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

Breather Solutions of the Cubic Klein-Gordon Equation

Dominic Scheider D. Scheider Karlsruhe Institute of Technology Institute for Analysis Englerstraße 2 D-76131 Karlsruhe, Germany [email protected]
Abstract.

We obtain real-valued, time-periodic and radially symmetric solutions of the cubic Klein-Gordon equation

t2UΔU+m2U=Γ(x)U3on ×3,\partial_{t}^{2}U-\Delta U+m^{2}U=\Gamma(x)U^{3}\quad\text{on }\mathbb{R}\times\mathbb{R}^{3},

which are weakly localized in space. Various families of such “breather” solutions are shown to bifurcate from any given nontrivial stationary solution. The construction of weakly localized breathers in three space dimensions is, to the author’s knowledge, a new concept and based on the reformulation of the cubic Klein-Gordon equation as a system of coupled nonlinear Helmholtz equations involving suitable conditions on the far field behavior.

Key words and phrases:
Klein-Gordon equation, breather, bifurcation, Helmholtz equation
2010 Mathematics Subject Classification:
Primary: 35L71, 35B32, Secondary: 35B10, 35J05

1. Introduction and Main Results

We construct real-valued solutions U(t,x)U(t,x) of the cubic Klein-Gordon equation

(1) t2UΔU+m2U=Γ(x)U3on ×3\partial_{t}^{2}U-\Delta U+m^{2}U=\Gamma(x)\>U^{3}\qquad\text{on }\mathbb{R}\times\mathbb{R}^{3}

where ΓLrad(3)Cloc1(3)\Gamma\in L^{\infty}_{\mathrm{rad}}(\mathbb{R}^{3})\cap C^{1}_{\mathrm{loc}}(\mathbb{R}^{3}) and m>0m>0 is a (mass) parameter. Here we restrict ourselves to the case of three space dimensions which is the most relevant one for applications in physics and which allows to use the tools established in [19]. Throughout, the notations 1,2,3,,Δ,D2\partial_{1,2,3},\nabla,\Delta,D^{2} refer to differential operators acting on the space variables. The solutions we aim to construct are polychromatic, that is, they take the form

(2) U(t,x)=u0(x)+k=12cos(ωkt)uk(x)=keiωktuk(x)\displaystyle U(t,x)=u_{0}(x)+\sum_{k=1}^{\infty}2\,\cos(\omega kt)u_{k}(x)=\sum_{k\in\mathbb{Z}}\mathrm{e}^{\mathrm{i}\omega kt}u_{k}(x)
whereukX1={uCrad(3,)|(1+||2)12u<}Lrad4(3),uk=uk\displaystyle\text{where}\>\>u_{k}\in X_{1}=\left\{u\in C_{\mathrm{rad}}(\mathbb{R}^{3},\mathbb{R})\big{|}\>\left\lVert(1+|\cdot|^{2})^{\frac{1}{2}}u\right\rVert_{\infty}<\infty\right\}\subseteq L^{4}_{\mathrm{rad}}(\mathbb{R}^{3}),\>\>u_{-k}=u_{k}
and (for simplicity)ω>m.\displaystyle\text{and (for simplicity)}\>\>\omega>m.

Such solutions are periodic in time and localized as well as radially symmetric in space. They are sometimes referred to as breather solutions, c.f. the “Sine-Gordon breather” in [1], equation (28). The construction of breather solutions is of particular interest since, as indicated in a study [7] on perturbations of the Sine-Gordon breather, Birnir, McKean and Weinstein conjecture that “for the general nonlinear wave equation [author’s note: in 1+1 dimensions], breathing […] takes place only for isolated nonlinearities”, see  [7, p.1044]. This conjecture is supported by recent existence results for breathers for the 1+1 dimensional wave equation with specific, carefully designed potentials which we comment on below. Our results, however, indicate that the situation might be entirely different for weakly localized breathers for the Klein-Gordon equation in 1+3 dimensions, in the sense that such breather solutions are abundant even in “simple” settings.

We will find breather solutions of (1) with uk0u_{k}\not\equiv 0 for at least two distinct integers k0k\in\mathbb{N}_{0} by rewriting it into an infinite system of stationary equations for the functions uku_{k}. Indeed, inserting (2), a short and formal calculation leads to

(3a) Δu0+m2u0\displaystyle-\Delta u_{0}+m^{2}\,u_{0} =Γ(x)(𝐮𝐮𝐮)0,\displaystyle=\Gamma(x)\>\left(\mathbf{u}\star\mathbf{u}\star\mathbf{u}\right)_{0},
(3b) Δuk(ω2k2m2)uk\displaystyle-\Delta u_{k}-(\omega^{2}k^{2}-m^{2})u_{k} =Γ(x)(𝐮𝐮𝐮)kfor k{0}.\displaystyle=\Gamma(x)\>\left(\mathbf{u}\star\mathbf{u}\star\mathbf{u}\right)_{k}\qquad\text{for }k\in\mathbb{Z}\setminus\{0\}.

In fact, (3b) includes (3a), but we intend to separate the “Schrödinger” equation characterized by 0σ(Δ+m2)0\not\in\sigma(-\Delta+m^{2}) from the infinite number of “Helmholtz” equations characterized by 0σ(Δ(ω2k2m2))0\in\sigma(-\Delta-(\omega^{2}k^{2}-m^{2})), k0k\neq 0. Our construction of breathers for (1) relies on new methods for such Helmholtz equations introduced in [19] which exploit the so-called far field properties of their solutions and lead to a rich bifurcation structure. These methods will be sketched only briefly in the main body of this paper; more details will be given in Section 4 at the end (which can be read independently).

The solutions we obtain bifurcate from any given stationary (radial) solution w0X1w_{0}\in X_{1}, w00w_{0}\not\equiv 0 of the Klein-Gordon equation (1). That is, w0w_{0} solves the stationary nonlinear Schrödinger equation

(4) Δw0+m2w0=Γ(x)w03on 3;\displaystyle-\Delta w_{0}+m^{2}\,w_{0}=\Gamma(x)\>w_{0}^{3}\quad\text{on }\mathbb{R}^{3};

regarding existence of such w0w_{0}, cf. Remark 1 (b). Let us remark briefly that all (distributional) solutions of (4) in X1Lrad4(3)X_{1}\subseteq L^{4}_{\text{rad}}(\mathbb{R}^{3}) are twice differentiable by elliptic regularity. In order to make bifurcation theory work, we impose the following nondegeneracy assumption:

(5) q0X1,Δq0+q0=3Γ(x)w02q0 on 3implies q00.\displaystyle q_{0}\in X_{1},\>\>-\Delta q_{0}+q_{0}=3\Gamma(x)\>w_{0}^{2}\>q_{0}\text{ on }\mathbb{R}^{3}\qquad\text{implies }\qquad q_{0}\equiv 0.

We comment on this assumption in Remark 1 (c) below. In particular, (5) and our main result presented next hold if Γ\Gamma is constant and w0w_{0} is a (positive) ground state of (4). We now present our main result.

Theorem 1.

Let ΓLrad(3)Cloc1(3)\Gamma\in L^{\infty}_{\mathrm{rad}}(\mathbb{R}^{3})\cap C^{1}_{\mathrm{loc}}(\mathbb{R}^{3}), ω>m>0\omega>m>0 and assume there is some stationary solution U0(t,x)=w0(x)U^{0}(t,x)=w_{0}(x), w00w_{0}\not\equiv 0 of the cubic Klein-Gordon equation (1), i.e. w0X1w_{0}\in X_{1} solving (4). Assume further that w0w_{0} is nondegenerate in the sense of (5). Then for every ss\in\mathbb{N} there exist an open interval JsJ_{s}\subseteq\mathbb{R} with 0Js0\in J_{s} and a family (Uα)αJsC2(,X1)(U^{\alpha})_{\alpha\in J_{s}}\subseteq C^{2}(\mathbb{R},X_{1}) with the following properties:

  • (i)

    All UαU^{\alpha} are time-periodic, twice continuously differentiable classical solutions of (1) of the polychromatic form (2),

    Uα(t,x)=u0α(x)+k=12cos(ωkt)ukα(x).\displaystyle U^{\alpha}(t,x)=u_{0}^{\alpha}(x)+\sum_{k=1}^{\infty}2\,\cos(\omega kt)u_{k}^{\alpha}(x).
  • (ii)

    The map α(ukα)k0\alpha\mapsto(u_{k}^{\alpha})_{k\in\mathbb{N}_{0}} is smooth in the topology of 1(0,X1)\ell^{1}(\mathbb{N}_{0},X_{1}) with

    ddα|α=0ukα0if and only ifk=s\displaystyle\frac{\mathrm{d}}{\mathrm{d}\alpha}\bigg{|}_{\alpha=0}u_{k}^{\alpha}\not\equiv 0\quad\text{if and only if}\quad k=s

    (“excitation of the s-th mode”). In particular, for sufficiently small α0\alpha\neq 0, these solutions are non-stationary. Moreover, for different values of ss, the families of solutions mutually differ close to U0U^{0}.

  • (iii)

    If we assume additionally Γ(x)0\Gamma(x)\neq 0 for almost all x3x\in\mathbb{R}^{3}, then every nonstationary polychromatic solution UαU^{\alpha} possesses infinitely many nonvanishing modes ukαu_{k}^{\alpha}.

Remark 1.
  • (a)

    We require continuity of Γ\Gamma since we use the functional analytic framework of [19]. The existence and continuity of Γ\nabla\Gamma will be exploited in proving that UαU^{\alpha} is twice differentiable. This assumption as well as Γ0\Gamma\neq 0 almost everywhere in (iii) might be relaxed; however, this study does not aim at the most general setting for the coefficients but rather focuses on the introduction of the setup for the existence result.

  • (b)

    The existence of stationary solutions of the Klein-Gordon equation (1) resp. of solutions to (4) can be guaranteed under additional assumptions on Γ\Gamma. We refer to [18], Theorem I.2 and Remarks I.5, I.6 by Lions for positive (ground state) solutions and to Theorems 2.1 of [5][4] by Bartsch and Willem for bound states.

  • (c)

    In some special cases, nondegeneracy properties like (5) have been verified, e.g. by Bates and Shi [6] in Theorem 5.4 (6), or by Wei [25] in Lemma 4.1, both assuming that w0w_{0} is a ground state solution of (4) in the autonomous case with constant positive Γ\Gamma. It should be pointed out that, although the quoted results discuss nondegeneracy in a setting on the Hilbert space H1(3)H^{1}(\mathbb{R}^{3}), the statements can be adapted to the topology of X1X_{1}, as we will demonstrate in Lemma 1.

  • (d)

    The assumption ω>m\omega>m on the frequency ensures that the stationary system (3) contains only one equation of Schrödinger type. This avoids further nondegeneracy assumptions on higher modes, which would not be covered by the previously mentioned results in the literature.

  • (e)

    The above result provides, locally, a multitude of families of breathers bifurcating from every given stationary solution characterized by different values of ss, ω\omega and possibly certain asymptotic parameters, see Remark 2 below.
    It would be natural, further, to ask for the global bifurcation picture given some trivial family 𝒯={(w0,λ)|λ}\mathcal{T}=\{(w_{0},\lambda)\,|\,\lambda\in\mathbb{R}\}. (Here λ\lambda\in\mathbb{R} denotes a bifurcation parameter which in our case is not visible in the differential equation and thus will be properly introduced later.) Typically, global bifurcation theorems state that a maximal bifurcating continuum of solutions (U,λ)(U,\lambda) emanating from 𝒯\mathcal{T} at (w0,λ0)(w_{0},\lambda_{0}) is unbounded unless it returns to 𝒯\mathcal{T} at some point (w0,λ0)(w_{0},\lambda_{0}^{\prime}), λ0λ0\lambda_{0}^{\prime}\neq\lambda_{0}. In the former (desirable) case, however, a satisfactory characterization of global bifurcation structures should provide a criterion whether or not unboundedness results from another stationary solution w1w0w_{1}\neq w_{0} with {(w1,λ)|λ}\{(w_{1},\lambda)\,|\,\lambda\in\mathbb{R}\} belonging to the maximal continuum. Since it is not obvious at all whether and how such a criterion might be derived within our framework, we focus on the local result, which already adds new aspects to the state of knowledge about the existence of breather solutions summarized next.

1.1. An Overview of Literature

Polychromatic Solutions

The results in Theorem 1 can and should be compared with recent findings on breather (that is to say, time-periodic and spatially localized) solutions of the wave equation with periodic potentials V(x),q(x)=cV(x)0V(x),q(x)=c\cdot V(x)\geq 0,

(6) V(x)t2Ux2U+q(x)U=Γ(x)U3on ×.\displaystyle V(x)\partial_{t}^{2}U-\partial_{x}^{2}U+q(x)U=\Gamma(x)U^{3}\qquad\text{on }\mathbb{R}\times\mathbb{R}.

Such breather solutions have been constructed by Schneider et al., see Theorem 1.1 in [8], and Hirsch and Reichel, see Theorem 1.3 in [13], respectively. In brief, the main differences to the results in this article are that the authors of [8][13] consider a setting in one space dimension and obtain strongly spatially localized solutions, which requires a comparably huge technical effort. We give some details: Both existence results are established using a polychromatic ansatz, which reduces the time-dependent equation to an infinite set of stationary problems with periodic coefficients, see [8], p. 823, resp. [13], equation (1.2). The authors of [8] apply spatial dynamics and center manifold reduction; their ansatz is based on a very explicit choice of the coefficients q,V,Γq,V,\Gamma. The approach in [13] incorporates more general potentials and nonlinearities and is based on variational techniques. It provides ground state solutions, which are possibly “large” - in contrast to our local bifurcation methods, which only yield solutions close to a given stationary one as described in Theorem 1, i.e. with a typically “small” time-dependent contribution.
Periodicity of the potentials in (6) is explicitly required since it leads to the occurrence of spectral gaps when analyzing the associated differential operators of the stationary equations. In contrast to the Helmholtz methods introduced here, the authors both of [8] and of [13] strive to construct the potentials in such way that 0 lies in the aforementioned spectral gaps, and moreover that the distance between 0 and the spectra has a positive lower bound. This is realized by assuming a certain “roughness” of the potentials, referring to the step potential defined in Theorem 1.1 of [8] and to the assumptions (P1)-(P3) in [13] which allow potentials with periodic spikes modeled by Dirac delta distributions, periodic step potentials or some specific, non-explicit potentials in Hradr()H^{r}_{\text{rad}}(\mathbb{R}) with 1r<321\leq r<\frac{3}{2} (see [13], Lemma 2.8).

