Linear and orbital stability analysis for solitary-wave solutions of variable-coefficient scalar field equations
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 perturbation2020 Mathematics Subject Classification:
35L71, 35C07, 35P991. Introduction
We consider a semilinear, variable-coefficient scalar field equation of the form
(1.1) |
Our assumptions on the potential are
(1.2) |
The linear operator is assumed to be a perturbation of the Laplacian . More precisely, for a small parameter , we assume
(1.3) |
With , the energy functional
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 or the coefficients , , and .
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 of kink or solitary-wave type, i.e. an increasing stationary solution with as . Standard arguments then show that (1.1) is locally well-posed for initial data , where . In this context, is referred to as the energy space, and indeed, it is not hard to see (using in particular that ) that functions in this space have finite energy. Our goal is to study the stability of with respect to small perturbations in the energy space.
1.1. Motivation
One-dimensional kinks such as 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) |
In this constant-coefficient regime, it is standard that the assumptions (1.2) imply the existence of a kink solution connecting and . (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 , satisfies
(1.5) |
We find in Theorem 1.1 that and are close in an appropriate norm.
Orbital stability of 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 . In particular, the two most studied versions of (1.4) are the equation with nonlinearity , 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 , 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 , and abusing notation by writing and , we have
(1.6) |
The hypotheses (1.3) imply
(1.7) |
for some .
Our first result is the existence of a stationary kink:
Theorem 1.1.
Assume that is not an -eigenvalue of the operator . Then, for sufficiently small, there exists a solution to
(1.8) |
This solution can be written , where solves (1.5) and
Unlike , which satisfies and , our static kink 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 and , it is possible to show has exponential asymptotics at as in [46, Theorem 1.1], but we do not explore the details here.
Our next result concerns the linearized operator around . Writing , the perturbation satisfies
Adding to both sides, and defining the linear operator and the nonlinearity , the equation for can be written as a nonlinear Klein-Gordon equation:
(1.9) |
We are most interested in situations where the spectrum of , the operator corresponding to the constant-coefficient kink, is known exactly. We then have , and we ask how the perturbation changes the spectral properties of .
The spectrum of is given by
where is a possibly empty, increasing collection of positive, simple eigenvalues. The eigenfunction corresponding to is exactly , the translation invariance mode.
As expected, discrete eigenvalues will drift to nearby discrete eigenvalues of under the perturbation. A more delicate question is whether an extra discrete eigenvalue emerges from the essential spectrum. This aspect is especially relevant when has a threshold resonance, i.e. a function satisfying , as is the case for both the and sine-Gordon equations. We derive a criterion in terms of and the coefficients and that governs whether the resonance drifts into a discrete eigenvalue.
Our results on the spectrum of are collected in the following theorem:
Theorem 1.2.
Let and be as defined above. There exists a universal such that:
-
(a)
The spectrum is real, the essential spectrum , and lies in the -neighborhood of .
-
(b)
For every eigenvalue with eigenvector , there is a corresponding . The eigenvalue is real, simple, and satisfies . Also, if
then has the same sign as . The eigenfunction of corresponding to satisfies . Furthermore, for suitable normalizations of and , we have
for a universal constant .
-
(c)
If is a simple resonance of , and
then there exists a discrete eigenvalue of with . The eigenfunction also satisfies and
for suitable normalizations of and .
If
then there are no eigenvalues of in , and is non-resonant.
-
(d)
If is not a resonance or an embedded eigenvalue of , then the same is true of , and there are no eigenvalues of in .
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 , , , and , via a first-order approximation for . (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 and have a profound impact on the stability properties of the kink. At the very least, any proof of asymptotic stability or instability for 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 with respect to . Our spectral analysis is based on the well-known method of finding solutions to the eigenvalue equation which decay at , and studying the Evans function (see e.g. [15, 25, 39, 26, 27]) which is related to the Wronskian of and . The key property is that the Wronskian is zero when is an eigenvalue or resonance of . The slow decay of our coefficients and (as well as ) rules out tools such as the Gap Lemma (see [2, 17]) which would allow one to analytically continue the Evans function past the threshold , but which requires exponential decay of the coefficients.
