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

Linear and orbital stability analysis for solitary-wave solutions of variable-coefficient scalar field equations

Mashael Alammari Department of Mathematical Sciences, Florida Institute of Technology, Melbourne, FL 32901 [email protected]  and  Stanley Snelson Department of Mathematical Sciences, Florida Institute of Technology, Melbourne, FL 32901 [email protected]
Abstract.

We study general semilinear scalar-field equations on the real line with variable coefficients in the linear terms. These coefficients are uniformly small, but slowly decaying, perturbations of a constant-coefficient operator. We are motivated by the question of how these perturbations of the equation may change the stability properties of kink solutions (one-dimensional topological solitons). We prove existence of a stationary kink solution in our setting, and perform a detailed spectral analysis of the corresponding linearized operator, based on perturbing the linearized operator around the constant-coefficient kink. We derive a formula that allows us to check whether a discrete eigenvalue emerges from the essential spectrum under this perturbation. Known examples suggest that this extra eigenvalue may have an important influence on the long-time dynamics in a neighborhood of the kink. We also establish orbital stability of solitary-wave solutions in the variable-coefficient regime, despite the possible presence of negative eigenvalues in the linearization.

Key words and phrases:
Scalar-field equations, solitary waves, variable coefficients, spectral perturbation
2020 Mathematics Subject Classification:
35L71, 35C07, 35P99

1. Introduction

We consider a semilinear, variable-coefficient scalar field equation of the form

(1.1) t2u[a(x)x2u+b(x)xu+c(x)u]+F(u)=0,x.\partial_{t}^{2}u-\left[a(x)\partial_{x}^{2}u+b(x)\partial_{x}u+c(x)u\right]+F^{\prime}(u)=0,\quad x\in\mathbb{R}.

Our assumptions on the potential FF are

(1.2) FC3(),F(a)=F(a+)=0 for some a<a+,F(a±)=0,F′′(a±)=m2>0,F(s)>0,s(a,a+).\begin{split}F\in C^{3}(\mathbb{R}),\quad F(a_{-})=F(a_{+})=0\text{ for some }a_{-}<a_{+},\\ F^{\prime}(a_{\pm})=0,\quad F^{\prime\prime}(a_{\pm})=m^{2}>0,\quad F(s)>0,s\in(a_{-},a_{+}).\end{split}

The linear operator a(x)x2+b(x)x+c(x)a(x)\partial_{x}^{2}+b(x)\partial_{x}+c(x) is assumed to be a perturbation of the 1D1D Laplacian x2\partial_{x}^{2}. More precisely, for a small parameter δ>0\delta>0, we assume

(1.3) |a1|+|xa|+|b|+|c|L1()+|a1|+|xa|+|b|+|c|L()δ.\||a-1|+|\partial_{x}a|+|b|+|c|\|_{L^{1}(\mathbb{R})}+\||a-1|+|\partial_{x}a|+|b|+|c|\|_{L^{\infty}(\mathbb{R})}\leq\delta.

With ω(x)=exp(xb(z)/a(z)dz)\omega(x)=\exp(\int_{-\infty}^{x}b(z)/a(z)\,\mathrm{d}z), the energy functional

E(u):=ω(x)a(x)(12(tu)2+12a(xu)212cu2+F(u))dx,E(u):=\int_{\mathbb{R}}\frac{\omega(x)}{a(x)}\left(\frac{1}{2}(\partial_{t}u)^{2}+\frac{1}{2}a(\partial_{x}u)^{2}-\frac{1}{2}cu^{2}+F(u)\right)\,\mathrm{d}x,

is formally conserved under the flow of (1.1). We note that equation (1.1) is not invariant under translations, and that we make no parity assumptions on FF or the coefficients aa, bb, and cc.

We are interested in the long-time behavior of solutions to (1.1). Our first result (Theorem 1.1) is the existence of a stationary solution TT of kink or solitary-wave type, i.e. an increasing stationary solution with T(x)a±T(x)\to a_{\pm} as x±x\to\pm\infty. Standard arguments then show that (1.1) is locally well-posed for initial data (u,tu)|t=0HT1()×L2()(u,\partial_{t}u)|_{t=0}\in H^{1}_{T}(\mathbb{R})\times L^{2}(\mathbb{R}), where HT1()={φ:φTH1()}H_{T}^{1}(\mathbb{R})=\{\varphi:\varphi-T\in H^{1}(\mathbb{R})\}. In this context, HT1()×L2()H^{1}_{T}(\mathbb{R})\times L^{2}(\mathbb{R}) is referred to as the energy space, and indeed, it is not hard to see (using in particular that |ω/a1|δ|\omega/a-1|\lesssim\delta) that functions in this space have finite energy. Our goal is to study the stability of TT with respect to small perturbations in the energy space.

1.1. Motivation

One-dimensional kinks such as T(x)T(x) are the simplest examples of topological solitons, and thus are an important model for physical phenomena arising in areas such as quantum field theory, condensed matter physics, and cosmology, among others. (See [48, 37, 28, 38] for some physics-oriented discussions.) Understanding their stability has proven to be a difficult mathematical challenge. The majority of work focuses on the constant-coefficient version of (1.1),

(1.4) t2ux2u+F(u)=0.\partial_{t}^{2}u-\partial_{x}^{2}u+F^{\prime}(u)=0.

In this constant-coefficient regime, it is standard that the assumptions (1.2) imply the existence of a kink solution connecting aa_{-} and a+a_{+}. (Convenient proofs of this fact may be found in [33, Lemma 1.1] or [23, Proposition 2.1].) This constant-coefficient stationary kink, which we denote by SS, satisfies

(1.5) S′′+F′′(S)=0,limxS(x)=a,limxS(x)=a+.-S^{\prime\prime}+F^{\prime\prime}(S)=0,\quad\lim_{x\to-\infty}S(x)=a_{-},\lim_{x\to\infty}S(x)=a_{+}.

We find in Theorem 1.1 that TT and SS are close in an appropriate norm.

Orbital stability of SS in the constant-coefficient setting has been known for some time [20], and we extend this to our setting in Theorem 1.3. Asymptotic stability of kinks is more subtle, and depends on the specific choice of potential FF. In particular, the two most studied versions of (1.4) are the ϕ4\phi^{4} equation with nonlinearity F(u)=u3uF^{\prime}(u)=u^{3}-u, which is known to be asymptotically stable with respect to odd perturbations [32] and conjectured to be asymptotically stable in general; and the sine-Gordon equation with nonlinearity F(u)=sin(u)F^{\prime}(u)=\sin(u), which is not asymptotically stable, at least with respect to perturbations in the energy space. (See Section 1.3 for more on these examples.)

Our motivation is to understand the effect of linear perturbations of the equation (1.4) on the stability properties of kink solutions. On the one hand, given that (1.4) is in some sense an idealized model, it is important on physical grounds to understand whether stability properties of kink solutions persist under perturbations of the equation. There is also reason to expect such perturbations to have a nontrivial qualitative impact on the stability analysis (rather than simply adding a small error term) in some situations. As we explain below, this is connected with the possibility that a discrete eigenvalue may emerge from the essential spectrum of the linearized operator around the kink.

1.2. Main results

Before stating our main theorems, we make a technically convenient change of variables in (1.1) that will be in effect throughout this article. Letting y=0xa1/2(z)dzy=\int_{0}^{x}a^{-1/2}(z)\,\mathrm{d}z, and abusing notation by writing u(t,y)=u(t,x(y))u(t,y)=u(t,x(y)) and b(y)=b(x(y))a1/2(x(y))ddya1/2(x(y))b(y)=b(x(y))-a^{-1/2}(x(y))\frac{d}{dy}a^{1/2}(x(y)), we have

(1.6) t2u[y2u+b(y)yu+c(y)u]+F(u)=0.\partial_{t}^{2}u-[\partial_{y}^{2}u+b(y)\partial_{y}u+c(y)u]+F^{\prime}(u)=0.

The hypotheses (1.3) imply

(1.7) |b|+|c|L1()+|b|+|c|L()C0δ,\||b|+|c|\|_{L^{1}(\mathbb{R})}+\||b|+|c|\|_{L^{\infty}(\mathbb{R})}\leq C_{0}\delta,

for some C0>0C_{0}>0.

Our first result is the existence of a stationary kink:

Theorem 1.1.

Assume that 0 is not an L2()L^{2}(\mathbb{R})-eigenvalue of the operator y2by(cF′′(S))-\partial_{y}^{2}-b\partial_{y}-(c-F^{\prime\prime}(S)). Then, for δ>0\delta>0 sufficiently small, there exists a solution TT to

(1.8) T′′b(y)Tc(y)T+F(T)=0,limyT(y)=a,limyT(y)=a+.-T^{\prime\prime}-b(y)T^{\prime}-c(y)T+F^{\prime}(T)=0,\quad\lim_{y\to-\infty}T(y)=a_{-},\lim_{y\to\infty}T(y)=a_{+}.

This solution can be written T(y)=S(y)+Sb(y)T(y)=S(y)+S_{b}(y), where SS solves (1.5) and

SbW1,1()+SbW1,()Cδ.\|S_{b}\|_{W^{1,1}(\mathbb{R})}+\|S_{b}\|_{W^{1,\infty}(\mathbb{R})}\leq C\delta.

Unlike SS, which satisfies |S(x)a±|emx|S(x)-a_{\pm}|\lesssim e^{\mp mx} and |S(x)|em|x||S^{\prime}(x)|\lesssim e^{-m|x|}, our static kink TT does not necessarily posess exponential tails. This behavior is reminiscent of some higher-order, constant-coefficient field theories that do not fit into the assumptions (1.2) (see e.g. [29]). Under additional exponential decay assumptions on bb and cc, it is possible to show TT has exponential asymptotics at ±\pm\infty as in [46, Theorem 1.1], but we do not explore the details here.

Our next result concerns the linearized operator around TT. Writing u(t,y)=T(y)+φ(t,y)u(t,y)=T(y)+\varphi(t,y), the perturbation φ\varphi satisfies

t2φy2φbyφcφ=F(T)F(T+φ).\partial_{t}^{2}\varphi-\partial_{y}^{2}\varphi-b\partial_{y}\varphi-c\varphi=F^{\prime}(T)-F^{\prime}(T+\varphi).

Adding F′′(T)φF^{\prime\prime}(T)\varphi to both sides, and defining the linear operator T=y2byc+F′′(T)\mathcal{L}_{T}=-\partial_{y}^{2}-b\partial_{y}-c+F^{\prime\prime}(T) and the nonlinearity 𝒩(T,φ)=F(T)F(T+φ)+F′′(T)φ=O(φ2)\mathcal{N}(T,\varphi)=F^{\prime}(T)-F^{\prime}(T+\varphi)+F^{\prime\prime}(T)\varphi=O(\varphi^{2}), the equation for φ\varphi can be written as a nonlinear Klein-Gordon equation:

(1.9) t2φ+Tφ=𝒩(T,φ).\partial_{t}^{2}\varphi+\mathcal{L}_{T}\varphi=\mathcal{N}(T,\varphi).

We are most interested in situations where the spectrum of S=y2+F′′(S)\mathcal{L}_{S}=-\partial_{y}^{2}+F^{\prime\prime}(S), the operator corresponding to the constant-coefficient kink, is known exactly. We then have T=Sbyc+F′′(T)F′′(S)\mathcal{L}_{T}=\mathcal{L}_{S}-b\partial_{y}-c+F^{\prime\prime}(T)-F^{\prime\prime}(S), and we ask how the perturbation byc+F′′(T)F′′(S)-b\partial_{y}-c+F^{\prime\prime}(T)-F^{\prime\prime}(S) changes the spectral properties of S\mathcal{L}_{S}.

The L2()L^{2}(\mathbb{R}) spectrum of S\mathcal{L}_{S} is given by

σ(S)={0,λ1,,λn}[m2,),\sigma(\mathcal{L}_{S})=\{0,\lambda_{1},\ldots,\lambda_{n}\}\cup[m^{2},\infty),

where λ1,,λn\lambda_{1},\ldots,\lambda_{n} is a possibly empty, increasing collection of positive, simple eigenvalues. The eigenfunction corresponding to 0 is exactly SS^{\prime}, the translation invariance mode.

As expected, discrete eigenvalues λi\lambda_{i} will drift to nearby discrete eigenvalues λi\lambda_{i}^{\prime} of T\mathcal{L}_{T} under the perturbation. A more delicate question is whether an extra discrete eigenvalue emerges from the essential spectrum. This aspect is especially relevant when S\mathcal{L}_{S} has a threshold resonance, i.e. a function RL()L2()R\in L^{\infty}(\mathbb{R})\setminus L^{2}(\mathbb{R}) satisfying SR=m2R\mathcal{L}_{S}R=m^{2}R, as is the case for both the ϕ4\phi^{4} and sine-Gordon equations. We derive a criterion in terms of RR and the coefficients bb and cc that governs whether the resonance drifts into a discrete eigenvalue.

Our results on the spectrum of T\mathcal{L}_{T} are collected in the following theorem:

Theorem 1.2.

Let T\mathcal{L}_{T} and S\mathcal{L}_{S} be as defined above. There exists a universal c0>0c_{0}>0 such that:

  1. (a)

    The spectrum σ(T)\sigma(\mathcal{L}_{T}) is real, the essential spectrum σess(T)=σess(S)=[m2,)\sigma_{ess}(\mathcal{L}_{T})=\sigma_{ess}(\mathcal{L}_{S})=[m^{2},\infty), and σ(T)\sigma(\mathcal{L}_{T}) lies in the c0δc_{0}\delta-neighborhood of σ(S)\sigma(\mathcal{L}_{S}).

  2. (b)

    For every eigenvalue λσd(S)\lambda\in\sigma_{d}(\mathcal{L}_{S}) with eigenvector YλY_{\lambda}, there is a corresponding λσd(T)\lambda^{\prime}\in\sigma_{d}(\mathcal{L}_{T}). The eigenvalue λ\lambda^{\prime} is real, simple, and satisfies |λλ|c0δ|\lambda-\lambda^{\prime}|\leq c_{0}\delta. Also, if

    A:=Yλ[(F′′(T)F′′(S)c)YλbyYλ]dy0,A:=\int_{\mathbb{R}}Y_{\lambda}[(F^{\prime\prime}(T)-F^{\prime\prime}(S)-c)Y_{\lambda}-b\partial_{y}Y_{\lambda}]\,\mathrm{d}y\neq 0,

    then λλ\lambda^{\prime}-\lambda has the same sign as AA. The eigenfunction YλY_{\lambda^{\prime}} of T\mathcal{L}_{T} corresponding to λ\lambda^{\prime} satisfies |Yλ(y)|+|Yλ(y)|em2λ|y||Y_{\lambda^{\prime}}(y)|+|Y_{\lambda^{\prime}}^{\prime}(y)|\lesssim e^{-\sqrt{m^{2}-\lambda^{\prime}}|y|}. Furthermore, for suitable normalizations of YλY_{\lambda} and YλY_{\lambda^{\prime}}, we have

    em2λ|y|Yλ(y)em2λ|y|Yλ(y)L()Cδ,\|e^{\sqrt{m^{2}-\lambda^{\prime}}|y|}Y_{\lambda^{\prime}}(y)-e^{\sqrt{m^{2}-\lambda}|y|}Y_{\lambda}(y)\|_{L^{\infty}(\mathbb{R})}\leq C\delta,

    for a universal constant C>0C>0.

  3. (c)

    If m2m^{2} is a simple resonance of S\mathcal{L}_{S}, and

    R[(F′′(T)F′′(S)c)RbyR]dy<0,\int_{\mathbb{R}}R[(F^{\prime\prime}(T)-F^{\prime\prime}(S)-c)R-b\partial_{y}R]\,\mathrm{d}y<0,

    then there exists a discrete eigenvalue λ\lambda of T\mathcal{L}_{T} with 0<m2λ<c0δ0<m^{2}-\lambda<c_{0}\delta. The eigenfunction YλY_{\lambda} also satisfies |Yλ(y)|+|Yλ(y)|em2λ|y||Y_{\lambda}(y)|+|Y_{\lambda}^{\prime}(y)|\lesssim e^{-\sqrt{m^{2}-\lambda}|y|} and

    Yλ(y)em2λ|y|R(y)L()C(k+δ)em2λ|y|,\|Y_{\lambda}(y)-e^{-\sqrt{m^{2}-\lambda}|y|}R(y)\|_{L^{\infty}(\mathbb{R})}\leq C(k+\delta)e^{-\sqrt{m^{2}-\lambda}|y|},

    for suitable normalizations of YλY_{\lambda} and RR.