Let us summarize that the methods for constructing breather solutions of (6) outlined above can handle periodic potentials but require irregularity, are very restrictive concerning the form of the potentials and involve a huge technical effort in analyzing spectral properties based on Floquet-Bloch theory. The Helmholtz ansatz presented in this article provides a technically elegant and short approach suitable for constant potentials; in the context of breather solutions, it is new in the sense that it provides breathers with slow decay, it provides breathers on the full space 3\mathbb{R}^{3}, and it provides breathers for simple (constant) potentials.

The Klein-Gordon Equation as a Cauchy Problem

Possibly due to its relevance in physics, there is a number of classical results in the literature concerning the nonlinear Klein-Gordon equation. The fundamental difference to the results in this article is that the vast majority of these concerns the Cauchy problem of the Klein-Gordon equation, i.e.

(7) t2UΔU+m2U=±U3on [0,)×3U(0,x)=f(x),tU(0,x)=g(x) on 3\begin{split}&\partial_{t}^{2}U-\Delta U+m^{2}\,U=\pm U^{3}\quad\text{on }[0,\infty)\times\mathbb{R}^{3}\\ &U(0,x)=f(x),\>\partial_{t}U(0,x)=g(x)\quad\text{ on }\mathbb{R}^{3}\end{split}

for suitable initial data f,g:3f,g:\mathbb{R}^{3}\to\mathbb{R}. Usually, the dependence of the nonlinearity on UU is much more general (allowing also derivatives of UU) and the space dimension is not restricted to N=3N=3. On the other hand, most results in the literature only concern the autonomous case, which is why we set in this discussion Γ±1\Gamma\equiv\pm 1.

An overview of the state of knowledge towards the end of the 1970s can be found e.g. in [24] by Strauss, who discusses among other topics global existence (Theorem 1.1), regularity and uniqueness (Theorem 1.2), blow-up (Theorem 1.4) and scattering (Theorem 4.1). In the first-mentioned result, which is originally due to Jörgens, global existence of distributional solutions with locally as well as globally finite energy

EB[U(t,)]=12B|tU(t,x)|2+|U(t,x)|2+m2|U(t,x)|2dx+14B|U(t,x)|4dx,B3\displaystyle E_{B}[U(t,\,\cdot\,)]=\frac{1}{2}\int_{B}|\partial_{t}U(t,x)|^{2}+|\nabla U(t,x)|^{2}+m^{2}|U(t,x)|^{2}\>\mathrm{d}x+\frac{1}{4}\int_{B}|U(t,x)|^{4}\>\mathrm{d}x,\>\>B\subseteq\mathbb{R}^{3}

is proved provided Γ1\Gamma\equiv-1. Following a classical strategy for evolution problems, local existence is shown by means of a fixed point iteration, and global existence can be obtained by an iteration argument based on energy conservation. For Γ+1\Gamma\equiv+1, Theorem 1.4 due to Keller and Levine demonstrates the existence of blow-up solutions.
During the following decade, Klainerman [16, 15] and Shatah [22, 23] independently developed new techniques leading to significant improvements in the study of uniqueness questions and of the asymptotic behavior of solutions as tt\to\infty. These results work in settings with high regularity and admit more general nonlinearities with growth assumptions for small arguments, which includes the cubic one as a special case. In particular, Klainerman and Shatah prove the convergence to solutions of the free Klein-Gordon equation and show uniform decay rates of solutions as tt\to\infty. In the case of a cubic nonlinearity, these results only apply if the space dimension is at least 22. This is why, more recently, the question of corresponding uniqueness and convergence properties for cubic nonlinearities in N=1N=1 space dimensions has attracted attention; we wish to mention at least some of the related papers. For explicit choices of the cubic nonlinearity, there are results by Moriyama and by Delort, see Theorem 1.1 of [20] resp. Théorèmes 1.2, 1.3 in [9]. Only the latter result allows a nonlinearity of the form ±U3\pm U^{3} not containing derivatives (see [9], Remarque 1.4); however, the initial data are assumed to have compact support. Global existence, uniqueness, decay rates and scattering exclusively for the nonlinearity ±U3\pm U^{3} can be found in Corollary 1.2 of [12] by Hayashi and Naumkin.

The relation to our results is not straightforward since the bifurcation methods automatically provide solutions UαU^{\alpha} which exist globally in time irrespective of the sign (or even of a possible xx-dependence) of Γ\Gamma and which do not decay as tt\to\infty, and there is no special emphasis on the role of the initial values Uα(0,x),Uα(0,x)U^{\alpha}(0,x),\nabla U^{\alpha}(0,x) along the bifurcating branches. Our methods instead focus on several global properties of the solutions Uα(t,x)U^{\alpha}(t,x) such as periodicity in time and localization as well as decay rates in space, i.e. the defining properties of breathers.

1.2. Research Perspectives

Apart from bifurcation methods, nonlinear Helmholtz equations and systems can also be discussed in a “dual” variational framework as introduced by Evéquoz and Weth [10]. This might offer another way to analyze the system (3) leading to “large” breathers in the sense that they are not close to a given stationary solution as the ones constructed in Theorem 1. Furthermore, such an ansatz might be a promising step towards extensions to non-constant, e.g. periodic potentials.

2. The Proof of Theorem 1

2.1. The Functional-Analytic Setting

We look for polychromatic solutions as in (2) with coefficients 𝐮=(uk)k𝒳1\mathbf{u}=(u_{k})_{k\in\mathbb{Z}}\in\mathcal{X}_{1} where

𝒳1:=sym1(,X1):={(uk)k|uk=ukX1,(uk)k𝒳1:=kukX1<}.\displaystyle\mathcal{X}_{1}:=\ell^{1}_{\text{sym}}(\mathbb{Z},X_{1}):=\left\{(u_{k})_{k\in\mathbb{Z}}\>\bigg{|}\>u_{k}=u_{-k}\in X_{1},\left\lVert(u_{k})_{k\in\mathbb{Z}}\right\rVert_{\mathcal{X}_{1}}:=\sum_{k\in\mathbb{Z}}\left\lVert u_{k}\right\rVert_{X_{1}}<\infty\right\}.

The Banach space X1X_{1} has been defined in (2); it prescribes a decay rate which is the natural one for solutions of Helmholtz equations as in (3b), see also Section 4. Throughout, we denote by 𝐰=(δk,0w0)k=(,0,w0,0,)\mathbf{w}=(\delta_{k,0}w_{0})_{k\in\mathbb{Z}}=(...,0,w_{0},0,...) the stationary solution with w0X1Cloc2(3)w_{0}\in X_{1}\cap C^{2}_{\mathrm{loc}}(\mathbb{R}^{3}) fixed according to equation (4). We will find polychromatic solutions of (1) by solving the countably infinite Schrödinger-Helmholtz system (3a), (3b), which is equivalent to (1), (2) on a formal level; for details including convergence of the polychromatic sum in (2), see Proposition 3.

Our strategy is then as follows:

  • \triangleright

    Intending to apply bifurcation techniques, we have to analyze the linearized version of the infinite-dimensional system (3a), (3b), which resembles the one of the two-component system discussed by the author in [19]. We therefore summarize, for the reader’s convenience, a collection of results concerning the linearized setting in Proposition 2.

  • \triangleright

    We then present a suitable setup for bifurcation theory; in particular, we introduce a bifurcation parameter which is not visible in the differential equation but appears in the so-called far field of the functions uku_{k}, more specifically a phase parameter in the leading-order contribution as |x||x|\to\infty.

  • \triangleright

    The aforementioned fact that solutions of (3a), (3b) obtained in this setting provide polychromatic, classical solutions of the Klein-Gordon equation (1) will be proved as a part of Proposition 3 below. Indeed, regarding differentiability, we will see that the choice of suitable asymptotic conditions will ensure uniform convergence and hence smoothness properties of the infinite sums defining the polychromatic states.

  • \triangleright

    Finally, in Proposition 4, we essentially verify the assumptions of the Crandall-Rabinowitz Bifurcation Theorem.

After that, we are able to give a very short proof of Theorem 1. The auxiliary results will be proved in Section 3. The final Section 4 provides some more details on the theory of linear Helmholtz equations in X1X_{1}.

Throughout, we denote the convolution in 3\mathbb{R}^{3} by the symbol \ast and use \star in the convolution algebra 1\ell^{1}. Extending the notation defined above, for q0q\geq 0, we let

Xq:={uCrad(3,)|uXq<}\displaystyle X_{q}:=\left\{u\in C_{\text{rad}}(\mathbb{R}^{3},\mathbb{R})\,|\,\left\lVert u\right\rVert_{X_{q}}<\infty\right\} with uXq:=supx3(1+|x|2)q/2|u(x)|,\displaystyle\text{with }\left\lVert u\right\rVert_{X_{q}}:=\sup_{x\in\mathbb{R}^{3}}(1+|x|^{2})^{q/2}|u(x)|,
𝒳q:=sym1(,Xq)\displaystyle\mathcal{X}_{q}:=\ell^{1}_{\text{sym}}(\mathbb{Z},X_{q}) with 𝐮𝒳q:=(uk)k𝒳q:=kukXq.\displaystyle\text{with }\left\lVert\mathbf{u}\right\rVert_{\mathcal{X}_{q}}:=\left\lVert(u_{k})_{k}\right\rVert_{\mathcal{X}_{q}}:=\sum_{k\in\mathbb{Z}}\left\lVert u_{k}\right\rVert_{X_{q}}.
Proposition 1.

The convolution of sequences 𝐮(1),𝐮(2),𝐮(3)𝒳1\mathbf{u}^{(1)},\mathbf{u}^{(2)},\mathbf{u}^{(3)}\in\mathcal{X}_{1} is well-defined in a pointwise sense and satisfies 𝐮(1)𝐮(2)𝐮(3)𝒳3\mathbf{u}^{(1)}\star\mathbf{u}^{(2)}\star\mathbf{u}^{(3)}\in\mathcal{X}_{3}. Moreover, we have the estimate

𝐮(1)𝐮(2)𝐮(3)𝒳3𝐮(1)𝒳1𝐮(2)𝒳1𝐮(3)𝒳1.\displaystyle\left\lVert\mathbf{u}^{(1)}\star\mathbf{u}^{(2)}\star\mathbf{u}^{(3)}\right\rVert_{\mathcal{X}_{3}}\leq\left\lVert\mathbf{u}^{(1)}\right\rVert_{\mathcal{X}_{1}}\left\lVert\mathbf{u}^{(2)}\right\rVert_{\mathcal{X}_{1}}\left\lVert\mathbf{u}^{(3)}\right\rVert_{\mathcal{X}_{1}}.

We rewrite the system (3a), (3b) using 𝐮=𝐰+𝐯\mathbf{u}=\mathbf{w}+\mathbf{v} with 𝐰=(,0,w0,0,)\mathbf{w}=(...,0,w_{0},0,...); then,

(8) Δvk(ω2k2m2)vk=Γ(x)[((𝐰+𝐯)(𝐰+𝐯)(𝐰+𝐯))kδk,0w03]on 3.\displaystyle-\Delta v_{k}-(\omega^{2}k^{2}-m^{2})\,v_{k}=\Gamma(x)\cdot\left[\left((\mathbf{w}+\mathbf{v})\star(\mathbf{w}+\mathbf{v})\star(\mathbf{w}+\mathbf{v})\right)_{k}-\delta_{k,0}w_{0}^{3}\right]\quad\text{on }\mathbb{R}^{3}.

We will find solutions of this system of differential equations by solving instead a system of coupled convolution equations which, for k{0,±s}k\not\in\{0,\pm s\}, have the form vk=μkτk[fk]v_{k}=\mathcal{R}_{\mu_{k}}^{\tau_{k}}[f_{k}]. Here fkf_{k} represents the right-hand side of (8), μk:=ω2k2m2\mu_{k}:=\omega^{2}k^{2}-m^{2}, and the coefficients τk(0,π)\tau_{k}\in(0,\pi) will have to be chosen properly according to a nondegeneracy condition. The convolution operators

μτ=sin(||μ+τ)4πsin(τ)||:X3X1(μ>0,  0<τ<π)\mathcal{R}_{\mu}^{\tau}=\frac{\sin(|\,\cdot\,|\sqrt{\mu}+\tau)}{4\pi\sin(\tau)|\,\cdot\,|}\>\ast\>:X_{3}\to X_{1}\qquad(\mu>0,\>\>0<\tau<\pi)

can be viewed as resolvent-type operators for the Helmholtz equation (Δμ)v=f(-\Delta-\mu)v=f on 3\mathbb{R}^{3} involving an asymptotic condition on the far field of the solution vv, namely

|x|v(x)sin(|x|μ+τ)+O(1|x|)as|x|.|x|\>v(x)\sim\sin(|x|\sqrt{\mu}+\tau)+O\left(\frac{1}{|x|}\right)\quad\text{as}\quad|x|\to\infty.

Such conditions are required since the homogeneous Helmholtz equation (Δμ)v=0(-\Delta-\mu)v=0 has smooth nontrivial solutions in X1X_{1} (known as Herglotz waves), which are all multiples of

Ψ~μ(x):=sin(|x|μ)4π|x|(x0).\tilde{\Psi}_{\mu}(x):=\frac{\sin(|x|\sqrt{\mu})}{4\pi|x|}\quad(x\neq 0).

We refer to Section 4, more precisely Lemma 5, for details; the case τ=0\tau=0 requires a larger technical effort and is presented in Lemma 6. This involves linear functionals α(μ),β(μ)X1\alpha^{(\mu)},\beta^{(\mu)}\in X_{1}^{\prime} which, essentially, yield the coefficients of the sine resp. cosine terms in the asymptotic expansion above. Relying on these tools and notations, we summarize the relevant facts on the linearized versions of the Helmholtz equations (3b) in the following Proposition.

Proposition 2.

