=1NT0NT(u(z,0)u¯(z))𝑑z,\delta M:=\frac{1}{NT}\int_{0}^{NT}\left({u}(z,0)-\bar{u}(z)\right)dz,

we expect that the associated nearby solution u(x,t){u}(x,t) should satisfy

u(x,t)u¯(x(δM)t)+δM{u}(x,t)\approx\bar{u}\left(x-\left(\delta M\right)t\right)+\delta M

for t1t\gg 1. Note that for each fixed NN this is exactly the long-time dynamics predicted by Proposition 1.6 As discussed in the introduction, however, the result of Proposition 1.6 lacks uniformity with respect to the period of the perturbations in both the rate of decay of perturbations and the allowable size of initial perturbations. In this section, we use the linear results in Section 3 above, which were designed specifically to be uniform in the perturbation’s period, to establish our main result Theorem 1.7.

To this end, let u(x,t){u}(x,t) be a solution (in the same co-moving frame as u¯\bar{u}) with initial data u(x,0)LN1LN2{u}(x,0)\in L^{1}_{N}\cap L^{2}_{N} which is close (in LN2L^{2}_{N}) to u¯\bar{u} and consider a nonlinear perturbation of the form

(4.1) v(x,t):=u(x+(δM)t1Nγ(t)ψ(x,t),t)δMu¯(x)v(x,t):={u}\left(x+\left(\delta M\right)t-\frac{1}{N}\gamma(t)-\psi(x,t),t\right)-\delta M-\bar{u}(x)

where γ:+\gamma:{\mathbb{R}}_{+}\to{\mathbb{R}} and ψ:×+\psi:{\mathbb{R}}\times{\mathbb{R}}_{+}\to{\mathbb{R}} are functions to be determined later666Note though that we assume ψ(,t)LN2\psi(\cdot,t)\in L^{2}_{N} for each t0t\geq 0.. Note, in particular, that integrating (4.1) over [0,NT][0,NT] gives

0NTu(z,t)𝑑z=0NTu¯(z)𝑑z+NT(δM)+0NTv(z,t)𝑑z\int_{0}^{NT}{u}(z,t)dz=\int_{0}^{NT}\bar{u}(z)dz+NT\left(\delta M\right)+\int_{0}^{NT}v(z,t)dz

which, since the integral of u(,t){u}(\cdot,t) over [0,NT][0,NT] is conserved by the flow of (1.1), implies that

(4.2) 0NTv(z,t)𝑑z=0\int_{0}^{NT}v(z,t)dz=0

for all t0t\geq 0 for which it is defined by the choice of δM\delta M above.

With the above decomposition in hand, we next derive equations that must be satisfied by the perturbation vv and the modulation functions γ\gamma and ψ\psi.

Proposition 4.1.

The triple (v,γ,ψ)(v,\gamma,\psi) satisfies

(4.3) (t)((1ψx)v+1Nγu¯+ψu¯)=x𝒩,(\partial_{t}-\mathcal{L})\left((1-\psi_{x})v+\frac{1}{N}\gamma\bar{u}^{\prime}+\psi\bar{u}^{\prime}\right)=\partial_{x}\mathcal{N},

where

x𝒩=x𝒬+x+(ψxv),𝒬=12v2,\partial_{x}\mathcal{N}=\partial_{x}\mathcal{Q}+\partial_{x}\mathcal{R}+\mathcal{L}(\psi_{x}v),\quad\mathcal{Q}=-\frac{1}{2}v^{2},