    If

    R[(F′′(T)F′′(S)c)RbyR]dy>0,\int_{\mathbb{R}}R[(F^{\prime\prime}(T)-F^{\prime\prime}(S)-c)R-b\partial_{y}R]\,\mathrm{d}y>0,

    then there are no eigenvalues of T\mathcal{L}_{T} in [m2c0δ,m2][m^{2}-c_{0}\delta,m^{2}], and m2m^{2} is non-resonant.

  4. (d)

    If λ=m2\lambda=m^{2} is not a resonance or an embedded eigenvalue of S\mathcal{L}_{S}, then the same is true of T\mathcal{L}_{T}, and there are no eigenvalues of T\mathcal{L}_{T} in [m2c0δ,m2][m^{2}-c_{0}\delta,m^{2}].

Part (a) of this theorem is standard, and included for clarity of exposition. Part (b) is arguably not surprising, but its proof (see Section 3) is a useful warm-up for parts (c) and (d). We also remark that the formulas in this theorem may be replaced with (more cumbersome, but in some sense more elementary) formulas that depend only on FF, SS, bb, and cc, via a first-order approximation for F′′(T)F′′(S)F^{\prime\prime}(T)-F^{\prime\prime}(S). (See (3.7) and (3.11).)

The possible extra eigenvalue as in Theorem 1.2(c) is one of our primary motivations for performing this perturbation analysis. In general, eigenvalues lying in between 0 and m2m^{2} have a profound impact on the stability properties of the kink. At the very least, any proof of asymptotic stability or instability for TT would likely need to account for this extra eigenvalue in some way.

It should be noted that we are outside the realm of analytic perturbation theory, since we do not assume any continuity of the coefficients b,cb,c with respect to δ\delta. Our spectral analysis is based on the well-known method of finding solutions U±λU_{\pm\infty}^{\lambda} to the eigenvalue equation TUλ=λUλ\mathcal{L}_{T}U^{\lambda}=\lambda U^{\lambda} which decay at ±\pm\infty, and studying the Evans function (see e.g. [15, 25, 39, 26, 27]) which is related to the Wronskian of UλU_{\infty}^{\lambda} and UλU_{-\infty}^{\lambda}. The key property is that the Wronskian is zero when λ\lambda is an eigenvalue or resonance of T\mathcal{L}_{T}. The slow decay of our coefficients bb and cc (as well as F′′(T)F′′(S)F^{\prime\prime}(T)-F^{\prime\prime}(S)) rules out tools such as the Gap Lemma (see [2, 17]) which would allow one to analytically continue the Evans function past the threshold λ=m2\lambda=m^{2}, but which requires exponential decay of the coefficients.

Our last main result establishes the orbital stability of TT:

Theorem 1.3.

There exists an ε>0\varepsilon>0, depending on δ\delta, such that for any initial data (u,tu)|t=0=(T+v1,v2)(u,\partial_{t}u)\Big{|}_{t=0}=(T+v_{1},v_{2}) for (v1,v2)L2×H1(v_{1},v_{2})\in L^{2}\times H^{1} with

(v1,v2)H1()×L2()<ε,\|(v_{1},v_{2})\|_{H^{1}(\mathbb{R})\times L^{2}(\mathbb{R})}<\varepsilon,

the corresponding solution uu to (1.6) exists globally in time, and satisfies

uTH1()+tuL2()Cε,\|u-T\|_{H^{1}(\mathbb{R})}+\|\partial_{t}u\|_{L^{2}(\mathbb{R})}\leq C\varepsilon,

for some CC depending on δ\delta and FF.

The proof is based on classical energy arguments, but must contend with the lack of translation invariance.

1.3. Examples

1.3.1. ϕ4\phi^{4} model

The choice of a double-well potential F(u)=14(1u2)2F(u)=\frac{1}{4}(1-u^{2})^{2} leads to the ϕ4\phi^{4} model

(1.10) t2ux2u=uu3.\partial_{t}^{2}u-\partial_{x}^{2}u=u-u^{3}.

Standard references on this equation include [44, 6, 32, 41]. In this case, the kink solution S(x)=tanh(x/2)S(x)=\tanh(x/\sqrt{2}) is known explicitly, and the linearization S=x2+(3S21)\mathcal{L}_{S}=-\partial_{x}^{2}+(3S^{2}-1) has spectrum equal to

σ(S)={0,32}[2,).\sigma(\mathcal{L}_{S})=\left\{0,\frac{3}{2}\right\}\cup[2,\infty).

The odd eigenfunction Y3/2=tanh(x/2)sech(x/2)Y_{3/2}=\tanh(x/\sqrt{2})\ \mbox{sech}(x/\sqrt{2}) corresponding to λ=32\lambda=\frac{3}{2} is known as the internal oscillation mode. The operator S\mathcal{L}_{S} also possesses an even resonance R=2tanh2(x/2)sech2(x/2)R=2\tanh^{2}(x/\sqrt{2})-\ \mbox{sech}^{2}(x/\sqrt{2}) at the threshold λ=2\lambda=2.

The ϕ4\phi^{4} kink is asymptotically stable with respect to odd perturbations in the energy space, by the important work of Kowalczyk-Martel-Muñoz [32]. When working in the odd energy space, the even translation invariance mode at λ=0\lambda=0 and the even resonance do not play any role, but the internal oscillation mode has a dramatic effect on the dynamics. The method of [32] involved projecting φ\varphi onto Y3/2Y_{3/2} and the continuous spectrum, and carefully tracking the interaction between these two parts induced by the nonlinear terms of (1.9). A delicate coupling between the internal oscillation mode and the continuous part leads to the dissipation of energy away from a neighborhood of the kink.

Asymptotic stability with respect to odd perturbations was extended to a variable-coefficient version of (1.10) by the second named author in [46], though the coefficients were less general than those considered here (only a second-order perturbation, which was taken to be even and exponentially decaying). The symmetry assumption means that any eigenvalue emerging from the essential spectrum would be even, and therefore can be ignored.

It remains an important open question whether this kink is asymptotically stable with respect to general perturbations. Our Theorem 1.2 implies that for certain choices of b,cb,c in (1.6), the bottom of the continuous spectrum is non-resonant and there are no extra discrete eigenvalues. Such a version of (1.6) could serve as an interesting test case for the ϕ4\phi^{4} asymptotic stability problem, especially if one is convinced that the threshold resonance is an important source of difficulties.

1.3.2. Sine-Gordon equation

The choice F(u)=1cos(u)F(u)=1-\cos(u) results in the sine-Gordon equation:

t2ux2u=sin(u).\partial_{t}^{2}u-\partial_{x}^{2}u=-\sin(u).

This equation arises in the study of superconductivity as well as of surfaces with constant negative curvature, among other areas. (See e.g. [22, 7, 8] for background on this equation.)

The explicit static kink is given by S(x)=4arctan(ex)S(x)=4\arctan(e^{x}). The equation, which is completely integrable, possesses other special solutions including breathers and wobbling kinks [7, 43]. The presence of these wobbling kinks (periodic-in-time, spatially localized perturbations of the kink) implies that SS is not asymptotically stable in the energy space. (However, see [5] for an asymptotic stability result in a different topology, and [1], which identified an infinite-codimensional manifold of initial data near the kink for which asymptotic stability in the energy space does hold.) With S=x2+cos(S)\mathcal{L}_{S}=-\partial_{x}^{2}+\cos(S) the linearization around SS, it is known that

σ(S)={0}[1,),\sigma(\mathcal{L}_{S})=\{0\}\cup[1,\infty),

The failure of asymptotic stability in the energy space is consistent with the absense of an internal oscillation mode, which rules out the mechanism of stability observed for the ϕ4\phi^{4} model in [32]. However, there is an odd resonance R(x)=tanh(x)R(x)=\tanh(x) at the bottom of the continuous spectrum. Our Theorem 1.2 gives conditions under which the variable-coefficient version of sine-Gordon possesses a discrete eigenvalue λ\lambda with 0<1λ10<1-\lambda\ll 1. In this case, one may ask whether the new odd eigenfunction behaves sufficiently like an internal oscillation mode that a stability mechanism like the one mentioned above comes into force. We plan to explore this question in a future article.

Somewhat different perturbed forms of the sine-Gordon equation have been considered in, e.g., [13, 14, 9, 16]. The general belief is that breathers and wobbles are non-generic phenomena, so one may conjecture that some dense set of coefficients satisfying (1.7) lead to asymptotic stability.

1.3.3. Other examples

Let us briefly mention some other models whose variable-coefficient counterparts are included in our setting: the P(ϕ)2P(\phi)_{2} theory [37], the double-sine-Gordon equation [4], and certain higher-order field theories [28], i.e. potentials equal to a polynomial of even degree, which in some cases satisfies the assumptions (1.2) and other cases not.

1.4. Related work

The asymptotic stability of kinks in scalar field equations such as (1.4) is an active area of inquiry. In addition to the results mentioned above, we should mention the recent work of Kowalczyk-Martel-Muñoz-Van Den Bosch [33], which proved asymptotic stability for a general class of scalar-field models satisfying a condition on the potential FF that, in particular, rules out internal oscillation modes and threshold resonances. In the setting of odd perturbations, Delort-Masmoudi [12] established explicit decay rates for odd perturbations of the ϕ4\phi^{4} kink on time scales of order ε4\varepsilon^{-4}, where ε\varepsilon is the size of the initial perturbation. Let us also mention asymptotic stability results by Komech-Kopylova [30, 31] for kink solutions of relativistic Ginzburg-Landau equations, which are of the form (1.4) with additional assumptions of the flatness of FF at a±a_{\pm}.

This class of questions is a partial motivation for the closely related subject of scattering theory for NLKG equations similar to (1.9). See [10, 11, 36, 47, 35, 34, 18] and the references therein.

The operator S\mathcal{L}_{S} is (up to subtraction by m2Im^{2}I) a Schrödinger operator with rapidly decaying potential. There is a well-established theory of spectral perturbation of Schrödinger and related operators, see e.g. the review [45] for an overview. Works that specifically address perturbation of threshold resonances include [24, 3, 19, 40]. As mentioned above, aspects such as the slow decay of coefficients and lack of continuous dependence on δ\delta make it convenient to perform the perturbation “by hand” in our setting, rather than apply an abstract theorem or existing result.

1.5. Outline of the paper

In Section 2, we prove the existence of the stationary solution TT. In Section 3, we perform a spectral perturbation analysis of the linearized operator around the kink, and in Section 4, we establish orbital stability of TT. Appendix A contains some useful lemmas on the global solvability of second-order ODE systems.

2. Stationary solution

First, we recall the existence of the static kink in the constant-coefficient case, which can be found by explicitly integrating the equation S′′=F′′(S)S^{\prime\prime}=F^{\prime\prime}(S). We quote from [33, Lemma 1.1]:

Lemma 2.1.

Under the assumptions (1.2) on FF, there is a solution SC4()S\in C^{4}(\mathbb{R}) to the stationary equation

S′′+F(S)=0,-S^{\prime\prime}+F^{\prime}(S)=0,

with S>0S^{\prime}>0 and Sa±S\to a_{\pm} as y±y\to\pm\infty. Furthermore, SS and SS^{\prime} satisfy

|S(x)a±|Cemy,|S(x)|Cem|y|,|S(x)-a_{\pm}|\leq Ce^{\mp my},\quad|S^{\prime}(x)|\leq Ce^{-m|y|},

and the energy of SS is finite:

[S(x)2+F(S(x))]dx<.\int_{\mathbb{R}}[S^{\prime}(x)^{2}+F(S(x))]\,\mathrm{d}x<\infty.

We now prove the existence of a static kink T(y)T(y) for our equation (1.6):

Proof of Theorem 1.1.

Let SS be the stationary solution to S′′+F(S)=0-S^{\prime\prime}+F^{\prime}(S)=0 guaranteed by Lemma 2.1. Making the ansatz T=S+SbT=S+S_{b}, we have the following equation for SbS_{b}:

(2.1) Sb′′bSbcSb=bS+cSF(S+Sb)+F(S)=bS+cSF′′(S)Sb𝒩(S,Sb),\begin{split}-S_{b}^{\prime\prime}-bS_{b}^{\prime}-cS_{b}&=bS^{\prime}+cS-F^{\prime}(S+S_{b})+F^{\prime}(S)\\ &=bS^{\prime}+cS-F^{\prime\prime}(S)S_{b}-\mathcal{N}(S,S_{b}),\end{split}

where 𝒩(S,Sb)=F(S+Sb)F(S)F′′(S)Sb\mathcal{N}(S,S_{b})=F^{\prime}(S+S_{b})-F^{\prime}(S)-F^{\prime\prime}(S)S_{b}. Defining

b=y2b(y)yc(y)+F′′(S)(y)=Sb(y)yc(y),\mathcal{L}_{b}=-\partial_{y}^{2}-b(y)\partial_{y}-c(y)+F^{\prime\prime}(S)(y)=\mathcal{L}_{S}-b(y)\partial_{y}-c(y),

equation (2.1) becomes

(2.2) bSb=bS+cS𝒩(S,Sb).\mathcal{L}_{b}S_{b}=bS^{\prime}+cS-\mathcal{N}(S,S_{b}).

We can find solutions Y,YY_{-\infty},Y_{\infty} both satisfying bY±=0\mathcal{L}_{b}Y_{\pm\infty}=0, with limyY=0\lim_{y\to-\infty}Y_{-\infty}=0 and limyY=0\lim_{y\to\infty}Y_{\infty}=0. In more detail, bY=0\mathcal{L}_{b}Y=0 may be written as the linear system 𝐘=(M1+M2(y))𝐘\mathbf{Y}^{\prime}=(M_{1}+M_{2}(y))\mathbf{Y}, with 𝐘=(Y,Y)\mathbf{Y}=(Y,Y^{\prime}), and

M1=(01m20),M2(y)=(00c(y)+F′′(S)(y)m2b(y)).M_{1}=\left(\begin{array}[]{cc}0&1\\ m^{2}&0\end{array}\right),\quad M_{2}(y)=\left(\begin{array}[]{cc}0&0\\ -c(y)+F^{\prime\prime}(S)(y)-m^{2}&-b(y)\end{array}\right).

Lemma A.2 below implies existence of YY_{\infty} and YY_{-\infty}. In particular, YY_{\infty} and YY_{-\infty} are linearly independent, since otherwise there would be a nontrivial solution in L2L^{2} to bY=0\mathcal{L}_{b}Y=0, contradicting our assumption that 0 is not an eigenvalue.

Define the Green’s function

G(y,w):=1W𝐘(y){Y(y)Y(w),y<w,Y(y)Y(w),wy,G(y,w):=\frac{1}{W_{\mathbf{Y}}(y)}\begin{cases}Y_{-\infty}(y)Y_{\infty}(w),&y<w,\\ Y_{\infty}(y)Y_{-\infty}(w),&w\leq y,\end{cases}

where W𝐘(y)=det(𝐘,𝐘)W_{\mathbf{Y}}(y)=\det(\mathbf{Y}_{-\infty},\mathbf{Y}_{\infty}). Abel’s formula implies W𝐘(y)=W𝐘(0)exp(0yb(z)dz)W_{\mathbf{Y}}(y)=W_{\mathbf{Y}}(0)\exp(\int_{0}^{y}b(z)\,\mathrm{d}z), which for δ>0\delta>0 sufficiently small, is bounded uniformly away from 0.