Let w0X1w_{0}\in X_{1} be a solution of equation (4) with ΓLrad(3)Cloc(3)\Gamma\in L^{\infty}_{\mathrm{rad}}(\mathbb{R}^{3})\cap C_{\mathrm{loc}}(\mathbb{R}^{3}) and ω>m>0\omega>m>0; define μk:=ω2k2m2\mu_{k}:=\omega^{2}k^{2}-m^{2}. For every k{0}k\in\mathbb{Z}\setminus\{0\}, there exists (up to a multiplicative constant) a unique nontrivial and radially symmetric solution qkX1q_{k}\in X_{1} of

(9a) Δqkμkqk=3Γ(x)w02(x)qkon 3.\displaystyle-\Delta q_{k}-\mu_{k}\,q_{k}=3\,\Gamma(x)w_{0}^{2}(x)\>q_{k}\qquad\text{on }\mathbb{R}^{3}.
It is twice continuously differentiable and satisfies, for some ck0c_{k}\neq 0 and σk[0,π)\sigma_{k}\in[0,\pi),
(9b) qk(x)=cksin(|x|μk+σk)|x|+O(1|x|2)as |x|.\displaystyle q_{k}(x)=c_{k}\cdot\frac{\sin(|x|\,\sqrt{\mu_{k}}+\sigma_{k})}{|x|}+O\left(\frac{1}{|x|^{2}}\right)\quad\text{as }|x|\to\infty.

The equations (9a), (9b) are equivalent to the convolution identities

{qk=3μkσk[Γw02qk]=3(μk[Γw02qk]+cot(σk)~μk[Γw02qk])if σk(0,π),qk=3μkπ/2[Γw02qk]+(α(μk)(qk)+β(μk)(qk))Ψ~μkif σk=0.\displaystyle\begin{cases}q_{k}=3\>\mathcal{R}_{\mu_{k}}^{\sigma_{k}}[\Gamma w_{0}^{2}\,q_{k}]=3\>\left(\mathcal{R}_{\mu_{k}}[\Gamma w_{0}^{2}\,q_{k}]+\cot(\sigma_{k})\tilde{\mathcal{R}}_{\mu_{k}}[\Gamma w_{0}^{2}\,q_{k}]\right)&\text{if }\sigma_{k}\in(0,\pi),\\ q_{k}=3\>\mathcal{R}_{\mu_{k}}^{\pi/2}[\Gamma w_{0}^{2}\,q_{k}]+\left(\alpha^{(\mu_{k})}(q_{k})+\beta^{(\mu_{k})}(q_{k})\right)\cdot\tilde{\Psi}_{\mu_{k}}&\text{if }\sigma_{k}=0.\end{cases}

For all kk\in\mathbb{Z}, cos(σk)β(μk)(qk)=sin(σk)α(μk)(qk)\cos(\sigma_{k})\,\beta^{(\mu_{k})}(q_{k})=\sin(\sigma_{k})\,\alpha^{(\mu_{k})}(q_{k}).

The existence statement and the asymptotic properties in (9) can be proved using the Prüfer transformation, see [19], Proposition 6; the statements in the second part are consequences of Lemmas 5 and 6 in the final Section 4. For these results to apply we have assumed initially that Γ\Gamma is continuous and bounded, whence 3Γw02X23\,\Gamma w_{0}^{2}\in X_{2}.

We now present the general assumptions valid throughout the following construction and the proof of Theorem 1. We let σk\sigma_{k} for k{0}k\in\mathbb{Z}\setminus\{0\} as in Proposition 2 above and fix ss\in\mathbb{N}, recalling that we aim to “excite the ss-th mode” in the sense of Theorem 1 (ii). With this, let us introduce

(10) τ±s:=σ±s,τk:={π4if σkπ4,3π4if σk=π4 for k{0,±s},\tau_{\pm s}:=\sigma_{\pm s},\qquad\tau_{k}:=\begin{cases}\frac{\pi}{4}&\text{if }\sigma_{k}\neq\frac{\pi}{4},\\ \frac{3\pi}{4}&\text{if }\sigma_{k}=\frac{\pi}{4}\end{cases}\quad\text{ for }k\in\mathbb{Z}\setminus\{0,\pm s\},

see also Remark 2 (b). Thus in particular τkσk\tau_{k}\neq\sigma_{k} for k{0,±s}k\in\mathbb{Z}\setminus\{0,\pm s\}, and we conclude from the uniqueness statement in Proposition 2 the nondegeneracy property

(11a) k{0,±s},qX1,q=3μkτk[Γw02q]q0;k\in\mathbb{Z}\setminus\{0,\pm s\},\quad q\in X_{1},\quad q=3\>\mathcal{R}_{\mu_{k}}^{\tau_{k}}[\Gamma w_{0}^{2}\,q]\qquad\Rightarrow\qquad q\equiv 0;
for the 0-th mode, using the resolvent 𝒫μ0=(Δ+μ0)1:X3X1\mathcal{P}_{\mu_{0}}=(-\Delta+\mu_{0})^{-1}:X_{3}\to X_{1} (see Lemma 7), the corresponding property is assumed in (5):
(11b) qX1,q=3𝒫μ0[Γw02q]q0.q\in X_{1},\quad q=3\>\mathcal{P}_{\mu_{0}}[\Gamma w_{0}^{2}\,q]\qquad\Rightarrow\qquad q\equiv 0.

We now introduce a map the zeros of which provide solutions of the system (8). Throughout, we use the shorthand notation 𝐮=𝐯+𝐰\mathbf{u}=\mathbf{v}+\mathbf{w} for 𝐯𝒳1\mathbf{v}\in\mathcal{X}_{1} and the stationary solution 𝐰=(,0,w0,0,)\mathbf{w}=(...,0,w_{0},0,...). As above, we have to distinguish the cases τs(0,π)\tau_{s}\in(0,\pi) and τs=0\tau_{s}=0. (In the following, please recall that we consider some fixed s0s\neq 0.) For 0<τ±s<π0<\tau_{\pm s}<\pi, we introduce F:𝒳1×𝒳1F:\>\mathcal{X}_{1}\times\mathbb{R}\to\mathcal{X}_{1} via

(12a) F(𝐯,λ)k:=vk{𝒫μ0[Γ(𝐮𝐮𝐮)0Γw03]k=0,μsπ/2[Γ(𝐮𝐮𝐮)±s]+(cot(τ±s)λ)Ψ~μs[Γ(𝐮𝐮𝐮)±s]k=±s,μkτk[Γ(𝐮𝐮𝐮)k]else.F(\mathbf{v},\lambda)_{k}:=v_{k}-\begin{cases}\mathcal{P}_{\mu_{0}}\left[\Gamma\>\left(\mathbf{u}\star\mathbf{u}\star\mathbf{u}\right)_{0}-\Gamma\>w_{0}^{3}\right]&k=0,\\ \mathcal{R}_{\mu_{s}}^{\pi/2}\left[\Gamma\>\left(\mathbf{u}\star\mathbf{u}\star\mathbf{u}\right)_{\pm s}\right]\\ \qquad+(\cot(\tau_{\pm s})-\lambda)\tilde{\Psi}_{\mu_{s}}\ast\left[\Gamma\>\left(\mathbf{u}\star\mathbf{u}\star\mathbf{u}\right)_{\pm s}\right]&k=\pm s,\\ \mathcal{R}_{\mu_{k}}^{\tau_{k}}\left[\Gamma\>\left(\mathbf{u}\star\mathbf{u}\star\mathbf{u}\right)_{k}\right]&\text{else}.\end{cases}
Similarly, if σs=0\sigma_{s}=0, we define G:𝒳1×𝒳1G:\>\mathcal{X}_{1}\times\mathbb{R}\to\mathcal{X}_{1} by
(12b) G(𝐯,λ)k:=vk{𝒫μ0[Γ(𝐮𝐮𝐮)0Γw03]k=0,μsπ/2[Γ(𝐮𝐮𝐮)±s]+(1λ)(α(μs)(v±s)+β(μs)(v±s))Ψ~μsk=±s,μkτk[Γ(𝐮𝐮𝐮)k]else.G(\mathbf{v},\lambda)_{k}:=v_{k}-\begin{cases}\mathcal{P}_{\mu_{0}}\left[\Gamma\>\left(\mathbf{u}\star\mathbf{u}\star\mathbf{u}\right)_{0}-\Gamma\>w_{0}^{3}\right]&k=0,\\ \mathcal{R}_{\mu_{s}}^{\pi/2}\left[\Gamma\>\left(\mathbf{u}\star\mathbf{u}\star\mathbf{u}\right)_{\pm s}\right]\\ \qquad+(1-\lambda)\left(\alpha^{(\mu_{s})}(v_{\pm s})+\beta^{(\mu_{s})}(v_{\pm s})\right)\tilde{\Psi}_{\mu_{s}}&k=\pm s,\\ \mathcal{R}_{\mu_{k}}^{\tau_{k}}\left[\Gamma\>\left(\mathbf{u}\star\mathbf{u}\star\mathbf{u}\right)_{k}\right]&\text{else}.\end{cases}

The following result collects some basic properties of the maps FF and GG and the polychromatic states related to their zeros.

Proposition 3.

Let ss\in\mathbb{N} and (τk)k(\tau_{k})_{k\in\mathbb{Z}} be chosen as in (10). The maps F,G:𝒳1×𝒳1F,G:\mathcal{X}_{1}\times\mathbb{R}\to\mathcal{X}_{1} are well-defined and smooth with F(𝟎,λ)=G(𝟎,λ)=𝟎F(\mathbf{0},\lambda)=G(\mathbf{0},\lambda)=\mathbf{0} for all λ\lambda\in\mathbb{R}. Further, if F(𝐯,λ)=𝟎F(\mathbf{v},\lambda)=\mathbf{0} resp. G(𝐯,λ)=𝟎G(\mathbf{v},\lambda)=\mathbf{0} for some 𝐯𝒳1,λ\mathbf{v}\in\mathcal{X}_{1},\lambda\in\mathbb{R}, then 𝐯\mathbf{v} solves the stationary system (8) and

U(t,x):=w0(x)+v0(x)+k=12cos(ωkt)vk(x)(t,x3)\displaystyle U(t,x):=w_{0}(x)+v_{0}(x)+\sum_{k=1}^{\infty}2\,\cos(\omega kt)v_{k}(x)\qquad(t\in\mathbb{R},x\in\mathbb{R}^{3})

defines a twice continuously differentiable, classical solution UC2(,X1)U\in C^{2}(\mathbb{R},X_{1}) of the Klein-Gordon equation (1).

Again, the proof can be found in Section 3. We will even show that UC(,X1)U\in C^{\infty}(\mathbb{R},X_{1}). For the derivatives of FF resp. GG with respect to the Banach space component 𝐯𝒳1\mathbf{v}\in\mathcal{X}_{1}, we will verify the following explicit formulas: Letting 𝐪𝒳1\mathbf{q}\in\mathcal{X}_{1} and abbreviating 𝐮:=𝐯+𝐰\mathbf{u}:=\mathbf{v}+\mathbf{w},

(13a) (DF(𝐯,λ)[𝐪])k=qk{3𝒫μ0[Γ(𝐪𝐮𝐮)0]k=0,3μsπ/2[Γ(𝐪𝐮𝐮)±s]+3(cot(τs)λ)Ψ~μs[Γ(𝐪𝐮𝐮)±s]k=±s,3μkτk[Γ(𝐪𝐮𝐮)k]else;\displaystyle(DF(\mathbf{v},\lambda)[\mathbf{q}])_{k}=q_{k}-\begin{cases}3\>\mathcal{P}_{\mu_{0}}[\Gamma\left(\mathbf{q}\star\mathbf{u}\star\mathbf{u}\right)_{0}]&k=0,\\ 3\>\mathcal{R}_{\mu_{s}}^{\pi/2}\left[\Gamma\left(\mathbf{q}\star\mathbf{u}\star\mathbf{u}\right)_{\pm s}\right]&\\ \>+3\>(\cot(\tau_{s})-\lambda)\tilde{\Psi}_{\mu_{s}}\ast\left[\Gamma\left(\mathbf{q}\star\mathbf{u}\star\mathbf{u}\right)_{\pm s}\right]&k=\pm s,\\ 3\>\mathcal{R}_{\mu_{k}}^{\tau_{k}}\left[\Gamma\left(\mathbf{q}\star\mathbf{u}\star\mathbf{u}\right)_{k}\right]&\text{else};\end{cases}
(13b) (DG(𝐯,λ)[𝐪])k=qk{3𝒫μ0[Γ(𝐪𝐮𝐮)0]k=0,3μsπ/2[Γ(𝐪𝐮𝐮)±s]+(1λ)(α(μs)(q±s)+β(μs)(q±s))Ψ~μsk=±s,3μkτk[Γ(𝐪𝐮𝐮)k]else.\displaystyle(DG(\mathbf{v},\lambda)[\mathbf{q}])_{k}=q_{k}-\begin{cases}3\>\mathcal{P}_{\mu_{0}}[\Gamma\left(\mathbf{q}\star\mathbf{u}\star\mathbf{u}\right)_{0}]&k=0,\\ 3\>\mathcal{R}_{\mu_{s}}^{\pi/2}\left[\Gamma\left(\mathbf{q}\star\mathbf{u}\star\mathbf{u}\right)_{\pm s}\right]&\\ \>+(1-\lambda)\!\left(\alpha^{(\mu_{s})}(q_{\pm s})+\beta^{(\mu_{s})}(q_{\pm s})\right)\!\tilde{\Psi}_{\mu_{s}}&k=\pm s,\\ 3\>\mathcal{R}_{\mu_{k}}^{\tau_{k}}\left[\Gamma\left(\mathbf{q}\star\mathbf{u}\star\mathbf{u}\right)_{k}\right]&\text{else}.\end{cases}
Remark 2.
  • (a)

    As earlier announced, we now see that the bifurcation parameter λ\lambda appears only in the asymptotic expansions of the ss-th components v±sv_{\pm s} of the solutions and not in the differential equation (1). This is different from [19] where the bifurcation parameter takes the role of a coupling parameter of the Helmholtz system.

