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

Dispersive estimates for 1D matrix Schrödinger operators with threshold resonance

Yongming Li Department of Mathematics
Texas A&M University
College Station, TX 77843, USA
liyo0008@tamu.edu
Abstract.

We establish dispersive estimates and local decay estimates for the time evolution of non-self-adjoint matrix Schrödinger operators with threshold resonances in one space dimension. In particular, we show that the decay rates in the weighted setting are the same as in the regular case after subtracting a finite rank operator corresponding to the threshold resonances. Such matrix Schrödinger operators naturally arise from linearizing a focusing nonlinear Schrödinger equation around a solitary wave. It is known that the linearized operator for the 1D focusing cubic NLS equation exhibits a threshold resonance. We also include an observation of a favorable structure in the quadratic nonlinearity of the evolution equation for perturbations of solitary waves of the 1D focusing cubic NLS equation.

The author was partially supported by NSF grants DMS-1954707 and DMS-2235233.

1. Introduction

In this article, we establish dispersive estimates and local decay estimates for the (non-self-adjoint) matrix Schrödinger operators

=0+𝒱=[x2+μ00x2μ]+[V1V2V2V1]on L2()×L2(),\mathcal{H}={\mathcal{H}}_{0}+\mathcal{V}=\begin{bmatrix}-\partial_{x}^{2}+\mu&0\\ 0&\partial_{x}^{2}-\mu\end{bmatrix}+\begin{bmatrix}-V_{1}&-V_{2}\\ V_{2}&V_{1}\end{bmatrix}\quad\text{on $L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})$}, (1.1)

where μ\mu is a positive constant and V1V_{1}, V2V_{2} are real-valued sufficiently decaying potentials. The operator \mathcal{H} is closed on the domain D()=H2()×H2()D(\mathcal{H})=H^{2}({\mathbb{R}})\times H^{2}({\mathbb{R}}).

These matrix operators arise when linearizing a focusing nonlinear Schrödinger equation around a solitary wave. By our assumptions on V1V_{1} and V2V_{2}, Weyl’s criterion implies that the essential spectrum of \mathcal{H} is the same as that of 0\mathcal{H}_{0}, given by (,μ][μ,)(-\infty,-\mu]\cup[\mu,\infty). As a core assumption in this paper, we suppose that the edges ±μ\pm\mu of the essential spectrum are irregular in the sense of Definition 4.4. This implies that there exist non-trivial bounded solutions to the equation Ψ±=±μΨ±\mathcal{H}\vec{\Psi}_{\pm}=\pm\mu\vec{\Psi}_{\pm}, see Lemma 4.5. The dispersive estimates for \mathcal{H} when the thresholds ±μ\pm\mu are regular have been obtained in Section 7-8 of the paper by Krieger-Schlag [KS06], building on the scattering theory developed by Buslaev-Perel’man [BP95]. See also the recent work of Collot-Germain [CG23]. Our proof is instead based on the unifying approach to resolvent expansions first initiated by Jensen-Nenciu [JN01], and then further refined in Erdogan-Schlag [ES06] for matrix Schrödinger operators. We also adopt techniques from Erdogan-Green [EG21], where the authors prove similar dispersive estimates for one-dimensional Dirac operators.

1.1. Motivation

Our interest in developing dispersive estimates for (1.1) stems from the asymptotic stability problem for solitary wave solutions to nonlinear Schrödinger (NLS) equations. The NLS equation

itψ+x2ψ+F(|ψ|2)ψ=0,whereψ:t×x,\begin{split}i\partial_{t}\psi+\partial_{x}^{2}\psi+F(|\psi|^{2})\psi&=0,{\ \ \text{where}\ \ }\psi\colon{\mathbb{R}}_{t}\times{\mathbb{R}}_{x}\rightarrow{\mathbb{C}},\end{split} (1.2)

appears in many important physical contexts such as the propagation of a laser beam, the envelope description of water waves in an ideal fluid, or the propagation of light waves in nonlinear optical fibers. See, e.g., Sulem-Sulem [SS07] for physics background.

Under certain general conditions on the nonlinearity F()F(\cdot) (see, e.g., [BL83]), the equation (1.2) admits a parameterized family of localized, finite energy, traveling solitary waves of the form ψ(t,x)=eitα2ϕ(x;α)\psi(t,x)=e^{it\alpha^{2}}\phi(x;\alpha), where ϕ(;α)\phi(\cdot;\alpha) is a ground state, i.e., a positive, decaying, real-valued solution to the (nonlinear) elliptic equation

x2ϕ+α2ϕ=F(ϕ2)ϕ.-\partial_{x}^{2}\phi+\alpha^{2}\phi=F(\phi^{2})\phi. (1.3)

The existence and uniqueness of these ground state solutions are well-understood, see, e.g., [BL83], [Kwo89].

The solitary wave solutions (or simply, solitons) are of importance due to the special role they play for the long-time dynamics of the Cauchy problem (1.2). Consequently, over the last few decades there has been a significant interest in the study of stability (or instability) of such solitary waves under small perturbations. The primary notion of stability is that of orbital stability, and it is by now well-understood for the NLS equation. The pioneering works in this direction were due to Cazenave-Lions [CL82], Shatah-Strauss [SS85], and Weinstein [Wei85]; see also [GSS87] for the general theory. On the other hand, a stronger notion of stability is that of asymptotic stability. There are two general approaches for the asymptotic stability problem. The first approach is to use integrability techniques, when the underlying partial differential equation is completely integrable and inverse scattering is available. A second approach is perturbative, which means that one studies the dynamics of the nonlinear flow in the neighborhood of the solitary wave, on a restricted set of the initial data. Generally, one starts by decomposing the perturbed solution into a sum of a solitary wave and a dispersive remainder term. For the perturbative approach, dispersive estimates for the linear flow are key.

Let us briefly describe the perturbative approach for the NLS equation. To keep our exposition short, we will not take into account any modulation aspects related to the Galilean invariance of the equation. For small α>0\alpha>0, consider the perturbation ansatz ψ(t,x)=eitα2(ϕ(x)+u(t,x))\psi(t,x)=e^{it\alpha^{2}}(\phi(x)+u(t,x)) with the ground state ϕ()=ϕ(;α)\phi(\cdot)=\phi(\cdot;\alpha) and the dispersive remainder term u(t,x)u(t,x). The linearization of (1.2) around the solitary wave eitα2ϕ(x)e^{it\alpha^{2}}\phi(x) then leads to the following nonlinear partial differential equation

itu=(x2+α2V)u+Wu¯+N,i\partial_{t}u=(-\partial_{x}^{2}+\alpha^{2}-V)u+W\overline{u}+N,

where N=N(ϕ,u,u¯)N=N(\phi,u,\overline{u}) is nonlinear in the variables (u,u¯)(u,\overline{u}), and V=F(ϕ2)+F(ϕ2)ϕ2V=F(\phi^{2})+F^{\prime}(\phi^{2})\phi^{2} and W=F(ϕ2)ϕ2W=F^{\prime}(\phi^{2})\phi^{2} are real-valued potentials related to the ground state ϕ\phi. Equivalently, the above equation can be recast as a system for the vector U:=(u,u¯)U:=(u,\overline{u})^{\top}, which is given by

itUU=𝒩,i\partial_{t}U-\mathcal{H}U=\mathcal{N}, (1.4)

where 𝒩\mathcal{N} is a nonlinear term, and \mathcal{H} is a matrix Schrödinger operator of the form (1.1) with the parameters μ=α2\mu=\alpha^{2}, V1=VV_{1}=V, and V2=WV_{2}=W.

For the study of asymptotic stability of solitary waves for NLS, it is thus crucial to fully understand the spectral properties of the matrix operator \mathcal{H} as well as to derive dispersive estimates for the linear evolution operator eite^{it\mathcal{H}}. One of the key steps in a perturbative analysis is to prove that the dispersive remainder (1.4) decays to zero in a suitable topology. Let us consider for example, the 1D focusing NLS with a pure power nonlinearity, i.e.

itψ+x2ψ+|ψ|2σψ=0,whereσ>0.i\partial_{t}\psi+\partial_{x}^{2}\psi+|\psi|^{2\sigma}\psi=0,{\ \ \text{where}\ \ }\sigma>0. (1.5)

The ground state ϕ(x;1)\phi(x;1) has an explicit formula for all σ>0\sigma>0 given by

ϕ(x;1)=(σ+1)12σsech1σ(σx),\phi(x;1)=(\sigma+1)^{\frac{1}{2\sigma}}\operatorname{sech}^{\frac{1}{\sigma}}(\sigma x), (1.6)

and the linearized operator around eitϕ(x;1)e^{it}\phi(x;1) takes the form

σ=[x2(σ+1)2sech2(σx)+1σ(σ+1)sech2(σx)σ(σ+1)sech2(σx)x2+(σ+1)2sech2(σx)1].\mathcal{H}_{\sigma}=\begin{bmatrix}-\partial_{x}^{2}-(\sigma+1)^{2}\operatorname{sech}^{2}(\sigma x)+1&-\sigma(\sigma+1)\operatorname{sech}^{2}(\sigma x)\\ \sigma(\sigma+1)\operatorname{sech}^{2}(\sigma x)&\partial_{x}^{2}+(\sigma+1)^{2}\operatorname{sech}^{2}(\sigma x)-1\end{bmatrix}.

For monomial nonlinearities, we may obtain ϕ(x;α)\phi(x;\alpha) from rescaling by ϕ(x;α)=α1σϕ(αx,1)\phi(x;\alpha)=\alpha^{\frac{1}{\sigma}}\phi(\alpha x,1). The matrix operators when linearizing around eitα2ϕ(x;α)e^{it\alpha^{2}}\phi(x;\alpha) are also equivalent to the matrix operator σ\mathcal{H}_{\sigma} by rescaling. The spectra for these matrix operators were investigated in [CGNT08]; see also Section 9 of [KS06]. For σ2\sigma\geq 2, Krieger-Schlag [KS06] were able to construct finite co-dimensional center-stable manifolds around the solitary waves and prove asymptotic stability using dispersive and Strichartz estimates developed for the evolution operator eite^{it\mathcal{H}}. However, for the completely integrable case (σ=1\sigma=1), it was shown in [CGNT08] that the matrix operator 1\mathcal{H}_{1} exhibits the threshold resonance Ψ(x)=(tanh2(x),sech2(x))\Psi(x)=\big{(}\tanh^{2}(x),-\operatorname{sech}^{2}(x)\big{)}^{\top} at λ=1\lambda=1. The dispersive estimates developed in [KS06] do not apply in this case. Furthermore, we note that a key assumption in the papers [BP95], [GS05], [KS06], [CG23] is that the linearized matrix operator \mathcal{H} does not possess threshold resonances at the edges of the essential spectrum. In these “generic” (regular) cases, it can be shown that the evolution operator enjoy improved decay estimates in weighted spaces; see, e.g., Proposition 8.1 in [KS06]. Thus, a meaningful motivation for this paper is to prove dispersive estimates in the presence of threshold resonances under some general spectral assumptions on the matrix operator \mathcal{H}, which are applicable to the 1D cubic NLS case (σ=1\sigma=1). We will discuss this particular case briefly in Section 1.4.

1.2. Main result

We are now in the position to state the main result of this paper. We begin by specifying some spectral assumptions on \mathcal{H}.

Assumption 1.1.
  1. (A1)

    σ3𝒱-\sigma_{3}\mathcal{V} is a positive matrix, where σ3\sigma_{3} is one of the Pauli matrices (c.f. (1.9)),

  2. (A2)

    L:=x2+μV1+V2L_{-}:=-\partial_{x}^{2}+\mu-V_{1}+V_{2} is non-negative,

  3. (A3)

    there exists β>0\beta>0 such that |V1(x)|+|V2(x)|e(2μ+β)|x||V_{1}(x)|+|V_{2}(x)|\lesssim e^{-(\sqrt{2\mu}+\beta)|x|} for all xx\in{\mathbb{R}},

  4. (A4)

    there are no embedded eigenvalues in (,μ)(μ,)(-\infty,-\mu)\cup(\mu,\infty).

Under these assumptions, we recall the general spectral theory for \mathcal{H} from [ES06].111The results in Section 2 of [ES06] are stated for dimension 3, but they in fact hold for all dimensions. Moreover, only a polynomial decay on V1V_{1} and V2V_{2} is assumed in [ES06]. See also [HL07, Theorem 1.3].

Lemma 1.2.

[ES06, Lemma 3] Suppose Assumption 1.1 holds. The essential spectrum of \mathcal{H} equals (,μ][μ,)(-\infty,-\mu]\cup[\mu,\infty). Moreover,

spec()=spec()=spec()¯=spec(),\operatorname{spec}(\mathcal{H})=-\operatorname{spec}(\mathcal{H})=\overline{\operatorname{spec}(\mathcal{H})}=\operatorname{spec}(\mathcal{H}^{*}), (1.7)

and spec()i\operatorname{spec}(\mathcal{H})\subset{\mathbb{R}}\cup i{\mathbb{R}}. The discrete spectrum of \mathcal{H} consists of eigenvalues {zj}j=1N\{z_{j}\}_{j=1}^{N}, 0N<0\leq N<\infty, of finite multiplicity. For each zj0z_{j}\neq 0, the algebraic and geometric multiplicities coincide and Ran(zj)\operatorname{Ran}(\mathcal{H}-z_{j}) is closed. The zero eigenvalue has finite algebraic multiplicity, i.e., the generalized eigenspace k=1ker(k)\cup_{k=1}^{\infty}\ker(\mathcal{H}^{k}) has finite dimension. In fact, there is a finite m1m\geq 1 so that ker(k)=ker(k+1)\ker(\mathcal{H}^{k})=\ker(\mathcal{H}^{k+1}) for all kmk\geq m.

The symmetry (1.7) is due to the following commutation properties of \mathcal{H},

=σ3σ3,=σ1σ1,\mathcal{H}^{*}=\sigma_{3}\mathcal{H}\sigma_{3},\qquad-\mathcal{H}=\sigma_{1}\mathcal{H}\sigma_{1}, (1.8)

with the Pauli matrices

σ1=[0110],σ2=[0ii0],σ3=[1001].\sigma_{1}=\begin{bmatrix}0&1\\ 1&0\end{bmatrix},\quad\sigma_{2}=\begin{bmatrix}0&-i\\ i&0\end{bmatrix},\quad\sigma_{3}=\begin{bmatrix}1&0\\ 0&-1\end{bmatrix}. (1.9)

As a core assumption in this paper, we impose that the thresholds ±μ\pm\mu of the essential spectrum are irregular.

Assumption 1.3.
  1. (A5)

    The thresholds ±μ\pm\mu are irregular in the sense of Definition 4.4. This implies that there exist non-trivial bounded solutions Ψ±=(Ψ1±,Ψ2±)\vec{\Psi}_{\pm}=(\Psi_{1}^{\pm},\Psi_{2}^{\pm})^{\top} to the equation Ψ±=±μΨ±\mathcal{H}\vec{\Psi}_{\pm}=\pm\mu\vec{\Psi}_{\pm}.