For the inverse operator ηG(,w)η(w)dw\eta\mapsto\int_{\mathbb{R}}G(\cdot,w)\eta(w)\,\mathrm{d}w, we have the following useful bounds. First,

(2.3) |G(y,w)η(w)dw|=|Y(y)yY(w)W𝐘(w)η(w)dw+Y(y)yY(w)W𝐘(w)η(w)dw|CηL()(emyyemwdw+emyyemwdw)CηL(),\begin{split}\left|\int_{\mathbb{R}}G(y,w)\eta(w)\,\mathrm{d}w\right|&=\left|Y_{\infty}(y)\int_{-\infty}^{y}\frac{Y_{-\infty}(w)}{W_{\mathbf{Y}}(w)}\eta(w)\,\mathrm{d}w+Y_{-\infty}(y)\int_{y}^{\infty}\frac{Y_{\infty}(w)}{W_{\mathbf{Y}}(w)}\eta(w)\,\mathrm{d}w\right|\\ &\leq C\|\eta\|_{L^{\infty}(\mathbb{R})}\left(e^{-my}\int_{-\infty}^{y}e^{mw}\,\mathrm{d}w+e^{my}\int_{y}^{\infty}e^{-mw}\,\mathrm{d}w\right)\\ &\leq C\|\eta\|_{L^{\infty}(\mathbb{R})},\end{split}

for all yy\in\mathbb{R}. We also have

(2.4) |G(y,w)η(w)dwdy|C(emyyemw|η(w)|dwdy+emyyemw|η(w)|dwdy).\left|\int_{\mathbb{R}}\int_{\mathbb{R}}G(y,w)\eta(w)\,\mathrm{d}w\,\mathrm{d}y\right|\leq C\left(\int_{\mathbb{R}}e^{-my}\int_{-\infty}^{y}e^{mw}|\eta(w)|\,\mathrm{d}w\,\mathrm{d}y+\int_{\mathbb{R}}e^{my}\int_{y}^{\infty}e^{-mw}|\eta(w)|\,\mathrm{d}w\,\mathrm{d}y\right).

For the first term on the right, we integrate by parts to obtain

emyyemw|η(w)|dwdy=yem(wy)m|η(w)|dw|y=y=emymemy|η(y)|dy.\int_{\mathbb{R}}e^{-my}\int_{-\infty}^{y}e^{mw}|\eta(w)|\,\mathrm{d}w\,\mathrm{d}y=\int_{-\infty}^{y}\frac{e^{m(w-y)}}{-m}|\eta(w)|\,\mathrm{d}w\Big{|}_{y=-\infty}^{y=\infty}-\int_{\mathbb{R}}\frac{e^{-my}}{-m}e^{my}|\eta(y)|\,\mathrm{d}y.

If ηL1()\eta\in L^{1}(\mathbb{R}), then since em(wy)1e^{m(w-y)}\leq 1, the boundary term at -\infty vanishes, and the boundary term at \infty is bounded by 1mηL1()\frac{1}{m}\|\eta\|_{L^{1}(\mathbb{R})}. After applying a similar calculation to the last term in (2.4), we conclude

(2.5) G(,w)η(w)dwL1()CηL1(),\left\|\int_{\mathbb{R}}G(\cdot,w)\eta(w)\,\mathrm{d}w\right\|_{L^{1}(\mathbb{R})}\leq C\|\eta\|_{L^{1}(\mathbb{R})},

for a constant depending on mm and the coefficients b,cb,c. The estimates (2.3) and (2.5) clearly hold also if we replace G(y,w)G(y,w) with |G(y,w)||G(y,w)|.

In addition, using |Y±(y)|emy|Y_{\pm\infty}^{\prime}(y)|\lesssim e^{\mp my}, estimates similar to (2.3) and (2.5) imply

(2.6) yG(y,w)η(w)dwL1()CηL1(),yG(y,w)η(w)dwL()CηL().\left\|\partial_{y}\int_{\mathbb{R}}G(y,w)\eta(w)\,\mathrm{d}w\right\|_{L^{1}(\mathbb{R})}\leq C\|\eta\|_{L^{1}(\mathbb{R})},\quad\left\|\partial_{y}\int_{\mathbb{R}}G(y,w)\eta(w)\,\mathrm{d}w\right\|_{L^{\infty}(\mathbb{R})}\leq C\|\eta\|_{L^{\infty}(\mathbb{R})}.

Now we write the equation (2.2) for SbS_{b} as

(2.7) Sb(y)=(𝒯Sb)(y):=g(y)G(y,w)𝒩(S,Sb)dw,S_{b}(y)=(\mathcal{T}S_{b})(y):=g(y)-\int_{\mathbb{R}}G(y,w)\mathcal{N}(S,S_{b})\,\mathrm{d}w,

where

g(y)=G(y,w)[b(w)S(w)+c(w)S(w)]dw.g(y)=\int_{\mathbb{R}}G(y,w)[b(w)S^{\prime}(w)+c(w)S(w)]\,\mathrm{d}w.

We want to find a fixed point for 𝒯\mathcal{T} in the space X:=L1()L()X:=L^{1}(\mathbb{R})\cap L^{\infty}(\mathbb{R}) with norm X:=L1()+L()\|\cdot\|_{X}:=\|\cdot\|_{L^{1}(\mathbb{R})}+\|\cdot\|_{L^{\infty}(\mathbb{R})}. From (2.3) and (2.5), we have

gXC(bS+cSL()+bS+cSL1())C0δ,\|g\|_{X}\leq C\left(\|bS^{\prime}+cS\|_{L^{\infty}(\mathbb{R})}+\|bS^{\prime}+cS\|_{L^{1}(\mathbb{R})}\right)\leq C_{0}\delta,

since SS and SS^{\prime} are bounded and b+cXδ\|b+c\|_{X}\lesssim\delta. For the nonlinear term, since FF is C3C^{3} on [a,a+][a_{-},a_{+}], there is some K>0K>0 such that

(2.8) |𝒩(S,η)|=|F(S+η)F(S)F′′(S)η|Kη2,|\mathcal{N}(S,\eta)|=|F^{\prime}(S+\eta)-F^{\prime}(S)-F^{\prime\prime}(S)\eta|\leq K\eta^{2},

globally in yy. This gives

(2.9) |G(y,w)𝒩(S,η)(w)dw|K|G(y,w)|η2(w)dw\begin{split}\left|\int_{\mathbb{R}}G(y,w)\mathcal{N}(S,\eta)(w)\,\mathrm{d}w\right|&\leq K\int_{\mathbb{R}}|G(y,w)|\eta^{2}(w)\,\mathrm{d}w\end{split}

and

|G(y,w)𝒩(S,η)(w)dwdy|K(emyyemwη2(w)dw+emyyemwη2(w)dw)dy,\begin{split}\left|\int_{\mathbb{R}}\int_{\mathbb{R}}G(y,w)\mathcal{N}(S,\eta)(w)\,\mathrm{d}w\,\mathrm{d}y\right|&\leq K\int_{\mathbb{R}}\left(e^{-my}\int_{-\infty}^{y}e^{mw}\eta^{2}(w)\,\mathrm{d}w+e^{my}\int_{y}^{\infty}e^{-mw}\eta^{2}(w)\,\mathrm{d}w\right)\,\mathrm{d}y,\end{split}

so that the estimates (2.3) and (2.5) imply

G(y,w)𝒩(S,η)(w)dwXCKη2XCKηX2,\left\|\int_{\mathbb{R}}G(y,w)\mathcal{N}(S,\eta)(w)\,\mathrm{d}w\right\|_{X}\leq CK\|\eta^{2}\|_{X}\leq CK\|\eta\|_{X}^{2},

after applying the standard interpolation L22LL1\|\cdot\|_{L^{2}}^{2}\leq\|\cdot\|_{L^{\infty}}\|\cdot\|_{L^{1}}.

With C0C_{0} such that gXC0δ\|g\|_{X}\leq C_{0}\delta, define 𝒜:={ηX,ηX2C0δ}\mathcal{A}:=\{\eta\in X,\|\eta\|_{X}\leq 2C_{0}\delta\}. For any η𝒜\eta\in\mathcal{A}, the above estimates imply

𝒯ηX\displaystyle\|\mathcal{T}\eta\|_{X} =gG(,w)𝒩(S,η)(w)dwXgX+CKηX2C0δ+Cδ2,\displaystyle=\left\|g-\int_{\mathbb{R}}G(\cdot,w)\mathcal{N}(S,\eta)(w)\,\mathrm{d}w\right\|_{X}\leq\|g\|_{X}+CK\|\eta\|_{X}^{2}\leq C_{0}\delta+C\delta^{2},

so for δ<C0/C\delta<C_{0}/C, we have Tη𝒜T\eta\in\mathcal{A}. Next, for η1,η2𝒜\eta_{1},\eta_{2}\in\mathcal{A}, we have from Taylor’s Theorem that

F(S+η1)=F(S+η2)+F′′(S+η1)(η1η2)+12F′′′(ξy)(η1η2)2,F^{\prime}(S+\eta_{1})=F^{\prime}(S+\eta_{2})+F^{\prime\prime}(S+\eta_{1})(\eta_{1}-\eta_{2})+\frac{1}{2}F^{\prime\prime\prime}(\xi_{y})(\eta_{1}-\eta_{2})^{2},

for some ξy[a,a+]\xi_{y}\in[a_{-},a_{+}] depending on yy. Using this in 𝒩(S,η1)𝒩(S,η2)\mathcal{N}(S,\eta_{1})-\mathcal{N}(S,\eta_{2}), we have

|𝒩(S,η1)𝒩(S,η2)|=|F(S+η1)F(S+η2)F′′(S)(η1η2)|=|[F′′(S+η1)F′′(S)](η1η2)+12F′′′(ξy)(η1η2)2||max[a,a+]|F′′′(s)η1η1η2|+12|F′′′(ξy)|(η1η2)2Kδ|η1η2|,\begin{split}|\mathcal{N}(S,\eta_{1})-\mathcal{N}(S,\eta_{2})|&=|F^{\prime}(S+\eta_{1})-F^{\prime}(S+\eta_{2})-F^{\prime\prime}(S)(\eta_{1}-\eta_{2})|\\ &=\left|[F^{\prime\prime}(S+\eta_{1})-F^{\prime\prime}(S)](\eta_{1}-\eta_{2})+\frac{1}{2}F^{\prime\prime\prime}(\xi_{y})(\eta_{1}-\eta_{2})^{2}\right|\\ &\leq|\max_{[a_{-},a_{+}]}|F^{\prime\prime\prime}(s)||\eta_{1}||\eta_{1}-\eta_{2}|+\frac{1}{2}|F^{\prime\prime\prime}(\xi_{y})|(\eta_{1}-\eta_{2})^{2}\\ &\leq K\delta|\eta_{1}-\eta_{2}|,\end{split}

for some K>0K>0. By (2.3) and (2.5) we have

(2.10) (𝒯η1)(y)(𝒯η2)(y)X=G(y,w)[𝒩(S,η1)𝒩(S,η2)](w)dwXCKδη1η2X,\begin{split}\|(\mathcal{T}\eta_{1})(y)-(\mathcal{T}\eta_{2})(y)\|_{X}&=\left\|\int_{\mathbb{R}}G(y,w)[\mathcal{N}(S,\eta_{1})-\mathcal{N}(S,\eta_{2})](w)\,\mathrm{d}w\right\|_{X}\\ &\leq CK\delta\|\eta_{1}-\eta_{2}\|_{X},\end{split}

as above. The constant CK>0CK>0 depends on mm and the C3C^{3} norm of FF. For δ\delta sufficiently small, we conclude 𝒯\mathcal{T} is a contraction on 𝒜\mathcal{A}, and a unique solution SbS_{b} to (2.7) exists in 𝒜\mathcal{A}.

To derive the bounds on SbS_{b}^{\prime}, we differentiate equation (2.7) and use the derivative bounds (2.6) and the Taylor estimate (2.8):

SbX=yG(y,w)[bS+cSN(S,Sb)]dwXbS+cSN(S,Sb)Xδ+KSb2Xδ.\begin{split}\|S_{b}^{\prime}\|_{X}&=\left\|\partial_{y}\int_{\mathbb{R}}G(y,w)[bS^{\prime}+cS-N(S,S_{b})]\,\mathrm{d}w\right\|_{X}\leq\|bS^{\prime}+cS-N(S,S_{b})\|_{X}\lesssim\delta+K\|S_{b}^{2}\|_{X}\lesssim\delta.\end{split}

The proof of Theorem 1.1 also provides the following approximation for Sb=TSS_{b}=T-S: since Sb=g+G(y,w)𝒩(S,Sb)(w)dwS_{b}=g+\int_{\mathbb{R}}G(y,w)\mathcal{N}(S,S_{b})(w)\,\mathrm{d}w, the estimates (2.3), (2.5) imply

(2.11) SbG~(,w)[bS+cS](w)dwXG~(,w)𝒩(S,Sb)(w)dwXSbX2δ2,\begin{split}\left\|S_{b}-\int_{\mathbb{R}}\tilde{G}(\cdot,w)[bS^{\prime}+cS](w)\,\mathrm{d}w\right\|_{X}&\leq\left\|\int_{\mathbb{R}}\tilde{G}(\cdot,w)\mathcal{N}(S,S_{b})(w)\,\mathrm{d}w\right\|_{X}\\ &\lesssim\|S_{b}\|_{X}^{2}\lesssim\delta^{2},\end{split}

with G~=G\tilde{G}=G if 0 is not an eigenvalue of b=Sbyc\mathcal{L}_{b}=\mathcal{L}_{S}-b\partial_{y}-c, and G~=Gλ\tilde{G}=G_{\lambda} otherwise, with λ\lambda chosen such that |λ|δ|\lambda|\lesssim\delta, so that λSbλGλ(y,w)Sb(w)dw\lambda S_{b}-\lambda\int_{\mathbb{R}}G_{\lambda}(y,w)S_{b}(w)\,\mathrm{d}w are O(δ2)O(\delta^{2}).

3. Perturbation of the spectrum

We consider the spectrum of

(3.1) T:=y2byc+F′′(T),\mathcal{L}_{T}:=-\partial_{y}^{2}-b\partial_{y}-c+F^{\prime\prime}(T),

where TT is the stationary solution guaranteed by Theorem 1.1. Defining d=c+F′′(T)F′′(S)d=-c+F^{\prime\prime}(T)-F^{\prime\prime}(S), we have T=Sby+d\mathcal{L}_{T}=\mathcal{L}_{S}-b\partial_{y}+d. By the C3C^{3} regularity of FF, we have |F′′(T)F′′(S)|K|Sb||F^{\prime\prime}(T)-F^{\prime\prime}(S)|\leq K|S_{b}|, and Theorem 1.1 implies dL1()+dL()δ\|d\|_{L^{1}(\mathbb{R})}+\|d\|_{L^{\infty}(\mathbb{R})}\lesssim\delta. With (2.11), we can also write a first-order approximation for dd as follows:

(3.2) d(y)=c(y)+F′′′(S)(y)G~(y,w)[bS+cS](w)dw+ε(y),d(y)=-c(y)+F^{\prime\prime\prime}(S)(y)\int_{\mathbb{R}}\tilde{G}(y,w)[bS^{\prime}+cS](w)\,\mathrm{d}w+\varepsilon(y),

with G~\tilde{G} as in (2.11) and ε(y)=o(Sb(y))\varepsilon(y)=o(S_{b}(y)).

Our goal is to investigate how the spectrum of S\mathcal{L}_{S} changes under the perturbation by+d-b\partial_{y}+d. Since T\mathcal{L}_{T} is self-adjoint with respect to the inner product

f,gω:=ωfgdy,\langle f,g\rangle_{\omega}:=\int_{\mathbb{R}}\omega fg\,\mathrm{d}y,

with ω=yb(z)dz\omega=\int_{-\infty}^{y}b(z)\,\mathrm{d}z, the spectrum σ(T)\sigma(\mathcal{L}_{T}) is real. Since the perturbation is relatively \mathcal{L}-compact, we have σess(T)=σess(S)\sigma_{ess}(\mathcal{L}_{T})=\sigma_{ess}(\mathcal{L}_{S}). (See, e.g. [21, Chapter 14].) Given our upper bounds on bb and dd, it is standard that σ(T)\sigma(\mathcal{L}_{T}) lies in the c0δc_{0}\delta neighborhood of σ(S)\sigma(\mathcal{L}_{S}), for some c0>0c_{0}>0. (An elementary argument to this effect can be found in the proof of Theorem 3.1 in [46].) This already establishes part (a) of Theorem 1.2.

To analyze the eigenvalue problem, we write the equation (Sλ)Yλ=0(\mathcal{L}_{S}-\lambda)Y^{\lambda}=0 in vector form:

(𝐘λ)(y)=((01m2λ0)+(00F′′(S)m20))𝐘λ(y).(\mathbf{Y}^{\lambda})^{\prime}(y)=\left(\left(\begin{array}[]{cc}0&1\\ m^{2}-\lambda&0\end{array}\right)+\left(\begin{array}[]{cc}0&0\\ F^{\prime\prime}(S)-m^{2}&0\end{array}\right)\right)\mathbf{Y}^{\lambda}(y).