and

\displaystyle\mathcal{R} =ψtv1Nγvε[x(ψx1ψxvx)+ψx1ψxx(11ψxvx)]δ(ψx1ψxvx)\displaystyle=-\psi_{t}v-\frac{1}{N}\gamma^{\prime}v-\varepsilon\left[\partial_{x}\left(\frac{\psi_{x}}{1-\psi_{x}}v_{x}\right)+\frac{\psi_{x}}{1-\psi_{x}}\partial_{x}\left(\frac{1}{1-\psi_{x}}v_{x}\right)\right]-\delta\left(\frac{\psi_{x}}{1-\psi_{x}}v_{x}\right)
δ[x2(ψx1ψxvx)+x(ψx1ψxx(11ψxvx))+ψx1ψxx(11ψxx(11ψxvx))]\displaystyle\qquad-\delta\left[\partial_{x}^{2}\left(\frac{\psi_{x}}{1-\psi_{x}}v_{x}\right)+\partial_{x}\left(\frac{\psi_{x}}{1-\psi_{x}}\partial_{x}\left(\frac{1}{1-\psi_{x}}v_{x}\right)\right)+\frac{\psi_{x}}{1-\psi_{x}}\partial_{x}\left(\frac{1}{1-\psi_{x}}\partial_{x}\left(\frac{1}{1-\psi_{x}}v_{x}\right)\right)\right]
ε[ψx1ψxx(ψx1ψxu¯)+x(ψx21ψxu¯)+ψx21ψxu¯′′]δ(ψx21ψxu¯)\displaystyle\qquad-\varepsilon\left[\frac{\psi_{x}}{1-\psi_{x}}\partial_{x}\left(\frac{\psi_{x}}{1-\psi_{x}}\bar{u}^{\prime}\right)+\partial_{x}\left(\frac{\psi_{x}^{2}}{1-\psi_{x}}\bar{u}^{\prime}\right)+\frac{\psi_{x}^{2}}{1-\psi_{x}}\bar{u}^{\prime\prime}\right]-\delta\left(\frac{\psi_{x}^{2}}{1-\psi_{x}}\bar{u}^{\prime}\right)
δ[11ψxx(ψx1ψxx(ψx1ψxu¯))+ψx1ψxx2(ψx1ψxu¯)+ψx1ψxx(ψx1ψxu¯′′)\displaystyle\qquad-\delta\left[\frac{1}{1-\psi_{x}}\partial_{x}\left(\frac{\psi_{x}}{1-\psi_{x}}\partial_{x}\left(\frac{\psi_{x}}{1-\psi_{x}}\bar{u}^{\prime}\right)\right)+\frac{\psi_{x}}{1-\psi_{x}}\partial_{x}^{2}\left(\frac{\psi_{x}}{1-\psi_{x}}\bar{u}^{\prime}\right)+\frac{\psi_{x}}{1-\psi_{x}}\partial_{x}\left(\frac{\psi_{x}}{1-\psi_{x}}\bar{u}^{\prime\prime}\right)\right.
+x2(ψx21ψxu¯)+x(ψx21ψxu¯′′)+ψx21ψxu¯′′′].\displaystyle\qquad+\left.\partial_{x}^{2}\left(\frac{\psi_{x}^{2}}{1-\psi_{x}}\bar{u}^{\prime}\right)+\partial_{x}\left(\frac{\psi_{x}^{2}}{1-\psi_{x}}\bar{u}^{\prime\prime}\right)+\frac{\psi_{x}^{2}}{1-\psi_{x}}\bar{u}^{\prime\prime\prime}\right].
Proof.

The proof of the above is a relatively routine, yet long calculation. For completeness, we present the details in Appendix A. ∎

Our goal is now to obtain a closed nonlinear iteration scheme by integrating (4.3) and exploiting the decomposition of the linear solution operator ete^{\mathcal{L}t} provided in (3.4). To motivate this, we first provide an informal description of how to identify the modulation functions γ\gamma and ψ\psi. To this end, note that using Duhamel’s formula we can rewrite (4.3) as the equivalent implicit integral equation

(1ψx(x,t))v(x,t)+1Nγ(t)u¯(x)+ψ(x,t)u¯(x)=etv(x,0)+0te(ts)x𝒩(x,s)ds,\left(1-\psi_{x}(x,t)\right)v(x,t)+\frac{1}{N}\gamma(t)\bar{u}^{\prime}(x)+\psi(x,t)\bar{u}^{\prime}(x)=e^{\mathcal{L}t}v(x,0)+\int_{0}^{t}e^{\mathcal{L}(t-s)}\partial_{x}\mathcal{N}(x,s)ds,

where here we have taken the initial data γ(0)=0\gamma(0)=0, ψ(,0)0\psi(\cdot,0)\equiv 0 and v(x,0)=u(x,0)δMu¯(x)v(x,0)={u}(x,0)-\delta M-\bar{u}(x). Recalling that (3.4) implies the linear solution operator can be decomposed as

etv(x)\displaystyle e^{\mathcal{L}t}v(x) =u¯(x)[1Nψadj,vLN2+sp,N(t)v(x)phasemodulationt(1N0NTv(z)𝑑z)wavespeedcorrection]\displaystyle=\bar{u}^{\prime}(x)\left[\underbrace{\frac{1}{N}\left<\psi^{adj},v\right>_{L^{2}_{N}}+s_{p,N}(t)v(x)}_{\rm phase~modulation}-t\underbrace{\left(\frac{1}{N}\int_{0}^{NT}v(z)dz\right)}_{\rm wave~speed~correction}\right]
+1N0NTv(z)𝑑zmassmodulation+S~N(t)v(x)fasterdecayingresidual,\displaystyle\quad+\underbrace{\frac{1}{N}\int_{0}^{NT}v(z)dz}_{\rm mass~modulation}+\underbrace{\widetilde{S}_{N}(t)v(x)}_{\rm faster~decaying~residual},

it follows that we can remove the principle (i.e., slowest decaying) part of the nonlinear perturbation by implicitly defining