  2. (A6)

    The vanishing (bilateral)-Laplace transform condition holds

    𝔏[V2Ψ1++V1Ψ2+](±2μ)=e2μ(V2Ψ1++V1Ψ2+)(y)dy=0.\mathfrak{L}[V_{2}\Psi_{1}^{+}+V_{1}\Psi_{2}^{+}](\pm\sqrt{2\mu})=\int_{-\infty}^{\infty}e^{\mp\sqrt{2\mu}}(V_{2}\Psi_{1}^{+}+V_{1}\Psi_{2}^{+})(y)\,\mathrm{d}y=0. (1.10)

For details about the characterization of the threshold functions Ψ\vec{\Psi}, we refer the reader to Definition 4.4 and Lemma 4.5 in Section 4. Due to the commutation identity (1.8), we have the relation Ψ+=σ1Ψ\vec{\Psi}_{+}=\sigma_{1}\vec{\Psi}_{-}. We emphasize that assumption (A6) is used to infer that (non-trivial) bounded solutions Ψ±=(Ψ1±,Ψ2±)\vec{\Psi}_{\pm}=(\Psi_{1}^{\pm},\Psi_{2}^{\pm}) to the equation Ψ±=±μΨ±\mathcal{H}\vec{\Psi}_{\pm}=\pm\mu\vec{\Psi}_{\pm} satisfy Ψ1+=Ψ2L()L2()\Psi_{1}^{+}=\Psi_{2}^{-}\in L^{\infty}({\mathbb{R}})\setminus L^{2}({\mathbb{R}}).

Let Pd:L2()×L2()L2()×L2()P_{\mathrm{d}}\colon L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})\to L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}}) be the Riesz projection corresponding to the discrete spectrum of \mathcal{H}, and define Ps:=IPdP_{\mathrm{s}}:=I-P_{\mathrm{d}}. We now state the main theorem of this article.

Theorem 1.4.

Suppose assumptions (A1) – (A6) hold, and let Ψ=(Ψ1,Ψ2)\vec{\Psi}=(\Psi_{1},\Psi_{2}) be the L()×L()L2()×L2()L^{\infty}({\mathbb{R}})\times L^{\infty}({\mathbb{R}})\setminus L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}}) distributional solution to

Ψ=μΨ,\mathcal{H}\vec{\Psi}=\mu\vec{\Psi}, (1.11)

with the normalization

limx(|Ψ1(x)|2+|Ψ1(x)|2)=2.\lim_{x\to\infty}\left(|\Psi_{1}(x)|^{2}+|\Psi_{1}(-x)|^{2}\right)=2. (1.12)

Then, for any f=(f1,f2)𝒮()×𝒮()\vec{f}=(f_{1},f_{2})\in\mathcal{S}({\mathbb{R}})\times\mathcal{S}({\mathbb{R}}), we have

  1. (1)

    the unweighted dispersive estimate

    eitPsfL()×L()|t|12fL1()×L1(),|t|1,\left\|e^{it\mathcal{H}}P_{\mathrm{s}}\vec{f}\,\right\|_{L^{\infty}({\mathbb{R}})\times L^{\infty}({\mathbb{R}})}\lesssim|t|^{-\frac{1}{2}}\left\|\vec{f}\,\right\|_{L^{1}({\mathbb{R}})\times L^{1}({\mathbb{R}})},\quad\forall\,|t|\geq 1, (1.13)
  2. (2)

    and the weighted dispersive estimate

    x2(eitPsFt)fL()×L()|t|32x2fL1()×L1(),|t|1,\left\|\langle x\rangle^{-2}(e^{it\mathcal{H}}P_{s}-F_{t})\vec{f}\,\right\|_{L^{\infty}({\mathbb{R}})\times L^{\infty}({\mathbb{R}})}\lesssim|t|^{-\frac{3}{2}}\left\|\langle x\rangle^{2}\vec{f}\,\right\|_{L^{1}({\mathbb{R}})\times L^{1}({\mathbb{R}})},\quad\forall\,|t|\geq 1, (1.14)

    where

    Ftf:=eitμ4πitσ3Ψ,fΨeitμ4πitσ3σ1Ψ,fσ1Ψ.F_{t}\vec{f}:=\frac{e^{it\mu}}{\sqrt{-4\pi it}}\langle\sigma_{3}\vec{\Psi},\vec{f}\,\rangle\vec{\Psi}-\frac{e^{-it\mu}}{\sqrt{4\pi it}}\langle\sigma_{3}\sigma_{1}\vec{\Psi},{\vec{f}}\,\rangle\sigma_{1}\vec{\Psi}. (1.15)

We proceed with some remarks on the main theorem:

  1. (1)

    The estimate (1.14) is an analogue of the weighted dispersive estimates obtained by Goldberg [Gol07] for the scalar Schödinger operator H=x2+VH=-\partial_{x}^{2}+V on the real line for non-generic potentials VV; see [Gol07, Theorem 2]. The local decay estimate (1.14) shows that the bulk of the free wave eitPse^{it\mathcal{H}}P_{\mathrm{s}} enjoys improved local decay at the integrable rate 𝒪(|t|32)\mathcal{O}(|t|^{-\frac{3}{2}}), and that the slow 𝒪(|t|12)\mathcal{O}(|t|^{-\frac{1}{2}}) local decay can be pinned down to the contribution of the finite rank operator FtF_{t}. Such sharp information can be useful for nonlinear asymptotic stability problems, see also Section 1.4 below.

  2. (2)

    We make some comments on the spectral hypotheses. The assumptions (A1)–(A4) are known to be satisfied by the linearized operator around the solitary wave for the 1D focusing power-type NLS (1.5). In the case of the 1D focusing cubic NLS (σ=1\sigma=1), the linearized operator 1\mathcal{H}_{1} satisfies the assumptions (A1)–(A6); see Section 1.4.1 below. More generally, in Lemma 4.5, we show for matrix operators \mathcal{H} of the form (1.1) satisfying assumptions (A1)–(A6) that the edges ±μ\pm\mu of the essential spectrum of \mathcal{H} cannot be eigenvalues, and that the non-trivial bounded solutions Ψ±=(Ψ1±,Ψ2±)\vec{\Psi}_{\pm}=(\Psi_{1}^{\pm},\Psi_{2}^{\pm})^{\top} to Ψ±=±μΨ±\mathcal{H}\vec{\Psi}_{\pm}=\pm\mu\vec{\Psi}_{\pm} belong to LL2L^{\infty}\setminus L^{2} since Ψ1(x)\Psi_{1}(x) has a non-zero limit as x±x\to\pm\infty. In this sense, we characterize the solutions Ψ±\vec{\Psi}_{\pm} as threshold resonances. However, it is not yet clear to the author whether assumption (A6) is strictly needed to show that non-trivial bounded solutions Ψ±\vec{\Psi}_{\pm} to Ψ±=±μΨ±\mathcal{H}\vec{\Psi}_{\pm}=\pm\mu\vec{\Psi}_{\pm} cannot be eigenfunctions. Moreover, an inspection of the proof of Lemma 4.5 reveals that the strong exponential decay assumption (A3) and the vanishing condition assumption (A6) are only used in a Volterra integral equation argument. In all other proofs, we only use some polynomial decay of the potentials V1V_{1} and V2V_{2}.

  3. (3)

    It might be possible to prove Theorem 1.4 using the scattering theory developed by [BP95]. However, one major difficulty for this approach is due to the fact that the matrix Wronskian associated with the vector Jost solutions is not invertible at the origin for cases where the matrix operators \mathcal{H} exhibit threshold resonances. Hence, the vector-valued distorted Fourier basis functions are not immediately well-defined at zero frequency. See Corollary 5.21 and Section 6 in [KS06] for further details.

1.3. Previous works

In this subsection, we collect references related to dispersive estimates for Schrödinger operators and to the study of the stability of solitary waves.

For dispersive estimates for the matrix Schrödinger operator \mathcal{H}, we refer to Section 5-9 of [KS06] in dimension 1, and to [ES06, Mar11, Gre12, EG13, Top17] in higher dimensions. A comprehensive study on the spectral theory for the matrix operator arising from pure-power type NLS is given in [CGNT08]. See also [Vou10, CHS11, MMS20] for related analytical and numerical studies. For dispersive estimates for the scalar Schrödinger operators, pioneering works include [Mur82, JSS91, Wed00], and we refer to [ES04, GS04, Sch05, Sch06, Gol07, Gol10, Miz11, Gre12, EGG14, GG15, Bec16, GG17] for a sample of recent works. Finally, we mention the papers [BGW85, JN01] on resolvent expansions for the scalar Schrödinger operator.

On the general well-posedness theory for the NLS Cauchy problem (1.2), we refer to the pioneering works [GV79, Kat87, Tsu87]. Results on the orbital stability (or instability) of solitary waves for the NLS equation were first obtained by [BC81, CL82, SS85, Wei85, Wei86], and a general theory was established in [GSS87]. Subsequent developments for general nonlinearities were due to [Gri88, Gri90, Oht95, CP03, Mae08]. Regarding the asymptotic stability of solitary waves, the first results were due to Buslaev-Perel’man [BP92, BP95]. Subsequent works in this direction were due to [GS05, KS06, Bec08, Sch09, Cuc14, Mar23, CG23]. For surveys on the stability of solitary waves, we refer to the reviews [KMM17, CM21] and the monographs [Caz03, SS07].

1.4. On the solitary wave for the 1D focusing cubic NLS

In this subsection, we present two observations related to the asymptotic stability problem for the solitary wave of the 1D focusing cubic NLS. First, we verify that the assumption (A6) holds for the linearized operator around the solitary wave of the 1D focusing cubic NLS. Second, we use the local decay estimate (1.14) to shed some light on the leading order structure of the quadratic nonlinearity in the perturbation equation for the solitary wave of the 1D focusing cubic NLS.

We note that a proof for the asymptotic stability problem has been given by Cuccagna-Pelinovsky [CP14] via inverse scattering techniques. On the other hand, a perturbative proof that does not explicitly rely on the integrable structure has not yet appeared in the literature to the best of the author’s knowledge. We now briefly discuss the evolution equation for perturbations of the solitary wave for the 1D focusing cubic NLS. To keep our exposition short, we do not discuss the modulation aspects for the solitary wave. For simplicity, consider the perturbation ansatz

ψ(t,x)=eit(Q(x)+u(t,x))\psi(t,x)=e^{it}(Q(x)+u(t,x))

for the equation (1.5) (σ=1\sigma=1). The ground state has the explicit formula

Q(x):=ϕ(x;1)=2sech(x).Q(x):=\phi(x;1)=\sqrt{2}\operatorname{sech}(x).

The evolution equation for the perturbation in vector form u=(u1,u2):=(u,u¯)\vec{u}=(u_{1},u_{2}):=(u,\bar{u}) is given by

itu1u=𝒬(u)+𝒞(u),\begin{split}&i\partial_{t}\vec{u}-\mathcal{H}_{1}\vec{u}=\mathcal{Q}({\vec{u}})+\mathcal{C}(\vec{u}),\end{split} (1.16)

where

1=0+𝒱1=[x2+100x21]+[4sech2(x)2sech2(x)2sech2(x)4sech2(x)],\mathcal{H}_{1}=\mathcal{H}_{0}+\mathcal{V}_{1}=\begin{bmatrix}-\partial_{x}^{2}+1&0\\ 0&\partial_{x}^{2}-1\end{bmatrix}+\begin{bmatrix}-4\operatorname{sech}^{2}(x)&-2\operatorname{sech}^{2}(x)\\ 2\operatorname{sech}^{2}(x)&4\operatorname{sech}^{2}(x)\end{bmatrix}, (1.17)

and

𝒬(u):=[Qu122Qu1u2Qu22+2Qu1u2],and𝒞(u):=[u12u2u1u22].\mathcal{Q}({\vec{u}}):=\begin{bmatrix}-Qu_{1}^{2}-2Qu_{1}u_{2}\\ Qu_{2}^{2}+2Qu_{1}u_{2}\end{bmatrix},{\ \ \text{and}\ \ }\mathcal{C}({\vec{u}}):=\begin{bmatrix}-u_{1}^{2}u_{2}\\ u_{1}u_{2}^{2}\end{bmatrix}. (1.18)

Recall from [CGNT08] that the matrix operator 1\mathcal{H}_{1} has the essential spectrum (,1][1,)(-\infty,-1]\cup[1,\infty), and a four-dimensional generalized nullspace

𝒩g(1)=span{[QQ],[(1+xx)Q(1+xx)Q],[xQxQ],[xQxQ]},\mathcal{N}_{\mathrm{g}}(\mathcal{H}_{1})=\operatorname*{span}\left\{\begin{bmatrix}Q\\ -Q\end{bmatrix},\begin{bmatrix}(1+x\partial_{x})Q\\ (1+x\partial_{x})Q\end{bmatrix},\begin{bmatrix}\partial_{x}Q\\ \partial_{x}Q\end{bmatrix},\begin{bmatrix}xQ\\ -xQ\end{bmatrix}\right\}, (1.19)

as well as a threshold resonance at +1+1 given by

ΨΨ+:=[Ψ1Ψ2]=[112Q212Q2]=[tanh2(x)sech2(x)].\vec{\Psi}\equiv\vec{\Psi}_{+}:=\begin{bmatrix}\Psi_{1}\\ \Psi_{2}\end{bmatrix}=\begin{bmatrix}1-\frac{1}{2}Q^{2}\\ -\frac{1}{2}Q^{2}\end{bmatrix}=\begin{bmatrix}\tanh^{2}(x)\\ -\operatorname{sech}^{2}(x)\end{bmatrix}. (1.20)

By symmetry, there is also a threshold resonance function at 1-1 given by

Ψ=σ1Ψ+=[sech2(x)tanh2(x)].\vec{\Psi}_{-}=\sigma_{1}\vec{\Psi}_{+}=\begin{bmatrix}-\operatorname{sech}^{2}(x)\\ \tanh^{2}(x)\end{bmatrix}. (1.21)

The eigenfunctions listed in (1.19) are related to the underlying symmetries for the NLS equation. Note that we have normalized the resonance function Ψ\vec{\Psi} to satisfy the condition (1.12) stated in Theorem 1.4.

1.4.1. On assumption (A6) for the 1D focusing cubic NLS

Our first observation is that the assumption (A6) is satisfied by the matrix operator 1\mathcal{H}_{1}.

Lemma 1.5.

Let V1(x)=4sech2(x)V_{1}(x)=4\operatorname{sech}^{2}(x), V2(x)=2sech2(x)V_{2}(x)=2\operatorname{sech}^{2}(x), and (Ψ1(x),Ψ2(x))=(tanh2(x),sech2(x))(\Psi_{1}(x),\Psi_{2}(x))=(\tanh^{2}(x),-\operatorname{sech}^{2}(x)). Then, we have

e±2y(V2(y)Ψ1(y)+V1(y)Ψ2(y))dy=0.\int_{{\mathbb{R}}}e^{\pm\sqrt{2}y}\big{(}V_{2}(y)\Psi_{1}(y)+V_{1}(y)\Psi_{2}(y)\big{)}\,\mathrm{d}y=0. (1.22)
Proof.

We denote the (two-sided) Laplace transform by

𝔏[f](s)=esyf(y)dy,s,\mathfrak{L}[f](s)=\int_{-\infty}^{\infty}e^{-sy}f(y)\,\mathrm{d}y,\quad s\in{\mathbb{C}}, (1.23)

which is formally related to the Fourier transform by

𝔏[f](s)=2π[f](is).\mathfrak{L}[f](s)=\sqrt{2\pi}\mathcal{F}[f](is).