Our last main result establishes the orbital stability of :
Theorem 1.3.
There exists an , depending on , such that for any initial data for with
the corresponding solution to (1.6) exists globally in time, and satisfies
for some depending on and .
The proof is based on classical energy arguments, but must contend with the lack of translation invariance.
1.3. Examples
1.3.1. model
The choice of a double-well potential leads to the model
(1.10) |
Standard references on this equation include [44, 6, 32, 41]. In this case, the kink solution is known explicitly, and the linearization has spectrum equal to
The odd eigenfunction corresponding to is known as the internal oscillation mode. The operator also possesses an even resonance at the threshold .
The 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 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 onto 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 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 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 results in the sine-Gordon equation:
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 . 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 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 the linearization around , it is known that
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 model in [32]. However, there is an odd resonance 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 with . 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.
1.3.3. Other examples
Let us briefly mention some other models whose variable-coefficient counterparts are included in our setting: the 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 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 kink on time scales of order , where 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 at .
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 is (up to subtraction by ) 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 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 . 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 . 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 . We quote from [33, Lemma 1.1]:
Lemma 2.1.
Under the assumptions (1.2) on , there is a solution to the stationary equation
with and as . Furthermore, and satisfy
and the energy of is finite:
We now prove the existence of a static kink for our equation (1.6):
Proof of Theorem 1.1.
Let be the stationary solution to guaranteed by Lemma 2.1. Making the ansatz , we have the following equation for :
(2.1) |
where . Defining
equation (2.1) becomes
(2.2) |
We can find solutions both satisfying , with and . In more detail, may be written as the linear system , with , and
Lemma A.2 below implies existence of and . In particular, and are linearly independent, since otherwise there would be a nontrivial solution in to , contradicting our assumption that is not an eigenvalue.
Define the Green’s function
where . Abel’s formula implies , which for sufficiently small, is bounded uniformly away from 0.
For the inverse operator , we have the following useful bounds. First,
(2.3) |
for all . We also have
(2.4) |
For the first term on the right, we integrate by parts to obtain
If , then since , the boundary term at vanishes, and the boundary term at is bounded by . After applying a similar calculation to the last term in (2.4), we conclude
(2.5) |
for a constant depending on and the coefficients . The estimates (2.3) and (2.5) clearly hold also if we replace with .
We want to find a fixed point for in the space with norm . From (2.3) and (2.5), we have
since and are bounded and . For the nonlinear term, since is on , there is some such that
(2.8) |
globally in . This gives
(2.9) |
and
so that the estimates (2.3) and (2.5) imply
after applying the standard interpolation .
With such that , define . For any , the above estimates imply
so for , we have . Next, for , we have from Taylor’s Theorem that
for some depending on . Using this in , we have
for some . By (2.3) and (2.5) we have
(2.10) |
as above. The constant depends on and the norm of . For sufficiently small, we conclude is a contraction on , and a unique solution to (2.7) exists in .
3. Perturbation of the spectrum
We consider the spectrum of
(3.1) |
where is the stationary solution guaranteed by Theorem 1.1. Defining , we have . By the regularity of , we have , and Theorem 1.1 implies . With (2.11), we can also write a first-order approximation for as follows:
(3.2) |
with as in (2.11) and .
Our goal is to investigate how the spectrum of changes under the perturbation . Since is self-adjoint with respect to the inner product
with , the spectrum is real. Since the perturbation is relatively -compact, we have . (See, e.g. [21, Chapter 14].) Given our upper bounds on and , it is standard that lies in the neighborhood of , for some . (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 in vector form:
For any , Lemma A.2(a) implies there exist satisfying , and
(3.3) |
with . For , we also obtain the integral representations
(3.4) |
For the operator , we similarly apply Lemma A.2(a) with to obtain solving , with the same boundary conditions (3.3), and for ,
(3.5) |
First, we prove a suitable approximation lemma for and for nearby values of :
Lemma 3.1.