  • (b)

    The choice of the parameters τk\tau_{k} in equation (10) is far from unique. Indeed, one could instead consider any configuration satisfying

    τk=τkσk for all k{±s},{τk|k{±s}}(δ,πδ)\displaystyle\tau_{k}=\tau_{-k}\neq\sigma_{k}\>\text{ for all }k\in\mathbb{Z}\setminus\{\pm s\},\qquad\{\tau_{k}\,|\,k\in\mathbb{Z}\setminus\{\pm s\}\}\subseteq(\delta,\pi-\delta)

    for some δ(0,π2)\delta\in\left(0,\frac{\pi}{2}\right). The former condition is required for the nondegeneracy statement (11a), and the latter will be used to obtain uniform decay estimates in the proof of Proposition 3, see Lemma 3.
    However, as in [19], the question whether another choice of τk\tau_{k} leads to different bifurcating families is still open. Hence we discuss only the explicit choice in (10).

In the so-established framework, we intend to apply the Crandall-Rabinowitz Bifurcation Theorem. The next result shows that its assumptions are satisfied.

Proposition 4 (Simplicity and transversality).

Let ss\in\mathbb{N} and (τk)k(\tau_{k})_{k\in\mathbb{Z}} be chosen as in (10). The linear operator DF(𝟎,0):𝒳1𝒳1DF(\mathbf{0},0):\mathcal{X}_{1}\to\mathcal{X}_{1} is 1-1-Fredholm with a kernel of the form

kerDF(𝟎,0)=span{𝐪}where qk0 if and only if k=±s.\displaystyle\ker DF(\mathbf{0},0)=\mathrm{span}\,\{\mathbf{q}\}\qquad\text{where }q_{k}\neq 0\text{ if and only if }k=\pm s.

Moreover, the transversality condition is satisfied, that is,

λDF(𝟎,0)[𝐪]ranDF(𝟎,0).\displaystyle\partial_{\lambda}DF(\mathbf{0},0)[\mathbf{q}]\not\in\mathrm{ran}\,DF(\mathbf{0},0).

A corresponding statement holds true for DG(𝟎,0):𝒳1𝒳1DG(\mathbf{0},0):\mathcal{X}_{1}\to\mathcal{X}_{1}.

2.2. The Proof of Theorem 1

Let us fix some ss\in\mathbb{N}, and choose (τk)k(\tau_{k})_{k\in\mathbb{Z}} as in (10). We introduce the trivial family 𝒯:={(𝟎,λ)𝒳1×|λ}\mathcal{T}:=\{(\mathbf{0},\lambda)\in\mathcal{X}_{1}\times\mathbb{R}\>|\>\lambda\in\mathbb{R}\}.

Step 1: Proof of (i).

By Proposition 3, the maps FF resp. GG are smooth and vanish on the trivial family 𝒯\mathcal{T}. In view of Proposition 4, the Crandall-Rabinowitz Theorem shows that (𝟎,0)𝒯(\mathbf{0},0)\in\mathcal{T} is a bifurcation point for F(𝐯,λ)=0F(\mathbf{v},\lambda)=0 resp. G(𝐯,λ)=0G(\mathbf{v},\lambda)=0 and provides an open interval JsJ_{s}\subseteq\mathbb{R} containing 0 and a smooth curve

Js𝒳1×,α(𝐯α,λα)=((vkα)k,λα)\displaystyle J_{s}\to\mathcal{X}_{1}\times\mathbb{R},\qquad\alpha\mapsto(\mathbf{v}^{\alpha},\lambda^{\alpha})=\left((v_{k}^{\alpha})_{k\in\mathbb{Z}},\lambda^{\alpha}\right)

of zeros of FF resp. GG (we do not denote its dependence on ss) with 𝐯0=𝟎,λ0=0\mathbf{v}^{0}=\mathbf{0},\lambda^{0}=0 as well as ddα|α=0𝐯α=𝐪\frac{\mathrm{d}}{\mathrm{d}\alpha}\big{|}_{\alpha=0}\mathbf{v}^{\alpha}=\mathbf{q} where 𝐪\mathbf{q} is a nontrivial element of the kernel of DF(𝟎,0)DF(\mathbf{0},0) resp. DG(𝟎,0)DG(\mathbf{0},0). We let 𝐮α:=𝐯α+𝐰\mathbf{u}^{\alpha}:=\mathbf{v}^{\alpha}+\mathbf{w} and define polychromatic states UαU^{\alpha} as in (i). Then UαU^{\alpha} is a classical solution of the cubic Klein-Gordon equation (1) due to Proposition 3 since F(𝐯α,λα)=0F(\mathbf{v}^{\alpha},\lambda^{\alpha})=0 resp. G(𝐯α,λα)=0G(\mathbf{v}^{\alpha},\lambda^{\alpha})=0. By their very definition, the solutions UαU^{\alpha} are time-periodic with period 2π/ω2\pi/\omega (maybe less). This proves (i).

Step 2: Proof of (ii).

Since FF resp. GG are smooth, so is the map Js𝒳1×,α(𝐯α,λα)J_{s}\to\mathcal{X}_{1}\times\mathbb{R},\>\alpha\mapsto(\mathbf{v}^{\alpha},\lambda^{\alpha}). By Proposition 4, qk0q_{k}\neq 0 if and only if k=±sk=\pm s, which implies that only the ±s\pm s-th components of

ddα|α=0𝐮α=ddα|α=0𝐯α=𝐪\displaystyle\frac{\mathrm{d}}{\mathrm{d}\alpha}\bigg{|}_{\alpha=0}\mathbf{u}^{\alpha}=\frac{\mathrm{d}}{\mathrm{d}\alpha}\bigg{|}_{\alpha=0}\mathbf{v}^{\alpha}=\mathbf{q}

do not vanish. For sufficiently small nonzero values of α\alpha, the solutions UαU^{\alpha} are thus nonstationary. In particular, the direction of bifurcation changes when changing the value of ss, and the associated bifurcating curves are, at least locally, mutually different.

Step 3: Proof of (iii).

We show finally that, under the additional assumption that Γ(x)0\Gamma(x)\neq 0 for almost all x3x\in\mathbb{R}^{3}, every non-stationary solution

Uα(t,x)=w0(x)+v0α(x)+k=12cos(ωkt)vkα(x)\displaystyle U^{\alpha}(t,x)=w_{0}(x)+v_{0}^{\alpha}(x)+\sum_{k=1}^{\infty}2\cos(\omega kt)\>v_{k}^{\alpha}(x)

in fact possesses infinitely many nontrivial coefficients vkαv_{k}^{\alpha}. Indeed, assuming the contrary, we can choose a maximal r>0r>0 (since UαU^{\alpha} is non-stationary) with vrα0v_{r}^{\alpha}\not\equiv 0 or equivalently urα=vrα+wr=vrα0u_{r}^{\alpha}=v_{r}^{\alpha}+w_{r}=v_{r}^{\alpha}\not\equiv 0. But then,

v3rα=l+m+n=3rμ3rτ3r[Γulαumαunα]=μ3rτ3r[Γ(vrα)3]0\displaystyle v_{3r}^{\alpha}=\sum_{l+m+n=3r}\mathcal{R}_{\mu_{3r}}^{\tau_{3r}}[\Gamma\>u_{l}^{\alpha}\,u_{m}^{\alpha}\,u_{n}^{\alpha}]=\mathcal{R}_{\mu_{3r}}^{\tau_{3r}}[\Gamma\>(v_{r}^{\alpha})^{3}]\not\equiv 0

since the convolution identity implies Δv3rαμ3rv3rα=Γ(vrα)3-\Delta v_{3r}^{\alpha}-\mu_{3r}v_{3r}^{\alpha}=\Gamma\>(v_{r}^{\alpha})^{3}, and Γ(vrα)30\Gamma\>(v_{r}^{\alpha})^{3}\not\equiv 0 since Γ(x)0\Gamma(x)\neq 0 almost everywhere by assumption. This contradicts the maximality of rr. \square

2.3. The Proof of Remark 1 (c)

Finally, as announced in Remark 1 (c), we verify the nondegeneracy assumption (11b) resp. (5) for constant positive Γ\Gamma.

Lemma 1 (Nondegeneracy, à la Bates and Shi [6]).

Let ΓΓ0\Gamma\equiv\Gamma_{0} for some Γ0>0\Gamma_{0}>0, and assume that w0Crad2(3)w_{0}\in C^{2}_{\mathrm{rad}}(\mathbb{R}^{3}) is a radially symmetric solution of (4) the profile of which satisfies w0(r)>0w_{0}(r)>0, w0(r)<0w_{0}^{\prime}(r)<0 for all r>0r>0, and both w0(r)w_{0}(r) and w0(r)w_{0}^{\prime}(r) decay exponentially as rr\to\infty. Then the nondegeneracy property (5) holds, i.e. for any radial, twice differentiable q0X1q_{0}\in X_{1}

Δq0+q0=3Γ0w02q0 on 3implies q00.\displaystyle-\Delta q_{0}+q_{0}=3\Gamma_{0}\>w_{0}^{2}\>q_{0}\text{ on }\mathbb{R}^{3}\qquad\text{implies }\qquad q_{0}\equiv 0.

This can be proved closely following the line of argumentation by Bates and Shi [6], Theorem 5.4 (6). The main difference is that they state the nondegeneracy result as a spectral property of the operator Δ+m2+3Γ0w02:H2(3)L2(3)-\Delta+m^{2}+3\Gamma_{0}w_{0}^{2}:\>H^{2}(\mathbb{R}^{3})\to L^{2}(\mathbb{R}^{3}) whereas we cannot use the Hilbert space setting but discuss solutions in X1X_{1}. However, the technique of Bates and Shi (and also of Wei’s proof in [25]) is based on an expansion at a fixed radius r>0r>0 in terms of the eigenfunctions of the Laplace-Beltrami operator on L2(𝕊2)L^{2}(\mathbb{S}^{2}). This provides coefficients depending on rr, and the conclusions are obtained from the analysis of these profiles on an ODE level using results due to Kwong and Zhang [17]. These ideas apply in the topology of X1X_{1} in the very same way; for details, cf. [21], (proof of) Lemma 4.11.

3. Proofs of the Auxiliary Results

Proof of Proposition 1.
Let 𝐮(j)=(uk(j))k𝒳1\mathbf{u}^{(j)}=(u^{(j)}_{k})_{k\in\mathbb{Z}}\in\mathcal{X}_{1} for j=1,2,3j=1,2,3. We find the following chain of inequalities

𝐮(1)𝐮(2)𝐮(3)𝒳3=k(𝐮(1)𝐮(2)𝐮(3))kX3\displaystyle\left\lVert\mathbf{u}^{(1)}\star\mathbf{u}^{(2)}\star\mathbf{u}^{(3)}\right\rVert_{\mathcal{X}_{3}}=\sum_{k\in\mathbb{Z}}\left\lVert(\mathbf{u}^{(1)}\star\mathbf{u}^{(2)}\star\mathbf{u}^{(3)})_{k}\right\rVert_{X_{3}}
kl,m,nl+m+n=kul(1)um(2)un(3)X3\displaystyle\quad\leq\sum_{k\in\mathbb{Z}}\sum_{\begin{subarray}{c}l,m,n\in\mathbb{Z}\\ l+m+n=k\end{subarray}}\left\lVert u_{l}^{(1)}\,u_{m}^{(2)}\,u_{n}^{(3)}\right\rVert_{X_{3}}
kl,m,nl+m+n=kul(1)X1um(2)X1un(3)X1\displaystyle\quad\leq\sum_{k\in\mathbb{Z}}\sum_{\begin{subarray}{c}l,m,n\in\mathbb{Z}\\ l+m+n=k\end{subarray}}\left\lVert u_{l}^{(1)}\right\rVert_{X_{1}}\,\left\lVert u_{m}^{(2)}\right\rVert_{X_{1}}\,\left\lVert u_{n}^{(3)}\right\rVert_{X_{1}}
=(ul(1)X1)l(um(2)X1)m(un(3)X1)n1()\displaystyle\quad=\left\lVert\left(\left\lVert u_{l}^{(1)}\right\rVert_{X_{1}}\right)_{l\in\mathbb{Z}}\star\left(\left\lVert u_{m}^{(2)}\right\rVert_{X_{1}}\right)_{m\in\mathbb{Z}}\star\left(\left\lVert u_{n}^{(3)}\right\rVert_{X_{1}}\right)_{n\in\mathbb{Z}}\right\rVert_{\ell^{1}(\mathbb{Z})}
(ul(1)X1)l1()(um(2)X1)m1()(un(3)X1)n1()\displaystyle\quad\leq\left\lVert\left(\left\lVert u_{l}^{(1)}\right\rVert_{X_{1}}\right)_{l\in\mathbb{Z}}\right\rVert_{\ell^{1}(\mathbb{Z})}\left\lVert\left(\left\lVert u_{m}^{(2)}\right\rVert_{X_{1}}\right)_{m\in\mathbb{Z}}\right\rVert_{\ell^{1}(\mathbb{Z})}\left\lVert\left(\left\lVert u_{n}^{(3)}\right\rVert_{X_{1}}\right)_{n\in\mathbb{Z}}\right\rVert_{\ell^{1}(\mathbb{Z})}
=𝐮(1)𝒳1𝐮(2)𝒳1𝐮(3)𝒳1,\displaystyle\quad=\left\lVert\mathbf{u}^{(1)}\right\rVert_{\mathcal{X}_{1}}\left\lVert\mathbf{u}^{(2)}\right\rVert_{\mathcal{X}_{1}}\left\lVert\mathbf{u}^{(3)}\right\rVert_{\mathcal{X}_{1}},

where finally Young’s inequality for convolutions in 1()\ell^{1}(\mathbb{Z}) has been applied. Since the latter term is finite, we infer 𝐮(1)𝐮(2)𝐮(3)𝒳3\mathbf{u}^{(1)}\star\mathbf{u}^{(2)}\star\mathbf{u}^{(3)}\in\mathcal{X}_{3}. ∎

Proof of Proposition 3.
Step 1: Decay estimates

The proof of Proposition 3 requires convergence properties in order to handle the infinite series in the definition of U(t,x)U(t,x), which we first provide in the following two lemmas.

Lemma 2.