(4.4) {γ(t)ψadj,v(,0)LN2+0tψadj,x𝒩(,s)LN2𝑑sψ(x,t)sp,N(t)v(x,0)+0tsp,N(ts)x𝒩(x,s)ds,\left\{\begin{aligned} &\gamma(t)\sim\left<\psi^{adj},v(\cdot,0)\right>_{L^{2}_{N}}+\int_{0}^{t}\left<\psi^{adj},\partial_{x}\mathcal{N}(\cdot,s)\right>_{L^{2}_{N}}ds\\ &\psi(x,t)\sim s_{p,N}(t)v(x,0)+\int_{0}^{t}s_{p,N}(t-s)\partial_{x}\mathcal{N}(x,s)ds,\end{aligned}\right.

where here \sim indicates equality for t1t\geq 1. This choice then yields the implicit description

(4.5) v(x,t)ψx(x,t)v(x,t)+S~N(t)v(x,0)+0tS~N(ts)x𝒩(x,s)dsv(x,t)\sim\psi_{x}(x,t)v(x,t)+\widetilde{S}_{N}(t)v(x,0)+\int_{0}^{t}\widetilde{S}_{N}(t-s)\partial_{x}\mathcal{N}(x,s)ds

involving only the faster decaying residual component of the linear solution operator.

To make the above choices consistent (in short time) with the prescribed initial data, we simply interpolate between the initial data and the long-time choices prescribed in (4.4)-(4.5) above. To this end, let χ(t)\chi(t) be a smooth cutoff function that is zero for t1/2t\leq 1/2 and one for t1t\geq 1, and define the modulation functions γ\gamma and ψ\psi implicitly for all t0t\geq 0 as

(4.6) {γ(t)=χ(t)[ψadj,v(,0)LN2+0tψadj,x𝒩(,s)LN2𝑑s]ψ(x,t)=χ(t)[sp,N(t)v(x,0)+0tsp,N(ts)x𝒩(x,s)ds],\left\{\begin{aligned} &\gamma(t)=\chi(t)\left[\left<\psi^{adj},v(\cdot,0)\right>_{L^{2}_{N}}+\int_{0}^{t}\left<\psi^{adj},\partial_{x}\mathcal{N}(\cdot,s)\right>_{L^{2}_{N}}ds\right]\\ &\psi(x,t)=\chi(t)\left[s_{p,N}(t)v(x,0)+\int_{0}^{t}s_{p,N}(t-s)\partial_{x}\mathcal{N}(x,s)ds\right],\end{aligned}\right.

which now leaves the implicit description

(4.7) v(x,t)\displaystyle v(x,t) =(1χ(t))[etv(x,0)+0te(ts)x𝒩(s)ds]\displaystyle=\left(1-\chi(t)\right)\left[e^{\mathcal{L}t}v(x,0)+\int_{0}^{t}e^{\mathcal{L}(t-s)}\partial_{x}\mathcal{N}(s)ds\right]
+χ(t)(ψx(x,t)v(x,t)+S~N(t)v(x,0)+0tS~N(ts)x𝒩(x,s)ds)\displaystyle\qquad+\chi(t)\left(\psi_{x}(x,t)v(x,t)+\widetilde{S}_{N}(t)v(x,0)+\int_{0}^{t}\widetilde{S}_{N}(t-s)\partial_{x}\mathcal{N}(x,s)ds\right)

to be satisfied in LN2L^{2}_{N} for all t0t\geq 0. Observe, however, that there is an inherent loss of derivatives associated with the system (4.6)-(4.7). For example, attempting to control the LN2L^{2}_{N} norm of v(,t)v(\cdot,t) via the implicit equation (4.7) requires control of at least the HN4H^{4}_{N} derivative of vv (as well as various derivatives of γ\gamma and ψ\psi). This loss of derivatives may be compensated by the fact that the damping in (1.1) corresponds to the highest-order spatial derivative, which allows high Sobolev norms of vv to be slaved to low Sobolev norms of vv plus sufficient control on the modulation functions γ\gamma and ψ\psi. This is the content of the following technical result.

Proposition 4.2 (Nonlinear Damping).

Fix NN\in{\mathbb{N}} and suppose that the nonlinear perturbation vv defined in (4.1) satisfies v(,0)HN5v(\cdot,0)\in H^{5}_{N}. Then there exist positive constants θ\theta, CC and ε0\varepsilon_{0} (independent of NN) such that if vv, ψ\psi and γ\gamma solve the system (4.6)-(4.7) on [0,τ][0,\tau] for some τ>0\tau>0 and

supt[0,τ]((v,ψx)HN5+ψt(t)HN4+|γ(t)|)ε0\sup_{t\in[0,\tau]}\left(\left\|(v,\psi_{x})\right\|_{H^{5}_{N}}+\left\|\psi_{t}(t)\right\|_{H^{4}_{N}}+\left|\gamma^{\prime}(t)\right|\right)\leq\varepsilon_{0}

then, for all 0tτ0\leq t\leq\tau we have

v(t)HN52\displaystyle\|v(t)\|_{H^{5}_{N}}^{2} Ceθtv(0)HN52\displaystyle\leq Ce^{-\theta t}\|v(0)\|_{H^{5}_{N}}^{2}
+C0teθ(ts)(v(s)LN22+ψx(s)HN62+ψt(s)HN32+|γt(s)|2)𝑑s.\displaystyle\quad+C\int_{0}^{t}e^{-\theta(t-s)}\left(\|v(s)\|_{L^{2}_{N}}^{2}+\|\psi_{x}(s)\|_{H^{6}_{N}}^{2}+\|\psi_{t}(s)\|_{H^{3}_{N}}^{2}+|\gamma_{t}(s)|^{2}\right)ds.