Assume . For any compact subset of , there exists a constant such that for any , there holds
where .
Proof.
We prove only the first estimate, as the second follows by a similar argument.
Since , the mean value theorem applied to and implies, after a straightforward calculation, the inequalities
for a constant depending on . Since and is uniformly bounded on , we therefore have .
For , since is bounded on , we have , for a constant depending only on .
Define the integral kernel
From the exponential decay of we see that
is bounded by a constant depending only on and . Lemma A.1 then implies
(3.6) |
Since for a constant depending only on , the proof is complete. ∎
Now, we are ready to derive a result that governs the direction in which eigenvalues of drift under the perturbation:
Theorem 3.2.
Assume . For any eigenvalue of with eigenfunction , there exists a simple, real eigenvalue of with . Furthermore, we have the following expansion for :
In particular, if
then has the same sign as .
Remark.
Using the formula (3.2), one can show that has the same sign as
(3.7) |
Proof.
With solving (3.4), since is a simple eigenvalue, we have , for constants . Let . From our construction, it is clear that decays exponentially at a rate .
For near let and let be the solutions to (3.5) as above. From Lemma 3.1 and our assumption that , we can write
(3.8) |
with . Denote the Wronskian
By Abel’s formula, is independent of . We focus on and apply (3.8) to obtain
(3.9) |
In the second line, we used that and are parallel, and in the last line, we used and . Since
we can use (3.8) again to write
and
Feeding these expressions into (3.9), we obtain
From the approximation and the exponential decay of we have, for an eigenvalue of ,
which implies the first-order expansion for in the statement of the theorem. ∎
Now, we analyze the threshold resonance , which is an function solving . First, we prove a modified version of Lemma 3.1 for the borderline case . This will be useful in tracking how a threshold resonance of translates to the spectrum of . Writing in vector form as above, Lemma A.2(b) implies
(3.10) |
Lemma 3.3.
Assume . For any , there exists a constant such that for any , there holds
where .
Proof.
The proof is similar to Lemma 3.1, with the difference that satisfies the modified integral equation (3.10). From (3.10) and (3.5), we have
with
and defined as in the proof of Lemma 3.1, with replacing .
We claim that . Indeed, applying the mean value theorem to gives
or . Next, Taylor’s Theorem implies , with . We have and , since . This gives , with , or
Plugging these inequalities into the definition of and using decay of gives the desired estimate.
In the following theorem, we make the (mild) assumption that the limits at of are nonzero.
Theorem 3.4.
(a) Assume that is a simple resonance for , i.e. that there exists with . Then there exists depending only on the function , such that if and
then there exists a discrete eigenvalue of with . If
then there is no discrete eigenvalue of in a neighborhood of the essential spectrum, i.e. the discrete spectrum consists of the same number of eigenvalues as .
(b) On the other hand, if is nonresonant and not an eigenvalue of , then for is sufficiently small, cannot be a resonance or an eigenvalue of , and there is no eigenvalue of in a neighborhood of the essential spectrum.
Remark.
As above, using (3.2), the quantity has the same sign as
(3.11) |
Proof.
With solving (3.10), we have , for constants .
Our first step is to analyze the unperturbed Wronskian . With near and , by abuse of notation, we write . Applying Lemma 3.3 with , we may write
with . By the equation satisfied by , the Wronskian is independent of . Proceeding as in the proof of Theorem 3.2, we have as before (see (3.9))
(3.12) |
A direct calculation shows , which gives, since ,
(3.13) |
Note that all integrals converge, since , , and are all uniformly bounded.
In the last expression of (3.13), we note that the first term is proportional to . Indeed, since has non-zero limits as , there exist (independent of ) such that if . As a result, for any , one has . On the other hand, we have . It is also clear that, since , the second term on the right in (3.13) is . To sum up, we have shown
(3.14) |
with for some independent of .