By direct computation,

(V1Ψ2+V2Ψ1)(x)=2sech2(x)6sech4(x),(V_{1}\Psi_{2}+V_{2}\Psi_{1})(x)=2\operatorname{sech}^{2}(x)-6\operatorname{sech}^{4}(x),

and

sech4(x)=23sech2(x)16x2(sech2(x)).\operatorname{sech}^{4}(x)=\frac{2}{3}\operatorname{sech}^{2}(x)-\frac{1}{6}\partial_{x}^{2}(\operatorname{sech}^{2}(x)). (1.24)

Recall from [LS21, Corollary 5.7] that as equalities in 𝒮()\mathcal{S}({\mathbb{R}}),

[sech2](ξ)=π2ξsinh(π2ξ).\mathcal{F}[\operatorname{sech}^{2}](\xi)=\sqrt{\frac{\pi}{2}}\frac{\xi}{\sinh(\tfrac{\pi}{2}\xi)}. (1.25)

Hence, using the basic property [x2f](ξ)=ξ2[f](ξ)\mathcal{F}[-\partial_{x}^{2}f](\xi)=\xi^{2}\mathcal{F}[f](\xi) and (1.24), we obtain

[sech4](ξ)=16π2ξ(4+ξ2)sinh(π2ξ).\mathcal{F}[\operatorname{sech}^{4}](\xi)=\frac{1}{6}\sqrt{\frac{\pi}{2}}\frac{\xi(4+\xi^{2})}{\sinh(\tfrac{\pi}{2}\xi)}. (1.26)

As complex functions, we recall that sinh(iz)=isin(z)\sinh(iz)=i\sin(z) and that zzsin(z)z\mapsto\frac{z}{\sin(z)} is analytic222to be pedantic, there is a removable singularity at z=0z=0 which we can remove by setting the function zsin(z)\frac{z}{\sin(z)} equal to 11 at z=0z=0. in the strip {s+iσ:s(π,π),σ}\{s+i\sigma:s\in(-\pi,\pi),\,\sigma\in{\mathbb{R}}\}. Thus, by analytic continuation,

𝔏[V1Ψ2+V2Ψ1](s)=2π(2[sech2](is)6[sech4](is))=πs(2+s2)sin(πs2),\mathfrak{L}[V_{1}\Psi_{2}+V_{2}\Psi_{1}](s)=\sqrt{2\pi}\left(2\,\mathcal{F}[\operatorname{sech}^{2}](is)-6\mathcal{F}[\operatorname{sech}^{4}](is)\right)=\frac{\pi s(-2+s^{2})}{\sin(\tfrac{\pi s}{2})},

for any ss\in{\mathbb{C}} with (s)(2,2)\Re(s)\in(-2,2), which in particular proves the vanishing condition (1.22). ∎

The other assumptions (A1)–(A5) for 1\mathcal{H}_{1} are also satisfied by either checking directly or invoking the results from Section 9 in [KS06].

1.4.2. Null structure for perturbations of the solitary wave of the 1D focusing cubic NLS

Due to the slow local decay of the Schrödinger waves in the presence of a threshold resonance, the spatially localized quadratic nonlinearity in (1.16) may pose significant difficulties for proving decay of small solutions to (1.16). The weighted dispersive estimate (1.14) shows that the slow local decay is only due to the finite rank projection FtF_{t}. To shed some light on the expected leading order behavior of the quadratic nonlinearity 𝒬(u)\mathcal{Q}({\vec{u}}) in (1.16), it is instructive to insert a free Schrödinger wave

ufree(t):=eitPsf,{\vec{u}}_{\mathrm{free}}(t):=e^{-it\mathcal{H}}P_{\mathrm{s}}{\vec{f}},

for some fixed f𝒮()×𝒮(){\vec{f}}\in\mathcal{S}({\mathbb{R}})\times\mathcal{S}({\mathbb{R}}). By Theorem 1.4, we have

ufree(t)=ceitt[Ψ1Ψ2]+c+eitt[Ψ2Ψ1]+r(t),{\vec{u}}_{\mathrm{free}}(t)=c_{-}\frac{e^{-it}}{\sqrt{t}}\begin{bmatrix}\Psi_{1}\\ \Psi_{2}\end{bmatrix}+c_{+}\frac{e^{it}}{\sqrt{t}}\begin{bmatrix}\Psi_{2}\\ \Psi_{1}\end{bmatrix}+{\vec{r}}(t), (1.27)

with

c=14πiσ3Ψ,f,c+=14πiσ3σ1Ψ,f,c_{-}=\frac{1}{\sqrt{-4\pi i}}\langle\sigma_{3}\vec{\Psi},{\vec{f}}\,\rangle,\quad c_{+}=-\frac{1}{\sqrt{4\pi i}}\langle\sigma_{3}\sigma_{1}\vec{\Psi},{\vec{f}}\,\rangle, (1.28)

and where the remainder r(t){\vec{r}}(t) satisfies

x2r(t)Lx()×Lx()|t|32x2fLx1()×Lx1().\left\|\langle x\rangle^{-2}{\vec{r}}(t)\right\|_{L_{x}^{\infty}({\mathbb{R}})\times L_{x}^{\infty}({\mathbb{R}})}\lesssim|t|^{-\frac{3}{2}}\left\|\langle x\rangle^{2}{\vec{f}}\,\right\|_{L_{x}^{1}({\mathbb{R}})\times L_{x}^{1}({\mathbb{R}})}. (1.29)

Thus, owing to the spatial localization of the quadratic nonlinearity, we have

𝒬(ufree(t))=c+2e2itt𝒬1(Ψ)+c+ct𝒬2(Ψ)+c2e2itt𝒬3(Ψ)+𝒪L(|t|2),\mathcal{Q}({\vec{u}}_{\mathrm{free}}(t))=\frac{c_{+}^{2}e^{2it}}{t}\mathcal{Q}_{1}(\vec{\Psi})+\frac{c_{+}c_{-}}{t}\mathcal{Q}_{2}(\vec{\Psi})+\frac{c_{-}^{2}e^{-2it}}{t}\mathcal{Q}_{3}(\vec{\Psi})+\mathcal{O}_{L^{\infty}}(|t|^{-2}), (1.30)

where

𝒬1(Ψ)\displaystyle\mathcal{Q}_{1}(\vec{\Psi}) =[QΨ222QΨ1Ψ2QΨ12+2QΨ1Ψ2],\displaystyle=\begin{bmatrix}-Q\Psi_{2}^{2}-2Q\Psi_{1}\Psi_{2}\\ Q\Psi_{1}^{2}+2Q\Psi_{1}\Psi_{2}\end{bmatrix}, (1.31)
𝒬2(Ψ)\displaystyle\mathcal{Q}_{2}(\vec{\Psi}) =[2QΨ1Ψ22Q(Ψ12+Ψ22)2QΨ1Ψ2+2Q(Ψ12+Ψ22)],\displaystyle=\begin{bmatrix}-2Q\Psi_{1}\Psi_{2}-2Q(\Psi_{1}^{2}+\Psi_{2}^{2})\\ 2Q\Psi_{1}\Psi_{2}+2Q(\Psi_{1}^{2}+\Psi_{2}^{2})\end{bmatrix}, (1.32)
𝒬3(Ψ)\displaystyle\mathcal{Q}_{3}(\vec{\Psi}) =σ1𝒬1(Ψ)=[QΨ122QΨ1Ψ2QΨ22+2QΨ1Ψ2].\displaystyle=-\sigma_{1}\mathcal{Q}_{1}(\vec{\Psi})=\begin{bmatrix}-Q\Psi_{1}^{2}-2Q\Psi_{1}\Psi_{2}\\ Q\Psi_{2}^{2}+2Q\Psi_{1}\Psi_{2}\end{bmatrix}. (1.33)

Due to the critical 𝒪(|t|1)\mathcal{O}(|t|^{-1}) decay of the leading order terms on the right-hand side of (1.30), it is instructive to analyze the long-time behavior of small solutions to the inhomogeneous matrix Schrödinger equation with such a source term

{itusrc1usrc=Ps(c+2e2itt𝒬1(Ψ)+c+ct𝒬2(Ψ)+c2e2itt𝒬3(Ψ)),t1,usrc(1)=0.\left\{\begin{aligned} i\partial_{t}{\vec{u}}_{\mathrm{src}}-\mathcal{H}_{1}{\vec{u}}_{\mathrm{src}}&=P_{\mathrm{s}}\left(\frac{c_{+}^{2}e^{2it}}{t}\mathcal{Q}_{1}(\vec{\Psi})+\frac{c_{+}c_{-}}{t}\mathcal{Q}_{2}(\vec{\Psi})+\frac{c_{-}^{2}e^{-2it}}{t}\mathcal{Q}_{3}(\vec{\Psi})\right),\quad t\geq 1,\\ {\vec{u}}_{\mathrm{src}}(1)&=\vec{0}.\end{aligned}\right. (1.34)

To this end, it will be useful to exploit a special conjugation identity for the matrix Schrödinger operator 1\mathcal{H}_{1}. It was recently pointed out by Martel, see [Mar23, Section 2.3], that the matrix operator 1\mathcal{H}_{1} can be conjugated to the flat matrix Schrödinger operator 0\mathcal{H}_{0}. By first conjugating 1\mathcal{H}_{1} with the unitary matrix 𝒥=12[1i1i]\mathcal{J}=\frac{1}{\sqrt{2}}\begin{bmatrix}1&i\\ 1&-i\end{bmatrix}, we obtain the equivalent matrix Schrödinger operator

1=i𝒥11𝒥:=[0LL+0]=0+𝒲:=[0x2+1x210]+[02sech2(x)6sech2(x)0].\mathcal{L}_{1}=-i\mathcal{J}^{-1}\mathcal{H}_{1}\mathcal{J}:=\begin{bmatrix}0&L_{-}\\ -L_{+}&0\end{bmatrix}=\mathcal{L}_{0}+\mathcal{W}:=\begin{bmatrix}0&-\partial_{x}^{2}+1\\ \partial_{x}^{2}-1&0\end{bmatrix}+\begin{bmatrix}0&-2\operatorname{sech}^{2}(x)\\ 6\operatorname{sech}^{2}(x)&0\end{bmatrix}.

Introducing the operator

𝒟:=[0(x2+1)S2S2L+0],whereS:=QxQ1=x+tanh(x),\mathcal{D}:=\begin{bmatrix}0&(-\partial_{x}^{2}+1)S^{2}\\ -S^{2}L_{+}&0\end{bmatrix},{\ \ \text{where}\ \ }S:=Q\cdot\partial_{x}\cdot Q^{-1}=\partial_{x}+\tanh(x), (1.35)

one has the conjugation identity (see also [CGNT08, Section 3.4])

𝒟1=0𝒟.\mathcal{D}\mathcal{L}_{1}=\mathcal{L}_{0}\mathcal{D}. (1.36)

We then transfer the above identity to the matrix operator \mathcal{H} by setting 𝒟~:=𝒥𝒟𝒥1\widetilde{\mathcal{D}}:=\mathcal{J}\mathcal{D}\mathcal{J}^{-1} to obtain the conjugation identity

𝒟~1=0𝒟~.\widetilde{\mathcal{D}}\mathcal{H}_{1}=\mathcal{H}_{0}\widetilde{\mathcal{D}}. (1.37)

Moreover, it can be checked directly that 𝒟~η=0\widetilde{\mathcal{D}}\vec{\eta}=0 for any generalized eigenfunction η𝒩g(1)\vec{\eta}\in\mathcal{N}_{\mathrm{g}}(\mathcal{H}_{1}), and this implies that 𝒟~Pd0\widetilde{\mathcal{D}}P_{\mathrm{d}}\equiv 0, which is equivalent to saying that 𝒟~=𝒟~Ps\widetilde{\mathcal{D}}=\widetilde{\mathcal{D}}P_{\mathrm{s}}. Hence, by applying the transformation 𝒟~\widetilde{\mathcal{D}} to the equation (1.34), we obtain the transformed equation

itvsrc0vsrc=𝒟~(c+2e2itt𝒬1(Ψ)+c+ct𝒬2(Ψ)+c2e2itt𝒬3(Ψ)),i\partial_{t}{\vec{v}}_{\mathrm{src}}-\mathcal{H}_{0}{\vec{v}}_{\mathrm{src}}=\widetilde{\mathcal{D}}\left(\frac{c_{+}^{2}e^{2it}}{t}\mathcal{Q}_{1}(\vec{\Psi})+\frac{c_{+}c_{-}}{t}\mathcal{Q}_{2}(\vec{\Psi})+\frac{c_{-}^{2}e^{-2it}}{t}\mathcal{Q}_{3}(\vec{\Psi})\right), (1.38)

where vsrc:=𝒟~usrc{\vec{v}}_{\mathrm{src}}:=\widetilde{\mathcal{D}}{\vec{u}}_{\mathrm{src}} is the transformed variable. Note that the above equation features the flat operator 0\mathcal{H}_{0} on the left. The Duhamel formula for vsrc(t){\vec{v}}_{\mathrm{src}}(t) at times t1t\geq 1 reads

vsrc(t)=i1tei(ts)0𝒟~(c+2e2iss𝒬1(Ψ)+c+cs𝒬2(Ψ)+c2e2iss𝒬3(Ψ))ds.{\vec{v}}_{\mathrm{src}}(t)=-i\int_{1}^{t}e^{-i(t-s)\mathcal{H}_{0}}\widetilde{\mathcal{D}}\left(\frac{c_{+}^{2}e^{2is}}{s}\mathcal{Q}_{1}(\vec{\Psi})+\frac{c_{+}c_{-}}{s}\mathcal{Q}_{2}(\vec{\Psi})+\frac{c_{-}^{2}e^{-2is}}{s}\mathcal{Q}_{3}(\vec{\Psi})\right)\,\mathrm{d}s. (1.39)

The flat, self-adjoint, matrix operator 0\mathcal{H}_{0} has the benefit that the semigroup eit0e^{-it\mathcal{H}_{0}} can be represented in terms of the standard Fourier transform by the formula

(eit0g)(x)=12πeit(ξ2+1)g1^(ξ)eixξdξe¯1+12πeit(ξ2+1)g2^(ξ)eixξdξe¯2,\left(e^{-it\mathcal{H}_{0}}{\vec{g}}\right)(x)=\frac{1}{\sqrt{2\pi}}\int_{{\mathbb{R}}}e^{-it(\xi^{2}+1)}\widehat{g_{1}}(\xi)e^{ix\xi}\,\mathrm{d}\xi\,\underline{e}_{1}+\frac{1}{\sqrt{2\pi}}\int_{{\mathbb{R}}}e^{it(\xi^{2}+1)}\widehat{g_{2}}(\xi)e^{ix\xi}\,\mathrm{d}\xi\,\underline{e}_{2}, (1.40)

where g=(g1,g2){\vec{g}}=(g_{1},g_{2})^{\top} and e¯1,e¯2\underline{e}_{1},\underline{e}_{2} are the standard unit vectors in 2{\mathbb{R}}^{2}. The profile of vsrc(t){\vec{v}}_{\mathrm{src}}(t) is given by

fsrc(t):=eit0vsrc(t).{\vec{f}}_{\mathrm{src}}(t):=e^{it\mathcal{H}_{0}}{\vec{v}}_{\mathrm{src}}(t). (1.41)