The convolution operators μτ:X3X1\mathcal{R}_{\mu}^{\tau}:X_{3}\to X_{1} satisfy for τ(0,π)\tau\in(0,\pi) and μ>0\mu>0

fX3μτ[f]X1Csin(τ)(1+1μ)fX3,μτ[f]L4(3)Cμ4sin(τ)fL43(3).\qquad\forall\>f\in X_{3}\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\begin{split}&\left\lVert\mathcal{R}_{\mu}^{\tau}[f]\right\rVert_{X_{1}}\leq\frac{C}{\sin(\tau)}\>\left(1+\frac{1}{\sqrt{\mu}}\right)\>\cdot\left\lVert f\right\rVert_{X_{3}},\\ &\left\lVert\mathcal{R}_{\mu}^{\tau}[f]\right\rVert_{L^{4}(\mathbb{R}^{3})}\leq\frac{C}{\sqrt[4]{\mu}\>\sin(\tau)}\cdot\left\lVert f\right\rVert_{L^{\frac{4}{3}}(\mathbb{R}^{3})}.\end{split}

The fact that a power of μ\mu appears in the denominator is crucial since it will finally provide the convergence and regularity of the polychromatic sums where μ=μk=ω2k2m2\mu=\mu_{k}=\omega^{2}k^{2}-m^{2} for kk\in\mathbb{Z}.
The proof of Lemma 2 relies, via rescaling, on the respective estimates for μ=1\mu=1. These can be found in [19], pp. 1038–1039 for the X3X_{3}-X1X_{1} estimate and in [10], Theorem 2.1 for the L4/3L^{4/3}-L4L^{4} estimate.

Lemma 3.

Let ΓLrad(3)Cloc1(3)\Gamma\in L^{\infty}_{\mathrm{rad}}(\mathbb{R}^{3})\cap C^{1}_{\mathrm{loc}}(\mathbb{R}^{3}) and assume 𝐮=(uk)k𝒳1\mathbf{u}=(u_{k})_{k\in\mathbb{Z}}\in\mathcal{X}_{1} is a sequence of Cloc2C^{2}_{\mathrm{loc}} functions which satisfy the following system of convolution equations:

uk=μkτk[Γ(𝐮𝐮𝐮)k]for all k with |k|>s\displaystyle u_{k}=\mathcal{R}_{\mu_{k}}^{\tau_{k}}[\Gamma\>(\mathbf{u}\star\mathbf{u}\star\mathbf{u})_{k}]\qquad\qquad\qquad\text{for all }k\in\mathbb{Z}\text{ with }|k|>s

where μk=ω2k2m2\mu_{k}=\omega^{2}k^{2}-m^{2} and τk(δ,πδ)\tau_{k}\in(\delta,\pi-\delta) for some ω>m,δ(0,π2)\omega>m,\delta\in\left(0,\frac{\pi}{2}\right). Then there holds:

  • (i)

    For every α0\alpha\geq 0, there exists a constant Cα0C_{\alpha}\geq 0 with

    ukL4(3)+Γ(𝐮𝐮𝐮)kL4(3)Cα(k2+1)α2(k).\displaystyle\left\lVert u_{k}\right\rVert_{L^{4}(\mathbb{R}^{3})}+\left\lVert\Gamma\>(\mathbf{u}\star\mathbf{u}\star\mathbf{u})_{k}\right\rVert_{L^{4}(\mathbb{R}^{3})}\leq C_{\alpha}\cdot(k^{2}+1)^{-\frac{\alpha}{2}}\qquad(k\in\mathbb{Z}).
  • (ii)

    For every ball B=BR(0)3B=B_{R}(0)\subseteq\mathbb{R}^{3} and α0\alpha\geq 0 there exists a constant Dα(B)0D_{\alpha}(B)\geq 0 with

    |uk(x)|+|uk(x)|+|D2uk(x)|Dα(B)(k2+1)α2(k,xB).\displaystyle|u_{k}(x)|+|\nabla u_{k}(x)|+|D^{2}u_{k}(x)|\leq D_{\alpha}(B)\cdot(k^{2}+1)^{-\frac{\alpha}{2}}\qquad(k\in\mathbb{Z},x\in B).
  • (iii)

    For every α0\alpha\geq 0, there exists a constant Eα0E_{\alpha}\geq 0 with

    ukX1Eα(k2+1)α2(k).\displaystyle\left\lVert u_{k}\right\rVert_{X_{1}}\leq E_{\alpha}\cdot(k^{2}+1)^{-\frac{\alpha}{2}}\qquad(k\in\mathbb{Z}).

The proof of Lemma 3 can also be found in detail in the author’s PhD thesis [21], Lemma 4.13. We present here the most important step and summarize the remainder briefly, since it is mainly based on the application of (standard) elliptic regularity estimates.

As ukX1Cloc2(3)u_{k}\in X_{1}\cap C^{2}_{\text{loc}}(\mathbb{R}^{3}) for all kk\in\mathbb{Z} by assumption, it is straightforward to find constants as in the lemma for a finite number of elements us,,usu_{-s},...,u_{s}. Hence it is sufficient to study those kk\in\mathbb{Z} with |k|>s|k|>s; for these, we have μk=k2ω2m2cs(k2+1)\mu_{k}=k^{2}\omega^{2}-m^{2}\geq c_{s}(k^{2}+1) for some positive cs>0c_{s}>0 depending on the parameters ω\omega and mm. The decay estimates of arbitrary order in kk we aim to prove essentially go back to the L4/3L^{4/3}-L4L^{4} scaling property stated in Lemma 2 above. Indeed, due to δ<τk<πδ\delta<\tau_{k}<\pi-\delta, it provides C1=C1(Γ,δ,ω,m,s)0C_{1}=C_{1}(\left\lVert\Gamma\right\rVert_{\infty},\delta,\omega,m,s)\geq 0 with

(14) ukL4(3)C1(k2+1)14(𝐮𝐮𝐮)kL43(3)for all k.\displaystyle\left\lVert u_{k}\right\rVert_{L^{4}(\mathbb{R}^{3})}\leq\frac{C_{1}}{(k^{2}+1)^{\frac{1}{4}}}\>\left\lVert(\mathbf{u}\star\mathbf{u}\star\mathbf{u})_{k}\right\rVert_{L^{\frac{4}{3}}(\mathbb{R}^{3})}\qquad\text{for all }k\in\mathbb{Z}.

With that, assuming k(k2+1)α2ukL4(3)<\sum_{k\in\mathbb{Z}}(k^{2}+1)^{\frac{\alpha}{2}}\left\lVert u_{k}\right\rVert_{L^{4}(\mathbb{R}^{3})}<\infty for some α0\alpha\geq 0 (which is trivially satisfied for α=0\alpha=0 since 𝐮𝒳1\mathbf{u}\in\mathcal{X}_{1}), one can iterate as follows

k(k2+1)α+1/22ukL4(3)(14)C1k(k2+1)α2(𝐮𝐮𝐮)kL43(3)\displaystyle\sum_{k\in\mathbb{Z}}(k^{2}+1)^{\frac{\alpha+1/2}{2}}\>\left\lVert u_{k}\right\rVert_{L^{4}(\mathbb{R}^{3})}\overset{\eqref{eq_proof-scale-k}}{\leq}C_{1}\>\sum_{k\in\mathbb{Z}}(k^{2}+1)^{\frac{\alpha}{2}}\>\left\lVert(\mathbf{u}\star\mathbf{u}\star\mathbf{u})_{k}\right\rVert_{L^{\frac{4}{3}}(\mathbb{R}^{3})}
C1kl+m+n=k((l+m+n)2+1)α2ulL4(3)umL4(3)unL4(3)\displaystyle\quad\leq C_{1}\>\sum_{k\in\mathbb{Z}}\sum_{l+m+n=k}((l+m+n)^{2}+1)^{\frac{\alpha}{2}}\left\lVert u_{l}\right\rVert_{L^{4}(\mathbb{R}^{3})}\,\left\lVert u_{m}\right\rVert_{L^{4}(\mathbb{R}^{3})}\,\left\lVert u_{n}\right\rVert_{L^{4}(\mathbb{R}^{3})}
2αC1kl+m+n=k[(l2+1)α2ulL4(3)(m2+1)α2umL4(3)(n2+1)α2unL4(3)]\displaystyle\quad\leq 2^{\alpha}\>C_{1}\>\sum_{k\in\mathbb{Z}}\sum_{l+m+n=k}\left[(l^{2}+1)^{\frac{\alpha}{2}}\left\lVert u_{l}\right\rVert_{L^{4}(\mathbb{R}^{3})}\,(m^{2}+1)^{\frac{\alpha}{2}}\left\lVert u_{m}\right\rVert_{L^{4}(\mathbb{R}^{3})}\,(n^{2}+1)^{\frac{\alpha}{2}}\left\lVert u_{n}\right\rVert_{L^{4}(\mathbb{R}^{3})}\right]
=2αC1(k(k2+1)α2ukL4(3))3\displaystyle\quad=2^{\alpha}\>C_{1}\>\left(\sum_{k\in\mathbb{Z}}(k^{2}+1)^{\frac{\alpha}{2}}\left\lVert u_{k}\right\rVert_{L^{4}(\mathbb{R}^{3})}\right)^{3}
<.\displaystyle\quad<\infty.

This shows the first part of the estimate in (i), and the second part follows by combining the former with the interpolation estimate

ulumunL4(3)\displaystyle\left\lVert u_{l}u_{m}u_{n}\right\rVert_{L^{4}(\mathbb{R}^{3})} ulL12(3)umL12(3)unL12(3)\displaystyle\leq\left\lVert u_{l}\right\rVert_{L^{12}(\mathbb{R}^{3})}\left\lVert u_{m}\right\rVert_{L^{12}(\mathbb{R}^{3})}\left\lVert u_{n}\right\rVert_{L^{12}(\mathbb{R}^{3})}
[ulL4(3)umL4(3)unL4(3)]13[ulumun]23\displaystyle\leq\left[\left\lVert u_{l}\right\rVert_{L^{4}(\mathbb{R}^{3})}\left\lVert u_{m}\right\rVert_{L^{4}(\mathbb{R}^{3})}\left\lVert u_{n}\right\rVert_{L^{4}(\mathbb{R}^{3})}\right]^{\frac{1}{3}}\>\Big{[}\left\lVert u_{l}\right\rVert_{\infty}\left\lVert u_{m}\right\rVert_{\infty}\left\lVert u_{n}\right\rVert_{\infty}\Big{]}^{\frac{2}{3}}
[ulL4(3)umL4(3)unL4(3)]13𝐮𝒳12.\displaystyle\leq\left[\left\lVert u_{l}\right\rVert_{L^{4}(\mathbb{R}^{3})}\left\lVert u_{m}\right\rVert_{L^{4}(\mathbb{R}^{3})}\left\lVert u_{n}\right\rVert_{L^{4}(\mathbb{R}^{3})}\right]^{\frac{1}{3}}\>\left\lVert\mathbf{u}\right\rVert^{2}_{\mathcal{X}_{1}}.

The local estimate in (ii) can be derived from the global L4L^{4} bounds in (i) using elliptic regularity, which first provides estimates in Wloc2,4(3)W^{2,4}_{\text{loc}}(\mathbb{R}^{3}) and then is suitable Hölder spaces. The estimate (iii) in the X1X_{1} norm essentially uses the explicit representations (given fX3f\in X_{3})

μτ[f](x)=3sin(|xy|μk+τk)4π|xy|sin(τk)f(y)dy\displaystyle\mathcal{R}_{\mu}^{\tau}[f](x)=\int_{\mathbb{R}^{3}}\frac{\sin(|x-y|\sqrt{\mu_{k}}+\tau_{k})}{4\pi|x-y|\sin(\tau_{k})}\cdot f(y)\>\mathrm{d}y
=sin(|x|μk+τk)|x|sin(τk)0|x|sin(rμk)rμkf(r)r2dr+sin(|x|μk)|x|sin(τk)|x|sin(rμk+τk)rμkf(r)r2dr.\displaystyle=\frac{\sin(|x|\sqrt{\mu_{k}}+\tau_{k})}{|x|\sin(\tau_{k})}\int_{0}^{|x|}\frac{\sin(r\sqrt{\mu_{k}})}{r\sqrt{\mu_{k}}}f(r)\>r^{2}\>\mathrm{d}r+\frac{\sin(|x|\sqrt{\mu_{k}})}{|x|\sin(\tau_{k})}\int_{|x|}^{\infty}\frac{\sin(r\sqrt{\mu_{k}}+\tau_{k})}{r\sqrt{\mu_{k}}}f(r)\>r^{2}\>\mathrm{d}r.

Starting here, Hölder’s inequality and (i) yield (iii); again, for details, cf. [21].

Step 2: Mapping properties of FF resp. GG.

For λ\lambda\in\mathbb{R} and 𝐯𝒳1\mathbf{v}\in\mathcal{X}_{1}, we set 𝐮:=𝐰+𝐯\mathbf{u}:=\mathbf{w}+\mathbf{v} and recall the defining equations (12a) and (12b):

F(𝐯,λ)k:=vk{𝒫μ0[Γ(𝐮𝐮𝐮)0Γw03]k=0,μsπ/2[Γ(𝐮𝐮𝐮)±s]+(cot(τ±s)λ)Ψ~μs[Γ(𝐮𝐮𝐮)±s]k=±s,μkτk[Γ(𝐮𝐮𝐮)k]else;\displaystyle F(\mathbf{v},\lambda)_{k}:=v_{k}-\begin{cases}\mathcal{P}_{\mu_{0}}\left[\Gamma\>\left(\mathbf{u}\star\mathbf{u}\star\mathbf{u}\right)_{0}-\Gamma\>w_{0}^{3}\right]&k=0,\\ \mathcal{R}_{\mu_{s}}^{\pi/2}\left[\Gamma\>\left(\mathbf{u}\star\mathbf{u}\star\mathbf{u}\right)_{\pm s}\right]\\ \qquad+(\cot(\tau_{\pm s})-\lambda)\tilde{\Psi}_{\mu_{s}}\ast\left[\Gamma\>\left(\mathbf{u}\star\mathbf{u}\star\mathbf{u}\right)_{\pm s}\right]&k=\pm s,\\ \mathcal{R}_{\mu_{k}}^{\tau_{k}}\left[\Gamma\>\left(\mathbf{u}\star\mathbf{u}\star\mathbf{u}\right)_{k}\right]&\text{else};\end{cases}
G(𝐯,λ)k:=vk{𝒫μ0[Γ(𝐮𝐮𝐮)0Γw03]k=0,μsπ/2[Γ(𝐮𝐮𝐮)±s]+(1λ)(α(μs)(v±s)+β(μs)(v±s))Ψ~μsk=±s,μkτk[Γ(𝐮𝐮𝐮)k]else.\displaystyle G(\mathbf{v},\lambda)_{k}:=v_{k}-\begin{cases}\mathcal{P}_{\mu_{0}}\left[\Gamma\>\left(\mathbf{u}\star\mathbf{u}\star\mathbf{u}\right)_{0}-\Gamma\>w_{0}^{3}\right]&k=0,\\ \mathcal{R}_{\mu_{s}}^{\pi/2}\left[\Gamma\>\left(\mathbf{u}\star\mathbf{u}\star\mathbf{u}\right)_{\pm s}\right]\\ \qquad+(1-\lambda)\left(\alpha^{(\mu_{s})}(v_{\pm s})+\beta^{(\mu_{s})}(v_{\pm s})\right)\tilde{\Psi}_{\mu_{s}}&k=\pm s,\\ \mathcal{R}_{\mu_{k}}^{\tau_{k}}\left[\Gamma\>\left(\mathbf{u}\star\mathbf{u}\star\mathbf{u}\right)_{k}\right]&\text{else}.\end{cases}

Our main concern will be convergence of the infinite sums related to the space 𝒳1=sym1(,X1)\mathcal{X}_{1}=\ell^{1}_{\mathrm{sym}}(\mathbb{Z},X_{1}). Noticing that FF and GG only differ in the ±s\pm s-th component, and that the scalar parameter λ\lambda only appears as a multiplicative factor, we solely discuss smoothness of the map F(,λ):𝒳1𝒳1F(\,\cdot\,,\lambda):\mathcal{X}_{1}\to\mathcal{X}_{1} with λ\lambda\in\mathbb{R} fixed.