The proof of the above estimate is by now standard, and can be found (for perturbations in L2()L^{2}({\mathbb{R}})) in [3, Proposition 3.4]. The result in [3] directly adapts to the current situation and is hence omitted.

4.2 Nonlinear Iteration

With the above nonlinear preliminaries, we now complete the proof of Theorem 1.7. Associated to the solution (v,γt,ψx,ψt)(v,\gamma_{t},\psi_{x},\psi_{t}) of (4.6)-(4.1) we define, so long as it is finite, the function

ζ(t):=sup0st(v(s)HN52+ψt(s)HN52+ψx(s)HN62+|γt(s)|)1/2(1+s)1/4.\zeta(t):=\sup_{0\leq s\leq t}\left(\|v(s)\|_{H^{5}_{N}}^{2}+\|\psi_{t}(s)\|_{H^{5}_{N}}^{2}+\|\psi_{x}(s)\|_{H^{6}_{N}}^{2}+|\gamma_{t}(s)|\right)^{1/2}(1+s)^{1/4}.

Using the linear estimates from Section 3 as well as the above nonlinear preparations, we now establish an estimate on ζ\zeta that will establish both the global existence of the nearby solution u{u} and the modulation functions, but will also establish our main stability result.

Proposition 4.3.

Under the hypotheses of Theorem 1.7, there exist positive constants C,ε>0C,\varepsilon>0, both independent of NN, such that if v(,0)v(\cdot,0) satisfies

E0:=v(,0)LN1HN5εandζ(τ)εE_{0}:=\|v(\cdot,0)\|_{L^{1}_{N}\cap H^{5}_{N}}\leq\varepsilon~~{\rm and}~~\zeta(\tau)\leq\varepsilon

for some τ>0\tau>0, then we have

ζ(t)C(E0+ζ(t)2)\zeta(t)\leq C\left(E_{0}+\zeta(t)^{2}\right)

for all 0tτ0\leq t\leq\tau.

Proof.

From Proposition 4.1, we note that

(4.8) 𝒩LN1HN1C((v,ψx)HN52+ψtHN12+|γt|2)Cζ(t)2(1+t)1/2\left\|\mathcal{N}\right\|_{L^{1}_{N}\cap H^{1}_{N}}\leq C\left(\|(v,\psi_{x})\|_{H^{5}_{N}}^{2}+\|\psi_{t}\|_{H^{1}_{N}}^{2}+|\gamma_{t}|^{2}\right)\leq C\zeta(t)^{2}\left(1+t\right)^{-1/2}

for some constant C>0C>0 independent of NN. Using the linear estimates in Proposition 3.1, as well as the conservative structure of the nonlinearity in (4.3), it follows that

v(t)LN2\displaystyle\left\|v(t)\right\|_{L^{2}_{N}} v(t)ψx(t)LN2+CE0(1+t)1/4+C0t(1+ts)3/4𝒩(s)LN1LN2𝑑s\displaystyle\leq\left\|v(t)\psi_{x}(t)\right\|_{L^{2}_{N}}+CE_{0}\left(1+t\right)^{-1/4}+C\int_{0}^{t}(1+t-s)^{-3/4}\left\|\mathcal{N}(s)\right\|_{L^{1}_{N}\cap L^{2}_{N}}ds
ζ(t)2(1+t)1/2+CE0(1+t)1/4+Cζ(t)20t(1+ts)3/4(1+s)1/2𝑑s\displaystyle\leq\zeta(t)^{2}(1+t)^{-1/2}+CE_{0}(1+t)^{-1/4}+C\zeta(t)^{2}\int_{0}^{t}(1+t-s)^{-3/4}(1+s)^{-1/2}ds
C(E0+ζ(t)2)(1+t)1/4,\displaystyle\leq C\left(E_{0}+\zeta(t)^{2}\right)(1+t)^{-1/4},

where, again, the above constant C>0C>0 is independent of NN. Similarly, from (4.6) we have for each 161\leq\ell\leq 6

xψx(x,t)=χ(t)[x+1sp,N(t)v(x,0)+0tx+1sp,N(ts)x𝒩(x,s)ds]\partial_{x}^{\ell}\psi_{x}(x,t)=\chi(t)\left[\partial_{x}^{\ell+1}s_{p,N}(t)v(x,0)+\int_{0}^{t}\partial_{x}^{\ell+1}s_{p,N}(t-s)\partial_{x}\mathcal{N}(x,s)ds\right]

and hence

ψxHN6\displaystyle\left\|\psi_{x}\right\|_{H^{6}_{N}} CE0(1+t)1/4+C0t(1+ts)3/4x𝒩(,s)LN2𝑑s\displaystyle\leq CE_{0}(1+t)^{-1/4}+C\int_{0}^{t}\left(1+t-s\right)^{-3/4}\left\|\partial_{x}\mathcal{N}(\cdot,s)\right\|_{L^{2}_{N}}ds
C(E0+ζ(t)2)(1+t)1/4\displaystyle\leq C\left(E_{0}+\zeta(t)^{2}\right)(1+t)^{-1/4}

again where the constant C>0C>0 is independent of NN. A completely analogous calculation gives