Setting

𝒟~𝒬j(Ψ)=:(Gj,1,Gj,2)for 1j3,\widetilde{\mathcal{D}}\mathcal{Q}_{j}(\vec{\Psi})=:(G_{j,1},G_{j,2})^{\top}{\ \ \text{for}\ \ }1\leq j\leq 3,

we have for times t1t\geq 1 that

[fsrc(t)](ξ)=c+21teis(ξ2+3)sG1,1^(ξ)dse¯1+c+c1teis(ξ2+1)sG2,1^(ξ)dse¯1+c21teis(ξ21)sG3,1^(ξ)dse¯1+c+21teis(ξ21)sG1,2^(ξ)dse¯2+c+c1teis(ξ2+1)sG2,2^(ξ)dse¯2+c21teis(ξ2+3)sG3,2^(ξ)dse¯2.\begin{split}&\mathcal{F}[{\vec{f}}_{\mathrm{src}}(t)](\xi)\\ &=c_{+}^{2}\int_{1}^{t}\frac{e^{is(\xi^{2}+3)}}{s}\widehat{G_{1,1}}(\xi)\,\mathrm{d}s\,\underline{e}_{1}+c_{+}c_{-}\int_{1}^{t}\frac{e^{is(\xi^{2}+1)}}{s}\widehat{G_{2,1}}(\xi)\,\mathrm{d}s\,\underline{e}_{1}+c_{-}^{2}\int_{1}^{t}\frac{e^{is(\xi^{2}-1)}}{s}\widehat{G_{3,1}}(\xi)\,\mathrm{d}s\,\underline{e}_{1}\\ &\quad+c_{+}^{2}\int_{1}^{t}\frac{e^{-is(\xi^{2}-1)}}{s}\widehat{G_{1,2}}(\xi)\,\mathrm{d}s\,\underline{e}_{2}+c_{+}c_{-}\int_{1}^{t}\frac{e^{-is(\xi^{2}+1)}}{s}\widehat{G_{2,2}}(\xi)\,\mathrm{d}s\,\underline{e}_{2}+c_{-}^{2}\int_{1}^{t}\frac{e^{-is(\xi^{2}+3)}}{s}\widehat{G_{3,2}}(\xi)\,\mathrm{d}s\,\underline{e}_{2}.\end{split} (1.42)

The uniform-in-time boundedness in LξL_{\xi}^{\infty} of the Fourier transform of the profile [fsrc(t)](ξ)\mathcal{F}[{\vec{f}}_{\mathrm{src}}(t)](\xi) is related to recovering the free decay rate for vsrc(t){\vec{v}}_{\mathrm{src}}(t). However, in view of the critical decay of the integrand, this requires favorable time oscillations. Observe that the above terms with time phases e±is(ξ2+1)e^{\pm is(\xi^{2}+1)}, e±is(ξ2+3)e^{\pm is(\xi^{2}+3)} are non-stationary for any ss\in{\mathbb{R}} which implies that they have a better decay rate using integration by parts in the variable ss. On the other hand, the terms with the phases e±is(ξ21)e^{\pm is(\xi^{2}-1)} are stationary at the points ξ=±1\xi=\pm 1. Thus, it is important to know if the Fourier coefficients G3,1^(±1)\widehat{G_{3,1}}(\pm 1) and G1,2^(±1)\widehat{G_{1,2}}(\pm 1) vanish. Indeed, this is true due to the following lemma.

Lemma 1.6.

It holds that

G3,1^(±1)=G1,2^(±1)=0.\widehat{G_{3,1}}(\pm 1)=\widehat{G_{1,2}}(\pm 1)=0. (1.43)
Proof.

First, to ease notation, we write

𝒟~=i2[(D1D2)(D1D2)(D1+D2)(D1+D2)],\widetilde{\mathcal{D}}=\frac{i}{2}\begin{bmatrix}(-D_{1}-D_{2})&(D_{1}-D_{2})\\ (-D_{1}+D_{2})&(D_{1}+D_{2})\end{bmatrix}, (1.44)

where

D1:=(x2+1)S2=(x2+1)(x+tanh(x))(x+tanh(x)),D2:=S2L+=(x+tanh(x))(x+tanh(x))(x26sech2(x)+1).\begin{split}&D_{1}:=(-\partial_{x}^{2}+1)S^{2}=(-\partial_{x}^{2}+1)(\partial_{x}+\tanh(x))(\partial_{x}+\tanh(x)),\\ &D_{2}:=S^{2}L_{+}=(\partial_{x}+\tanh(x))(\partial_{x}+\tanh(x))(-\partial_{x}^{2}-6\operatorname{sech}^{2}(x)+1).\end{split} (1.45)

Since σ1𝒟~=𝒟~σ1\sigma_{1}\widetilde{\mathcal{D}}=-\widetilde{\mathcal{D}}\sigma_{1} and 𝒬3(Ψ)=σ1𝒬1(Ψ)\mathcal{Q}_{3}(\vec{\Psi})=-\sigma_{1}\mathcal{Q}_{1}(\vec{\Psi}) (c.f. (1.33)), it follows that G3,1G1,2G_{3,1}\equiv G_{1,2} as functions. Note that

G3,1=i2(D1(QΨ12)+D1(QΨ22)+2D1(2QΨ1Ψ2)+D2(QΨ12)D2(QΨ22)),G_{3,1}=\frac{i}{2}\left(D_{1}(Q\Psi_{1}^{2})+D_{1}(Q\Psi_{2}^{2})+2D_{1}(2Q\Psi_{1}\Psi_{2})+D_{2}(Q\Psi_{1}^{2})-D_{2}(Q\Psi_{2}^{2})\right), (1.46)

where

(QΨ12)(x)=2sech(x)tanh4(x),(QΨ1Ψ2)(x)=2sech3(x)tanh2(x),(QΨ22)(x)=2sech5(x).\begin{split}(Q\Psi_{1}^{2})(x)&=\sqrt{2}\operatorname{sech}(x)\tanh^{4}(x),\\ (Q\Psi_{1}\Psi_{2})(x)&=-\sqrt{2}\operatorname{sech}^{3}(x)\tanh^{2}(x),\\ (Q\Psi_{2}^{2})(x)&=\sqrt{2}\operatorname{sech}^{5}(x).\end{split}

By using the trigonometric identity sech2(x)+tanh2(x)=1\operatorname{sech}^{2}(x)+\tanh^{2}(x)=1, we may simplify the expression for G3,1G_{3,1} into

G3,1(x)=i22(D1(sech(x)6sech3(x)+6sech5(x))+D2(sech(x)2sech3(x))).G_{3,1}(x)=\frac{i\sqrt{2}}{2}\Big{(}D_{1}\big{(}\operatorname{sech}(x)-6\operatorname{sech}^{3}(x)+6\operatorname{sech}^{5}(x)\big{)}+D_{2}\big{(}\operatorname{sech}(x)-2\operatorname{sech}^{3}(x)\big{)}\Big{)}.

By patient direct computation, we find

F1(x):=D1(sech(x)6sech3(x)+6sech5(x))=192sech3(x)3456sech5(x)+9720sech7(x)6720sech9(x)\begin{split}F_{1}(x)&:=D_{1}\big{(}\operatorname{sech}(x)-6\operatorname{sech}^{3}(x)+6\operatorname{sech}^{5}(x)\big{)}\\ &=192\operatorname{sech}^{3}(x)-3456\operatorname{sech}^{5}(x)+9720\operatorname{sech}^{7}(x)-6720\operatorname{sech}^{9}(x)\end{split} (1.47)

and

F2(x):=D2(sech(x)2sech3(x))=48sech3(x)264sech5(x)+240sech7(x).\begin{split}F_{2}(x):=D_{2}\big{(}\operatorname{sech}(x)-2\operatorname{sech}^{3}(x)\big{)}=48\operatorname{sech}^{3}(x)-264\operatorname{sech}^{5}(x)+240\operatorname{sech}^{7}(x).\end{split} (1.48)

Moreover, using the identities

(x2sech)(x)=sech(x)2sech3(x),(x4sech)(x)=sech(x)20sech3(x)+24sech5(x),(x6sech)(x)=sech(x)182sech3(x)+840sech5(x)720sech7(x),(x8sech)(x)=sech(x)1640sech3(x)+23184sech5(x)60480sech7(x)+40320sech9(x),\begin{split}(\partial_{x}^{2}\operatorname{sech})(x)&=\operatorname{sech}(x)-2\operatorname{sech}^{3}(x),\\ (\partial_{x}^{4}\operatorname{sech})(x)&=\operatorname{sech}(x)-20\operatorname{sech}^{3}(x)+24\operatorname{sech}^{5}(x),\\ (\partial_{x}^{6}\operatorname{sech})(x)&=\operatorname{sech}(x)-182\operatorname{sech}^{3}(x)+840\operatorname{sech}^{5}(x)-720\operatorname{sech}^{7}(x),\\ (\partial_{x}^{8}\operatorname{sech})(x)&=\operatorname{sech}(x)-1640\operatorname{sech}^{3}(x)+23184\operatorname{sech}^{5}(x)-60480\operatorname{sech}^{7}(x)+40320\operatorname{sech}^{9}(x),\\ \end{split} (1.49)

we obtain

F1(x)=16(x2+3x43x6+x8)sech(x)=16(x2+1)3(x2)sech(x),F_{1}(x)=-\frac{1}{6}\left(-\partial_{x}^{2}+3\partial_{x}^{4}-3\partial_{x}^{6}+\partial_{x}^{8}\right)\operatorname{sech}(x)=-\frac{1}{6}(-\partial_{x}^{2}+1)^{3}(-\partial_{x}^{2})\operatorname{sech}(x), (1.50)

and

F2(x)=13(x2+2x4x6)sech(x)=13(x2+1)2(x2)sech(x).F_{2}(x)=\frac{1}{3}\left(-\partial_{x}^{2}+2\partial_{x}^{4}-\partial_{x}^{6}\right)\operatorname{sech}(x)=\frac{1}{3}(-\partial_{x}^{2}+1)^{2}(-\partial_{x}^{2})\operatorname{sech}(x). (1.51)

Thus, using the property [x2f]=ξ2[f](ξ)\mathcal{F}[-\partial_{x}^{2}f]=\xi^{2}\mathcal{F}[f](\xi) and the fact that

sech^(ξ)=π2sech(πξ2),\widehat{\operatorname{sech}}(\xi)=\sqrt{\frac{\pi}{2}}\operatorname{sech}\left(\frac{\pi\xi}{2}\right),

we compute that

G3,1^(ξ)=i22(F1^(ξ)+F2^(ξ))=iπ12(ξ21)ξ2(ξ2+1)2sech(πξ2),\widehat{G_{3,1}}(\xi)=\frac{i\sqrt{2}}{{2}}\left(\widehat{F_{1}}(\xi)+\widehat{F_{2}}(\xi)\right)=-\frac{i\sqrt{\pi}}{12}(\xi^{2}-1)\xi^{2}(\xi^{2}+1)^{2}\operatorname{sech}\left(\frac{\pi\xi}{2}\right), (1.52)

which implies (1.43) as claimed. ∎

Remark 1.7.

We determined the identities (1.47) – (1.51) with the aid of the Wolfram Mathematica software.

The above lemma shows that the localized quadratic resonant terms are well-behaved for the nonlinear perturbation equation (1.16). The presence of this null structure is potentially a key ingredient for a perturbative proof of the asymptotic stability of the solitary wave solutions to the 1D focusing cubic NLS. We end this subsection with the following closing remark.

Remark 1.8.

The motivation for analyzing the quadratic nonlinearity in the perturbation equation (1.16) and for uncovering the null structure for the localized quadratic resonant terms in Lemma 1.6 is due to the recent work by Lührmann-Schlag [LS21], where the authors investigate the asymptotic stability of kink solutions to the 1D sine-Gordon equation under odd perturbations. In [LS21], the authors employ a similar conjugation identity like the one we used in (1.37) to transform the scalar Schrödinger operator H1:=x22sech2(x)H_{1}:=-\partial_{x}^{2}-2\operatorname{sech}^{2}(x) to the flat operator H0:=x2H_{0}:=-\partial_{x}^{2} for the perturbation equation. In fact, it is easy to check that one has the conjugation identity SH1=H0SSH_{1}=H_{0}S, where S=x+tanh(x)S=\partial_{x}+\tanh(x). Moreover, an analogue of Lemma 1.6 on the non-resonant property for the localized quadratic resonant terms in the perturbation equation for the sine-Gordon kink was first obtained in [LLSS23, Remark 1.2]. This remarkable null structure for the sine-Gordon model played a key role in the asymptotic stability proof in [LS21]. In [LS23], the same authors obtained long-time decay estimates for even perturbation of the soliton of the 1D focusing cubic Klein-Gordon equation. The absence of the null structure in the nonlinearity of the perturbation equation in the focusing cubic Klein-Gordon model is a major obstruction to full co-dimension one asymptotic stability result under even perturbations.

Our short discussion on the effects of the threshold resonance on the quadratic term for (1.16) suggests that the localized quadratic resonant terms are well-behaved for the perturbation equation in the 1D cubic NLS model. However, note that a full perturbative proof of the asymptotic stability problem for this model has to encompass the modulation theory associated to the moving solitary wave, and take into account the long-range (modified) scattering effects due to the non-localized cubic nonlinearities in the perturbation equation. We point out that Collot-Germain [CG23] recently obtained general such asymptotic stability results for solitary waves for 1D nonlinear Schrödinger equations under the assumption that the linearized matrix Schrödinger operator does not exhibit threshold resonances.

1.5. Organization of the article

The remaining sections of this paper are devoted to the proof of Theorem 1.4. In Section 2, we state a few stationary phase lemmas, which will be heavily utilized in Sections 5 and 6, and we will also provide an analogue of Theorem 1.4 for the free matrix operator 0\mathcal{H}_{0}. In Section 3, we employ the symmetric resolvent expansion following the framework in [ES06], and in Section 4, we carefully extract the leading operators for these resolvent expansions. A characterization of the threshold resonance is stated in Lemma 4.5 under the spectral assumptions (A1)–(A6). Then, in Section 5, we prove dispersive estimates for the evolution operator eite^{it\mathcal{H}} in the low energy regime. The approach taken in Section 5 largely follows the techniques employed in [EG21] for one-dimensional Dirac operators. In Section 6, we prove dispersive estimates for the remaining energy regimes and finish the proof of Theorem 1.4.

1.6. Notation

For any f=(f1,f2),g=(g1,g2)L2()×L2(){\vec{f}}=(f_{1},f_{2})^{\top},\ {\vec{g}}=(g_{1},g_{2})^{\top}\in L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}}), we use the inner product

f,g:=fgdx=(f¯1g1+f¯2g2)dx,wheref:=(f¯1,f¯2).\langle{\vec{f}},{\vec{g}}\rangle:=\int_{{\mathbb{R}}}\vec{f}^{*}\vec{g}\ \mathrm{d}x=\int_{{\mathbb{R}}}\left(\bar{f}_{1}g_{1}+\bar{f}_{2}g_{2}\right)\mathrm{d}x,{\ \ \text{where}\ \ }\vec{f}^{*}:=(\bar{f}_{1},\bar{f}_{2}). (1.53)

The Schwartz space is denoted by 𝒮()\mathcal{S}({\mathbb{R}}) and we use the weighted L2L^{2}-spaces

Xσ:=xσL2()×xσL2(),fXσ:=xσfL2()×L2(),whereσ.X_{\sigma}:=\langle x\rangle^{-\sigma}L^{2}({\mathbb{R}})\times\langle x\rangle^{-\sigma}L^{2}({\mathbb{R}}),\quad\|{\vec{f}}\|_{X_{\sigma}}:=\|\langle x\rangle^{\sigma}{\vec{f}}\|_{L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})},{\ \ \text{where}\ \ }\sigma\in{\mathbb{R}}. (1.54)

Note that for any α>β>0\alpha>\beta>0, one has the continuous inclusions

XαXβX0=L2()×L2()XβXα,X_{\alpha}\subset X_{\beta}\subset X_{0}=L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})\subset X_{-\beta}\subset X_{-\alpha}, (1.55)

and the duality Xα=XαX_{\alpha}^{*}=X_{-\alpha}. Our convention for the Fourier transform is

[f](ξ)=f^(ξ)=12πeixξf(x)dx,1[f](x)=fˇ(x)=12πeixξf(ξ)dξ.\mathcal{F}[f](\xi)={\hat{f}}(\xi)=\frac{1}{\sqrt{2\pi}}\int_{\mathbb{R}}e^{-ix\xi}f(x)\mathrm{d}x,\quad\mathcal{F}^{-1}[f](x)={\check{f}}(x)=\frac{1}{\sqrt{2\pi}}\int_{\mathbb{R}}e^{ix\xi}f(\xi)\mathrm{d}\xi.