The main tool is the following uniform norm estimate for the operators appearing in the components of FF. Recalling that τk{π4,3π4}\tau_{k}\in\{\frac{\pi}{4},\frac{3\pi}{4}\} for k0,±sk\neq 0,\pm s by (10), Lemma 2 above (for k0,±sk\neq 0,\pm s) as well as the continuity properties stated in Lemmas 4 and 7 (for k=±sk=\pm s and k=0k=0, respectively) provide a constant C0=C0(λ,τs,ω,m)>0C_{0}=C_{0}(\lambda,\tau_{s},\omega,m)>0 with

(15) μkτk(X3,X1)C0(k{±s}),μsπ/2(X3,X1)C02,(cot(τ±s)λ)Ψ~μs(X3,X1)C02,𝒫μ0(X3,X1)C0.\begin{split}&\left\lVert\mathcal{R}_{\mu_{k}}^{\tau_{k}}\right\rVert_{\mathcal{L}(X_{3},X_{1})}\leq C_{0}\quad(k\in\mathbb{Z}\setminus\{\pm s\}),\\ &\left\lVert\mathcal{R}_{\mu_{s}}^{\pi/2}\right\rVert_{\mathcal{L}(X_{3},X_{1})}\leq\frac{C_{0}}{2},\>\>\left\lVert(\cot(\tau_{\pm s})-\lambda)\>\tilde{\Psi}_{\mu_{s}}\ast\right\rVert_{\mathcal{L}(X_{3},X_{1})}\leq\frac{C_{0}}{2},\\ &\left\lVert\mathcal{P}_{\mu_{0}}\right\rVert_{\mathcal{L}(X_{3},X_{1})}\leq C_{0}.\end{split}

We now let 𝐯𝒳1\mathbf{v}\in\mathcal{X}_{1} and define 𝐮=𝐯+𝐰\mathbf{u}=\mathbf{v}+\mathbf{w}. Since Γ\Gamma is assumed to be continuous and bounded, Proposition 1 implies that Γ(𝐮𝐮𝐮)𝒳3\Gamma\>\left(\mathbf{u}\star\mathbf{u}\star\mathbf{u}\right)\in\mathcal{X}_{3}. Thus every component F(𝐯,λ)kF(\mathbf{v},\lambda)_{k} is a well-defined element of X1X_{1}, and we estimate

F(𝐯,λ)𝒳1\displaystyle\left\lVert F(\mathbf{v},\lambda)\right\rVert_{\mathcal{X}_{1}} =kF(𝐯,λ)kX1\displaystyle=\sum_{k\in\mathbb{Z}}\left\lVert F(\mathbf{v},\lambda)_{k}\right\rVert_{X_{1}}
(15)𝐯𝒳1+C0Γw03X3+C0kΓ(𝐮𝐮𝐮)kX3\displaystyle\overset{\eqref{eq_proof-C0}}{\leq}\left\lVert\mathbf{v}\right\rVert_{\mathcal{X}_{1}}+C_{0}\left\lVert\Gamma w_{0}^{3}\right\rVert_{X_{3}}+C_{0}\sum_{k\in\mathbb{Z}}\left\lVert\Gamma\left(\mathbf{u}\star\mathbf{u}\star\mathbf{u}\right)_{k}\right\rVert_{X_{3}}
Prop.1𝐯𝒳1+C0Γw0X13+C0Γ𝐮𝒳13.\displaystyle\!\!\!\overset{\text{Prop.}~{}\ref{prop_convolution-x1}}{\leq}\left\lVert\mathbf{v}\right\rVert_{\mathcal{X}_{1}}+C_{0}\left\lVert\Gamma\right\rVert_{\infty}\left\lVert w_{0}\right\rVert_{X_{1}}^{3}+C_{0}\left\lVert\Gamma\right\rVert_{\infty}\left\lVert\mathbf{u}\right\rVert_{\mathcal{X}_{1}}^{3}.

This is finite, hence F(𝐯,λ)𝒳1F(\mathbf{v},\lambda)\in\mathcal{X}_{1} as asserted. Since F(,λ)F(\,\cdot\,,\lambda) is a combination of continuous linear operators and polynomials in the convolution algebra, essentially the same estimates can be used to show differentiability (to arbitrary order); one thus obtains in particular (13a).

Step 3: Solution properties of uk(x)u_{k}(x).

First of all, recalling that 𝐰=(,0,w0,0,)\mathbf{w}=(...,0,w_{0},0,...) and hence (𝐰𝐰𝐰)k=δk,0w03(\mathbf{w}\star\mathbf{w}\star\mathbf{w})_{k}=\delta_{k,0}\,w_{0}^{3} for kk\in\mathbb{Z}, one can immediately see that F(𝟎,λ)=G(𝟎,λ)=𝟎F(\mathbf{0},\lambda)=G(\mathbf{0},\lambda)=\mathbf{0} for all λ\lambda\in\mathbb{R}. Let us now assume that F(𝐯,λ)=0F(\mathbf{v},\lambda)=0 resp. G(𝐯,λ)=0G(\mathbf{v},\lambda)=0 for some 𝐯𝒳1\mathbf{v}\in\mathcal{X}_{1} and λ\lambda\in\mathbb{R}. Again, we define 𝐮:=𝐯+𝐰\mathbf{u}:=\mathbf{v}+\mathbf{w}, and summarize

u0w0\displaystyle u_{0}-w_{0} =v0=𝒫μ0[Γ(𝐮𝐮𝐮)0Γw03],\displaystyle=v_{0}=\mathcal{P}_{\mu_{0}}\left[\Gamma\left(\mathbf{u}\star\mathbf{u}\star\mathbf{u}\right)_{0}-\Gamma\>w_{0}^{3}\right],
u±s\displaystyle u_{\pm s} =v±s=μsπ/2[Γ(𝐮𝐮𝐮)±s]+{(cot(τs)λ)Ψ~μs[Γ(𝐮𝐮𝐮)±s],(1λ)(α(μs)(v±s)+β(μs)(v±s))Ψ~μs,\displaystyle=v_{\pm s}=\mathcal{R}_{\mu_{s}}^{\pi/2}\left[\Gamma\left(\mathbf{u}\star\mathbf{u}\star\mathbf{u}\right)_{\pm s}\right]+\begin{cases}(\cot(\tau_{s})-\lambda)\tilde{\Psi}_{\mu_{s}}\ast\left[\Gamma\left(\mathbf{u}\star\mathbf{u}\star\mathbf{u}\right)_{\pm s}\right],\\ (1-\lambda)\left(\alpha^{(\mu_{s})}(v_{\pm s})+\beta^{(\mu_{s})}(v_{\pm s})\right)\tilde{\Psi}_{\mu_{s}},\end{cases}
uk\displaystyle u_{k} =vk=μkτk[Γ(𝐮𝐮𝐮)k](k{0,±s}).\displaystyle=v_{k}=\mathcal{R}_{\mu_{k}}^{\tau_{k}}\left[\Gamma\left(\mathbf{u}\star\mathbf{u}\star\mathbf{u}\right)_{k}\right]\qquad(k\in\mathbb{Z}\setminus\{0,\pm s\}).

By choice of τk\tau_{k} in equation (10), we observe in particular that the requirements of Lemma 3 are satisfied with any δ<π4\delta<\frac{\pi}{4}, which we will rely on throughout the subsequent steps. But first, according to Lemmas 4 and 7, vk,ukX1Cloc2(3)v_{k},u_{k}\in X_{1}\cap C^{2}_{\mathrm{loc}}(\mathbb{R}^{3}) satisfy the differential equations

Δvkμkvk=Γ(x)[(𝐮𝐮𝐮)kδk,0w03]on 3\displaystyle-\Delta v_{k}-\mu_{k}v_{k}=\Gamma(x)\>\big{[}(\mathbf{u}\star\mathbf{u}\star\mathbf{u})_{k}-\delta_{k,0}w_{0}^{3}\big{]}\quad\text{on }\mathbb{R}^{3}

or equivalently, in view of 𝐰=(,0,w0,0,)\mathbf{w}=(...,0,w_{0},0,...), of (4) and of μk=ω2k2m2\mu_{k}=\omega^{2}k^{2}-m^{2},

(16) Δuk(ω2k2m2)uk=Γ(x)(𝐮𝐮𝐮)kon 3.\displaystyle-\Delta u_{k}-(\omega^{2}k^{2}-m^{2})u_{k}=\Gamma(x)\>(\mathbf{u}\star\mathbf{u}\star\mathbf{u})_{k}\quad\text{on }\mathbb{R}^{3}.

We now define formally for t,x3t\in\mathbb{R},x\in\mathbb{R}^{3}

(17) U(t,x):=w0(x)+v0(x)+k=12cos(ωkt)vk(x)=keiωktuk(x).\displaystyle U(t,x):=w_{0}(x)+v_{0}(x)+\sum_{k=1}^{\infty}2\cos(\omega kt)\>v_{k}(x)=\sum_{k\in\mathbb{Z}}\mathrm{e}^{\mathrm{i}\omega kt}\>u_{k}(x).

Since by assumption 𝐮=𝐯+𝐰1(,X1)\mathbf{u}=\mathbf{v}+\mathbf{w}\in\ell^{1}(\mathbb{Z},X_{1}), the Weierstrass M-test asserts that the sum in (17) converges in X1X_{1} uniformly with respect to tt\in\mathbb{R}, and hence the map tU(t,)t\mapsto U(t,\,\cdot\,) is continuous as a map from \mathbb{R} to X1X_{1}. We next show stronger regularity properties of U(t,x)U(t,x).

Step 4: Differentiability of U(t,x)U(t,x).

We prove that the map tU(t,)t\mapsto U(t,\,\cdot\,), when interpreted as a map from \mathbb{R} to X1X_{1}, possesses two continuous time derivatives given by

tU(t,)=kiωkeiωktuk,t2U(t,)=kω2k2eiωktuk.\displaystyle\partial_{t}U(t,\,\cdot\,)=\sum_{k\in\mathbb{Z}}\mathrm{i}\omega k\>\mathrm{e}^{\mathrm{i}\omega kt}\>u_{k},\qquad\partial_{t}^{2}U(t,\,\cdot\,)=\sum_{k\in\mathbb{Z}}-\omega^{2}k^{2}\>\mathrm{e}^{\mathrm{i}\omega kt}\>u_{k}.

Indeed, term-by-term differentiation is justified since the sums above as well as in (17) converge in X1X_{1} uniformly with respect to time. This is a consequence of the Weierstraß M-test and the decay estimate in Lemma 3 (iii). Hence, as asserted, the map tU(t,)t\mapsto U(t,\,\cdot\,) is twice continuously differentiable as a map from \mathbb{R} to X1X_{1} - the same strategy yields in fact CC^{\infty} regularity in time.