ψtHN5C(E0+ζ(t)2)(1+t)1/4,\left\|\psi_{t}\right\|_{H^{5}_{N}}\leq C\left(E_{0}+\zeta(t)^{2}\right)(1+t)^{-1/4},

and an application of integration by parts yields

|γt(t)||ψadj,x𝒩(,t)LN2|C𝒩(,t)LN1Cζ(t)2(1+t)1/2,|\gamma_{t}(t)|\leq\left|\left<\psi^{adj},\partial_{x}\mathcal{N}(\cdot,t)\right>_{L^{2}_{N}}\right|\leq C\left\|\mathcal{N}(\cdot,t)\right\|_{L^{1}_{N}}\leq C\zeta(t)^{2}(1+t)^{-1/2},

where here we used an L1LL^{1}-L^{\infty} bound to control the inner product777Otherwise, one would get a 𝒪(N)\mathcal{O}(N)-growth from the xψadjLN2\|\partial_{x}\psi^{adj}\|_{L^{2}_{N}} term.. Using now the nonlinear damping result in Proposition 4.2, it follows that

v(t)HN52\displaystyle\|v(t)\|_{H^{5}_{N}}^{2} CE02eθt+C(E0+ζ(t)2)20teθ(ts)(1+s)1/2𝑑s\displaystyle\leq CE_{0}^{2}e^{-\theta t}+C\left(E_{0}+\zeta(t)^{2}\right)^{2}\int_{0}^{t}e^{-\theta(t-s)}(1+s)^{-1/2}ds
C(E0+ζ(t)2)2(1+t)1/2.\displaystyle\leq C\left(E_{0}+\zeta(t)^{2}\right)^{2}(1+t)^{-1/2}.

Noting that ζ(t)\zeta(t) is a non-decreasing function, it follows for all t(0,τ)t\in(0,\tau) that

(v(s)HN52+ψt(s)HN52+ψx(s)HN62+|γt(s)|)1/2(1+s)1/4C(E0+ζ(t))2,\left(\|v(s)\|_{H^{5}_{N}}^{2}+\|\psi_{t}(s)\|_{H^{5}_{N}}^{2}+\|\psi_{x}(s)\|_{H^{6}_{N}}^{2}+|\gamma_{t}(s)|\right)^{1/2}(1+s)^{1/4}\leq C\left(E_{0}+\zeta(t)\right)^{2},

valid for all s(0,t)s\in(0,t). Taking the supremum over all s(0,t)s\in(0,t) yields the desired result. ∎

With Proposition 4.3 established, the proof of Theorem 1.7 follows directly. Indeed, since ζ(t)\zeta(t) is continuous for so long as it remains small, it follows from Proposition 4.3 that if E0<14CE_{0}<\frac{1}{4C} then we have 0ζ(t)2CE00\leq\zeta(t)\leq 2CE_{0} for all t0t\geq 0. Noting that the constant C>0C>0 from Proposition 4.3 is independent of NN and setting

ψ~(x,t)=1Nγ(t)+ψ(x,t),\widetilde{\psi}(x,t)=\frac{1}{N}\gamma(t)+\psi(x,t),

the stability estimates (1.7) in Theorem 1.7 follow. Finally, using that ζ(t)2CE0\zeta(t)\leq 2CE_{0} for all t0t\geq 0 it follows from (4.8) and applying (3.9) to the implicit representation (4.6) that

ψ(,t)LCE0+CE020t(1+ts)1/2(1+s)1/2𝑑sCE0,\left\|\psi(\cdot,t)\right\|_{L^{\infty}}\leq CE_{0}+CE_{0}^{2}\int_{0}^{t}(1+t-s)^{-1/2}(1+s)^{-1/2}ds\leq CE_{0},

yielding the LL^{\infty} estimate (1.10).

Appendix A Proof of Nonlinear Perturbation Equations

In this appendix, we present the details of the derivation of the nonlinear perturbation equations in Proposition 4.1. To this end, first set

u~(x,t):=u(x+(δM)t1Nγ(t)ψ(x,t),t)\widetilde{u}(x,t):={u}\left(x+\left(\delta M\right)t-\frac{1}{N}\gamma(t)-\psi(x,t),t\right)

and note that

u~x=11ψxx(v+u¯)\widetilde{u}_{x}=\frac{1}{1-\psi_{x}}\partial_{x}(v+\bar{u})

and

u~t=vt+11ψx(1Nγ+ψt1NδM)x(v+u¯).\widetilde{u}_{t}=v_{t}+\frac{1}{1-\psi_{x}}\left(\frac{1}{N}\gamma^{\prime}+\psi_{t}-\frac{1}{N}\delta M\right)\partial_{x}(v+\bar{u}).