For any λm2\lambda\leq m^{2}, Lemma A.2(a) implies there exist Yλ,YλLloc()Y_{\infty}^{\lambda},Y_{-\infty}^{\lambda}\in L^{\infty}_{\rm loc}(\mathbb{R}) satisfying (Sλ)Y±λ=0(\mathcal{L}_{S}-\lambda)Y_{\pm\infty}^{\lambda}=0, and

(3.3) limy±e±ky𝐘±λ(y)=(1k),\lim_{y\to\pm\infty}e^{\pm ky}\mathbf{Y}_{\pm\infty}^{\lambda}(y)=\left(\begin{array}[]{c}1\\ \mp k\end{array}\right),

with k=m2λk=\sqrt{m^{2}-\lambda}. For λ<m2\lambda<m^{2}, we also obtain the integral representations

(3.4) eky𝐘λ=(1k)12y(F′′(S)m2)Yλ(w)ekw((e2k(yw)1)/ke2k(yw)+1)dweky𝐘λ=(1k)12y(F′′(S)m2)Yλ(w)ekw((e2k(yw)1)/ke2k(yw)+1)dw.\begin{split}e^{ky}\mathbf{Y}_{\infty}^{\lambda}&=\left(\begin{array}[]{c}1\\ -k\end{array}\right)-\frac{1}{2}\int_{y}^{\infty}(F^{\prime\prime}(S)-m^{2})Y_{\infty}^{\lambda}(w)e^{kw}\left(\begin{array}[]{c}(e^{2k(y-w)}-1)/k\\ e^{2k(y-w)}+1\end{array}\right)\,\mathrm{d}w\\ e^{-ky}\mathbf{Y}_{-\infty}^{\lambda}&=\left(\begin{array}[]{c}1\\ k\end{array}\right)-\frac{1}{2}\int_{-\infty}^{y}(F^{\prime\prime}(S)-m^{2})Y_{-\infty}^{\lambda}(w)e^{-kw}\left(\begin{array}[]{c}(e^{2k(y-w)}-1)/k\\ e^{2k(y-w)}+1\end{array}\right)\,\mathrm{d}w.\end{split}

For the operator T\mathcal{L}_{T}, we similarly apply Lemma A.2(a) with V=F′′(S)m2+dV=F^{\prime\prime}(S)-m^{2}+d to obtain U±λU_{\pm\infty}^{\lambda} solving (Tλ)U±λ=0(\mathcal{L}_{T}-\lambda)U_{\pm\infty}^{\lambda}=0, with the same boundary conditions (3.3), and for λ<m2\lambda<m^{2},

(3.5) eky𝐔λ=(1k)12y[(F′′(S)m2+d)Uλb(Uλ)]ekw((e2k(yw)1)/ke2k(yw)+1)dweky𝐔λ=(1k)12y[(F′′(S)m2+d)Uλb(Uλ)]ekw((e2k(yw)1)/ke2k(yw)+1)dw.\begin{split}e^{ky}\mathbf{U}_{\infty}^{\lambda}&=\left(\begin{array}[]{c}1\\ -k\end{array}\right)-\frac{1}{2}\int_{y}^{\infty}[(F^{\prime\prime}(S)-m^{2}+d)U_{\infty}^{\lambda}-b(U_{\infty}^{\lambda})^{\prime}]e^{kw}\left(\begin{array}[]{c}(e^{2k(y-w)}-1)/k\\ e^{2k(y-w)}+1\end{array}\right)\,\mathrm{d}w\\ e^{-ky}\mathbf{U}_{-\infty}^{\lambda}&=\left(\begin{array}[]{c}1\\ k\end{array}\right)-\frac{1}{2}\int_{-\infty}^{y}[(F^{\prime\prime}(S)-m^{2}+d)U_{-\infty}^{\lambda}-b(U_{-\infty}^{\lambda})^{\prime}]e^{-kw}\left(\begin{array}[]{c}(e^{2k(y-w)}-1)/k\\ e^{2k(y-w)}+1\end{array}\right)\,\mathrm{d}w.\end{split}

First, we prove a suitable approximation lemma for Y±λY_{\pm\infty}^{\lambda} and U±λU_{\pm\infty}^{\lambda} for nearby values of λ\lambda:

Lemma 3.1.

Assume dL1()+bL1()1\|d\|_{L^{1}(\mathbb{R})}+\|b\|_{L^{1}(\mathbb{R})}\leq 1. For any compact subset BB of (,m2)(-\infty,m^{2}), there exists a constant C>0C>0 such that for any λ1,λ2B\lambda_{1},\lambda_{2}\in B, there holds

ek1y𝐘λ1ek2y𝐔λ2L([0,),2)C(|λ1λ2|+dL1()+bL1()),ek1y𝐘λ1ek2y𝐔λ2L((,0],2)C(|λ1λ2|+dL1()+bL1()),\begin{split}\|e^{k_{1}y}\mathbf{Y}_{\infty}^{\lambda_{1}}-e^{k_{2}y}\mathbf{U}_{\infty}^{\lambda_{2}}\|_{L^{\infty}([0,\infty),\mathbb{R}^{2})}&\leq C(|\lambda_{1}-\lambda_{2}|+\|d\|_{L^{1}(\mathbb{R})}+\|b\|_{L^{1}(\mathbb{R})}),\\ \|e^{-k_{1}y}\mathbf{Y}_{-\infty}^{\lambda_{1}}-e^{-k_{2}y}\mathbf{U}_{-\infty}^{\lambda_{2}}\|_{L^{\infty}((-\infty,0],\mathbb{R}^{2})}&\leq C(|\lambda_{1}-\lambda_{2}|+\|d\|_{L^{1}(\mathbb{R})}+\|b\|_{L^{1}(\mathbb{R})}),\end{split}

where ki=m2λik_{i}=\sqrt{m^{2}-\lambda_{i}}.

Proof.

We prove only the first estimate, as the second follows by a similar argument.

From (3.4) and (3.5) we have

ek1y𝐘λ1(y)ek2y𝐔λ2(y)=(0k2k1)12y[(F′′(S)m2)((e2k1(yw)1)/k1e2k1(yw)+1)Yλ1(w)ek1w((e2k2(yw)1)/k2e2k2(yw)+1)[(F′′(S)m2+d)Uλ2(w)b(Uλ2)ek2w]]dw=𝐉1(y)+𝐉2(y)12y(F′′(S)m2)((e2k1(yw)1)/k1e2k1(yw)+1)(ek1wYλ1(w)ek2wUλ2(w))dw,\begin{split}e^{k_{1}y}\mathbf{Y}_{\infty}^{\lambda_{1}}&(y)-e^{k_{2}y}\mathbf{U}_{\infty}^{\lambda_{2}}(y)\\ &=\left(\begin{array}[]{c}0\\ k_{2}-k_{1}\end{array}\right)-\frac{1}{2}\int_{y}^{\infty}\left[(F^{\prime\prime}(S)-m^{2})\left(\begin{array}[]{c}(e^{2k_{1}(y-w)}-1)/k_{1}\\ e^{2k_{1}(y-w)}+1\end{array}\right)Y_{\infty}^{\lambda_{1}}(w)e^{k_{1}w}\right.\\ &\qquad\qquad\left.-\left(\begin{array}[]{c}(e^{2k_{2}(y-w)}-1)/k_{2}\\ e^{2k_{2}(y-w)}+1\end{array}\right)[(F^{\prime\prime}(S)-m^{2}+d)U_{\infty}^{\lambda_{2}}(w)-b(U_{\infty}^{\lambda_{2}})^{\prime}e^{k_{2}w}]\right]\,\mathrm{d}w\\ &=\mathbf{J}_{1}(y)+\mathbf{J}_{2}(y)\\ &\quad-\frac{1}{2}\int_{y}^{\infty}(F^{\prime\prime}(S)-m^{2})\left(\begin{array}[]{c}(e^{2k_{1}(y-w)}-1)/k_{1}\\ e^{2k_{1}(y-w)}+1\end{array}\right)(e^{k_{1}w}Y_{\infty}^{\lambda_{1}}(w)-e^{k_{2}w}U_{\infty}^{\lambda_{2}}(w))\,\mathrm{d}w,\end{split}

where

𝐉1(y):=(0k2k1)12y(F′′(S)m2)ek2wUλ2(w)((e2k1(yw)1)/k1(e2k2(yw)1)/k2e2k1(yw)e2k2(yw))dw𝐉2(y):=12y((e2k2(yw)1)/k2e2k2(yw)+1)[dUλ2b(Uλ2)]ek2wdw.\begin{split}\mathbf{J}_{1}(y)&:=\left(\begin{array}[]{c}0\\ k_{2}-k_{1}\end{array}\right)\\ &\quad-\frac{1}{2}\int_{y}^{\infty}(F^{\prime\prime}(S)-m^{2})e^{k_{2}w}U_{\infty}^{\lambda_{2}}(w)\left(\begin{array}[]{c}(e^{2k_{1}(y-w)}-1)/k_{1}-(e^{2k_{2}(y-w)}-1)/k_{2}\\ e^{2k_{1}(y-w)}-e^{2k_{2}(y-w)}\end{array}\right)\,\mathrm{d}w\\ \mathbf{J}_{2}(y)&:=\frac{1}{2}\int_{y}^{\infty}\left(\begin{array}[]{c}(e^{2k_{2}(y-w)}-1)/k_{2}\\ e^{2k_{2}(y-w)}+1\end{array}\right)[dU_{\infty}^{\lambda_{2}}-b(U_{\infty}^{\lambda_{2}})^{\prime}]e^{k_{2}w}\,\mathrm{d}w.\end{split}

Since yw0y-w\leq 0, the mean value theorem applied to xe2x(yw)x\mapsto e^{2x(y-w)} and x(e2x(yw)1)/xx\mapsto(e^{2x(y-w)}-1)/x implies, after a straightforward calculation, the inequalities

|((e2k1(yw)1)/k1(e2k2(yw)1)/k2e2k1(yw)e2k2(yw))|C(1+|yw|)|k1k2|,\left|\left(\begin{array}[]{c}(e^{2k_{1}(y-w)}-1)/k_{1}-(e^{2k_{2}(y-w)}-1)/k_{2}\\ e^{2k_{1}(y-w)}-e^{2k_{2}(y-w)}\end{array}\right)\right|\leq C(1+|y-w|)|k_{1}-k_{2}|,

for a constant CC depending on BB. Since |F′′(S)m2|em|w||F^{\prime\prime}(S)-m^{2}|\leq e^{-m|w|} and ek2wUλ2(w)e^{k_{2}w}U_{\infty}^{\lambda_{2}}(w) is uniformly bounded on [0,)[0,\infty), we therefore have 𝐉1L([0,),2)C|k1k2|\|\mathbf{J}_{1}\|_{L^{\infty}([0,\infty),\mathbb{R}^{2})}\leq C|k_{1}-k_{2}|.

For 𝐉2\mathbf{J}_{2}, since ek2w𝐔λ2e^{k_{2}w}\mathbf{U}_{\infty}^{\lambda_{2}} is bounded on [0,)[0,\infty), we have 𝐉2L([0,),2)Cd+bL1()\|\mathbf{J}_{2}\|_{L^{\infty}([0,\infty),\mathbb{R}^{2})}\leq C\|d+b\|_{L^{1}(\mathbb{R})}, for a constant depending only on k2k_{2}.

Define the integral kernel

K(y,w)=12(F′′(S)m2)((e2k1(yw)1)/k10e2k1(yw)+10),K(y,w)=-\frac{1}{2}(F^{\prime\prime}(S)-m^{2})\left(\begin{array}[]{cc}(e^{2k_{1}(y-w)}-1)/k_{1}&0\\ e^{2k_{1}(y-w)}+1&0\end{array}\right),

From the exponential decay of F′′(S)m2F^{\prime\prime}(S)-m^{2} we see that

0sup0<y<wK(y,w)dw\int_{0}^{\infty}\sup_{0<y<w}\|K(y,w)\|\,\mathrm{d}w

is bounded by a constant depending only on k1k_{1} and mm. Lemma A.1 then implies

(3.6) ek1y𝐘λ1(y)ek2y𝐔λ2(y)L([0,),2)C𝐉1+𝐉2L([0,),2)C(|k1k2|+dL1()+bL1()).\|e^{k_{1}y}\mathbf{Y}_{\infty}^{\lambda_{1}}(y)-e^{k_{2}y}\mathbf{U}_{\infty}^{\lambda_{2}}(y)\|_{L^{\infty}([0,\infty),\mathbb{R}^{2})}\leq C\|\mathbf{J}_{1}+\mathbf{J}_{2}\|_{L^{\infty}([0,\infty),\mathbb{R}^{2})}\leq C(|k_{1}-k_{2}|+\|d\|_{L^{1}(\mathbb{R})}+\|b\|_{L^{1}(\mathbb{R})}).

Since |k1k2|C|λ1λ2||k_{1}-k_{2}|\leq C|\lambda_{1}-\lambda_{2}| for a constant depending only on KK, the proof is complete. ∎

Now, we are ready to derive a result that governs the direction in which eigenvalues of S\mathcal{L}_{S} drift under the perturbation:

Theorem 3.2.

Assume dL1()+bL1()δ1\|d\|_{L^{1}(\mathbb{R})}+\|b\|_{L^{1}(\mathbb{R})}\leq\delta\ll 1. For any eigenvalue λ<m2\lambda_{*}<m^{2} of S\mathcal{L}_{S} with eigenfunction YY_{*}, there exists a simple, real eigenvalue λ\lambda of T\mathcal{L}_{T} with |λλ|Cδ|\lambda-\lambda_{*}|\leq C\delta. Furthermore, we have the following expansion for λ\lambda:

λ=λ+Y[dYbY]dy(Y)2dy+O(δ2).\lambda=\lambda_{*}+\frac{\int_{\mathbb{R}}Y_{*}[dY_{*}-bY_{*}^{\prime}]\,\mathrm{d}y}{\int_{\mathbb{R}}(Y_{*})^{2}\,\mathrm{d}y}+O(\delta^{2}).

In particular, if

A:=Y[dYbY]dy0,A:=\int_{\mathbb{R}}Y_{*}[dY_{*}-bY_{*}^{\prime}]\,\mathrm{d}y\neq 0,

then λλ\lambda-\lambda_{*} has the same sign as AA.

Remark.

Using the formula (3.2), one can show that AA has the same sign as

(3.7) [Y2(y)(c(y)+F′′′(S)(y)G~(y,w)[bS+cS](w)dw)b(y)Y(y)Y(y)]dy.\int_{\mathbb{R}}\left[Y_{*}^{2}(y)\left(-c(y)+F^{\prime\prime\prime}(S)(y)\int_{\mathbb{R}}\tilde{G}(y,w)[bS^{\prime}+cS](w)\,\mathrm{d}w\right)-b(y)Y_{*}^{\prime}(y)Y_{*}(y)\right]\,\mathrm{d}y.
Proof.

With Y±λY_{\pm\infty}^{\lambda_{*}} solving (3.4), since λ\lambda_{*} is a simple eigenvalue, we have Y±λ=c±YY_{\pm\infty}^{\lambda_{*}}=c_{\pm}Y_{*}, for constants c±c_{\pm}. Let k=m2λk_{*}=\sqrt{m^{2}-\lambda_{*}}. From our construction, it is clear that YY_{*} decays exponentially at a rate Y(y)ek|y|Y_{*}(y)\lesssim e^{-k_{*}|y|}.