Now we turn to the perturbed operator . Let be the solutions to (3.5) as above. Applying Lemma 3.3 with , we have
(3.15) |
with .
With the Wronskian defined as in the proof of Theorem 3.2, we again write , and obtain
(3.16) |
Since satisfy , we have
Because , , and , the expression is integrable on , and we can use (3.15) to write
where the last integral converges and is since and . For the first integral on the right, we use Lemma 3.1 with and obtain
After applying a similar analysis to , the expression (3.16) becomes
With (3.14), this implies
For small enough, the expression inside the parentheses determines whether any zeroes of are present for . The bound with implies statement (a) of the theorem.
For statement (b), the assumption that is not a resonance or eigenvalue implies . The approximation (3.15) easily implies for small enough. ∎
4. Orbital stability
In this section, we prove orbital stability, i.e. that solutions starting close to are always close to some shifted version of .
Proof of Theorem 1.3.
For any solution of (1.6), the energy
is conserved, where , uniformly in . We also define the potential energy
A simple computation shows that
(4.1) |
where is the potential energy corresponding to the constant coefficient equation (1.4):
The idea is to use the (known) property that controls the distance between and , to show the corresponding fact for and . In more detail, for , define
for any in the energy space. We define in the analogous way. Proposition 1 of [20] proves the following: There exist such that
whenever .aaaProposition 1 in [20] is stated for where is a solution of (1.4), but an examination of the proof shows that the conclusion holds for any satisfying the hypotheses stated here. Note that
by Theorem 1.1. Using (4.1) twice, we then have
(4.2) |
To get to the last line, we used Sobolev embedding to write , and combined this term into the left-hand side.
Since as , there is some where the infimum defining is achieved. To save space, write , and similarly for and . We then have
(4.3) |
For small enough compared to and , this implies . By exchanging the roles of and in this calculation, we also obtain .
Next, combining (4.2) and (4.3),
(4.4) |
This inequality holds for such that . By above, we can ensure this condition by choosing .
Now, for a solution to (1.6) with and sufficiently small in , (4.4) implies that
The quantity is conserved in time. Calculations similar to (4.1) show that , and the last term may be combined into the left side. We finally have
This right-hand side is independent of , which implies the solution never leaves the neighborhood of as long as it exists. As above, for every , there is some at which the infimum defining is achieved. By standard arguments, this time-independent bound on combined with energy conservation implies the solution exists for all . ∎
Appendix A ODE Methods
In this section, we collect some convenient facts about the solvability and asymptotics of first-order systems on .
First, we have a standard lemma on vector-valued integral equations of Volterra type:
Lemma A.1.
For and , the Volterra equation
has a unique solution in , provided
(A.1) |
where is the operator norm of the matrix . This solution is given by the iteration
(A.2) |
with . This solution satisfies
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.
-
(a)
For , consider the system
(A.3) where
with . There exist solutions , defined on , such that
(A.4) and the bound holds for all , where the constant depends on and . These solutions also satisfy the integral equations
(A.5) -
(b)
For , assume in addition that and lie in . Then there exist solutions to (A.3) satisfying
as well as the integral equations
(A.6)
Proof.
(a) Note that the eigenvalues of are corresponding to eigenvectors . We will find a solution to the integral equation
(A.7) |
satisfying and . By direct calculation, such also solves (A.3), as well as the first integral equation in (A.5). Letting , (A.7) is equivalent to
(A.8) |
By diagonalizing , we obtain
With , we therefore have
and that
Lemma A.1 now implies a solution to (A.8) exists on , and is bounded by a constant, which implies the boundary condition (A.4) holds for , as well as the upper bound
where . Applying a similar argument with replacing , we can obtain a solution defined on with
For , we can write
with chosen so that and match our previous definition. This formula implies solves (A.3) and satisfies for negative also. By a similar method, we extend to the real line and obtain for all .
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 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 , , and 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 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 -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 . 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. solitary waves in a parabolic potential: existence, stability, and collisional dynamics. In A dynamical perspective on the 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 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 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 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.