Since u(x,t){u}(x,t) is a solution to (1.1), in the traveling coordinate frame xctx-ct, it follows that

(1ψx)vt+(1Nγ+ψt1NδM)x(v+u¯)cx(v+u¯)\displaystyle(1-\psi_{x})v_{t}+\left(\frac{1}{N}\gamma^{\prime}+\psi_{t}-\frac{1}{N}\delta M\right)\partial_{x}\left(v+\bar{u}\right)-c\partial_{x}\left(v+\bar{u}\right)
+εx(11ψxx(11ψxx(v+u¯)))+δx(11ψxx(v+u¯))\displaystyle\qquad+\varepsilon\partial_{x}\left(\frac{1}{1-\psi_{x}}\partial_{x}\left(\frac{1}{1-\psi_{x}}\partial_{x}\left(v+\bar{u}\right)\right)\right)+\delta\partial_{x}\left(\frac{1}{1-\psi_{x}}\partial_{x}\left(v+\bar{u}\right)\right)
+δx(11ψxx(11ψxx(11ψxx(v+u¯))))\displaystyle\qquad+\delta\partial_{x}\left(\frac{1}{1-\psi_{x}}\partial_{x}\left(\frac{1}{1-\psi_{x}}\partial_{x}\left(\frac{1}{1-\psi_{x}}\partial_{x}\left(v+\bar{u}\right)\right)\right)\right)
+(v+u¯+1NδM)x(v+u¯)=0\displaystyle\qquad+\left(v+\bar{u}+\frac{1}{N}\delta M\right)\partial_{x}\left(v+\bar{u}\right)=0

Clearly, the contributions from the δM\delta M terms cancel, which is a reflection of the Galilean invariance of (1.1). Now, subtracting off the profile equation (1.2) for u¯\bar{u} yields

vt(ψxv)t+(ψtv)x+ψtu¯+(1Nγv)x+1Nγu¯cvx\displaystyle v_{t}-(\psi_{x}v)_{t}+(\psi_{t}v)_{x}+\psi_{t}\bar{u}^{\prime}+\left(\frac{1}{N}\gamma^{\prime}v\right)_{x}+\frac{1}{N}\gamma^{\prime}\bar{u}^{\prime}-cv_{x}
+εx(11ψxx(11ψxx(v+u¯)))εu¯′′′+δx(11ψxx(v+u¯))δu¯′′\displaystyle\quad+\varepsilon\partial_{x}\left(\frac{1}{1-\psi_{x}}\partial_{x}\left(\frac{1}{1-\psi_{x}}\partial_{x}\left(v+\bar{u}\right)\right)\right)-\varepsilon\bar{u}^{\prime\prime\prime}+\delta\partial_{x}\left(\frac{1}{1-\psi_{x}}\partial_{x}\left(v+\bar{u}\right)\right)-\delta\bar{u}^{\prime\prime}
+δx(11ψxx(11ψxx(11ψxx(v+u¯))))δu¯′′′′\displaystyle\qquad+\delta\partial_{x}\left(\frac{1}{1-\psi_{x}}\partial_{x}\left(\frac{1}{1-\psi_{x}}\partial_{x}\left(\frac{1}{1-\psi_{x}}\partial_{x}\left(v+\bar{u}\right)\right)\right)\right)-\delta\bar{u}^{\prime\prime\prime\prime}
+12x(v2)+x(u¯v)=0\displaystyle\qquad+\frac{1}{2}\partial_{x}\left(v^{2}\right)+\partial_{x}\left(\bar{u}v\right)=0

or, equivalently888By adding and subtracting (t)(v+1Nγu¯+ψu¯)(\partial_{t}-\mathcal{L})\left(v+\frac{1}{N}\gamma\bar{u}^{\prime}+\psi\bar{u}^{\prime}\right).