For λ\lambda near λ\lambda_{*} let k=m2λk=\sqrt{m^{2}-\lambda} and let U±λU_{\pm\infty}^{\lambda} be the solutions to (3.5) as above. From Lemma 3.1 and our assumption that dL1()+bL1()δ\|d\|_{L^{1}(\mathbb{R})}+\|b\|_{L^{1}(\mathbb{R})}\lesssim\delta, we can write

(3.8) 𝐔±λ(y)=e±(kk)y𝐘±λ(y)+eky𝐄±(y),\begin{split}\mathbf{U}_{\pm\infty}^{\lambda}(y)&=e^{\pm(k_{*}-k)y}\mathbf{Y}_{\pm\infty}^{\lambda_{*}}(y)+e^{\mp ky}\mathbf{E}_{\pm\infty}(y),\end{split}

with 𝐄L([0,),2)+𝐄L((,0],2)δ\|\mathbf{E}_{\infty}\|_{L^{\infty}([0,\infty),\mathbb{R}^{2})}+\|\mathbf{E}_{-\infty}\|_{L^{\infty}((-\infty,0],\mathbb{R}^{2})}\lesssim\delta. Denote the Wronskian

W𝐔(λ,y)=det(𝐔λ(y),𝐔λ(y)).W_{\mathbf{U}}(\lambda,y)=\det(\mathbf{U}_{\infty}^{\lambda}(y),\mathbf{U}_{-\infty}^{\lambda}(y)).

By Abel’s formula, exp(yb(z)dz)W𝐔(λ,y)\exp(\int_{-\infty}^{y}b(z)\,\mathrm{d}z)W_{\mathbf{U}}(\lambda,y) is independent of yy. We focus on y=0y=0 and apply (3.8) to obtain

(3.9) W𝐔(λ,0)=det(𝐘λ(0)+𝐄(0),𝐘λ(0)+𝐄(0))=det(𝐘λ(0),𝐄(0))+det(𝐄(0),𝐘λ(0))+det(𝐄(0),𝐄(0))=det(𝐘λ(0),𝐔λ(0))+det(𝐔λ(0),𝐘λ(0))+O(δ2).\begin{split}W_{\mathbf{U}}(\lambda,0)&=\det(\mathbf{Y}_{\infty}^{\lambda_{*}}(0)+\mathbf{E}_{\infty}(0),\mathbf{Y}_{-\infty}^{\lambda_{*}}(0)+\mathbf{E}_{-\infty}(0))\\ &=\det(\mathbf{Y}_{\infty}^{\lambda_{*}}(0),\mathbf{E}_{-\infty}(0))+\det(\mathbf{E}_{\infty}(0),\mathbf{Y}_{-\infty}^{\lambda_{*}}(0))+\det(\mathbf{E}_{\infty}(0),\mathbf{E}_{-\infty}(0))\\ &=\det(\mathbf{Y}_{\infty}^{\lambda_{*}}(0),\mathbf{U}_{-\infty}^{\lambda}(0))+\det(\mathbf{U}_{\infty}^{\lambda}(0),\mathbf{Y}_{-\infty}^{\lambda_{*}}(0))+O(\delta^{2}).\end{split}

In the second line, we used that 𝐘λ\mathbf{Y}_{\infty}^{\lambda_{*}} and 𝐘λ\mathbf{Y}_{-\infty}^{\lambda_{*}} are parallel, and in the last line, we used 𝐄±(0)=𝐔±λ(0)𝐘±λ(0)\mathbf{E}_{\pm\infty}(0)=\mathbf{U}_{\pm\infty}^{\lambda}(0)-\mathbf{Y}_{\pm\infty}^{\lambda_{*}}(0) and |𝐄±(0)|δ|\mathbf{E}_{\pm\infty}(0)|\lesssim\delta. Since

det(𝐘±λ,𝐔λ)(y)=Y±λ[(d+λλ)Uλb(Uλ)],\det(\mathbf{Y}_{\pm\infty}^{\lambda_{*}},\mathbf{U}_{\mp\infty}^{\lambda})^{\prime}(y)=Y_{\pm\infty}^{\lambda_{*}}\left[(d+\lambda_{*}-\lambda)U_{\mp\infty}^{\lambda}-b(U_{\mp\infty}^{\lambda})^{\prime}\right],

we can use (3.8) again to write

det(𝐘λ(0),𝐔λ(0))=0Yλ[(d+λλ)Uλb(Uλ)]dw=0e(kk)wYλ[(d+λλ)Yλb(Yλ)]dw+0ekwYλ(w)[(d+λλ)E1bE2]dw,\begin{split}\det(\mathbf{Y}_{\infty}^{\lambda_{*}}(0),\mathbf{U}_{-\infty}^{\lambda}(0))&=\int_{-\infty}^{0}Y_{\infty}^{\lambda_{*}}\left[(d+\lambda_{*}-\lambda)U_{-\infty}^{\lambda}-b(U_{-\infty}^{\lambda})^{\prime}\right]\,\mathrm{d}w\\ &=\int_{-\infty}^{0}e^{(k-k_{*})w}Y_{\infty}^{\lambda_{*}}\left[(d+\lambda_{*}-\lambda)Y_{-\infty}^{\lambda_{*}}-b(Y_{-\infty}^{\lambda_{*}})^{\prime}\right]\,\mathrm{d}w\\ &\quad+\int_{-\infty}^{0}e^{kw}Y_{\infty}^{\lambda_{*}}(w)[(d+\lambda_{*}-\lambda)E_{-\infty}^{1}-bE_{-\infty}^{2}]\,\mathrm{d}w,\end{split}

and

det(𝐔λ,𝐘λ)(0)=0Yλ[(d+λλ)Uλb(Uλ)]dw=0e(kk)wYλ[(d+λλ)Yλb(Yλ)]dw+0ekwYλ(w)[(d+λλ)E1bE2]dw.\begin{split}\det(\mathbf{U}_{\infty}^{\lambda},\mathbf{Y}_{-\infty}^{\lambda_{*}})(0)&=\int_{0}^{\infty}Y_{-\infty}^{\lambda_{*}}\left[(d+\lambda_{*}-\lambda)U_{\infty}^{\lambda}-b(U_{\infty}^{\lambda})^{\prime}\right]\,\mathrm{d}w\\ &=\int_{0}^{\infty}e^{(k_{*}-k)w}Y_{-\infty}^{\lambda_{*}}\left[(d+\lambda_{*}-\lambda)Y_{\infty}^{\lambda_{*}}-b(Y_{\infty}^{\lambda_{*}})^{\prime}\right]\,\mathrm{d}w\\ &\quad+\int_{0}^{\infty}e^{-kw}Y_{-\infty}^{\lambda_{*}}(w)[(d+\lambda_{*}-\lambda)E_{\infty}^{1}-bE_{\infty}^{2}]\,\mathrm{d}w.\end{split}

Feeding these expressions into (3.9), we obtain

W𝐔(λ,0)=c+ce(kk)|w|[(d+λλ)(Y(w))2bY(w)Y(w)]dw+O(δ2).W_{\mathbf{U}}(\lambda,0)=c_{+}c_{-}\int_{-\infty}^{\infty}e^{(k_{*}-k)|w|}[(d+\lambda_{*}-\lambda)(Y_{*}(w))^{2}-bY_{*}(w)Y_{*}^{\prime}(w)]\,\mathrm{d}w+O(\delta^{2}).

From the approximation |e(kk)|w|1||kk||w|e|kk||w||e^{(k_{*}-k)|w|}-1|\leq|k_{*}-k||w|e^{|k_{*}-k||w|} and the exponential decay of YY_{*} we have, for λ\lambda an eigenvalue of T\mathcal{L}_{T},

0=eyb(z)dzc+c[(d+λλ)(Y(w))2bY(w)Y(w)]dw+O(δ2),0=e^{\int_{-\infty}^{y}b(z)\,\mathrm{d}z}c_{+}c_{-}\int_{-\infty}^{\infty}[(d+\lambda_{*}-\lambda)(Y_{*}(w))^{2}-bY_{*}(w)Y_{*}^{\prime}(w)]\,\mathrm{d}w+O(\delta^{2}),

which implies the first-order expansion for λ\lambda in the statement of the theorem. ∎

Now, we analyze the threshold resonance RR, which is an LL^{\infty} function solving SRm2R=0\mathcal{L}_{S}R-m^{2}R=0. First, we prove a modified version of Lemma 3.1 for the borderline case m2m^{2}. This will be useful in tracking how a threshold resonance of S\mathcal{L}_{S} translates to the spectrum of T\mathcal{L}_{T}. Writing (Sm2)Yλ=0(\mathcal{L}_{S}-m^{2})Y^{\lambda}=0 in vector form as above, Lemma A.2(b) implies

(3.10) 𝐘m2(y)=(10)y(F′′(S)m2)Ym2(yw1)dw.\mathbf{Y}_{\infty}^{m^{2}}(y)=\left(\begin{array}[]{c}1\\ 0\end{array}\right)-\int_{y}^{\infty}(F^{\prime\prime}(S)-m^{2})Y_{\infty}^{m^{2}}\left(\begin{array}[]{c}y-w\\ 1\end{array}\right)\,\mathrm{d}w.
Lemma 3.3.

Assume dL1()+bL1()1\|d\|_{L^{1}(\mathbb{R})}+\|b\|_{L^{1}(\mathbb{R})}\leq 1. For any λ0<m2\lambda_{0}<m^{2}, there exists a constant C>0C>0 such that for any λ(λ0,m2)\lambda\in(\lambda_{0},m^{2}), there holds

𝐘m2eky𝐔λL([0,),2)C(k+dL1()+bL1()),𝐘m2eky𝐔λL((,0],2)C(k+dL1()+bL1()),\begin{split}\|\mathbf{Y}_{\infty}^{m^{2}}-e^{ky}\mathbf{U}_{\infty}^{\lambda}\|_{L^{\infty}([0,\infty),\mathbb{R}^{2})}&\leq C(k+\|d\|_{L^{1}(\mathbb{R})}+\|b\|_{L^{1}(\mathbb{R})}),\\ \|\mathbf{Y}_{-\infty}^{m^{2}}-e^{-ky}\mathbf{U}_{-\infty}^{\lambda}\|_{L^{\infty}((-\infty,0],\mathbb{R}^{2})}&\leq C(k+\|d\|_{L^{1}(\mathbb{R})}+\|b\|_{L^{1}(\mathbb{R})}),\end{split}

where k=m2λk=\sqrt{m^{2}-\lambda}.

Proof.

The proof is similar to Lemma 3.1, with the difference that 𝐘m2\mathbf{Y}_{\infty}^{m^{2}} satisfies the modified integral equation (3.10). From (3.10) and (3.5), we have

𝐘m2(y)eky𝐔λ(y)=(0k)12y[2(F′′(S)m2)(yw1)Ym2(w)((e2k(yw)1)/ke2k(yw)+1)[(F′′(S)m2+d)Uλ(w)b(Uλ)ekw]]dw=𝐉1(y)+𝐉2(y)12y(F′′(S)m2)((e2k(yw)1)/ke2k(yw)+1)(Ym2(w)ekwUλ(w))dw,\begin{split}\mathbf{Y}_{\infty}^{m^{2}}(y)-e^{ky}\mathbf{U}_{\infty}^{\lambda}(y)&=\left(\begin{array}[]{c}0\\ k\end{array}\right)-\frac{1}{2}\int_{y}^{\infty}\left[2(F^{\prime\prime}(S)-m^{2})\left(\begin{array}[]{c}y-w\\ 1\end{array}\right)Y_{\infty}^{m^{2}}(w)\right.\\ &\qquad\qquad\left.-\left(\begin{array}[]{c}(e^{2k(y-w)}-1)/k\\ e^{2k(y-w)}+1\end{array}\right)[(F^{\prime\prime}(S)-m^{2}+d)U_{\infty}^{\lambda}(w)-b(U_{\infty}^{\lambda})^{\prime}e^{kw}]\right]\,\mathrm{d}w\\ &=\mathbf{J}_{1}(y)+\mathbf{J}_{2}(y)-\frac{1}{2}\int_{y}^{\infty}(F^{\prime\prime}(S)-m^{2})\left(\begin{array}[]{c}(e^{2k(y-w)}-1)/k\\ e^{2k(y-w)}+1\end{array}\right)(Y_{\infty}^{m^{2}}(w)-e^{kw}U_{\infty}^{\lambda}(w))\,\mathrm{d}w,\end{split}

with

𝐉1(y):=(0k)12y(F′′(S)m2)Ym2(w)(2(yw)(e2k(yw)1)/k2(e2k(yw)+1))dw,\mathbf{J}_{1}(y):=\left(\begin{array}[]{c}0\\ k\end{array}\right)-\frac{1}{2}\int_{y}^{\infty}(F^{\prime\prime}(S)-m^{2})Y_{\infty}^{m^{2}}(w)\left(\begin{array}[]{c}2(y-w)-(e^{2k(y-w)}-1)/k\\ 2-(e^{2k(y-w)}+1)\end{array}\right)\,\mathrm{d}w,

and 𝐉2(y)\mathbf{J}_{2}(y) defined as in the proof of Lemma 3.1, with λ\lambda replacing λ2\lambda_{2}.

We claim that 𝐉1L([0,),2)C|k|\|\mathbf{J}_{1}\|_{L^{\infty}([0,\infty),\mathbb{R}^{2})}\leq C|k|. Indeed, applying the mean value theorem to f(x)=e2x(yw)f(x)=e^{2x(y-w)} gives

|f(k)f(0)||k|sup0<x<k|f(x)||k|2|yw|e2x(yw)|k|2|yw|,|f(k)-f(0)|\leq|k|\sup_{0<x<k}|f^{\prime}(x)|\leq|k|2|y-w|e^{2x(y-w)}\leq|k|2|y-w|,

or |e2k(yw)1|2|k||yw||e^{2k(y-w)}-1|\leq 2|k||y-w|. Next, Taylor’s Theorem implies f(k)=f(0)+f(0)k+εf(k)=f(0)+f^{\prime}(0)k+\varepsilon, with |ε|12k2sup0<x<k|f′′(x)||\varepsilon|\leq\frac{1}{2}k^{2}\sup_{0<x<k}|f^{\prime\prime}(x)|. We have f(0)=2(yw)f^{\prime}(0)=2(y-w) and f′′(x)=4(yw)2e2x(yw)4(yw)2f^{\prime\prime}(x)=4(y-w)^{2}e^{2x(y-w)}\leq 4(y-w)^{2}, since yw<0y-w<0. This gives e2k(yw)1=2k(yw)+εe^{2k(y-w)}-1=2k(y-w)+\varepsilon, with |ε|2k2(yw)2|\varepsilon|\leq 2k^{2}(y-w)^{2}, or

|(e2k(yw)1)/k2(yw)|2k(yw)2.|(e^{2k(y-w)}-1)/k-2(y-w)|\leq 2k(y-w)^{2}.

Plugging these inequalities into the definition of 𝐉1\mathbf{J}_{1} and using decay of F′′(S)m2F^{\prime\prime}(S)-m^{2} gives the desired estimate.

The same calculation as in the proof of Lemma 3.1 implies that the boundedness property (A.1) is satisfied for this integral equation, and Lemma A.1 establishes the conclusion of the lemma. ∎

In the following theorem, we make the (mild) assumption that the limits at ±\pm\infty of R(y)R(y) are nonzero.

Theorem 3.4.

(a) Assume that m2m^{2} is a simple resonance for S\mathcal{L}_{S}, i.e. that there exists RL()L2()R\in L^{\infty}(\mathbb{R})\setminus L^{2}(\mathbb{R}) with SR=m2R\mathcal{L}_{S}R=m^{2}R. Then there exists δ>0\delta>0 depending only on the function RR, such that if bL1()+dL1()δ\|b\|_{L^{1}(\mathbb{R})}+\|d\|_{L^{1}(\mathbb{R})}\lesssim\delta and

R[dRbR]dw<0,\int_{\mathbb{R}}R[dR-bR^{\prime}]\,\mathrm{d}w<0,

then there exists a discrete eigenvalue λ\lambda of T\mathcal{L}_{T} with 0<m2λ<Cδ0<m^{2}-\lambda<C\delta. If

R[dRbR]dw>0,\int_{\mathbb{R}}R[dR-bR^{\prime}]\,\mathrm{d}w>0,

then there is no discrete eigenvalue of T\mathcal{L}_{T} in a neighborhood of the essential spectrum, i.e. the discrete spectrum σd(T)\sigma_{d}(\mathcal{L}_{T}) consists of the same number of eigenvalues as σd(S)\sigma_{d}(\mathcal{L}_{S}).

(b) On the other hand, if m2m^{2} is nonresonant and not an eigenvalue of S\mathcal{L}_{S}, then for δ\delta is sufficiently small, m2m^{2} cannot be a resonance or an eigenvalue of T\mathcal{L}_{T}, and there is no eigenvalue of T\mathcal{L}_{T} in a neighborhood of the essential spectrum.