Similarly, the local regularity estimate in Lemma 3 (ii) implies UC2(×B)U\in C^{2}(\mathbb{R}\times B) for every given ball B=BR(0)3B=B_{R}(0)\subseteq\mathbb{R}^{3} again via term-by-term differentiation. Since the radius of the ball BB is arbitrary, we conclude for tt\in\mathbb{R} and all x3x\in\mathbb{R}^{3}

[t2Δ+m2]U(t,x)\displaystyle\left[\partial_{t}^{2}-\Delta+m^{2}\right]U(t,x) =keiωkt[ω2k2Δ+m2]uk(x)\displaystyle=\sum_{k\in\mathbb{Z}}\mathrm{e}^{\mathrm{i}\omega kt}\>\left[-\omega^{2}k^{2}-\Delta+m^{2}\right]u_{k}(x)
=(16)keiωktΓ(x)l+m+n=kul(x)um(x)un(x)\displaystyle\!\!\overset{\eqref{eq_stationary-proof}}{=}\sum_{k\in\mathbb{Z}}\mathrm{e}^{\mathrm{i}\omega kt}\>\Gamma(x)\>\sum_{l+m+n=k}u_{l}(x)\,u_{m}(x)\,u_{n}(x)
=Γ(x)(leiωltul(x))(meiωmtum(x))(neiωntun(x))\displaystyle=\Gamma(x)\>\left(\sum_{l\in\mathbb{Z}}\mathrm{e}^{\mathrm{i}\omega lt}\>u_{l}(x)\right)\left(\sum_{m\in\mathbb{Z}}\mathrm{e}^{\mathrm{i}\omega mt}\>u_{m}(x)\right)\left(\sum_{n\in\mathbb{Z}}\mathrm{e}^{\mathrm{i}\omega nt}\>u_{n}(x)\right)
=Γ(x)U(t,x)3\displaystyle=\Gamma(x)\>U(t,x)^{3}

where the re-ordering of the summation is justified by absolute convergence of the sums. Thus UU is shown to be a classical solution of the Klein-Gordon equation (1). ∎

Proof of Proposition 4.
We prove the statement for the map FF and then comment on the aspects that differ in case of GG. Using formula (13a), we find for kk\in\mathbb{Z} and 𝐪𝒳1\mathbf{q}\in\mathcal{X}_{1}, recalling that wk=0w_{k}=0 for k{0}k\in\mathbb{Z}\setminus\{0\} and that μsτs=μsπ/2+cot(τs)Ψ~μs\mathcal{R}_{\mu_{s}}^{\tau_{s}}=\mathcal{R}_{\mu_{s}}^{\pi/2}+\cot(\tau_{s})\,\tilde{\Psi}_{\mu_{s}}\ast,

DF(𝟎,0)[𝐪]k=qk3μkτk[Γ(𝐪𝐰𝐰)k]=qk3μkτk[Γw02qk],\displaystyle DF(\mathbf{0},0)[\mathbf{q}]_{k}=q_{k}-3\>\mathcal{R}_{\mu_{k}}^{\tau_{k}}[\Gamma\>(\mathbf{q}\star\mathbf{w}\star\mathbf{w})_{k}]=q_{k}-3\>\mathcal{R}_{\mu_{k}}^{\tau_{k}}\left[\Gamma\>w_{0}^{2}\cdot q_{k}\right],
DF(𝟎,0)[𝐪]0=q03𝒫μ0[Γ(𝐪𝐰𝐰)0]=q03𝒫μ0[Γw02q0].\displaystyle DF(\mathbf{0},0)[\mathbf{q}]_{0}=q_{0}-3\>\mathcal{P}_{\mu_{0}}[\Gamma\>(\mathbf{q}\star\mathbf{w}\star\mathbf{w})_{0}]=q_{0}-3\>\mathcal{P}_{\mu_{0}}\left[\Gamma\>w_{0}^{2}\cdot q_{0}\right].

For 𝐪kerDF(𝟎,0)\mathbf{q}\in\ker DF(\mathbf{0},0), and in view of the choice of τk\tau_{k} in (10), the nondegeneracy properties (11) imply qk0q_{k}\equiv 0 for k,k±sk\in\mathbb{Z},k\neq\pm s. Since τ±s=σs\tau_{\pm s}=\sigma_{s} in (10), Proposition 2 guarantees the existence of a nontrivial solution qsX1q_{s}\in X_{1} of

(18) qs=3μsτs[Γw02qs]\displaystyle q_{s}=3\>\mathcal{R}_{\mu_{s}}^{\tau_{s}}\left[\Gamma\>w_{0}^{2}\cdot q_{s}\right]

which is unique up to a multiplicative factor. Hence kerDF(𝟎,0)\ker DF(\mathbf{0},0) has the asserted form. (We recall here that we consider the subspace of symmetric sequences, whence qs=qsq_{-s}=q_{s}.) Further, by Lemmas 4 and 7 in the final Section 4, the operators

X1X1,{qkqk3μkτk[Γw02qk](k0)q0q03𝒫μ0[Γw02q0]\displaystyle X_{1}\to X_{1},\qquad\begin{cases}q_{k}\mapsto q_{k}-3\>\mathcal{R}_{\mu_{k}}^{\tau_{k}}\left[\Gamma\>w_{0}^{2}\cdot q_{k}\right]&(k\neq 0)\\ q_{0}\mapsto q_{0}-3\>\mathcal{P}_{\mu_{0}}\left[\Gamma\>w_{0}^{2}\cdot q_{0}\right]\end{cases}

are linear compact perturbations of the identity and so kerDF(𝟎,0)\ker DF(\mathbf{0},0) is 1-1-Fredholm. In order to verify transversality, we compute for kk\in\mathbb{Z} and 𝐪kerDF(𝟎,0){𝟎}\mathbf{q}\in\ker DF(\mathbf{0},0)\setminus\{\mathbf{0}\}

λDF(𝟎,0)[𝐪]k={3Ψ~μs[Γw02qs],k=±s,0,else.\displaystyle\partial_{\lambda}DF(\mathbf{0},0)[\mathbf{q}]_{k}=\begin{cases}3\>\tilde{\Psi}_{\mu_{s}}\ast[\Gamma\>w_{0}^{2}\,q_{s}],&k=\pm s,\\ 0,&\text{else}.\end{cases}

Assuming for contradiction that λDF(𝟎,0)[𝐪]=DF(𝟎,0)[𝐩]\partial_{\lambda}DF(\mathbf{0},0)[\mathbf{q}]=DF(\mathbf{0},0)[\mathbf{p}] for some 𝐩𝒳1\mathbf{p}\in\mathcal{X}_{1}, we infer in particular that the component psp_{s} satisfies the convolution identity

(19) ps3μsτs[Γw02ps]=3Ψ~μs[Γw02qs]\displaystyle p_{s}-3\>\mathcal{R}_{\mu_{s}}^{\tau_{s}}\left[\Gamma\>w_{0}^{2}\cdot p_{s}\right]=3\>\tilde{\Psi}_{\mu_{s}}\ast[\Gamma\>w_{0}^{2}\cdot q_{s}]

and hence, following Lemmas 45

Δpsμsps=3Γ(x)w02(x)pson 3,\displaystyle-\Delta p_{s}-\mu_{s}p_{s}=3\>\Gamma(x)\>w_{0}^{2}(x)\,p_{s}\quad\text{on }\mathbb{R}^{3},

which is also nontrivially solved by qsq_{s} as a consequence of (18). Due to the uniqueness statement in Proposition 2, this implies that ps=cqsp_{s}=c\cdot q_{s} for some cc\in\mathbb{R}. But then, applying (18) to (19), we obtain Ψ~μs[Γw02qs]=0\tilde{\Psi}_{\mu_{s}}\ast[\Gamma\>w_{0}^{2}\cdot q_{s}]=0. Hence by the asymptotic expansion in Lemma 4

Γw02qs^(μs)=0\displaystyle\widehat{\Gamma w_{0}^{2}q_{s}}(\sqrt{\mu_{s}})=0

and therefore, due to qs=3μsτs[Γw02qs]q_{s}=3\>\mathcal{R}^{\tau_{s}}_{\mu_{s}}[\Gamma\>w_{0}^{2}\,q_{s}] and Lemma 5,

qs(x)=O(1|x|2) as |x|.\displaystyle q_{s}(x)=O\left(\frac{1}{|x|^{2}}\right)\text{ as }|x|\to\infty.

This contradicts Proposition 2 stating that the leading-order term as |x||x|\to\infty of a nontrivial solution qsq_{s} of Δqsμsqs=3Γ(x)w02(x)qs-\Delta q_{s}-\mu_{s}q_{s}=3\>\Gamma(x)\>w_{0}^{2}(x)\,q_{s} cannot vanish.

In the case τs=0\tau_{s}=0, we see as above that 𝐪kerDG(𝟎,0)\mathbf{q}\in\ker DG(\mathbf{0},0) if and only if qk=0q_{k}=0 for k±sk\neq\pm s, and that qs=qsq_{s}=q_{-s} can be chosen to be the (nontrivial) solution of

(20) qs=3μsπ/2[Γw02qs]+α(μs)(qs)Ψ~μswith β(μs)(qs)=0.\displaystyle q_{s}=3\>\mathcal{R}_{\mu_{s}}^{\pi/2}\left[\Gamma\>w_{0}^{2}\cdot q_{s}\right]+\alpha^{(\mu_{s})}(q_{s})\>\tilde{\Psi}_{\mu_{s}}\quad\text{with }\quad\beta^{(\mu_{s})}(q_{s})=0.

Similarly, kerDG(𝟎,0)\ker DG(\mathbf{0},0) is 1-1-Fredholm. We again assume for contradiction that there is 𝐩𝒳1\mathbf{p}\in\mathcal{X}_{1} with λDG(𝟎,0)[𝐪]=DG(𝟎,0)[𝐩]\partial_{\lambda}DG(\mathbf{0},0)[\mathbf{q}]=DG(\mathbf{0},0)[\mathbf{p}], which implies in particular

(21) ps3μsπ/2[Γw02ps](α(μs)(ps)+β(μs)(ps))Ψ~μs=α(μs)(qs)Ψ~μsp_{s}-3\>\mathcal{R}_{\mu_{s}}^{\pi/2}\left[\Gamma\>w_{0}^{2}\cdot p_{s}\right]-\left(\alpha^{(\mu_{s})}(p_{s})+\beta^{(\mu_{s})}(p_{s})\right)\tilde{\Psi}_{\mu_{s}}=\alpha^{(\mu_{s})}(q_{s})\tilde{\Psi}_{\mu_{s}}

with β(μs)(qs)=0\beta^{(\mu_{s})}(q_{s})=0. Thus, according to Lemma 4, psp_{s} solves the differential equation

Δpsμsps=3Γ(x)w02(x)pson 3,\displaystyle-\Delta p_{s}-\mu_{s}p_{s}=3\>\Gamma(x)\>w_{0}^{2}(x)\,p_{s}\quad\text{on }\mathbb{R}^{3},

which is also solved by qsq_{s}, see equation (20). As before, the uniqueness property in Proposition 2 implies ps=cqsp_{s}=c\cdot q_{s} for some cc\in\mathbb{R}, and inserting this into the identity (21), comparison with (20) yields α(μs)(qs)=0\alpha^{(\mu_{s})}(q_{s})=0. Since also β(μs)(qs)=0\beta^{(\mu_{s})}(q_{s})=0, we infer from the definition of the functionals α(μs),β(μs)\alpha^{(\mu_{s})},\beta^{(\mu_{s})} preceding Lemma 6 that, again, qs(x)=O(1/|x|2)q_{s}(x)=O(1/|x|^{2}), contradicting Proposition 2. ∎

4. Appendix: Stationary Linear Helmholtz and Schrödinger Equations

Given μ>0\mu>0, we study aspects of the solution theory of the linear equations

(22) Δu±μu=fon 3.\displaystyle-\Delta u\pm\mu u=f\qquad\text{on }\mathbb{R}^{3}.

In the case of a “++”, equation (22) is said to be a Schrödinger equation. Given any right-hand side fL2(3)f\in L^{2}(\mathbb{R}^{3}), a unique solution uH2(3)u\in H^{2}(\mathbb{R}^{3}) can be obtained by applying the resolvent (Δ+μ)1(-\Delta+\mu)^{-1}, which can be calculated explicitly by applying the Fourier transform

u=(Δ+μ)1f=3f^(ξ)|ξ|2+μei,ξdξ(2π)3/2.\displaystyle u=(-\Delta+\mu)^{-1}f=\int_{\mathbb{R}^{3}}\frac{\hat{f}(\xi)}{|\xi|^{2}+\mu}\>\mathrm{e}^{\mathrm{i}\left<\,\cdot\,,\xi\right>}\>\frac{\mathrm{d}\xi}{(2\pi)^{3/2}}.

In the case of a Helmholtz equation, i.e. of a “-” sign in (22), this is not possible since μ>0\mu>0 belongs to the essential spectrum of Δ-\Delta on 3\mathbb{R}^{3}. A well-established strategy to find solutions in spaces other than L2(3)L^{2}(\mathbb{R}^{3}) is known as Limiting Absorption Principle(s). The idea is to replace μ\mu by μ+iε\mu+\mathrm{i}\varepsilon, apply an L2L^{2}-resolvent, and pass to the limit ε0\varepsilon\to 0 in a suitable topology, i.e. formally

u=``limε0"(Δ(μ+iε))1f=``limε0"3f^(ξ)|ξ|2(μ+iε)ei,ξdξ(2π)3/2.\displaystyle u=``\lim_{\varepsilon\searrow 0}"\>(-\Delta-(\mu+\mathrm{i}\varepsilon))^{-1}f=``\lim_{\varepsilon\searrow 0}"\>\int_{\mathbb{R}^{3}}\frac{\hat{f}(\xi)}{|\xi|^{2}-(\mu+\mathrm{i}\varepsilon)}\>\mathrm{e}^{\mathrm{i}\left<\,\cdot\,,\xi\right>}\>\frac{\mathrm{d}\xi}{(2\pi)^{3/2}}.

Using tools from harmonic analysis, such a construction of solutions of linear inhomogeneous Helmholtz equations has been successfully done by Agmon [2] in weighted L2L^{2} spaces, and by Kenig, Ruiz and Sogge [14] as well as Gutiérrez [11] in certain pairs of LpL^{p} spaces. The resolvent-type operator is, then, for sufficiently nice ff, given by a convolution

u=ei||μ4π||f.\displaystyle u=\frac{\mathrm{e}^{\mathrm{i}|\,\cdot\,|\sqrt{\mu}}}{4\pi|\,\cdot\,|}\ast f.

Such studies are completed by characterizations of the so-called Herglotz waves, i.e. the solutions of the homogeneous equation Δuμu=0-\Delta u-\mu u=0 on the respective spaces, see e.g. [3].

We study the case of (real-valued, radial) functions fX3,uX1f\in X_{3},u\in X_{1} with the Banach spaces

Xq\displaystyle X_{q} :={vCrad(3)|vXq:=(1+||2)q2v<},q{1,3}.\displaystyle:=\left\{v\in C_{\mathrm{rad}}(\mathbb{R}^{3})\big{|}\>\left\lVert v\right\rVert_{X_{q}}:=\left\lVert(1+|\cdot|^{2})^{\frac{q}{2}}v\right\rVert_{\infty}<\infty\right\},\qquad q\in\{1,3\}.

These have been successfully applied in solving systems of cubic Helmholtz equations in [19]. Let us again point out that the decay rate prescribed in X1X_{1} is the natural one for solutions of Helmholtz equations on the full space 3\mathbb{R}^{3}. Such solutions of the Helmholtz equation

(23) Δuμu=fon 3\displaystyle-\Delta u-\mu u=f\qquad\text{on }\mathbb{R}^{3}

can be obtained using convolution operators with kernels Ψμ,Ψ~μ\Psi_{\mu},\tilde{\Psi}_{\mu} given by

Ψμ(x)=cos(|x|μ)4π|x|,Ψ~μ(x)=sin(|x|μ)4π|x|(x3{0}).\displaystyle\Psi_{\mu}(x)=\frac{\cos(|x|\sqrt{\mu})}{4\pi|x|},\qquad\tilde{\Psi}_{\mu}(x)=\frac{\sin(|x|\sqrt{\mu})}{4\pi|x|}\qquad(x\in\mathbb{R}^{3}\setminus\{0\}).