We denote by C>0C>0 an absolute constant whose value is allowed to change from line to line. In order to indicate that the constant depends on a parameter, say θ\theta, we will use the notation CθC_{\theta} or C(θ)C(\theta). For non-negative XX, YY we write XYX\lesssim Y if XCYX\leq CY. We use the Japanese bracket notation x=(1+x2)12\langle x\rangle=(1+x^{2})^{\frac{1}{2}} for xx\in{\mathbb{R}}. The standard tensors on 2{\mathbb{R}}^{2} are denoted by

e¯1=[10],e¯2=[01],e¯11=e¯1e¯1=[1000],e¯22=e¯2e¯2=[0001].\begin{split}\underline{e}_{1}=\begin{bmatrix}1\\ 0\end{bmatrix},\quad\underline{e}_{2}=\begin{bmatrix}0\\ 1\end{bmatrix},\quad\underline{e}_{11}=\underline{e}_{1}\underline{e}_{1}^{\top}=\begin{bmatrix}1&0\\ 0&0\end{bmatrix},\quad\underline{e}_{22}=\underline{e}_{2}\underline{e}_{2}^{\top}=\begin{bmatrix}0&0\\ 0&1\end{bmatrix}.\end{split} (1.56)

Acknowledgments. The author would like to thank his Ph.D. advisor Jonas Lührmann for suggesting the problem and patiently checking the manuscript. The author is grateful to Andrew Comech, Wilhelm Schlag, Gigliola Staffilani, and Ebru Toprak for helpful discussions.

2. Free matrix Schrödinger estimates

In this section, we derive dispersive estimates for the free evolution semigroup eit0e^{it\mathcal{H}_{0}}. We recall that the free matrix Schödinger operator

0=[x2+μ00x2μ],\mathcal{H}_{0}=\begin{bmatrix}-\partial_{x}^{2}+\mu&0\\ 0&\partial_{x}^{2}-\mu\end{bmatrix},

has a purely continuous spectrum

spec(0)=σac(0)=(,μ][μ,),\operatorname{spec}(\mathcal{H}_{0})=\sigma_{\mathrm{ac}}(\mathcal{H}_{0})=(-\infty,-\mu]\cup[\mu,\infty),

and the resolvent operator of 0\mathcal{H}_{0} is given by

(0λ)1=[R0(λμ)00R0(λμ)],λ(,μ][μ,),(\mathcal{H}_{0}-\lambda)^{-1}=\begin{bmatrix}R_{0}(\lambda-\mu)&0\\ 0&-R_{0}(-\lambda-\mu)\end{bmatrix},\quad\lambda\in{\mathbb{C}}\setminus(-\infty,-\mu]\cup[\mu,\infty), (2.1)

where R0R_{0} is the resolvent operator for the one-dimensional Laplacian, with an integral kernel given by

R0(ζ2)(x,y):=(2ζ2)1(x,y)=eiζ|xy|2iζ,ζ+,R_{0}(\zeta^{2})(x,y):=(-\partial^{2}-\zeta^{2})^{-1}(x,y)=\frac{-e^{i\zeta|x-y|}}{2i\zeta},\quad\zeta\in{\mathbb{C}}_{+}, (2.2)

where +{\mathbb{C}}_{+} is the upper half-plane. We obtain from the scalar resolvent theory due to Agmon [Agm75] that the limiting resolvent operators

(0(λ±i0))1=limε0(0(λ±iε))1,λ(,μ)(μ,),\big{(}\mathcal{H}_{0}-(\lambda\pm i0)\big{)}^{-1}=\lim_{{\varepsilon}\downarrow 0}\,\big{(}\mathcal{H}_{0}-(\lambda\pm i{\varepsilon})\big{)}^{-1},\quad\lambda\in(-\infty,-\mu)\cup(\mu,\infty),

are well defined as operators from XσXσX_{\sigma}\to X_{-\sigma} for any σ>12\sigma>\frac{1}{2}. Here, the matrix operator 0\mathcal{H}_{0} is self-adjoint and Stone’s formula applies:

eit0=12πi|λ|μeitλ[(0(λ+i0))1(0(λi0))1]dλ.e^{it\mathcal{H}_{0}}=\frac{1}{2\pi i}\int_{|\lambda|\geq\mu}e^{it\lambda}\left[\big{(}\mathcal{H}_{0}-(\lambda+i0)\big{)}^{-1}-\big{(}\mathcal{H}_{0}-(\lambda-i0)\big{)}^{-1}\right]\,\mathrm{d}\lambda. (2.3)

Let us focus on the spectrum on the positive semi-axis [μ,)[\mu,\infty), as the negative part can be treated using the symmetric properties of \mathcal{H} (c.f. Remark 3.3). By invoking the change of variables λλ=μ+z2\lambda\mapsto\lambda=\mu+z^{2} with 0<z<0<z<\infty, the kernel of eit0Ps+e^{it\mathcal{H}_{0}}P_{\mathrm{s}}^{+} is then given by

eit0Ps+(x,y)=eitμπi0eitz2z[(0(μ+z2+i0))1(0(μ+z2i0))1](x,y)dz.e^{it\mathcal{H}_{0}}P_{\mathrm{s}}^{+}(x,y)=\frac{e^{it\mu}}{\pi i}\int_{0}^{\infty}e^{itz^{2}}z\left[\big{(}\mathcal{H}_{0}-(\mu+z^{2}+i0)\big{)}^{-1}-\big{(}\mathcal{H}_{0}-(\mu+z^{2}-i0)\big{)}^{-1}\right](x,y)\,\mathrm{d}z.

Here, the notation Ps+P_{\mathrm{s}}^{+} means that we restrict the free evolution eit0e^{it\mathcal{H}_{0}} to the positive semi-axis in the integral representation (2.3). By (2.1) and (2.2), we have

(0(μ+z2±i0))1(x,y)=[±ie±iz|xy|2z00ez2+2μ|xy|2z2+2μ],0<z<,\big{(}\mathcal{H}_{0}-(\mu+z^{2}\pm i0)\big{)}^{-1}(x,y)=\begin{bmatrix}\frac{\pm ie^{\pm iz|x-y|}}{2z}&0\\ 0&-\frac{e^{-\sqrt{z^{2}+2\mu}|x-y|}}{2\sqrt{z^{2}+2\mu}}\end{bmatrix},\quad 0<z<\infty, (2.4)

and thus,

eit0Ps+(x,y)=eitμ2πeitz2eiz|xy|e¯11dz.e^{it\mathcal{H}_{0}}P_{\mathrm{s}}^{+}(x,y)=\frac{e^{it\mu}}{2\pi}\int_{\mathbb{R}}e^{itz^{2}}e^{iz|x-y|}\underline{e}_{11}\,\mathrm{d}z. (2.5)

Note that the above integral is to be understood in the principal value sense, due to the pole in (2.4). To this end, we recall the following standard stationary phase results. The first lemma is a direct consequence of the classic van der Corput lemma.

Lemma 2.1.

Let rr\in{\mathbb{R}}, and let ψ(z)\psi(z) be a compactly supported smooth function. Then for any |t|>0|t|>0,

|eitz2+izrψ(z)dz|C|t|12zψLz1().\left|\int_{{\mathbb{R}}}e^{itz^{2}+izr}\psi(z)\,\mathrm{d}z\right|\leq C|t|^{-\frac{1}{2}}\|\partial_{z}\psi\|_{L_{z}^{1}({\mathbb{R}})}. (2.6)

Moreover, if ψ(z)\psi(z) is supported away from zero, then for all |t|>0|t|>0,

|eitz2+izrψ(z)dz|C|t|32[z2+irz](ψz)Lz1().\left|\int_{{\mathbb{R}}}e^{itz^{2}+izr}\psi(z)\ \mathrm{d}z\right|\leq C|t|^{-\frac{3}{2}}\left\|[\partial_{z}^{2}+ir\partial_{z}]\big{(}\tfrac{\psi}{z}\big{)}\right\|_{L_{z}^{1}({\mathbb{R}})}. (2.7)
Proof.

The bound (2.6) follows from the van der Corput lemma (see e.g. [Ste93, VIII Proposition 2]) by observing that the phase ϕ(z)=z2+zrt\phi(z)=z^{2}+\frac{zr}{t} satisfies |z2ϕ(z)|=2>0|\partial_{z}^{2}\phi(z)|=2>0. The last bound follows by first integrating by parts

eitz2eizrψ(z)dz=12iteitz2z[eizrψ(z)z]dz=12iteitz2+izr[ir+z][ψ(z)z]dz,\int_{{\mathbb{R}}}e^{itz^{2}}e^{izr}\psi(z)\,\mathrm{d}z=-\frac{1}{2it}\int_{\mathbb{R}}e^{itz^{2}}\partial_{z}\left[e^{izr}\frac{\psi(z)}{z}\right]\mathrm{d}z=-\frac{1}{2it}\int_{\mathbb{R}}e^{itz^{2}+izr}[ir+\partial_{z}]\left[\frac{\psi(z)}{z}\right]\mathrm{d}z,

and then invoking the van der Corput lemma. ∎

We will also need the following sharper stationary phase lemma, which may be found in many text on oscillatory integrals with a Fresnel phase.

Lemma 2.2.

Let χ(z)\chi(z) be a smooth, non-negative, even cut-off function such that χ(z)=1\chi(z)=1 for z[1,1]z\in[-1,1] and χ(z)=0\chi(z)=0 for |z|2|z|\geq 2. For r,tr,t\in{\mathbb{R}}, define

Gt(r):=eitz2+izrχ(z2)dz.G_{t}(r):=\int_{\mathbb{R}}e^{itz^{2}+izr}\chi(z^{2})\ \mathrm{d}z. (2.8)

Then there exists C=C(χ(z2)W4,1())>0C=C\big{(}\|\chi(z^{2})\|_{W^{4,1}({\mathbb{R}})}\big{)}>0 such that for any rr\in{\mathbb{R}} and for any |t|>0|t|>0,

|Gt(r)πiteir24t|C|t|32r.\left|G_{t}(r)-\frac{\sqrt{\pi}}{{\sqrt{-it}}}e^{-i\frac{r^{2}}{4t}}\right|\leq C|t|^{-\frac{3}{2}}\langle r\rangle. (2.9)

Moreover, if r1,r20r_{1},r_{2}\geq 0, then

|Gt(r1+r2)πiteir124teir224t|C|t|32r1r2.\left|G_{t}(r_{1}+r_{2})-\frac{\sqrt{\pi}}{{\sqrt{-it}}}e^{-i\frac{r_{1}^{2}}{4t}}e^{-i\frac{r_{2}^{2}}{4t}}\right|\leq C|t|^{-\frac{3}{2}}\langle r_{1}\rangle\langle r_{2}\rangle. (2.10)
Proof.

First, the phase ϕ(z):=z2+zrt\phi(z):=z^{2}+\frac{zr}{t} has a critical point at z=r2tz_{*}=-\frac{r}{2t}\in{\mathbb{R}} with ϕ′′(z)=2>0\phi^{\prime\prime}(z)=2>0. We use Taylor expansion of ϕ(z)\phi(z) and shift the integral by the change of variables zz+zz\mapsto z+z^{*} to obtain

Gt(r)=eitϕ(z)χ(z2)dz=Reitϕ(z)+ϕ′′(z)(zz)2χ(z2)dz=eir24teitz2χ((z+z)2)dz.G_{t}(r)=\int_{{\mathbb{R}}}e^{it\phi(z)}\chi(z^{2})\,\mathrm{d}z=\int_{R}e^{it\phi(z^{*})+\phi^{\prime\prime}(z_{*})(z-z_{*})^{2}}\chi(z^{2})\,\mathrm{d}z=e^{-i\frac{r^{2}}{4t}}\int_{\mathbb{R}}e^{itz^{2}}\chi\big{(}(z+z_{*})^{2}\big{)}\,\mathrm{d}z. (2.11)

Using the Fourier transform of the free Schrödinger group and the Plancherel’s identity, we have

eitz2χ((z+z)2)dz=12iteiξ24tzξ[χ((z+z)2)](ξ)dξ=12itzξ[χ((z+z)2)](ξ)dξ+12it(eiξ24t1)zξ[χ((z+z)2)](ξ)dξ=2π2itχ(z2)+12it(eiξ24t1)eizξ[χ((z+z)2)](ξ)dξ.\begin{split}\int_{\mathbb{R}}e^{itz^{2}}\chi\big{(}(z+z_{*})^{2}\big{)}\,\mathrm{d}z&=\frac{1}{\sqrt{-2it}}\int_{\mathbb{R}}e^{-i\frac{\xi^{2}}{4t}}\mathcal{F}_{z\to\xi}\left[\chi\big{(}(z+z_{*})^{2}\big{)}\right](\xi)\,\mathrm{d}\xi\\ &=\frac{1}{\sqrt{-2it}}\int_{\mathbb{R}}\mathcal{F}_{z\to\xi}\left[\chi\big{(}(z+z_{*})^{2}\big{)}\right](\xi)\,\mathrm{d}\xi\\ &\qquad+\frac{1}{\sqrt{-2it}}\int_{\mathbb{R}}\Big{(}e^{-i\frac{\xi^{2}}{4t}}-1\Big{)}\mathcal{F}_{z\to\xi}\left[\chi\big{(}(z+z_{*})^{2}\big{)}\right](\xi)\,\mathrm{d}\xi\\ &=\frac{\sqrt{2\pi}}{\sqrt{-2it}}\chi(z_{*}^{2})+\frac{1}{\sqrt{-2it}}\int_{\mathbb{R}}\Big{(}e^{-i\frac{\xi^{2}}{4t}}-1\Big{)}e^{iz_{*}\xi}\mathcal{F}\left[\chi\big{(}(z+z_{*})^{2}\big{)}\right](\xi)\,\mathrm{d}\xi.\end{split}

Using the bound |eiξ24t1|C|t|1ξ2|e^{i\frac{\xi^{2}}{4t}}-1|\leq C|t|^{-1}\xi^{2} and the Hölder’s inequality, we bound the remainder term by

|12it(eiξ24t1)eizξ[χ((z+z)2)](ξ)dξ|C|t|32|ξ2[χ(z2)](ξ)|dξC|t|32χ(z2)W4,1()C|t|32.\begin{split}\left|\frac{1}{\sqrt{-2it}}\int_{\mathbb{R}}\Big{(}e^{-i\frac{\xi^{2}}{4t}}-1\Big{)}e^{iz_{*}\xi}\mathcal{F}\left[\chi\big{(}(z+z_{*})^{2}\big{)}\right](\xi)\,\mathrm{d}\xi\right|&\leq C|t|^{-\frac{3}{2}}\int_{\mathbb{R}}|\xi^{2}\mathcal{F}[\chi(z^{2})](\xi)|\,\mathrm{d}\xi\\ &\leq C|t|^{-\frac{3}{2}}\|\chi(z^{2})\|_{W^{4,1}({\mathbb{R}})}\leq C|t|^{-\frac{3}{2}}.\end{split}

Next, we use the fact that |1χ(z2)|C|z||1-\chi(z^{2})|\leq C|z| for all zz\in{\mathbb{R}} and for some C>0C>0 large enough so that

|1χ(z2)|C|z|C|t|1r.|1-\chi(z_{*}^{2})|\leq C|z_{*}|\leq C|t|^{-1}\langle r\rangle. (2.12)

Then (2.9) follows (2.11)–(2.12). Finally, we use the estimate (2.9) to obtain

|Gt(r1+r2)2π2itei(r1r2)24t|C|t|32r1r2C|t|1r1r2.\left|G_{t}(r_{1}+r_{2})-\frac{\sqrt{2\pi}}{{\sqrt{-2it}}}e^{-i\frac{(r_{1}-r_{2})^{2}}{4t}}\right|\leq C|t|^{-\frac{3}{2}}\langle r_{1}-r_{2}\rangle\leq C|t|^{-1}\langle r_{1}\rangle\langle r_{2}\rangle.