Remark.

As above, using (3.2), the quantity R[dRbR]dw\int_{\mathbb{R}}R[dR-bR^{\prime}]\,\mathrm{d}w has the same sign as

(3.11) [R2(y)(c(y)+F′′′(S)(y)G~(y,w)[bS+cS](w)dw)b(y)R(y)R(y)]dy\int_{\mathbb{R}}\left[R^{2}(y)\left(-c(y)+F^{\prime\prime\prime}(S)(y)\int_{\mathbb{R}}\tilde{G}(y,w)[bS^{\prime}+cS](w)\,\mathrm{d}w\right)-b(y)R^{\prime}(y)R(y)\right]\,\mathrm{d}y
Proof.

With Y±m2Y_{\pm\infty}^{m^{2}} solving (3.10), we have Y±m2=c±RY_{\pm\infty}^{m^{2}}=c_{\pm}R, for constants c±c_{\pm}.

Our first step is to analyze the unperturbed Wronskian W𝐘(λ,y)=det(𝐘λ,𝐘λ)W_{\mathbf{Y}}(\lambda,y)=\det(\mathbf{Y}_{\infty}^{\lambda},\mathbf{Y}_{-\infty}^{\lambda}). With λ\lambda near m2m^{2} and k=m2λk=\sqrt{m^{2}-\lambda}, by abuse of notation, we write W𝐘(k,y)=W𝐘(m2k2,y)W_{\mathbf{Y}}(k,y)=W_{\mathbf{Y}}(m^{2}-k^{2},y). Applying Lemma 3.3 with b=d=0b=d=0, we may write

𝐘±λ(y)=eky(Y±m2(y)+𝐄±),y,\mathbf{Y}_{\pm\infty}^{\lambda}(y)=e^{\mp ky}\mathbf{(}Y_{\pm\infty}^{m^{2}}(y)+\mathbf{E}_{\pm\infty}),\quad y\in\mathbb{R},

with 𝐄L([0,),2)+𝐄L((,0],2)k\|\mathbf{E}_{\infty}\|_{L^{\infty}([0,\infty),\mathbb{R}^{2})}+\|\mathbf{E}_{-\infty}\|_{L^{\infty}((-\infty,0],\mathbb{R}^{2})}\lesssim k. By the equation satisfied by Y±λY_{\pm\infty}^{\lambda}, the Wronskian W𝐘(λ,y)W_{\mathbf{Y}}(\lambda,y) is independent of yy. Proceeding as in the proof of Theorem 3.2, we have as before (see (3.9))

(3.12) W𝐘(k,0)=det(𝐘m2(0),𝐘λ(0))+det(𝐘λ(0),𝐘m2(0))+O(k2).\begin{split}W_{\mathbf{Y}}(k,0)&=\det(\mathbf{Y}_{\infty}^{m^{2}}(0),\mathbf{Y}_{-\infty}^{\lambda}(0))+\det(\mathbf{Y}_{\infty}^{\lambda}(0),\mathbf{Y}_{-\infty}^{m^{2}}(0))+O(k^{2}).\end{split}

A direct calculation shows det(𝐘±m2,𝐘λ)(y)=(m2λ)Y±m2Yλ\det(\mathbf{Y}_{\pm\infty}^{m^{2}},\mathbf{Y}_{\mp\infty}^{\lambda})^{\prime}(y)=(m^{2}-\lambda)Y_{\pm\infty}^{m^{2}}Y_{\mp\infty}^{\lambda}, which gives, since k2=m2λk^{2}=m^{2}-\lambda,

(3.13) W𝐘(k,0)=k20Ym2Yλdw+k20Ym2Yλdw+O(k2)=k20Ym2ekw(Ym2+E1)dw+k20Ym2ekw(Ym2+E1)dw+O(k2)=k2c+cek|w|R2dw+k2R(w)ek|w|(1{w<0}c+E1+1{w0}cE1)dw+O(k2).\begin{split}W_{\mathbf{Y}}(k,0)&=k^{2}\int_{-\infty}^{0}Y_{\infty}^{m^{2}}Y_{-\infty}^{\lambda}\,\mathrm{d}w+k^{2}\int_{0}^{\infty}Y_{-\infty}^{m^{2}}Y_{\infty}^{\lambda}\,\mathrm{d}w+O(k^{2})\\ &=k^{2}\int_{-\infty}^{0}Y_{\infty}^{m^{2}}e^{kw}(Y_{-\infty}^{m^{2}}+E_{-\infty}^{1})\,\mathrm{d}w+k^{2}\int_{0}^{\infty}Y_{-\infty}^{m^{2}}e^{-kw}(Y_{\infty}^{m^{2}}+E_{\infty}^{1})\,\mathrm{d}w+O(k^{2})\\ &=k^{2}c_{+}c_{-}\int_{-\infty}^{\infty}e^{-k|w|}R^{2}\,\mathrm{d}w\\ &\quad+k^{2}\int_{-\infty}^{\infty}R(w)e^{-k|w|}\left(1_{\{w<0\}}c_{+}E_{-\infty}^{1}+1_{\{w\geq 0\}}c_{-}E_{\infty}^{1}\right)\,\mathrm{d}w+O(k^{2}).\end{split}

Note that all integrals converge, since Y±m2Y_{\pm\infty}^{m^{2}}, e±kyY±λe^{\pm ky}Y_{\pm\infty}^{\lambda}, and e±kyE±1e^{\pm ky}E_{\pm\infty}^{1} are all uniformly bounded.

In the last expression of (3.13), we note that the first term is proportional to kk. Indeed, since RR has non-zero limits as y±y\to\pm\infty, there exist ζ,M>0\zeta,M>0 (independent of kk) such that R2(y)ζR^{2}(y)\geq\zeta if |y|M|y|\geq M. As a result, for any k(0,1)k\in(0,1), one has ek|w|R2dw2ζeM/k\int_{\mathbb{R}}e^{-k|w|}R^{2}\,\mathrm{d}w\geq 2\zeta e^{-M}/k. On the other hand, we have R2ek|w|dw2RL()2/k\int_{\mathbb{R}}R^{2}e^{-k|w|}\,\mathrm{d}w\leq 2\|R\|_{L^{\infty}(\mathbb{R})}^{2}/k. It is also clear that, since 𝐄±L()k\|\mathbf{E}_{\pm\infty}\|_{L^{\infty}(\mathbb{R})}\lesssim k, the second term on the right in (3.13) is O(k2)O(k^{2}). To sum up, we have shown

(3.14) W𝐘(k,0)=c+cA(k)+O(k2),W_{\mathbf{Y}}(k,0)=c_{+}c_{-}A(k)+O(k^{2}),

with A(k)A0kA(k)\geq A_{0}k for some A0>0A_{0}>0 independent of kk.

Now we turn to the perturbed operator T\mathcal{L}_{T}. Let U±λU_{\pm\infty}^{\lambda} be the solutions to (3.5) as above. Applying Lemma 3.3 with λ1=λ2=λ\lambda_{1}=\lambda_{2}=\lambda, we have

(3.15) 𝐔±λ(y)=𝐘±λ(y)+𝐄~±(y),\begin{split}\mathbf{U}_{\pm\infty}^{\lambda}(y)&=\mathbf{Y}_{\pm\infty}^{\lambda}(y)+\tilde{\mathbf{E}}_{\pm\infty}(y),\end{split}

with eky𝐄~L([0,),2)+eky𝐄~L((,0],2)dL1()+bL1()δ\|e^{ky}\tilde{\mathbf{E}}_{\infty}\|_{L^{\infty}([0,\infty),\mathbb{R}^{2})}+\|e^{-ky}\tilde{\mathbf{E}}_{-\infty}\|_{L^{\infty}((-\infty,0],\mathbb{R}^{2})}\lesssim\|d\|_{L^{1}(\mathbb{R})}+\|b\|_{L^{1}(\mathbb{R})}\lesssim\delta.

With the Wronskian W𝐔(λ,y)W_{\mathbf{U}}(\lambda,y) defined as in the proof of Theorem 3.2, we again write W𝐔(k,y)=W𝐔(m2k2,y)W_{\mathbf{U}}(k,y)=W_{\mathbf{U}}(m^{2}-k^{2},y), and obtain

(3.16) W𝐔(λ,0)=det(𝐘λ(0),𝐘λ(0))+det(𝐄~(0),𝐘λ(0))+det(𝐘λ(0),𝐄~(0))+det(𝐄~(0),𝐄~(0))=W𝐘(λ,0)+det(𝐄~(0),𝐘λ(0))+det(𝐘λ(0),𝐄~(0))+O(δ2).\begin{split}W_{\mathbf{U}}(\lambda,0)&=\det(\mathbf{Y}_{\infty}^{\lambda}(0),\mathbf{Y}_{-\infty}^{\lambda}(0))+\det(\tilde{\mathbf{E}}_{\infty}(0),\mathbf{Y}_{-\infty}^{\lambda}(0))+\det(\mathbf{Y}_{\infty}^{\lambda}(0),\tilde{\mathbf{E}}_{-\infty}(0))+\det(\tilde{\mathbf{E}}_{\infty}(0),\tilde{\mathbf{E}}_{-\infty}(0))\\ &=W_{\mathbf{Y}}(\lambda,0)+\det(\tilde{\mathbf{E}}_{\infty}(0),\mathbf{Y}_{-\infty}^{\lambda}(0))+\det(\mathbf{Y}_{\infty}^{\lambda}(0),\tilde{\mathbf{E}}_{-\infty}(0))+O(\delta^{2}).\end{split}

Since 𝐄~±=U±λY±λ\tilde{\mathbf{E}}_{\pm\infty}=U_{\pm\infty}^{\lambda}-Y_{\pm\infty}^{\lambda} satisfy 𝐄~±′′=(F′′(S)m2λ)𝐄~±b(U±λ)+dU±λ\tilde{\mathbf{E}}_{\pm\infty}^{\prime\prime}=(F^{\prime\prime}(S)-m^{2}-\lambda)\tilde{\mathbf{E}}_{\pm\infty}-b(U_{\pm\infty}^{\lambda})^{\prime}+dU_{\pm\infty}^{\lambda}, we have

det(𝐄~±,𝐘λ)(y)=[dU±λb(U±λ)]Yλ.\det(\tilde{\mathbf{E}}_{\pm\infty},\mathbf{Y}_{\mp\infty}^{\lambda})^{\prime}(y)=[dU_{\pm\infty}^{\lambda}-b(U_{\pm\infty}^{\lambda})^{\prime}]Y_{\mp\infty}^{\lambda}.

Because b,dL1()b,d\in L^{1}(\mathbb{R}), |Uλ|eky|U_{\infty}^{\lambda}|\lesssim e^{-ky}, and |Yλ|eky|Y_{-\infty}^{\lambda}|\lesssim e^{ky}, the expression [dUλb(Uλ)]Yλ\left[dU_{\infty}^{\lambda}-b(U_{\infty}^{\lambda})^{\prime}\right]Y_{-\infty}^{\lambda} is integrable on [0,)[0,\infty), and we can use (3.15) to write

det(𝐄~(0),𝐘λ(0))=0[dUλb(Uλ)]Yλdw=0[dYλb(Yλ)]Yλdw+0[dE~1bE~2]Yλdw,\begin{split}\det(\tilde{\mathbf{E}}_{\infty}(0),\mathbf{Y}_{-\infty}^{\lambda}(0))&=\int_{0}^{\infty}\left[dU_{\infty}^{\lambda}-b(U_{\infty}^{\lambda})^{\prime}\right]Y_{-\infty}^{\lambda}\,\mathrm{d}w\\ &=\int_{0}^{\infty}[dY_{\infty}^{\lambda}-b(Y_{\infty}^{\lambda})^{\prime}]Y_{-\infty}^{\lambda}\,\mathrm{d}w+\int_{0}^{\infty}[d\tilde{E}_{\infty}^{1}-b\tilde{E}_{\infty}^{2}]Y_{-\infty}^{\lambda}\,\mathrm{d}w,\end{split}

where the last integral converges and is O(δ2)O(\delta^{2}) since 𝐄~δeky\|\tilde{\mathbf{E}}_{\infty}\|\lesssim\delta e^{-ky} and YλekyY_{-\infty}^{\lambda}\lesssim e^{ky}. For the first integral on the right, we use Lemma 3.1 with d=b=0d=b=0 and obtain

0[dYλb(Yλ)]Yλdw=0[dYm2b(Ym2)]Ym2dw+O(δk).\int_{0}^{\infty}[dY_{\infty}^{\lambda}-b(Y_{\infty}^{\lambda})^{\prime}]Y_{-\infty}^{\lambda}\,\mathrm{d}w=\int_{0}^{\infty}[dY_{\infty}^{m^{2}}-b(Y_{\infty}^{m^{2}})^{\prime}]Y_{-\infty}^{m^{2}}\,\mathrm{d}w+O(\delta k).

After applying a similar analysis to det(𝐘λ(0),𝐄~(0))\det(\mathbf{Y}_{\infty}^{\lambda}(0),\tilde{\mathbf{E}}_{-\infty}(0)), the expression (3.16) becomes

W𝐔(k,0)=W𝐘(k,0)+c+c[d(R(w))2bR(w)R(w)]dw+O(δk)+O(δ2).W_{\mathbf{U}}(k,0)=W_{\mathbf{Y}}(k,0)+c_{+}c_{-}\int_{-\infty}^{\infty}[d(R(w))^{2}-bR(w)R^{\prime}(w)]\,\mathrm{d}w+O(\delta k)+O(\delta^{2}).

With (3.14), this implies

W𝐔(k,0)=c+c(A(k)+[d(R(w))2bR(w)R(w)]dw)+O(δk)+O(δ2).W_{\mathbf{U}}(k,0)=c_{+}c_{-}\left(A(k)+\int_{-\infty}^{\infty}[d(R(w))^{2}-bR(w)R^{\prime}(w)]\,\mathrm{d}w\right)+O(\delta k)+O(\delta^{2}).

For δ\delta small enough, the expression inside the parentheses determines whether any zeroes of W𝐔(k,0)W_{\mathbf{U}}(k,0) are present for k>0k>0. The bound A(k)A0kA(k)\geq A_{0}k with A0>0A_{0}>0 implies statement (a) of the theorem.

For statement (b), the assumption that m2m^{2} is not a resonance or eigenvalue implies W𝐘(0,0)0W_{\mathbf{Y}}(0,0)\neq 0. The approximation (3.15) easily implies W𝐔(0,0)=W𝐘(0,0)+O(δ)0W_{\mathbf{U}}(0,0)=W_{\mathbf{Y}}(0,0)+O(\delta)\neq 0 for δ\delta small enough. ∎

4. Orbital stability

In this section, we prove orbital stability, i.e. that solutions starting close to TT are always close to some shifted version of TT.

Proof of Theorem 1.3.

For any solution uu of (1.6), the energy

E(u)=[12(tu)2+12(yu)212cu2+F(u)]ω(y)dy,E(u)=\int_{\mathbb{R}}\left[\frac{1}{2}(\partial_{t}u)^{2}+\frac{1}{2}(\partial_{y}u)^{2}-\frac{1}{2}cu^{2}+F(u)\right]\omega(y)\,\mathrm{d}y,

is conserved, where ω(y)=exp(yb(z)dz)=1+O(δ)\omega(y)=\exp(\int_{-\infty}^{y}b(z)\,\mathrm{d}z)=1+O(\delta), uniformly in yy. We also define the potential energy

Ep(u)=[12(yu)212cu2+F(u)]ω(y)dy.E_{p}(u)=\int_{\mathbb{R}}\left[\frac{1}{2}(\partial_{y}u)^{2}-\frac{1}{2}cu^{2}+F(u)\right]\omega(y)\,\mathrm{d}y.