Here Ψμ,Ψ~μ\Psi_{\mu},\tilde{\Psi}_{\mu} are radial solutions of the homogeneous Helmholtz equation on 3{0}\mathbb{R}^{3}\setminus\{0\}. We notice that Ψ~μ\tilde{\Psi}_{\mu} extends to a smooth solution of Δuμu=0-\Delta u-\mu u=0 in X1X_{1} and it is, up to constant multiples, the only one. Moreover, the following holds:

Lemma 4 ([19], Proposition 4).

The convolution operators fΨμff\mapsto\Psi_{\mu}\ast f, fΨ~μff\mapsto\tilde{\Psi}_{\mu}\ast f are well-defined, linear and compact as operators from X3X_{3} to X1X_{1}. Moreover, given fX3f\in X_{3}, the functions w:=Ψμfw:=\Psi_{\mu}\ast f and w~:=Ψ~μf\tilde{w}:=\tilde{\Psi}_{\mu}\ast f belong to X1Cloc2(3)X_{1}\cap C^{2}_{\mathrm{loc}}(\mathbb{R}^{3}) and satisfy

Δwμw=fon 3,w(x)=4ππ2f^(μ)Ψμ(x)+O(1|x|2);\displaystyle-\Delta w-\mu w=f\quad\text{on }\mathbb{R}^{3},\qquad w(x)=4\pi\>\sqrt{\frac{\pi}{2}}\>\hat{f}(\sqrt{\mu})\>\>\Psi_{\mu}(x)+O\left(\frac{1}{|x|^{2}}\right);
Δw~μw~=0on 3,w~(x)=4ππ2f^(μ)Ψ~μ(x).\displaystyle-\Delta\tilde{w}-\mu\tilde{w}=0\quad\text{on }\mathbb{R}^{3},\qquad\tilde{w}(x)=4\pi\>\sqrt{\frac{\pi}{2}}\>\hat{f}(\sqrt{\mu})\>\>\tilde{\Psi}_{\mu}(x).

Here f^(μ)\hat{f}(\sqrt{\mu}) refers to the profile of the Fourier transform on 3\mathbb{R}^{3}. Working in a radial setting with strongly decaying inhomogeneities fX3f\in X_{3}, the properties in the previous Lemma (and in the following ones) can be verified immediately by explicit calculations and need not be derived from suitable Limiting Absorption Principles; for details, we refer to the earlier article [19].

The study of conditions guaranteeing uniqueness of solutions of (23) in X1X_{1} involves the characterization of Herglotz waves in X1X_{1}, which are all multiples of Ψ~μ\tilde{\Psi}_{\mu}. As in [19], inspired by the analysis of the so-called far field of solutions of Helmholtz equations in scattering theory, we impose asymptotic conditions governing the leading-order contribution of u(x)u(x) as |x||x|\to\infty. For τ(0,π)\tau\in(0,\pi), we introduce

μτ[f]=Ψμf+cot(τ)Ψ~μf=sin(||μ+τ)4πsin(τ)||f.\displaystyle\mathcal{R}_{\mu}^{\tau}[f]=\Psi_{\mu}\ast f+\cot(\tau)\>\tilde{\Psi}_{\mu}\ast f=\frac{\sin(|\,\cdot\,|\sqrt{\mu}+\tau)}{4\pi\sin(\tau)\>|\,\cdot\,|}\ast f.

Then, using the above Lemma 4, one obtains:

Lemma 5 ([19], Corollary 5).

Let τ(0,π)\tau\in(0,\pi) and μ>0\mu>0. Then the operator μτ:X3X1\mathcal{R}_{\mu}^{\tau}:X_{3}\to X_{1} is well-defined, linear and compact. Moreover, given fX3f\in X_{3}, we have u=μτ[f]u=\mathcal{R}_{\mu}^{\tau}[f] if and only if uCloc2u\in C^{2}_{\mathrm{loc}} with

Δuμu=fon 3,u(x)=csin(|x|μ+τ)|x|+O(1|x|2)as |x|\displaystyle-\Delta u-\mu u=f\quad\text{on }\mathbb{R}^{3},\qquad u(x)=c\cdot\frac{\sin(|x|\sqrt{\mu}+\tau)}{|x|}+O\left(\frac{1}{|x|^{2}}\right)\quad\text{as }|x|\to\infty

for some cc\in\mathbb{R}, and in this case c=1sin(τ)π2f^(μ)c=\frac{1}{\sin(\tau)}\>\sqrt{\frac{\pi}{2}}\>\hat{f}(\sqrt{\mu}).

Handling the case of far field conditions with τ=0\tau=0 is somewhat more delicate since the existence of the solution Ψ~μ\tilde{\Psi}_{\mu} (which satisfies exactly this condition) excludes an analogous uniqueness statement. For proving Theorem 1, the following setting is suitable. First, by the Hahn-Banach Theorem, we define continuous linear functionals α(μ),β(μ)X1\alpha^{(\mu)},\beta^{(\mu)}\in X^{\prime}_{1} with the property that, for uX1u\in X_{1} with

u(x)=αuΨ~μ(x)+βuΨμ(x)+O(1|x|2)as |x|,\displaystyle u(x)=\alpha_{u}\cdot\tilde{\Psi}_{\mu}(x)+\beta_{u}\cdot\Psi_{\mu}(x)+O\left(\frac{1}{|x|^{2}}\right)\quad\text{as }|x|\to\infty,

we have α(μ)(u)=αu\alpha^{(\mu)}(u)=\alpha_{u} and β(μ)(u)=βu\beta^{(\mu)}(u)=\beta_{u}, cf. [19], equation (13) and the following explanations. Then, the following analogue of Lemma 5 holds.

Lemma 6.

Given fX3f\in X_{3}, we have u=μπ/2[f]+(α(μ)(u)+β(μ)(u))Ψ~μu=\mathcal{R}_{\mu}^{\pi/2}[f]+(\alpha^{(\mu)}(u)+\beta^{(\mu)}(u))\cdot\tilde{\Psi}_{\mu} if and only if uCloc2u\in C^{2}_{\mathrm{loc}} with

Δuμu=fon 3,u(x)=csin(|x|μ)|x|+O(1|x|2)as |x|\displaystyle-\Delta u-\mu u=f\quad\text{on }\mathbb{R}^{3},\qquad u(x)=c\cdot\frac{\sin(|x|\sqrt{\mu})}{|x|}+O\left(\frac{1}{|x|^{2}}\right)\quad\text{as }|x|\to\infty

for some cc\in\mathbb{R}. In this case, β(μ)(u)=0\beta^{(\mu)}(u)=0.

These results will allow to handle the nonlinear Helmholtz equations in (3b); for the proofs, we refer to the corresponding parts of [19]. Nonlinear Schrödinger equations such as

(24) Δu+μu=fon 3\displaystyle-\Delta u+\mu u=f\qquad\text{on }\mathbb{R}^{3}

for some μ>0\mu>0 can also be discussed in a similar setting, which is certainly neither optimal nor most elegant but perfectly suitable for our purpose as another analogue of Lemma 4.

Lemma 7.

Let μ>0\mu>0. Then the operator

𝒫μ:X3X1,fe||μ4π||f\displaystyle\mathcal{P}_{\mu}:X_{3}\to X_{1},\quad f\mapsto\frac{\mathrm{e}^{-|\,\cdot\,|\sqrt{\mu}}}{4\pi|\,\cdot\,|}\ast f

is well-defined, linear and compact. Moreover, given fX3f\in X_{3}, we have u:=𝒫μ[f]X3Cloc2(3)u:=\mathcal{P}_{\mu}[f]\in X_{3}\cap C^{2}_{\mathrm{loc}}(\mathbb{R}^{3}), and uu is a solution in X1X_{1} of

Δu+μu=fon 3.\displaystyle-\Delta u+\mu u=f\quad\qquad\text{on }\mathbb{R}^{3}.

For details on the proof, which is similar to that of Lemma 4 but with less difficulties due to the strongly localized kernel, cf. [21], Lemma 4.10.

Let us remark that, in the Schrödinger case, we do not obtain a family of possible “resolvent-type” operators as 1τ=1+cot(τ)~1\mathcal{R}_{1}^{\tau}=\mathcal{R}_{1}+\cot(\tau)\tilde{\mathcal{R}}_{1}, 0<τ<π0<\tau<\pi, in the Helmholtz case. This is due to the fact that the homogeneous Schrödinger equation Δu+μu=0-\Delta u+\mu u=0 has no smooth and localized nontrivial solution in X1X_{1}. In particular, a major consequence in our study of Klein-Gordon breathers is that we have to impose nondegeneracy of w0w_{0} as an assumption rather than, as in the Helmholtz case, generate it by choosing an appropriate resolvent 1τ\mathcal{R}_{1}^{\tau}, as will be done in (10), (11a) below.

Acknowledgements

Funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) – Project-ID 258734477 – SFB 1173.

The construction of breather solutions of a similar wave-type equation is a major result of the author’s dissertation thesis and can be found partly verbatim in [21, Chapter 4]. Special thanks goes in particular to my PhD advisor Dr. Rainer Mandel who encouraged me to work on this topic and provided advice whenever asked.

References

  • [1] M. J. Ablowitz, D. J. Kaup, A. C. Newell, and H. Segur. Method for Solving the Sine-Gordon Equation. Phys. Rev. Lett., 30 :  1262–1264, 1973.
  • [2] S. Agmon. Spectral properties of Schrödinger operators and scattering theory. Annali della Scuola Normale Superiore di Pisa - Classe di Scienze, Ser. 4, 2  (2):  151–218, 1975.
  • [3] S. Agmon. A Representation Theorem for Solutions of the Helmholtz Equation and Resolvent Estimates for The Laplacian. Academic Press, 1990.
  • [4] T. Bartsch and M. Willem. Infinitely Many Nonradial Solutions of a Euclidean Scalar Field Equation. Journal of Functional Analysis, 117 (2):  447–460, 1993.
  • [5] T. Bartsch and M. Willem. Infinitely many radial solutions of a semilinear elliptic problem on n\mathbb{R}^{n}. Archive for Rational Mechanics and Analysis, 124 (3):  261–276, 1993.
  • [6] P. W. Bates and J. Shi. Existence and instability of spike layer solutions to singular perturbation problems. Journal of Functional Analysis, 196 (2):  211–264, 2002.
  • [7] B. Birnir, H. McKean, and A. Weinstein. The rigidity of sine-gordon breathers. Communications on Pure and Applied Mathematics, 47 (8):  1043–1051, 1994.
  • [8] C. Blank, M. Chirilus-Bruckner, V. Lescarret, and G. Schneider. Breather Solutions in Periodic Media. Communications in Mathematical Physics, (3):  815–841, 2011.
  • [9] J.-M. Delort. Existence globale et comportement asymptotique pour l’équation de Klein–Gordon quasi linéaire à données petites en dimension 1. Annales Scientifiques de l’École Normale Supérieure, 34  (1):  1–61, 2001.
  • [10] G. Evéquoz and T. Weth. Dual variational methods and nonvanishing for the nonlinear Helmholtz equation. Advances in Mathematics, 280 :  690–728, 2015.
  • [11] S. Gutiérrez. Non trivial Lq{L}^{q} solutions to the Ginzburg-Landau equation. Mathematische Annalen, 328 (1):  1–25, 2004.
  • [12] N. Hayashi and P. I. Naumkin. The initial value problem for the cubic nonlinear Klein-Gordon equation. Zeitschrift für angewandte Mathematik und Physik, 59  (6):  1002–1028, 2008.
  • [13] A. Hirsch and W. Reichel. Real-valued, time-periodic localized weak solutions for a semilinear wave equation with periodic potentials. Nonlinearity, 32  (4):  1408–1439, 2019.
  • [14] C. E. Kenig, A. Ruiz, and C. D. Sogge. Uniform Sobolev inequalities and unique continuation for second order constant coefficient differential operators. Duke Math. J., 55 (2):  329–347, 1987.
  • [15] S. Klainerman. Long-time behavior of solutions to nonlinear evolution equations. Archive for Rational Mechanics and Analysis, 78  (1):  73–98, 1982.
  • [16] S. Klainerman. Global existence of small amplitude solutions to nonlinear Klein-Gordon equations in four space-time dimensions. Communications on Pure and Applied Mathematics, 38  (5):  631–641, 1985.
  • [17] M. K. Kwong and L. Q. Zhang. Uniqueness of the positive solution of Δu+f(u)=0{\Delta}u+f(u)=0 in an annulus. Differential and Integral Equations, 4 (3):  583–599, 1991.
  • [18] P.-L. Lions. The concentration-compactness principle in the calculus of variations. The locally compact case, part 2. Annales de l’I.H.P. Analyse non linéaire, 1 (4):  223–283, 1984.
  • [19] R. Mandel and D. Scheider. Bifurcations of nontrivial solutions of a cubic Helmholtz system. Advances in Nonlinear Analysis, 9  (1):  1026–1045, 2019.
  • [20] K. Moriyama. Normal forms and global existence of solutions to a class of cubic nonlinear Klein-Gordon equations in one space dimension. Differential Integral Equations, 10  (3):  499–520, 1997.
  • [21] D. Scheider. On a Nonlinear Helmholtz System. PhD thesis, Karlsruher Institut für Technologie (KIT), 2019.
  • [22] J. Shatah. Global existence of small solutions to nonlinear evolution equations. Journal of Differential Equations, 46  (3):  409–425, 1982.
  • [23] J. Shatah. Normal forms and quadratic nonlinear Klein-Gordon equations. Communications on Pure and Applied Mathematics, 38  (5):  685–696, 1985.
  • [24] W. A. Strauss. Nonlinear invariant wave equations. In G. Velo and A. S. Wightman, editors, Invariant Wave Equations, pages 197–249. Springer Berlin Heidelberg, 1978.
  • [25] J. Wei. On the Construction of Single-Peaked Solutions to a Singularly Perturbed Semilinear Dirichlet Problem. Journal of Differential Equations, 129 (2):  315–333, 1996.