(t)(v+1Nγu¯+ψu¯)=(ψxv)t(ψtv+1Nγv)x12x(v2)\displaystyle(\partial_{t}-\mathcal{L})\left(v+\frac{1}{N}\gamma\bar{u}^{\prime}+\psi\bar{u}^{\prime}\right)=(\psi_{x}v)_{t}-\left(\psi_{t}v+\frac{1}{N}\gamma^{\prime}v\right)_{x}-\frac{1}{2}\partial_{x}\left(v^{2}\right)
ε[x(11ψxx(11ψxu¯))u¯′′′(ψxu¯)xx(ψxu¯′′)x]\displaystyle\qquad-\varepsilon\left[\partial_{x}\left(\frac{1}{1-\psi_{x}}\partial_{x}\left(\frac{1}{1-\psi_{x}}\bar{u}^{\prime}\right)\right)-\bar{u}^{\prime\prime\prime}-\left(\psi_{x}\bar{u}^{\prime}\right)_{xx}-\left(\psi_{x}\bar{u}^{\prime\prime}\right)_{x}\right]
ε[x(11ψxx(11ψxvx))vxxx]δ[x(11ψxu¯)u¯′′(ψxu¯)x]\displaystyle\qquad-\varepsilon\left[\partial_{x}\left(\frac{1}{1-\psi_{x}}\partial_{x}\left(\frac{1}{1-\psi_{x}}v_{x}\right)\right)-v_{xxx}\right]-\delta\left[\partial_{x}\left(\frac{1}{1-\psi_{x}}\bar{u}^{\prime}\right)-\bar{u}^{\prime\prime}-\left(\psi_{x}\bar{u}^{\prime}\right)_{x}\right]
δ[x(11ψxvx)vxx]δ[x(11ψxx(11ψxx(11ψxvx)))vxxxx]\displaystyle\qquad-\delta\left[\partial_{x}\left(\frac{1}{1-\psi_{x}}v_{x}\right)-v_{xx}\right]-\delta\left[\partial_{x}\left(\frac{1}{1-\psi_{x}}\partial_{x}\left(\frac{1}{1-\psi_{x}}\partial_{x}\left(\frac{1}{1-\psi_{x}}v_{x}\right)\right)\right)-v_{xxxx}\right]
δ[x(11ψxx(11ψxx(11ψxu¯)))u¯′′′′(ψxu¯)xxx(ψxu¯′′)xx(ψxu¯′′′)x].\displaystyle\qquad-\delta\left[\partial_{x}\left(\frac{1}{1-\psi_{x}}\partial_{x}\left(\frac{1}{1-\psi_{x}}\partial_{x}\left(\frac{1}{1-\psi_{x}}\bar{u}^{\prime}\right)\right)\right)-\bar{u}^{\prime\prime\prime\prime}-\left(\psi_{x}\bar{u}^{\prime}\right)_{xxx}-\left(\psi_{x}\bar{u}^{\prime\prime}\right)_{xx}-\left(\psi_{x}\bar{u}^{\prime\prime\prime}\right)_{x}\right].

Using the identity 11ψx=1+ψx1ψx\frac{1}{1-\psi_{x}}=1+\frac{\psi_{x}}{1-\psi_{x}} and rearranging now yields

(t)(v+1Nγu¯+ψu¯)=x𝒬+x+(ψxv)t.(\partial_{t}-\mathcal{L})\left(v+\frac{1}{N}\gamma\bar{u}^{\prime}+\psi\bar{u}^{\prime}\right)=\partial_{x}\mathcal{Q}+\partial_{x}\mathcal{R}+(\psi_{x}v)_{t}.

Adding and subtracting (ψxv)\mathcal{L}(\psi_{x}v) completes the proof.

Remark A.1.

One subtle point in our nonlinear analysis is the placement of the modulation functions in the decomposition (4.1). From the previous work [22, 23] and based on the linear analysis in Section 3, it may seem more natural to use the slightly different decomposition

(A.1) u~(xψ(x,t),t)=u¯(xt(δM)+1Nγ(t))+δM+v(xt(δM),t)\widetilde{u}\left(x-\psi(x,t),t\right)=\bar{u}\left(x-t\left(\delta M\right)+\frac{1}{N}\gamma(t)\right)+\delta M+v\left(x-t\left(\delta M\right),t\right)

where γ:+\gamma:{\mathbb{R}}_{+}\to{\mathbb{R}} and ψ:×+\psi:{\mathbb{R}}\times{\mathbb{R}}_{+}\to{\mathbb{R}} are functions to be determined as in the analysis above. However, one can readily check in the derivation of the nonlinear perturbation equations above that the decomposition (A.1) introduces terms that are

𝒪(1Nγ(t)+(δM)t)\mathcal{O}\left(\frac{1}{N}\gamma(t)+(\delta M)t\right)

and hence, in particular, are not decaying in time. While this is not a problem in the traditional fixed NN theory (with exponential semigroup bounds), it is insufficient to close a nonlinear iteration scheme when using the more delicate uniform algebraic bounds. This is reminiscent of the similar observation made in [10, Remark 2.4] in the context of localized perturbations of periodic waves in dissipative systems.