A simple computation shows that

(4.1) |Ep(ψ)E~p(ψ)|Cδ(ψL()2+E~p(ψ)),ψHT1(),|E_{p}(\psi)-\tilde{E}_{p}(\psi)|\leq C\delta\left(\|\psi\|_{L^{\infty}(\mathbb{R})}^{2}+\tilde{E}_{p}(\psi)\right),\quad\psi\in H^{1}_{T}(\mathbb{R}),

where E~p\tilde{E}_{p} is the potential energy corresponding to the constant coefficient equation (1.4):

E~p(ψ):=[12(tψ)2+12(yψ)2+F(ψ)]dy\tilde{E}_{p}(\psi):=\int_{\mathbb{R}}\left[\frac{1}{2}(\partial_{t}\psi)^{2}+\frac{1}{2}(\partial_{y}\psi)^{2}+F(\psi)\right]\,\mathrm{d}y

The idea is to use the (known) property that E~p(ψ)E~p(S)\tilde{E}_{p}(\psi)-\tilde{E}_{p}(S) controls the distance between ψ\psi and SS, to show the corresponding fact for EpE_{p} and TT. In more detail, for q>0q>0, define

dq(ψ,T)2:=infξ[(yψ(y)T(y+ξ))2+q(ψ(y)T(y+ξ))2]dy,d_{q}(\psi,T)^{2}:=\inf_{\xi\in\mathbb{R}}\int_{\mathbb{R}}[(\partial_{y}\psi(y)-T^{\prime}(y+\xi))^{2}+q(\psi(y)-T(y+\xi))^{2}]\,\mathrm{d}y,

for any ψ\psi in the energy space. We define dq(ψ,S)d_{q}(\psi,S) in the analogous way. Proposition 1 of [20] proves the following: There exist C,r,q>0C,r,q>0 such that

dq(ψ,S)2C(E~p(ψ)E~p(S)),d_{q}(\psi,S)^{2}\leq C(\tilde{E}_{p}(\psi)-\tilde{E}_{p}(S)),

whenever dq(ψ,S)rd_{q}(\psi,S)\leq r.aaaProposition 1 in [20] is stated for u(t,)u(t,\cdot) where uu is a solution of (1.4), but an examination of the proof shows that the conclusion holds for any ψ(y)\psi(y) satisfying the hypotheses stated here. Note that

|E~p(T)E~p(S)|=|[12(ySb)2+ySySb+F(S+Sb)F(S)]dy|SbH1()2+ySL2()2+FC1([a,a+]SbL1()δ,\begin{split}|\tilde{E}_{p}(T)-\tilde{E}_{p}(S)|&=\left|\int_{\mathbb{R}}\left[\frac{1}{2}(\partial_{y}S_{b})^{2}+\partial_{y}S\partial_{y}S_{b}+F(S+S_{b})-F(S)\right]\,\mathrm{d}y\right|\\ &\leq\|S_{b}\|_{H^{1}(\mathbb{R})}^{2}+\|\partial_{y}S\|_{L^{2}(\mathbb{R})}^{2}+\|F\|_{C^{1}([a_{-},a_{+}]}\|S_{b}\|_{L^{1}(\mathbb{R})}\\ &\lesssim\delta,\end{split}

by Theorem 1.1. Using (4.1) twice, we then have

(4.2) dq(ψ,S)2C(Ep(ψ)Ep(S))+Cδ(|ψ|+|T|L()2+E~p(ψ)+E~p(S))C(Ep(ψ)Ep(T))+Cδ(1+E~p(ψ)+E~p(S)).\begin{split}d_{q}(\psi,S)^{2}&\leq C\left(E_{p}(\psi)-E_{p}(S)\right)+C\delta\left(\||\psi|+|T|\|_{L^{\infty}(\mathbb{R})}^{2}+\tilde{E}_{p}(\psi)+\tilde{E}_{p}(S)\right)\\ &\leq C\left(E_{p}(\psi)-E_{p}(T)\right)+C\delta\left(1+\tilde{E}_{p}(\psi)+\tilde{E}_{p}(S)\right).\end{split}

To get to the last line, we used Sobolev embedding to write ψL()Cqdq(ψ,S)\|\psi\|_{L^{\infty}(\mathbb{R})}\leq C_{q}d_{q}(\psi,S), and combined this term into the left-hand side.

Since (u(t,y)S(y+ξ))2dy\int_{\mathbb{R}}(u(t,y)-S(y+\xi))^{2}\,\mathrm{d}y\to\infty as ξ±\xi\to\pm\infty, there is some ξ0\xi_{0}\in\mathbb{R} where the infimum defining dq(u(t,),S)d_{q}(u(t,\cdot),S) is achieved. To save space, write Tξ=T(y+ξ0)T_{\xi}=T(y+\xi_{0}), and similarly for SξS_{\xi} and Sb,ξS_{b,\xi}. We then have

(4.3) dq(ψ,T)2[(yψTξ)2+q(ψTξ)2]dy=[(yψSξ)2+q(ψSξ)2]dy+[(Sb,ξ)2+qSb,ξ22(yψSξ)ySb,ξ2q(ψSξ)Sb,ξ]dydq(ψ,S)2+CqSbH1()2+SbH1()ψSH1()dq(ψ,S)2+CqSbH1()2+SbH1()dq(ψ,S)22dq(ψ,S)2+Cδ2.\begin{split}d_{q}(\psi,T)^{2}&\leq\int_{\mathbb{R}}\left[(\partial_{y}\psi-T^{\prime}_{\xi})^{2}+q(\psi-T_{\xi})^{2}\right]\,\mathrm{d}y\\ &=\int_{\mathbb{R}}\left[(\partial_{y}\psi-S^{\prime}_{\xi})^{2}+q(\psi-S_{\xi})^{2}\right]\,\mathrm{d}y\\ &\quad+\int_{\mathbb{R}}\left[(S_{b,\xi}^{\prime})^{2}+qS_{b,\xi}^{2}-2(\partial_{y}\psi-S_{\xi}^{\prime})\partial_{y}S_{b,\xi}-2q(\psi-S_{\xi})S_{b,\xi}\right]\,\mathrm{d}y\\ &\leq d_{q}(\psi,S)^{2}+Cq\|S_{b}\|_{H^{1}(\mathbb{R})}^{2}+\|S_{b}\|_{H^{1}(\mathbb{R})}\|\psi-S\|_{H^{1}(\mathbb{R})}\\ &\leq d_{q}(\psi,S)^{2}+Cq\|S_{b}\|_{H^{1}(\mathbb{R})}^{2}+\|S_{b}\|_{H^{1}(\mathbb{R})}d_{q}(\psi,S)^{2}\\ &\leq 2d_{q}(\psi,S)^{2}+C\delta^{2}.\end{split}

For δ>0\delta>0 small enough compared to rr and qq, this implies dq(ψ,T)22dq(ψ,S)2+r/2d_{q}(\psi,T)^{2}\leq 2d_{q}(\psi,S)^{2}+r/2. By exchanging the roles of TT and SS in this calculation, we also obtain dq(ψ,S)22dq(ψ,T)+r/2d_{q}(\psi,S)^{2}\leq 2d_{q}(\psi,T)+r/2.

Next, combining (4.2) and (4.3),

(4.4) dq(ψ,T)2C(Ep(ψ)Ep(T))+Cδ(1+E~p(ψ)+E~p(S)).\begin{split}d_{q}(\psi,T)^{2}&\leq C\left(E_{p}(\psi)-E_{p}(T)\right)+C\delta\left(1+\tilde{E}_{p}(\psi)+\tilde{E}_{p}(S)\right).\end{split}

This inequality holds for ψ\psi such that dq(ψ,S)rd_{q}(\psi,S)\leq r. By above, we can ensure this condition by choosing dq(ψ,T)r/4d_{q}(\psi,T)\leq r/4.

Now, for a solution uu to (1.6) with dq(u(0,),T)r/4d_{q}(u(0,\cdot),T)\leq r/4 and tu(0,)\partial_{t}u(0,\cdot) sufficiently small in L2()L^{2}(\mathbb{R}), (4.4) implies that

dq(u(t,),T)2C(E(u(t,))E(T))+Cδ(1+E~p(u(t,))+E~p(S)).d_{q}(u(t,\cdot),T)^{2}\leq C(E(u(t,\cdot))-E(T))+C\delta\left(1+\tilde{E}_{p}(u(t,\cdot))+\tilde{E}_{p}(S)\right).

The quantity E(u(t,))E(u(t,\cdot)) is conserved in time. Calculations similar to (4.1) show that E~p(u)2Ep(u)+CδuL()2E(u)+Cδdq(u,T)2\tilde{E}_{p}(u)\leq 2E_{p}(u)+C\delta\|u\|_{L^{\infty}(\mathbb{R})}^{2}\leq E(u)+C\delta d_{q}(u,T)^{2}, and the last term may be combined into the left side. We finally have

dq(u(t,),T)2C(E(u0)E(T))+Cδ(1+E~p(S)).d_{q}(u(t,\cdot),T)^{2}\leq C(E(u_{0})-E(T))+C\delta(1+\tilde{E}_{p}(S)).

This right-hand side is independent of tt, which implies the solution uu never leaves the neighborhood of TT as long as it exists. As above, for every tt, there is some ξ=ξ(t)\xi=\xi(t) at which the infimum defining dq(u,T)d_{q}(u,T) is achieved. By standard arguments, this time-independent bound on dq(u(t,),T)d_{q}(u(t,\cdot),T) combined with energy conservation implies the solution uu exists for all t[0,)t\in[0,\infty). ∎

Appendix A ODE Methods

In this section, we collect some convenient facts about the solvability and asymptotics of 2×22\times 2 first-order systems on \mathbb{R}.

First, we have a standard lemma on vector-valued integral equations of Volterra type:

Lemma A.1.

For aa\in\mathbb{R} and 𝐔L([a,),2)\mathbf{U}\in L^{\infty}([a,\infty),\mathbb{R}^{2}), the Volterra equation

𝐙(y)=𝐔(y)+yK(y,w)𝐙(w)dw,\mathbf{Z}(y)=\mathbf{U}(y)+\int_{y}^{\infty}K(y,w)\mathbf{Z}(w)\,\mathrm{d}w,

has a unique solution in L([a,),2)L^{\infty}([a,\infty),\mathbb{R}^{2}), provided

(A.1) μ:=asupa<y<wK(y,w)dw<,\mu:=\int_{a}^{\infty}\sup_{a<y<w}\|K(y,w)\|\,\mathrm{d}w<\infty,

where \|\cdot\| is the operator norm of the matrix K(y,w)K(y,w). This solution is given by the iteration

(A.2) 𝐙(y)=𝐔(y)+n=1aai=1n1{yi1<yi}K(yi1,yi)𝐔(yn)dyndy1,\mathbf{Z}(y)=\mathbf{U}(y)+\sum_{n=1}^{\infty}\int_{a}^{\infty}\cdots\int_{a}^{\infty}\prod_{i=1}^{n}1_{\{y_{i-1}<y_{i}\}}K(y_{i-1},y_{i})\mathbf{U}(y_{n})\,\mathrm{d}y_{n}\cdots\,\mathrm{d}y_{1},

with y0=yy_{0}=y. This solution satisfies

𝐙L([a,),2)eμ𝐔L([a,),2).\|\mathbf{Z}\|_{L^{\infty}([a,\infty),\mathbb{R}^{2})}\leq e^{\mu}\|\mathbf{U}\|_{L^{\infty}([a,\infty),\mathbb{R}^{2})}.
Proof.

See [42, Lemma 2.4] for a proof of the corresponding fact for scalar-valued Volterra equations. The proof in the present vector-valued case is essentially the same, so we omit it. ∎

Next, we address a class of linear systems that arise from the eigenvalue problems in Sections 2 and 3:

Lemma A.2.
  1. (a)

    For k>0k>0, consider the system

    (A.3) 𝐘(y)=(M1+M2(y))𝐘(y),\mathbf{Y}^{\prime}(y)=(M_{1}+M_{2}(y))\mathbf{Y}(y),

    where

    M1=(01k20),M2(y)=(00V(y)b(y))M_{1}=\left(\begin{array}[]{cc}0&1\\ k^{2}&0\end{array}\right),\quad M_{2}(y)=\left(\begin{array}[]{cc}0&0\\ V(y)&-b(y)\end{array}\right)

    with V,bL1()V,b\in L^{1}(\mathbb{R}). There exist solutions 𝐘\mathbf{Y}_{-\infty}, 𝐘\mathbf{Y}_{\infty} defined on \mathbb{R}, such that

    (A.4) limyeky𝐘(y)=(1k),limyeky𝐘(y)=(1k),\lim_{y\to\infty}e^{ky}\mathbf{Y}_{\infty}(y)=\left(\begin{array}[]{c}1\\ -k\end{array}\right),\quad\lim_{y\to-\infty}e^{-ky}\mathbf{Y}_{-\infty}(y)=\left(\begin{array}[]{c}1\\ k\end{array}\right),

    and the bound |𝐘(y)|Ceky|\mathbf{Y}_{\infty}(y)|\leq Ce^{-ky} holds for all yy\in\mathbb{R}, where the constant depends on kk and V+bL1()\|V+b\|_{L^{1}(\mathbb{R})}. These solutions also satisfy the integral equations

    (A.5) 𝐘(y)=(1k)eky12y(V(w)YbY(w))(1k(ek(yw)ek(yw))ek(yw)+ek(yw))dw,𝐘(y)=(1k)eky+12y(V(w)YbY(w))(1k(ek(yw)ek(yw))ek(yw)+ek(yw))dw.\begin{split}\mathbf{Y}_{\infty}(y)&=\left(\begin{array}[]{c}1\\ -k\end{array}\right)e^{-ky}-\frac{1}{2}\int_{y}^{\infty}(V(w)Y_{\infty}-bY_{\infty}^{\prime}(w))\left(\begin{array}[]{c}\frac{1}{k}(e^{k(y-w)}-e^{-k(y-w)})\\ e^{k(y-w)}+e^{-k(y-w)}\end{array}\right)\,\mathrm{d}w,\\ \mathbf{Y}_{-\infty}(y)&=\left(\begin{array}[]{c}1\\ k\end{array}\right)e^{ky}+\frac{1}{2}\int_{-\infty}^{y}(V(w)Y_{-\infty}-bY_{-\infty}^{\prime}(w))\left(\begin{array}[]{c}\frac{1}{k}(e^{k(y-w)}-e^{-k(y-w)})\\ e^{k(y-w)}+e^{-k(y-w)}\end{array}\right)\,\mathrm{d}w.\end{split}
  2. (b)

    For k=0k=0, assume in addition that (1+|y|2)1/2V(1+|y|^{2})^{1/2}V and (1+|y|2)1/2b(1+|y|^{2})^{1/2}b lie in L1()L^{1}(\mathbb{R}). Then there exist solutions 𝐘,𝐘\mathbf{Y}_{-\infty},\mathbf{Y}_{\infty} to (A.3) satisfying

    limy𝐘±(y)=(10),\lim_{y\to\infty}\mathbf{Y}_{\pm\infty}(y)=\left(\begin{array}[]{c}1\\ 0\end{array}\right),

    as well as the integral equations

    (A.6) 𝐘(y)=(10)y(V(w)YbY(w))(yw1)dw,𝐘(y)=(10)+y(V(w)YbY(w))(yw1)dw.\begin{split}\mathbf{Y}_{\infty}(y)&=\left(\begin{array}[]{c}1\\ 0\end{array}\right)-\int_{y}^{\infty}(V(w)Y_{\infty}-bY_{\infty}^{\prime}(w))\left(\begin{array}[]{c}y-w\\ 1\end{array}\right)\,\mathrm{d}w,\\ \mathbf{Y}_{-\infty}(y)&=\left(\begin{array}[]{c}1\\ 0\end{array}\right)+\int_{-\infty}^{y}(V(w)Y_{-\infty}-bY_{-\infty}^{\prime}(w))\left(\begin{array}[]{c}y-w\\ 1\end{array}\right)\,\mathrm{d}w.\end{split}
Proof.