Thus, by the triangle inequality and the bound

|ei(r1r2)24teir124teir224t|=|eir124teir224t||eir1r22t1|C|t|1r1r2,\left|e^{-i\frac{(r_{1}-r_{2})^{2}}{4t}}-e^{-i\frac{r_{1}^{2}}{4t}}e^{-i\frac{r_{2}^{2}}{4t}}\right|=\left|e^{-i\frac{r_{1}^{2}}{4t}}e^{-i\frac{r_{2}^{2}}{4t}}\right|\left|e^{i\frac{r_{1}r_{2}}{2t}}-1\right|\leq C|t|^{-1}\langle r_{1}\rangle\langle r_{2}\rangle,

we conclude (2.10). ∎

Next, we prove the analogue of Theorem 1.4 for the free evolution. We emphasize that the free matrix Schrödinger operator 0\mathcal{H}_{0} has threshold resonances 0e¯1=μe¯1\mathcal{H}_{0}\underline{e}_{1}=\mu\underline{e}_{1} and 0e¯2=μe¯2\mathcal{H}_{0}\underline{e}_{2}=-\mu\underline{e}_{2}.

Proposition 2.3.

For any u=(u1,u2)𝒮()×𝒮()\vec{u}=(u_{1},u_{2})\in\mathcal{S}({\mathbb{R}})\times\mathcal{S}({\mathbb{R}}) and for any |t|1|t|\geq 1, we have

eit0Ps+uLx×Lx|t|12uLx1×Lx1,\left\|e^{it\mathcal{H}_{0}}P_{\mathrm{s}}^{+}\vec{u}\,\right\|_{L_{x}^{\infty}\times L_{x}^{\infty}}\lesssim|t|^{-\frac{1}{2}}\|\vec{u}\,\|_{L_{x}^{1}\times L_{x}^{1}}, (2.13)

and

x1(eit0Ps+Ft0)uLx×Lx`|t|32xuLx1×Lx1,\left\|\langle x\rangle^{-1}\left(e^{it\mathcal{H}_{0}}P_{\mathrm{s}}^{+}-F_{t}^{0}\right)\vec{u}\,\right\|_{L_{x}^{\infty}\times L_{x}^{\infty}`}\lesssim|t|^{-\frac{3}{2}}\|\langle x\rangle\vec{u}\,\|_{L_{x}^{1}\times L_{x}^{1}}, (2.14)

where

Ft0(x,y):=eitμ4πiteix24te¯1eiy24te¯1.F_{t}^{0}(x,y):=\frac{e^{it\mu}}{\sqrt{-4\pi it}}e^{-i\frac{x^{2}}{4t}}\underline{e}_{1}e^{-i\frac{y^{2}}{4t}}\underline{e}_{1}^{\top}. (2.15)
Proof.

We first begin by splitting the evolution operator into low and high energy parts333Symbols like χ(0μI)\chi(\mathcal{H}_{0}-\mu I) are only used in a formal way to represent the cut-off χ(z2)\chi(z^{2}) in the zz-integrals, where they arise.:

eit0Ps+(x,y)=eit0χ(0μI)Ps+(x,y)+eit0(1χ(0μI))Ps+(x,y)=eitμ2πeitz2+iz|xy|χ(z2)dze¯11+eitμ2πeitz2+iz|xy|(1χ(z2))dze¯11,\begin{split}e^{it\mathcal{H}_{0}}P_{\mathrm{s}}^{+}(x,y)&=e^{it\mathcal{H}_{0}}\chi(\mathcal{H}_{0}-\mu I)P_{\mathrm{s}}^{+}(x,y)+e^{it\mathcal{H}_{0}}(1-\chi(\mathcal{H}_{0}-\mu I))P_{\mathrm{s}}^{+}(x,y)\\ &=\frac{e^{it\mu}}{2\pi}\int_{\mathbb{R}}e^{itz^{2}+iz|x-y|}\chi(z^{2})\,\mathrm{d}z\underline{e}_{11}+\frac{e^{it\mu}}{2\pi}\int_{\mathbb{R}}e^{itz^{2}+iz|x-y|}(1-\chi(z^{2}))\,\mathrm{d}z\underline{e}_{11},\end{split} (2.16)

where χ(z)\chi(z) is a standard smooth, even, non-negative cut-off function satisfying χ(z)=1\chi(z)=1 for |z|1|z|\leq 1 and χ(z)=0\chi(z)=0 for |z|2|z|\geq 2.

In the high energy part in (2.16), following the ideas from [GS04] [Gol07], we prove the estimate

|eitz2+iz|xy|(1χ(z2))𝑑z|min{|t|12,|t|32xy}.\left|\int_{\mathbb{R}}e^{itz^{2}+iz|x-y|}(1-\chi(z^{2}))dz\right|\lesssim\min\{|t|^{-\frac{1}{2}},|t|^{-\frac{3}{2}}\langle x\rangle\langle y\rangle\}. (2.17)

For a more rigorous treatment, we instead use a truncated cutoff χL(z)=(1χ(z2))χ(z/L)\chi_{L}(z)=(1-\chi(z^{2}))\chi(z/L), where L1L\geq 1, and we prove the uniform estimate

supL1|eitz2+iz|xy|χL(z)dz|Cmin{|t|12,|t|32xy},\sup_{L\geq 1}\left|\int_{\mathbb{R}}e^{itz^{2}+iz|x-y|}\chi_{L}(z)\,\mathrm{d}z\right|\leq C\min\{|t|^{-\frac{1}{2}},|t|^{-\frac{3}{2}}\langle x\rangle\langle y\rangle\}, (2.18)

with a constant C>0C>0 independent of LL. This estimate will imply (2.17). Indeed for any |t|>0|t|>0, by the Plancherel’s identity, we have

supa|eitz2+iazχL(z)dz|=supa|1[eitz2+iaz](ξ)[χL(z)](ξ)dξ|C|t|12[χL]Lξ1().\sup_{a\in{\mathbb{R}}}\left|\int_{\mathbb{R}}e^{itz^{2}+iaz}\chi_{L}(z)\,\mathrm{d}z\right|=\sup_{a\in{\mathbb{R}}}\left|\int_{{\mathbb{R}}}\mathcal{F}^{-1}{[e^{itz^{2}+iaz}]}(\xi)\mathcal{F}{[\chi_{L}(z)]}(\xi)\,\mathrm{d}\xi\right|\leq C|t|^{-\frac{1}{2}}\|\mathcal{F}{[\chi_{L}]}\|_{L_{\xi}^{1}({\mathbb{R}})}.

Here, we use that the Fourier transform of the tempered distribution eitz2+iaze^{itz^{2}+iaz} has |t|12|t|^{-\frac{1}{2}} decay. Using the definition of χL\chi_{L}, the scaling properties of the Fourier transform, and Young’s convolution inequality, we obtain

[χL]Lξ1()[χ(z/L)]Lξ1()+[χ(z/L)]Lξ1()[χ(z2)]Lξ1()CL[χ](Lξ)Lξ1()=C[χ](ξ)Lξ1()CχW2,1()1.\begin{split}\|\mathcal{F}{[\chi_{L}]}\|_{L_{\xi}^{1}({\mathbb{R}})}&\leq\|\mathcal{F}[\chi(z/L)]\|_{L_{\xi}^{1}({\mathbb{R}})}+\|\mathcal{F}[\chi(z/L)]\|_{L_{\xi}^{1}({\mathbb{R}})}\|\mathcal{F}[\chi(z^{2})]\|_{L_{\xi}^{1}({\mathbb{R}})}\\ &\leq C\|L\mathcal{F}[\chi](L\xi)\|_{L_{\xi}^{1}({\mathbb{R}})}=C\|\mathcal{F}[\chi](\xi)\|_{L_{\xi}^{1}({\mathbb{R}})}\leq C\|\chi\|_{W^{2,1}({\mathbb{R}})}\lesssim 1.\end{split} (2.19)

For the high-energy weighted dispersive estimate, we use integration by parts to find that

|eitz2eiz|xy|)χL(z)dz|C|t|1|eitz2z(eiz|xy|z1χL(z))dz|.\begin{split}&\left|\int_{\mathbb{R}}e^{itz^{2}}e^{iz|x-y|)}\chi_{L}(z)\,\mathrm{d}z\right|\leq C|t|^{-1}\left|\int_{{\mathbb{R}}}e^{itz^{2}}\partial_{z}\left(e^{iz|x-y|}z^{-1}\chi_{L}(z)\right)\,\mathrm{d}z\right|.\end{split}

When the derivative falls onto eiz|xy|e^{iz|x-y|}, the weights xy\langle x\rangle\langle y\rangle appear, whereas the term z1χL(z)z^{-1}\chi_{L}(z) is smooth since χL\chi_{L} is compactly supported away from the interval [1,1][-1,1]. By following the previous argument, we conclude the 𝒪(|t|32xy)\mathcal{O}(|t|^{-\frac{3}{2}}\langle x\rangle\langle y\rangle) bound for (2.18) in the high-energy regime.

Next we turn to the low-energy estimates. For the low-energy unweighted estimate, we employ Lemma 2.1 to obtain

|eitz2+iz|xy|χ(z2)dz|C|t|12zχ(z2)L1()C|t|12.\left|\int_{{\mathbb{R}}}e^{itz^{2}+iz|x-y|}\chi(z^{2})\,\mathrm{d}z\right|\leq C|t|^{-\frac{1}{2}}\|\partial_{z}\chi(z^{2})\|_{L^{1}({\mathbb{R}})}\leq C|t|^{-\frac{1}{2}}. (2.20)

On the other hand, for the low-energy weighted estimate, we observe that by Lemma 2.2,

|eitz2+iz|xy|χ(z2)dz2π2iteix24teiy24t|C|t|32xy.\left|\int_{{\mathbb{R}}}e^{itz^{2}+iz|x-y|}\chi(z^{2})\,\mathrm{d}z-\frac{\sqrt{2\pi}}{\sqrt{-2it}}e^{-i\frac{x^{2}}{4t}}e^{-i\frac{y^{2}}{4t}}\right|\leq C|t|^{-\frac{3}{2}}\langle x\rangle\langle y\rangle.

Hence, using that e¯11=e¯1e¯1\underline{e}_{11}=\underline{e}_{1}\underline{e}_{1}^{\top}, we arrive at the kernel estimate

|eit0χ(0μ)Ps+(x,y)Ft0(x,y)|C|t|32xy,\left|e^{it\mathcal{H}_{0}}\chi(\mathcal{H}_{0}-\mu)P_{\mathrm{s}}^{+}(x,y)-F_{t}^{0}(x,y)\right|\leq C|t|^{-\frac{3}{2}}\langle x\rangle\langle y\rangle, (2.21)

where Ft0F_{t}^{0} is given by (2.15). Thus, by combining the high energy bounds (2.17) and the low energy bounds (2.20) - (2.21), we conclude the dispersive estimates (2.13) and (2.14). ∎

3. Symmetric resolvent identity

By assumption (A1), we can factorize the matrix potential

𝒱=σ3vv=v1v2,\mathcal{V}=-\sigma_{3}vv=v_{1}v_{2}, (3.1)

with

v1=σ3v:=[abba]andv2=v:=[abba],v_{1}=-\sigma_{3}v:=\begin{bmatrix}-a&-b\\ b&a\end{bmatrix}{\ \ \text{and}\ \ }v_{2}=v:=\begin{bmatrix}a&b\\ b&a\end{bmatrix},

where

a:=12(V1+V2+V1V2)andb:=12(V1+V2V1V2).a:=\frac{1}{2}\Big{(}\sqrt{V_{1}+V_{2}}+\sqrt{V_{1}-V_{2}}\Big{)}{\ \ \text{and}\ \ }b:=\frac{1}{2}\Big{(}\sqrt{V_{1}+V_{2}}-\sqrt{V_{1}-V_{2}}\Big{)}.

It will be helpful in later sections to keep in mind that

V1=a2+b2,V2=2ab.V_{1}=a^{2}+b^{2},\quad V_{2}=2ab. (3.2)

We denote the resolvent of =0+𝒱\mathcal{H}=\mathcal{H}_{0}+\mathcal{V} by (z)1(\mathcal{H}-z)^{-1} for zρ()z\in\rho(\mathcal{H}). The resolvent identity states that

(z)1=(I+(0z)1𝒱)1(0z)1,zρ(0)ρ().(\mathcal{H}-z)^{-1}=\big{(}I+(\mathcal{H}_{0}-z)^{-1}\mathcal{V}\big{)}^{-1}(\mathcal{H}_{0}-z)^{-1},\quad\forall z\in\rho(\mathcal{H}_{0})\cap\rho(\mathcal{H}).

This identity was used in [ES06] to establish that there is a limiting absorption principle for the resolvent of \mathcal{H} on the semi-axes (,μ)(μ,)(-\infty,-\mu)\cup(\mu,\infty) in the weighted L2L^{2}-spaces XσXσX_{\sigma}\rightarrow X_{-\sigma}, σ>12\sigma>\frac{1}{2}. Note that the lemma below applies in any spatial dimension.

Lemma 3.1.

([ES06, Lemma 4-Corollary 6], see also the proof in [KS06, Lemma 6.8]) Suppose assumptions (A1) – (A4) hold. Then, the following holds.

  1. (1)

    For σ>12\sigma>\frac{1}{2}, and |λ|>μ|\lambda|>\mu, the operator

    (0(λ±i0))1𝒱:XσXσ\big{(}\mathcal{H}_{0}-(\lambda\pm i0)\big{)}^{-1}\mathcal{V}:X_{-\sigma}\to X_{-\sigma} (3.3)

    is compact and I+(0(λ±i0))1𝒱I+\big{(}\mathcal{H}_{0}-(\lambda\pm i0)\big{)}^{-1}\mathcal{V} is boundedly invertible on XσX_{-\sigma}.

  2. (2)

    For σ>12\sigma>\frac{1}{2} and λ0>μ\lambda_{0}>\mu arbitrary, we have

    sup|λ|λ0,ε>0|λ|12((λ±iε))1XσXσ<.\sup_{|\lambda|\geq\lambda_{0},{\varepsilon}>0}|\lambda|^{\frac{1}{2}}\left\|\big{(}\mathcal{H}-(\lambda\pm i{\varepsilon})\big{)}^{-1}\right\|_{X_{\sigma}\to X_{-\sigma}}<\infty. (3.4)
  3. (3)

    For |λ|>μ|\lambda|>\mu, define

    ((λ±i0))1:=(I+(0(λ±i0))1𝒱)1(0(λ±i0))1.\big{(}\mathcal{H}-(\lambda\pm i0)\big{)}^{-1}:=\Big{(}I+\big{(}\mathcal{H}_{0}-(\lambda\pm i0)\big{)}^{-1}\mathcal{V}\Big{)}^{-1}\big{(}\mathcal{H}_{0}-(\lambda\pm i0)\big{)}^{-1}. (3.5)

    Then, as ε0{\varepsilon}\searrow 0,

    ((λ±iε))1((λ±i0))1XσXσ0\left\|\big{(}\mathcal{H}-(\lambda\pm i{\varepsilon})\big{)}^{-1}-\big{(}\mathcal{H}-(\lambda\pm i0)\big{)}^{-1}\right\|_{X_{\sigma}\to X_{-\sigma}}\longrightarrow 0 (3.6)

    for any σ>12\sigma>\frac{1}{2}.