References

  • [1] D. E. Bar and A. A. Nepomnyashchy. Stability of periodic waves governed by the modified Kawahara equation. Phys. D, 86(4):586–602, 1995.
  • [2] B. Barker, M. A. Johnson, P. Noble, L. M. Rodrigues, and K. Zumbrun. Stability of periodic Kuramoto-Sivashinsky waves. Appl. Math. Lett., 25(5):824–829, 2012.
  • [3] B. Barker, M. A. Johnson, P. Noble, L. M. Rodrigues, and K. Zumbrun. Nonlinear modulational stability of periodic traveling-wave solutions of the generalized Kuramoto-Sivashinsky equation. Phys. D, 258:11–46, 2013.
  • [4] H. S. Brown, I. G. Kevrekidis, and M. S. Jolly. A minimal model for spatio-temporal patterns in thin film flow. In Patterns and dynamics in reactive media (Minneapolis, MN, 1989), volume 37 of IMA Vol. Math. Appl., pages 11–31. Springer, New York, 1991.
  • [5] H.-C. Chang, E. Demekhin, and D. I. Kopelevich. Laminarizing effects of dispersion in an active-dissipative nonlinear medium. Physica D: Nonlinear Phenomena, 63:299–320, 1993.
  • [6] U. Frisch, Z.-S. She, and O. Thual. Viscoelastic behaviour of cellular solutions to the Kuramoto-Sivashinsky model. J. Fluid Mech., 168:221–240, 1986.
  • [7] D. M. G.I. Sivashinsky. On irregular wavy flow of a liquid down an vertical plane. Progr. Theoret. Phys., 63:2112–2114, 1980.
  • [8] M. Haragus, M. A. Johnson, and W. R. Perkins. Linear modulational and subharmonic dynamics of spectrally stable Lugiato-Lefever periodic waves. J. Differential Equations, 280:315–354, 2021.
  • [9] E. D. H.C. Chang. Complex wave dynamics on thin films, volume 14 of Stud. Interface Sci. Elsevier Science B.V., Amsterdam, 2002.
  • [10] M. A. Johnson, P. Noble, L. M. Rodrigues, and K. Zumbrun. Behavior of periodic solutions of viscous conservation laws under localized and nonlocalized perturbations. Inventiones Mathematicae, 197(1):115–213, 2014.
  • [11] M. A. Johnson, P. Noble, L. M. Rodrigues, and K. Zumbrun. Spectral stability of periodic wave trains of the Korteweg–de Vries/Kuramoto-Sivashinsky equation in the Korteweg–de Vries limit. Trans. Amer. Math. Soc., 367(3):2159–2212, 2015.
  • [12] M. A. Johnson and W. R. Perkins. Subharmonic dynamics of wave trains in reaction-diffusion systems. Phys. D, 422:132891, 11, 2021.
  • [13] M. A. Johnson and K. Zumbrun. Nonlinear stability of periodic traveling-wave solutions of viscous conservation laws in dimensions one and two. SIAM J. Appl. Dyn. Syst., 10(1):189–211, 2011.
  • [14] P. Kent and J. Elgin. Travelling-waves of the Kuramoto-Sivashinsky equation: period-multiplying bifurcations. Nonlinearity, 5(4):899–919, 1992.
  • [15] Y. Kuramoto. Chemical oscillations, waves, and turbulence. Springer-Verlag, Berlin, 1984.
  • [16] Y. Kuramoto and T. Tsuzuki. On the formation of dissipative structures in reaction-diffusion systems. Progr. Theoret. Phys., 1975.
  • [17] P. Noble and L. M. Rodrigues. Whitham’s modulation equations and stability of periodic wave solutions of the Korteweg-de Vries-Kuramoto-Sivashinsky equation. Indiana Univ. Math. J., 62(3):753–783, 2013.
  • [18] G. Schneider. Diffusive stability of spatial periodic solutions of the Swift-Hohenberg equation. Comm. Math. Phys., 178(3):679–702, 1996.
  • [19] G. Schneider. Nonlinear diffusive stability of spatially periodic solutions—abstract theorem and higher space dimensions. In Proceedings of the International Conference on Asymptotics in Nonlinear Diffusive Systems (Sendai, 1997), volume 8 of Tohoku Math. Publ., pages 159–167. Tohoku Univ., Sendai, 1998.
  • [20] G. Sivashinsky. Nonlinear analysis of hydrodynamic instability in laminar flame. i. derivation of basic equations. Acta. Astron., 4:1177–1206, 1977.
  • [21] G. Sivashinsky. Instabilities, pattern formation, and turbulence in flames. Annu. Rev. Fluid Mech., 15:179–199, 1983.
  • [22] M. Stanislavova and A. Stefanov. Asymptotic estimates and stability analysis of Kuramoto-Sivashinsky type models. J. Evol. Equ., 11(3):605–635, 2011.
  • [23] M. Stanislavova and A. Stefanov. Erratum to: Asymptotic estimates and stability analysis of Kuramoto-Sivashinsky type models [mr2827102]. J. Evol. Equ., 11(3):637–639, 2011.
  • [24] H. A. Win. Model equation of surface waves of viscous fluid down an inclined plane. J. Math. Kyoto Univ., 33(3):803–824, 1993.
  • [25] J. Yu and Y. Yang. Evolution of small periodic disturbances into roll waves in channel flow with internal dissipation. Stud. Appl. Math., 111(1):1–27, 2003.