(a) Note that the eigenvalues of M1M_{1} are ±k\pm k corresponding to eigenvectors (1k)\left(\begin{array}[]{c}1\\ \mp k\end{array}\right). We will find a solution to the integral equation

(A.7) 𝐘(y)=(1k)ekyyeM1(yw)M2(w)𝐘(w)dw,y,\mathbf{Y}_{\infty}(y)=\left(\begin{array}[]{c}1\\ -k\end{array}\right)e^{-ky}-\int_{y}^{\infty}e^{M_{1}(y-w)}M_{2}(w)\mathbf{Y}_{\infty}(w)\,\mathrm{d}w,\quad y\in\mathbb{R},

satisfying |𝐘(y)|Ceky|\mathbf{Y}_{\infty}(y)|\leq Ce^{-ky} and limyeky𝐘=(1k)\lim_{y\to\infty}e^{ky}\mathbf{Y}_{\infty}=\left(\begin{array}[]{c}1\\ -k\end{array}\right). By direct calculation, such 𝐘\mathbf{Y}_{\infty} also solves (A.3), as well as the first integral equation in (A.5). Letting 𝐙(y)=eky𝐘(y)\mathbf{Z}(y)=e^{ky}\mathbf{Y}_{\infty}(y), (A.7) is equivalent to

(A.8) 𝐙(y)=(1k)yeM1(yw)M2(w)ek(yw)𝐙(w)dw,y.\mathbf{Z}(y)=\left(\begin{array}[]{c}1\\ -k\end{array}\right)-\int_{y}^{\infty}e^{M_{1}(y-w)}M_{2}(w)e^{k(y-w)}\mathbf{Z}(w)\,\mathrm{d}w,\quad y\in\mathbb{R}.

By diagonalizing M1M_{1}, we obtain

eM1(yw)M2=12(1kV(ek(yw)ek(yw))1kb(ek(yw)ek(yw))V(ek(yw)+ek(yw))b(ek(yw)+ek(yw))).e^{M_{1}(y-w)}M_{2}=\frac{1}{2}\left(\begin{array}[]{cc}\frac{1}{k}V(e^{k(y-w)}-e^{-k(y-w)})&-\frac{1}{k}b(e^{k(y-w)}-e^{-k(y-w)})\\ V(e^{k(y-w)}+e^{-k(y-w)})&-b(e^{k(y-w)}+e^{-k(y-w)})\end{array}\right).

With K(y,w):=eM1(yw)M2(w)ek(yw)K(y,w):=e^{M_{1}(y-w)}M_{2}(w)e^{k(y-w)}, we therefore have

K(y,w)C(1+e2k(yw))(|V(w)|+|b(w)|),\|K(y,w)\|\leq C(1+e^{2k(y-w)})(|V(w)|+|b(w)|),

and that

0sup0<y<wK(y,w)dwC(VL1()+bL1()).\int_{0}^{\infty}\sup_{0<y<w}\|K(y,w)\|\,\mathrm{d}w\leq C(\|V\|_{L^{1}(\mathbb{R})}+\|b\|_{L^{1}(\mathbb{R})}).

Lemma A.1 now implies a solution to (A.8) exists on [0,)[0,\infty), and 𝐙L([0,),2)\|\mathbf{Z}\|_{L^{\infty}([0,\infty),\mathbb{R}^{2})} is bounded by a constant, which implies the boundary condition (A.4) holds for 𝐘\mathbf{Y}_{\infty}, as well as the upper bound

|𝐘(y)|Ceky,y0,|\mathbf{Y}_{\infty}(y)|\leq Ce^{-ky},\quad y\geq 0,

where 𝐘=(Y,Y)\mathbf{Y}_{\infty}=(Y_{\infty},Y_{\infty}^{\prime}). Applying a similar argument with y-y replacing yy, we can obtain a solution 𝐘\mathbf{Y}_{-\infty} defined on \mathbb{R} with

limyeky𝐘(y)=(1k)and|Y(y)|+|Y(y)|Ceky,y0.\lim_{y\to-\infty}e^{-ky}\mathbf{Y}_{-\infty}(y)=\left(\begin{array}[]{c}1\\ k\end{array}\right)\quad\text{and}\quad|Y_{-\infty}(y)|+|Y_{-\infty}^{\prime}(y)|\leq Ce^{ky},\quad y\leq 0.

For y<0y<0, we can write

Y(y)=c0Y(y)(0yexp(wb(z)dz)Y2(w)dw+c1),Y_{\infty}(y)=c_{0}Y_{-\infty}(y)\left(\int_{0}^{y}\frac{\exp\left(-\int_{-\infty}^{w}b(z)\,\mathrm{d}z\right)}{Y_{-\infty}^{2}(w)}\,\mathrm{d}w+c_{1}\right),

with c0,c1c_{0},c_{1} chosen so that Y(0)Y_{\infty}(0) and Y(0)Y_{\infty}^{\prime}(0) match our previous definition. This formula implies 𝐘=(Y,Y)\mathbf{Y}_{\infty}=(Y_{\infty},Y_{\infty}^{\prime}) solves (A.3) and satisfies |𝐘(y)|Ceky|\mathbf{Y}_{\infty}(y)|\leq Ce^{-ky} for negative yy also. By a similar method, we extend 𝐘\mathbf{Y}_{-\infty} to the real line and obtain |𝐘(y)|Ceky|\mathbf{Y}_{-\infty}(y)|\leq Ce^{ky} for all yy\in\mathbb{R}.

(b) In the case k=0k=0, we have

eM1(yw)M2(w)=(1yw01)(00V(y)b(y))=((yw)V(y)(yw)b(y)V(y)b(y)),e^{M_{1}(y-w)}M_{2}(w)=\left(\begin{array}[]{cc}1&y-w\\ 0&1\end{array}\right)\left(\begin{array}[]{cc}0&0\\ V(y)&-b(y)\end{array}\right)=\left(\begin{array}[]{cc}(y-w)V(y)&-(y-w)b(y)\\ V(y)&-b(y)\end{array}\right),

and the integral equation (A.7) reduces to

𝐘(y)=(10)y(V(w)YbY(w))(yw1)dw.\mathbf{Y}_{\infty}(y)=\left(\begin{array}[]{c}1\\ 0\end{array}\right)-\int_{y}^{\infty}(V(w)Y_{\infty}-bY_{\infty}^{\prime}(w))\left(\begin{array}[]{c}y-w\\ 1\end{array}\right)\,\mathrm{d}w.

Defining K(y,w)=eM1(yw)M2(w)K(y,w)=e^{M_{1}(y-w)}M_{2}(w) , we have

K(y,w)C1+(yw)2(|V(w)+|b(w)|),\|K(y,w)\|\leq C\sqrt{1+(y-w)^{2}}(|V(w)+|b(w)|),

and

0sup0<y<wK(y,w)dwC((1+|y|2)1/2VL1()+(1+|y|2)1/2bL1()).\int_{0}^{\infty}\sup_{0<y<w}\|K(y,w)\|\,\mathrm{d}w\leq C(\|(1+|y|^{2})^{1/2}V\|_{L^{1}(\mathbb{R})}+\|(1+|y|^{2})^{1/2}b\|_{L^{1}(\mathbb{R})}).

By Lemma A.1, a solution 𝐘\mathbf{Y}_{\infty} exists on [0,)[0,\infty), which also solves (A.3) by a direct calculation. Applying a similar method for 𝐘\mathbf{Y}_{-\infty} and extending both solutions to the real line proceeds as in the proof of (a). ∎

References

  • [1] M. A. Alejo, C. Muñoz, and J. M. Palacios. On the asymptotic stability of the sine-Gordon kink in the energy space. Preprint. ArXiv:2003.09358.
  • [2] J. Alexander, R. Gardner, and C. Jones. A topological invariant arising in the stability analysis of travelling waves. J. Reine Angew. Math., 410:167–212, 1990.
  • [3] C. Cacciapuoti, R. Carlone, and R. Figari. Perturbations of eigenvalues embedded at threshold: two-dimensional solvable models. J. Math. Phys., 52(8):083515, 12, 2011.
  • [4] D. K. Campbell, M. Peyrard, and P. Sodano. Kink-antikink interactions in the double sine-Gordon equation. Phys. D, 19(2):165–205, 1986.
  • [5] G. Chen, J. Liu, and B. Lu. Long-time asymptotics and stability for the sine-Gordon equation. Preprint. arXiv:2009.04260, 2020.
  • [6] S. Cuccagna. On asymptotic stability in 3D of kinks for the ϕ4\phi^{4} model. Trans. Amer. Math. Soc., 360(5):2581–2614, 2008.
  • [7] S. Cuenda, N. R. Quintero, and A. Sánchez. Sine-Gordon wobbles through Bäcklund transformations. Discrete Contin. Dyn. Syst. Ser. S, 4(5):1047–1056, 2011.
  • [8] J. Cuevas-Maraver, P. G. Kevrekidis, and F. Williams, editors. The sine-Gordon model and its applications, volume 10 of Nonlinear Systems and Complexity. Springer, Cham, 2014. From pendula and Josephson junctions to gravity and high-energy physics.
  • [9] A. D’Anna, M. De Angelis, and G. Fiore. Towards soliton solutions of a perturbed sine-Gordon equation. Rend. Accad. Sci. Fis. Mat. Napoli (4), 72:95–110, 2005.
  • [10] J.-M. Delort. Existence globale et comportement asymptotique pour l’équation de Klein-Gordon quasi linéaire à données petites en dimension 1. Ann. Sci. École Norm. Sup. (4), 34(1):1–61, 2001.
  • [11] J.-M. Delort, D. Fang, and R. Xue. Global existence of small solutions for quadratic quasilinear Klein-Gordon systems in two space dimensions. J. Funct. Anal., 211(2):288–323, 2004.
  • [12] J.-M. Delort and N. Masmoudi. Long time dispersive estimates for perturbations of a kink solution of one dimensional cubic wave equations. Preprint: hal-02862414v2, 2020.
  • [13] J. Denzler. Nonpersistence of breather families for the perturbed sine Gordon equation. Comm. Math. Phys., 158(2):397–430, 1993.
  • [14] G. Derks, A. Doelman, S. A. van Gils, and T. Visser. Travelling waves in a singularly perturbed sine-Gordon equation. Phys. D, 180(1-2):40–70, 2003.
  • [15] J. W. Evans. Nerve axon equations. IV. The stable and the unstable impulse. Indiana Univ. Math. J., 24(12):1169–1190, 1974/75.
  • [16] G. Fiore, G. Guerriero, A. Maio, and E. Mazziotti. On kinks and other travelling-wave solutions of a modified sine-Gordon equation. Meccanica, 50(8):1989–2006, 2015.
  • [17] R. A. Gardner and K. Zumbrun. The gap lemma and geometric criteria for instability of viscous shock profiles. Comm. Pure Appl. Math., 51(7):797–855, 1998.
  • [18] P. Germain and F. Pusateri. Quadratic Klein-Gordon equations with a potential in one dimension. Preprint. arXiv:2006.15688, 2020.
  • [19] F. Gesztesy and H. Holden. A unified approach to eigenvalues and resonances of Schrödinger operators using Fredholm determinants. J. Math. Anal. Appl., 123(1):181–198, 1987.
  • [20] D. B. Henry, J. F. Perez, and W. F. Wreszinski. Stability theory for solitary-wave solutions of scalar field equations. Comm. Math. Phys., 85(3):351–361, 1982.
  • [21] P. D. Hislop and I. M. Sigal. Introduction to spectral theory, volume 113 of Applied Mathematical Sciences. Springer-Verlag, New York, 1996. With applications to Schrödinger operators.
  • [22] V. G. Ivancevic and T. T. Ivancevic. Sine-Gordon solitons, kinks and breathers as physical models of nonlinear excitations in living cellular structures. J. Geom. Symmetry Phys., 31:1–56, 2013.
  • [23] J. Jendrej, M. Kowalczyk, and A. Lawrie. Dynamics of strongly interacting kink-antikink pairs for scalar fields on a line. Preprint. ArXiv:1911.02064, 2019.
  • [24] A. Jensen and M. Melgaard. Perturbation of eigenvalues embedded at a threshold. Proc. Roy. Soc. Edinburgh Sect. A, 132(1):163–179, 2002.
  • [25] C. K. R. T. Jones. Stability of the travelling wave solution of the FitzHugh-Nagumo system. Trans. Amer. Math. Soc., 286(2):431–469, 1984.
  • [26] T. Kapitula and B. Sandstede. Edge bifurcations for near integrable systems via Evans function techniques. SIAM J. Math. Anal., 33(5):1117–1143, 2002.
  • [27] T. Kapitula and B. Sandstede. Eigenvalues and resonances using the Evans function. Discrete Contin. Dyn. Syst., 10(4):857–869, 2004.
  • [28] A. Khare, I. C. Christov, and A. Saxena. Successive phase transitions and kink solutions in ϕ8{\phi}^{8}, ϕ10{\phi}^{10}, and ϕ12{\phi}^{12} field theories. Phys. Rev. E, 90:023208, Aug 2014.
  • [29] A. Khare and A. Saxena. Family of potentials with power law kink tails. J. Phys. A, 52(36):365401, 31, 2019.
  • [30] E. Kopylova and A. I. Komech. On asymptotic stability of kink for relativistic Ginzburg-Landau equations. Arch. Ration. Mech. Anal., 202(1):213–245, 2011.
  • [31] E. A. Kopylova and A. I. Komech. On asymptotic stability of moving kink for relativistic Ginzburg-Landau equation. Comm. Math. Phys., 302(1):225–252, 2011.
  • [32] M. Kowalczyk, Y. Martel, and C. Muñoz. Kink dynamics in the ϕ4\phi^{4} model: asymptotic stability for odd perturbations in the energy space. J. Amer. Math. Soc., 30(3):769–798, 2017.
  • [33] M. Kowalczyk, Y. Martel, C. Muñoz, and H. V. D. Bosch. A sufficient condition for asymptotic stability of kinks in general (1+1)(1+1)-scalar field models. Preprint. ArXiv:2008.01276.
  • [34] H. Lindblad, J. Luhrmann, W. Schlag, and A. Soffer. On modified scattering for 1D quadratic Klein-Gordon equations with non-generic potentials. Preprint. arXiv:2012.15191, 2020.
  • [35] H. Lindblad, J. Luhrmann, and A. Soffer. Decay and asymptotics for the 1D Klein-Gordon equation with variable coefficient cubic nonlinearities. Preprint. arXiv:1907.09922, 2019.
  • [36] H. Lindblad and A. Soffer. Scattering for the Klein-Gordon equation with quadratic and variable coefficient cubic nonlinearities. Trans. Amer. Math. Soc., 367(12):8861–8909, 2015.
  • [37] M. A. Lohe. Soliton structures in P(φ)2P{(\varphi)}_{2}. Phys. Rev. D, 20:3120–3130, Dec 1979.
  • [38] N. Manton and P. Sutcliffe. Topological solitons. Cambridge Monographs on Mathematical Physics. Cambridge University Press, Cambridge, 2004.
  • [39] R. L. Pego and M. I. Weinstein. Eigenvalues, and instabilities of solitary waves. Philos. Trans. Roy. Soc. London Ser. A, 340(1656):47–94, 1992.
  • [40] J. Rauch. Perturbation theory for eigenvalues and resonances of Schrödinger Hamiltonians. J. Functional Analysis, 35(3):304–315, 1980.
  • [41] R. M. Ross, P. G. Kevrekidis, D. K. Campbell, R. Decker, and A. Demirkaya. ϕ4\phi^{4} solitary waves in a parabolic potential: existence, stability, and collisional dynamics. In A dynamical perspective on the ϕ4\phi^{4} model, volume 26 of Nonlinear Syst. Complex., pages 213–234. Springer, Cham, 2019.
  • [42] W. Schlag, A. Soffer, and W. Staubach. Decay for the wave and Schrödinger evolutions on manifolds with conical ends. I. Trans. Amer. Math. Soc., 362(1):19–52, 2010.
  • [43] H. Segur. Wobbling kinks in φ4\varphi^{4} and sine-Gordon theory. J. Math. Phys., 24(6):1439–1443, 1983.
  • [44] H. Segur and M. D. Kruskal. Nonexistence of small-amplitude breather solutions in ϕ4\phi^{4} theory. Phys. Rev. Lett., 58(8):747–750, 1987.
  • [45] B. Simon. Schrödinger operators in the twentieth century. Journal of Mathematical Physics, 41(6):3523–3555, 2000.
  • [46] S. Snelson. Asymptotic stability for odd perturbations of the stationary kink in the variable-speed ϕ4\phi^{4} model. Transactions of the American Mathematical Society, 370(10):7437–7460, 2018.
  • [47] J. Sterbenz. Dispersive decay for the 1D Klein-Gordon equation with variable coefficient nonlinearities. Trans. Amer. Math. Soc., 368(3):2081–2113, 2016.
  • [48] T. Vachaspati. Kinks and domain walls. Cambridge University Press, New York, 2006. An introduction to classical and quantum solitons.