We recall the following spectral representation of eite^{it\mathcal{H}} from [ES06].

Lemma 3.2.

([ES06, Lemma 12]) Under assumptions (A1) – (A6), there is the representation

eit=12πi|λ|μeitλ[((λ+i0))1((λi0))1]dλ+jeitPzj,e^{it\mathcal{H}}=\frac{1}{2\pi i}\int_{|\lambda|\geq\mu}e^{it\lambda}\left[\big{(}\mathcal{H}-(\lambda+i0)\big{)}^{-1}-\big{(}\mathcal{H}-(\lambda-i0)\big{)}^{-1}\right]\,\mathrm{d}\lambda+\sum_{j}e^{it\mathcal{H}}P_{z_{j}}, (3.7)

where the sum runs over the entire discrete spectrum and PzjP_{z_{j}} is the Riesz projection corresponding to the eigenvalue zjz_{j}. The formula (3.7) and the convergence of the integral are to be understood in the sense that if ϕ,ψ[W2,2()×W2,2()][x1L2()×x1L2()]\phi,\psi\in[W^{2,2}({\mathbb{R}})\times W^{2,2}({\mathbb{R}})]\cap[\langle x\rangle^{-1-}L^{2}({\mathbb{R}})\times\langle x\rangle^{-1-}L^{2}({\mathbb{R}})], then

eitϕ,ψ=limR12πiR|λ|μeitλ[((λ+i0))1((λi0))1]ϕ,ψdλ+jeitPzjϕ,ψ,\begin{split}\langle e^{it\mathcal{H}}\phi,\psi\rangle&=\lim_{R\to\infty}\frac{1}{2\pi i}\int_{R\geq|\lambda|\geq\mu}e^{it\lambda}\left\langle\left[\big{(}\mathcal{H}-(\lambda+i0)\big{)}^{-1}-\big{(}\mathcal{H}-(\lambda-i0)\big{)}^{-1}\right]\phi,\psi\right\rangle\,\mathrm{d}\lambda\\ &\quad+\sum_{j}\langle e^{it\mathcal{H}}P_{z_{j}}\phi,\psi\rangle,\end{split} (3.8)

for all tt\in{\mathbb{R}}.

We write Ps=Ps++PsP_{\mathrm{s}}=P_{\mathrm{s}}^{+}+P_{\mathrm{s}}^{-}, where the signs ±\pm refer to the positive and negative halves of the essential spectrum (,μ][μ,)(-\infty,-\mu]\cup[\mu,\infty). In the following sections, we will focus on the analysis on the positive semi-axis part of the essential spectrum. We can treat the negative semi-axis of the essential spectrum by taking advantage of the symmetry properties of \mathcal{H}, see Remark 3.3 below. In view of the spectral representation of eite^{it\mathcal{H}} from Lemma 3.2, we use the change of variables λλ=μ+z2\lambda\mapsto\lambda=\mu+z^{2} with 0<z<0<z<\infty to write

eitPs+=eitμπi0eitz2z[((μ+z2+i0))1((μ+z2i0))1]dz.e^{it\mathcal{H}}P_{\mathrm{s}}^{+}=\frac{e^{it\mu}}{\pi i}\int_{0}^{\infty}e^{itz^{2}}z\left[\big{(}\mathcal{H}-(\mu+z^{2}+i0))^{-1}-(\mathcal{H}-(\mu+z^{2}-i0)\big{)}^{-1}\right]\,\mathrm{d}z.

For the upcoming dispersive estimates, it is convenient to first open up the domain of integration for the above integral to the entire real line by means of analytic continuation for the perturbed resolvent. Following the framework of Section 5 in [ES06], we introduce the operator

(z):=((μ+z2+i0))1,for z>0,(z):=((μ+z2i0))1=((μ+z2+i0))1¯,for z<0,\begin{split}&\mathcal{R}(z):=(\mathcal{H}-(\mu+z^{2}+i0))^{-1},\quad\text{for $z>0$},\\ &\mathcal{R}(z):=(\mathcal{H}-(\mu+z^{2}-i0))^{-1}=\overline{(\mathcal{H}-(\mu+z^{2}+i0))^{-1}},\quad\text{for $z<0$},\end{split} (3.9)

so that

eitPs+=eitμπieitz2z(z)dz.e^{it\mathcal{H}}P_{\mathrm{s}}^{+}=\frac{e^{it\mu}}{\pi i}\int_{\mathbb{R}}e^{itz^{2}}z\mathcal{R}(z)\,\mathrm{d}z. (3.10)

Here, the integral should be understood in the principal value sense due to the pole associated with the resolvent (z)\mathcal{R}(z) at the origin. We also set

0(z):=(0(μ+z2+i0))1,for z>0,0(z):=(0(μ+z2+i0))1¯,for z<0.\begin{split}&\mathcal{R}_{0}(z):=(\mathcal{H}_{0}-(\mu+z^{2}+i0))^{-1},\quad\text{for $z>0$},\\ &\mathcal{R}_{0}(z):=\overline{(\mathcal{H}_{0}-(\mu+z^{2}+i0))^{-1}},\quad\text{for $z<0$}.\end{split} (3.11)

In particular, with this definition, we have by (2.4) for all z{0}z\in{\mathbb{R}}\setminus\{0\} that

0(z)(x,y)=(0(μ+z2+i0))1(x,y)=[ieiz|xy|2z00ez2+2μ|xy|2z2+2μ].\mathcal{R}_{0}(z)(x,y)=(\mathcal{H}_{0}-(\mu+z^{2}+i0))^{-1}(x,y)=\begin{bmatrix}\frac{ie^{iz|x-y|}}{2z}&0\\ 0&-\frac{e^{-\sqrt{z^{2}+2\mu}|x-y|}}{2\sqrt{z^{2}+2\mu}}\end{bmatrix}. (3.12)

As in [ES06], we employ the symmetric resolvent identity

(z)=0(z)0(z)v1(M(z))1v20(z),\mathcal{R}(z)=\mathcal{R}_{0}(z)-\mathcal{R}_{0}(z)v_{1}(M(z))^{-1}v_{2}\mathcal{R}_{0}(z), (3.13)

where

M(z)=I+v20(z)v1,z{0}.M(z)=I+v_{2}\mathcal{R}_{0}(z)v_{1},\quad z\in{\mathbb{R}}\setminus\{0\}. (3.14)

By inserting the above identity, one checks that

eitPs+=eitμπieitz2z0(z)dzeitμπieitz2z0(z)v1(M(z))1v20(z)dz.e^{it\mathcal{H}}P_{\mathrm{s}}^{+}=\frac{e^{it\mu}}{\pi i}\int_{{\mathbb{R}}}e^{itz^{2}}z\mathcal{R}_{0}(z)\,\mathrm{d}z-\frac{e^{it\mu}}{\pi i}\int_{{\mathbb{R}}}e^{itz^{2}}z\mathcal{R}_{0}(z)v_{1}\big{(}M(z)\big{)}^{-1}v_{2}\mathcal{R}_{0}(z)\,\mathrm{d}z. (3.15)

In the next section, we will investigate the invertibility of the matrix operator M(z)M(z) near the origin. We give the following remark for the evolution operator in the negative part of the essential spectrum.

Remark 3.3.

Using the identities

=σ1σ1,𝒱=σ1𝒱σ1,\mathcal{H}=-\sigma_{1}\mathcal{H}\sigma_{1},\quad\mathcal{V}=-\sigma_{1}\mathcal{V}\sigma_{1}, (3.16)

we infer that

eitPs=σ1eitPs+σ1.e^{it\mathcal{H}}P_{\mathrm{s}}^{-}=\sigma_{1}e^{-it\mathcal{H}}P_{\mathrm{s}}^{+}\sigma_{1}. (3.17)

Furthermore, since these identities also hold for 0\mathcal{H}_{0}, the analogue of Proposition 2.3 for the weighted estimate of the free evolution eit0Pse^{it\mathcal{H}_{0}}P_{\mathrm{s}}^{-} is given by

x1(eit0PsF~t0)uLxC|t|32xuLx1,|t|1,\left\|\langle x\rangle^{-1}\left(e^{it\mathcal{H}_{0}}P_{\mathrm{s}}^{-}-\widetilde{F}_{t}^{0}\right){\vec{u}}\,\right\|_{L_{x}^{\infty}}\leq C|t|^{-\frac{3}{2}}\left\|\langle x\rangle{\vec{u}}\,\right\|_{L_{x}^{1}},\quad|t|\geq 1, (3.18)

where

F~t0(x,y):=eitμ4πiteix24te¯2eiy24te¯2.\widetilde{F}_{t}^{0}(x,y):=\frac{e^{-it\mu}}{\sqrt{4\pi it}}e^{i\frac{x^{2}}{4t}}\underline{e}_{2}e^{i\frac{y^{2}}{4t}}\underline{e}_{2}^{\top}. (3.19)

Note that F~t0=σ1Ft0σ1\widetilde{F}_{t}^{0}=\sigma_{1}F_{-t}^{0}\sigma_{1}.

4. Laurent expansion of the resolvent near the threshold

In this section we study asymptotic expansions of the perturbed resolvent operators near the thresholds of the essential spectrum, closely following the framework of the seminal paper [JN01] for the scalar Schrödinger operators H=x2+VH=-\partial_{x}^{2}+V on the real line. As specified in the introduction, we are interested in the irregular case, where the matrix Schrödinger operator \mathcal{H} exhibits a threshold resonance. See Definition 4.4 for a precise definition. This means that there exist globally bounded non-trivial solutions of Ψ=±μΨ\mathcal{H}\Psi=\pm\mu\Psi. In this context, we mention that the threshold regularity can also be characterized by the scattering theory introduced by [BP95]; see Lemma 5.20 of [KS06]. We begin with the terminology used in [Sch05].

Definition 4.1 (Absolutely bounded operators).

We say an operator A:L2()×L2()L2()×L2()A\colon L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})\to L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}}) with an integral kernel A(x,y)2×2A(x,y)\in{\mathbb{C}}^{2\times 2} is absolutely bounded if the operator with the kernel |A(x,y)|:=(|A(x,y)i,j|)i,j=122×2|A(x,y)|:=(|A(x,y)_{i,j}|)_{i,j=1}^{2}\in{\mathbb{R}}^{2\times 2} is bounded from L2()×L2()L2()×L2()L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})\to L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}}). In particular, Hilbert-Schmidt and finite rank operators are absolutely bounded.

To investigate the asymptotic expansions of the operator M(z)M(z) (c.f. (3.14)), we start with the following Taylor expansions of the free resolvent around the origin z=0z=0.

Lemma 4.2.

Let z0:=min{1,2μ}z_{0}:=\min\{1,\sqrt{2\mu}\}. For any 0<|z|<z00<|z|<z_{0}, we have the following expansion

0(z)(x,y)=i2ze¯11+𝒢0(x,y)+z𝒢1(x,y)+E(z)(x,y)\begin{split}\mathcal{R}_{0}(z)(x,y)&=\frac{i}{2z}\underline{e}_{11}+\mathcal{G}_{0}(x,y)+z\mathcal{G}_{1}(x,y)+E(z)(x,y)\\ \end{split} (4.1)

where

𝒢0(x,y)\displaystyle\mathcal{G}_{0}(x,y) :=[|xy|200e2μ|xy|22μ],\displaystyle:=\begin{bmatrix}-\frac{|x-y|}{2}&0\\ 0&-\frac{e^{-\sqrt{2\mu}|x-y|}}{2\sqrt{2\mu}}\end{bmatrix}, (4.2)
𝒢1(x,y)\displaystyle\mathcal{G}_{1}(x,y) :=[|xy|24i000],\displaystyle:=\begin{bmatrix}\frac{|x-y|^{2}}{4i}&0\\ 0&0\end{bmatrix}, (4.3)

and E(z)E(z) is an error term which satisfies the estimate

|z|k|zkE(z)(x,y)|Cμ,k|z|2x3+ky3+k,k=0,1,2,|z|^{k}\,|\partial_{z}^{k}E(z)(x,y)|\leq C_{\mu,k}\,|z|^{2}\langle x\rangle^{3+k}\langle y\rangle^{3+k},\quad\forall\ k=0,1,2, (4.4)

for any |z|<z0|z|<z_{0}.

Proof.

Recall from (3.12) that

0(z)(x,y)=[ieiz|xy|2z00ez2+2μ|xy|2z2+2μ].\mathcal{R}_{0}(z)(x,y)=\begin{bmatrix}\frac{ie^{iz|x-y|}}{2z}&0\\ 0&-\frac{e^{-\sqrt{z^{2}+2\mu}|x-y|}}{2\sqrt{z^{2}+2\mu}}\end{bmatrix}.

For 0<|z|<10<|z|<1, we have the Laurent expansion

ieiz|xy|2z=i2z+|xy|2+|xy|24iz+r1(z,|xy|),\frac{ie^{iz|x-y|}}{2z}=\frac{i}{2z}+\frac{-|x-y|}{2}+\frac{|x-y|^{2}}{4i}z+r_{1}(z,|x-y|), (4.5)

where the remainder term is

r1(z,|xy|):=i2zr~1(z,|xy|),r~1(z,|xy|):=(iz|xy|)32!01eisz|xy|(1s)2ds.r_{1}(z,|x-y|):=\frac{i}{2z}{\tilde{r}}_{1}(z,|x-y|),\quad{\tilde{r}}_{1}(z,|x-y|):=\frac{(iz|x-y|)^{3}}{2!}\int_{0}^{1}e^{isz|x-y|}(1-s)^{2}\,\mathrm{d}s.

By direct computation, for any x,yx,y\in{\mathbb{R}} and for any |z|<1|z|<1, we have the estimate

|z|k|zkr1(z,|xy|)||z|2x3+ky3+k,k=0,1,2.|z|^{k}\,|\partial_{z}^{k}\,r_{1}(z,|x-y|)|\lesssim|z|^{2}\langle x\rangle^{3+k}\langle y\rangle^{3+k},\quad k=0,1,2. (4.6)

In the lower component of the resolvent kernel, for |z|<2μ|z|<2\mu, we have the Taylor expansion

ez2+2μ|xy|2z2+2μ=e2μ|xy|22μ+r2(z,|xy|),-\frac{e^{-\sqrt{z^{2}+2\mu}|x-y|}}{2\sqrt{z^{2}+2\mu}}=-\frac{e^{-\sqrt{2\mu}|x-y|}}{2\sqrt{2\mu}}+r_{2}(z,|x-y|), (4.7)

where we denote the remainder term by

r2(z,|xy|):=z22!01(1s)(z2gμ)(sz,|xy|)ds,gμ(z,|xy|):=ez2+2μ|xy|2z2+2μ.r_{2}(z,|x-y|):=\frac{z^{2}}{2!}\int_{0}^{1}(1-s)(\partial_{z}^{2}\,g_{\mu})(sz,|x-y|)\,\mathrm{d}s,\quad g_{\mu}(z,|x-y|):=-\frac{e^{-\sqrt{z^{2}+2\mu}|x-y|}}{2\sqrt{z^{2}+2\mu}}.

Using the fact that for any η\eta\in{\mathbb{R}}, η:=(1+η2)12\langle\eta\rangle:=(1+\eta^{2})^{\frac{1}{2}}, one has the bounds

|ηkη1|Ckη1k|ηkη|Ckη1k,k=0,1,2,,|\partial_{\eta}^{k}\langle\eta\rangle^{-1}|\leq C_{k}\langle\eta\rangle^{-1-k}\qquad|\partial_{\eta}^{k}\langle\eta\rangle|\leq C_{k}\langle\eta\rangle^{1-k},\quad k=0,1,2,\ldots,

it follows that all derivatives of z2+2μ\sqrt{z^{2}+2\mu} and 2(z2+2μ)122(z^{2}+2\mu)^{-\frac{1}{2}} are uniformly bounded in zz up to a constant depending only on μ\mu and the number of derivatives. Therefore, by the Leibniz formula, we have the estimate

supz|zkgμ(z,|xy|)|Cμ,kxkyk,k=0,1,,4,\sup_{z\in{\mathbb{R}}}\left|\partial_{z}^{k}\,g_{\mu}(z,|x-y|)\right|\leq C_{\mu,k}\langle x\rangle^{k}\langle y\rangle^{k},\quad k=0,1,\ldots,4,

which in turn implies that

|z|k|zkr2(z,|xy|)||z|2x2+ky2+k,k=0,1,2.|z|^{k}\left|\partial_{z}^{k}\,r_{2}(z,|x-y|)\right|\lesssim|z|^{2}\langle x\rangle^{2+k}\langle y\rangle^{2+k},\quad k=0,1,2. (4.8)

Thus, by using (4.6) and (4.8), the error term given by

E(z)(x,y):=[r1(z,|xy|)00r2(z,|xy|)]E(z)(x,y):=\begin{bmatrix}r_{1}(z,|x-y|)&0\\ 0&r_{2}(z,|x-y|)\end{bmatrix} (4.9)

satisfies (4.4) as claimed. ∎

We insert the above asymptotic expansion into the operator M(z)=I+v20(z)v1M(z)=I+v_{2}\mathcal{R}_{0}(z)v_{1}. First, we have the transfer operator TT on L2()×L2()L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}}) with a kernel given by

T(x,y)=I+v2(x)𝒢0(x,y)v1(y).T(x,y)=I+v_{2}(x)\mathcal{G}_{0}(x,y)v_{1}(y). (4.10)

Note that TT is self-adjoint because

(v2𝒢0v1)=v1𝒢0v2=(vσ3)𝒢0v=v𝒢0(σ3v)=v2𝒢0v1.(v_{2}\mathcal{G}_{0}v_{1})^{*}=v_{1}^{*}\mathcal{G}_{0}v_{2}=(-v\sigma_{3})\mathcal{G}_{0}v=v\mathcal{G}_{0}(-\sigma_{3}v)=v_{2}\mathcal{G}_{0}v_{1}.

Since the potentials v1v_{1} and v2v_{2} have exponential decay by assumption (A3), it follows that v2𝒢0v1v_{2}\mathcal{G}_{0}v_{1} is a Hilbert-Schmidt operator on L2()×L2()L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}}). Hence, TT is a compact perturbation of the identity, and therefore the dimension of ker(T)\ker(T) is finite by the Fredholm alternative. Recalling the formulas for v1v_{1} and v2v_{2} from (3.1), we have the identity

v2e¯11v1=[a0b0][ab00]=[ab][ab].v_{2}\underline{e}_{11}v_{1}=-\begin{bmatrix}a&0\\ b&0\end{bmatrix}\begin{bmatrix}a&b\\ 0&0\end{bmatrix}=-\begin{bmatrix}a\\ b\end{bmatrix}\begin{bmatrix}a&b\end{bmatrix}. (4.11)

Next, we define the orthogonal projection onto the span of the vector (a,b)L2()×L2()(a,b)^{\top}\in L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}}) by

P[f1f2](x):=(a(y)f1(y)+b(y)f2(y))dya2+b2L1()[a(x)b(x)]=1V1L1()(a,b),f[a(x)b(x)].P\begin{bmatrix}f_{1}\\ f_{2}\end{bmatrix}(x):=\frac{\int_{{\mathbb{R}}}(a(y)f_{1}(y)+b(y)f_{2}(y))\,\mathrm{d}y}{\|a^{2}+b^{2}\|_{L^{1}({\mathbb{R}})}}\begin{bmatrix}a(x)\\ b(x)\end{bmatrix}=\frac{1}{\|V_{1}\|_{L^{1}({\mathbb{R}})}}\langle(a,b)^{\top},\vec{f}\,\rangle\begin{bmatrix}a(x)\\ b(x)\end{bmatrix}. (4.12)

Note that we use the identity (3.2) above. From (3.14), the contribution of the singular term i2ze¯11\frac{i}{2z}\underline{e}_{11} of 0(z)\mathcal{R}_{0}(z) to M(z)M(z) will be associated to the following integral operator with the kernel

i2zv2(x)e¯11v1(y)=i2z[a(x)b(x)][a(y)b(y)]=:g(z)P(x,y),\frac{i}{2z}v_{2}(x)\underline{e}_{11}v_{1}(y)=\frac{-i}{2z}\begin{bmatrix}a(x)\\ b(x)\end{bmatrix}\begin{bmatrix}a(y)&b(y)\end{bmatrix}=:g(z)P(x,y), (4.13)

where

g(z):=i2zV1L1().g(z):=-\frac{i}{2z}\|V_{1}\|_{L^{1}({\mathbb{R}})}. (4.14)

Lastly, we denote the orthogonal projection to the complement of the span of (a,b)(a,b)^{\top} by

Q:=IP.Q:=I-P. (4.15)

In summary, we have the following proposition.

Proposition 4.3.

Suppose |a(x)|,|b(x)|x5.5|a(x)|,|b(x)|\lesssim\langle x\rangle^{-5.5-}, and let z0:=min{1,2μ}z_{0}:=\min\{1,\sqrt{2\mu}\}. Then, for any 0<|z|<z00<|z|<z_{0}, we have

M(z)=g(z)P+T+zM1+2(z),\begin{split}M(z)&=g(z)P+T+zM_{1}+\mathcal{M}_{2}(z),\end{split} (4.16)

where M1M_{1} and 2(z)\mathcal{M}_{2}(z) are Hilbert-Schmidt operators on L2()×L2()L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}}) defined by

M1(x,y)\displaystyle M_{1}(x,y) :=v2(x)𝒢1(x,y)v1(y)=|xy|24i[a(x)b(x)][a(y)b(y)],\displaystyle:=v_{2}(x)\mathcal{G}_{1}(x,y)v_{1}(y)=\frac{|x-y|^{2}}{4i}\begin{bmatrix}a(x)\\ b(x)\end{bmatrix}\begin{bmatrix}a(y)&b(y)\end{bmatrix}, (4.17)
2(z)(x,y)\displaystyle\mathcal{M}_{2}(z)(x,y) :=v2(x)E(z)(x,y)v1(y),\displaystyle:=v_{2}(x)E(z)(x,y)v_{1}(y), (4.18)

with G1G_{1} and E(z)E(z) defined in Lemma 4.2. Moreover, the error term 2(z)\mathcal{M}_{2}(z) and its derivatives satisfy the absolute bound

|z|k|zk2(z)|L2()×L2()L2()×L2()|z|2,k=0,1,2,|z|^{k}\left\||\partial_{z}^{k}\mathcal{M}_{2}(z)|\right\|_{L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})\to L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})}\lesssim|z|^{2},\quad k=0,1,2, (4.19)

for all |z|<z0|z|<z_{0}.

Proof.

The identity on the right of (4.17) follows from (4.11). We recall that operators of the following type

U(x)xkykW(y)U(x)\langle x\rangle^{k}\langle y\rangle^{k}W(y)

are Hilbert-Schmidt operators on L2()L^{2}({\mathbb{R}}) whenever UU and WW are smooth potentials with polynomial decay |U(x)|,|W(x)|xk12|U(x)|,|W(x)|\lesssim\langle x\rangle^{-k-\frac{1}{2}-}, for kk\in{\mathbb{N}}. Hence, under the assumptions on a(x)a(x) and b(x)b(x), and using the fact that

|𝒢1(x,y)||xy|2x2y2,|\mathcal{G}_{1}(x,y)|\lesssim|x-y|^{2}\leq\langle x\rangle^{2}\langle y\rangle^{2},

it follows that M1M_{1} is Hilbert-Schmidt. The same argument can be applied to the error term 2(z)\mathcal{M}_{2}(z) and its derivatives using the remainder estimates in (4.4) and we are done. ∎

The next definition characterizes the regularity of the endpoint μ\mu of the essential spectrum.

Definition 4.4.
  1. (1)

    We say that the threshold μ\mu is a regular point of the spectrum of \mathcal{H} provided that the operator QTQQTQ is invertible on the subspace Q(L2()×L2())Q(L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})).

  2. (2)

    Suppose μ\mu is not a regular point. Let S1S_{1} be the Riesz projection onto the kernel of QTQQTQ, and we define D0=(Q(T+S1)Q)1D_{0}=(Q(T+S_{1})Q)^{-1}. Note that QD0QQD_{0}Q is an absolutely bounded operator on L2()×L2()L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}}). The proof for this follows from Lemma 8 of [Sch05] with minor changes. See also [GG15, Lemma 2.7].

Note that since we impose symmetry assumptions on the potential 𝒱\mathcal{V}, the thresholds μ\mu and μ-\mu are either both regular or irregular. The invertibility of QTQQTQ is related to the absence of distributional L()×L()L^{\infty}({\mathbb{R}})\times L^{\infty}({\mathbb{R}}) solutions to Ψ=μΨ\mathcal{H}\Psi=\mu\Psi. The following lemma establishes the equivalent definitions. See [JN01, Lemma 5.4] for the analogue in the scalar case.

Lemma 4.5.

Suppose assumptions (A1) – (A5) hold. Then the following holds.

  1. (1)

    Let ΦS1(L2()×L2()){0}\Phi\in S_{1}(L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}}))\setminus\{0\}. If Ψ=(Ψ1,Ψ2)\Psi=(\Psi_{1},\Psi_{2})^{\top} is defined by

    Ψ(x):=𝒢0[v1Φ](x)+c0e¯1,\Psi(x):=-\mathcal{G}_{0}[v_{1}\Phi](x)+c_{0}\underline{e}_{1}, (4.20)

    with

    c0=(a,b),TΦV1L1(),c_{0}=\frac{\langle(a,b)^{\top},T\Phi\rangle}{\|V_{1}\|_{L^{1}({\mathbb{R}})}}, (4.21)

    then

    Φ=v2Ψ,\Phi=v_{2}\Psi, (4.22)

    and ΨL()×L()\Psi\in L^{\infty}({\mathbb{R}})\times L^{\infty}({\mathbb{R}}) is a distributional solution to

    Ψ=μΨ.\mathcal{H}\Psi=\mu\Psi. (4.23)

    Furthermore, if additionally assumption (A6) holds, i.e.,

    c2,±:=122μe±2μy(V2(y)Ψ1(y)+V1(y)Ψ2(y))dy=0,c_{2,\pm}:=\frac{1}{2\sqrt{2\mu}}\int_{{\mathbb{R}}}e^{\pm\sqrt{2\mu}y}\big{(}V_{2}(y)\Psi_{1}(y)+V_{1}(y)\Psi_{2}(y)\big{)}\,\mathrm{d}y=0, (4.24)

    then

    limx±Ψ1(x)=c0c1,\lim_{x\to\pm\infty}\Psi_{1}(x)=c_{0}\mp c_{1}, (4.25)

    where

    c1:=12x(a(x),b(x)),Φ(x)=12x(a(x)Φ1(x)+b(x)Φ2(x))dx.c_{1}:=\frac{1}{2}\langle x(a(x),b(x))^{\top},\Phi(x)\rangle=\frac{1}{2}\int_{\mathbb{R}}x\big{(}a(x)\Phi_{1}(x)+b(x)\Phi_{2}(x)\big{)}\,\mathrm{d}x. (4.26)

    In particular,

    Ψ1L2().\Psi_{1}\notin L^{2}({\mathbb{R}}). (4.27)

    More precisely, the constants c0c_{0} and c1c_{1} cannot both be zero.

  2. (2)

    Conversely, suppose there exists ΨL()×L()\Psi\in L^{\infty}({\mathbb{R}})\times L^{\infty}({\mathbb{R}}) satisfying (4.23) in the distributional sense. Then

    Φ=v2ΨS1(L2()×L2()).\Phi=v_{2}\Psi\in S_{1}(L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})). (4.28)
  3. (3)

    Suppose assumptions (A1) – (A6) hold. Then, dimS1(L2()×L2())1\dim S_{1}(L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}}))\leq 1. In the case dimS1(L2()×L2())=1\dim S_{1}(L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}}))=1, i.e., S1(L2()×L2())=span{Φ}S_{1}(L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}}))=\operatorname*{span}\{\Phi\} for some Φ=(Φ1,Φ2)L2()×L2(){0}\Phi=(\Phi_{1},\Phi_{2})^{\top}\in L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})\setminus\{0\}, we have the following identities

    S1TPTS1\displaystyle S_{1}TPTS_{1} =|c0|2ΦL2()×L2()2V1L1()S1,\displaystyle=|c_{0}|^{2}\|\Phi\|_{L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})}^{-2}\|V_{1}\|_{L^{1}({\mathbb{R}})}S_{1}, (4.29)
    PTS1TP\displaystyle PTS_{1}TP =|c0|2ΦL2()×L2()2V1L1()P,\displaystyle=|c_{0}|^{2}\|\Phi\|_{L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})}^{-2}\|V_{1}\|_{L^{1}({\mathbb{R}})}P, (4.30)
    S1M1S1\displaystyle S_{1}M_{1}S_{1} =2i|c1|2ΦL2()×L2()2S1,\displaystyle=-2i|c_{1}|^{2}\|\Phi\|_{L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})}^{-2}S_{1}, (4.31)

    where the constants c0c_{0} and c1c_{1} are given by (4.21) and (4.26) respectively for this Φ\Phi.

Proof of (1).

Let Φ=(Φ1,Φ2)S1(L2()×L2())\Phi=(\Phi_{1},\Phi_{2})\in S_{1}(L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})) with Φ0\Phi\neq 0. Since S1(L2()×L2())S_{1}(L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})) is a subspace of Q(L2()×L2())Q(L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})), we have QΦ=ΦQ\Phi=\Phi. Using the fact that Φker(QTQ)\Phi\in\ker(QTQ) and the definition of TT (c.f (4.10)), we obtain

0=QTQΦ=(IP)TΦ=(I+v2𝒢0v1)ΦPTΦ.0=QTQ\Phi=(I-P)T\Phi=(I+v_{2}\mathcal{G}_{0}v_{1})\Phi-PT\Phi. (4.32)

Since (a,b)=v2e¯1(a,b)^{\top}=v_{2}\underline{e}_{1} and PP is the orthogonal projection onto the span of (a,b)(a,b)^{\top}, we have

PTΦ=(a,b),TΦV1L1()(a,b)=c0v2e¯1,PT\Phi=\frac{\langle(a,b)^{\top},T\Phi\rangle}{\|V_{1}\|_{L^{1}({\mathbb{R}})}}(a,b)^{\top}=c_{0}v_{2}\underline{e}_{1}, (4.33)

with c0c_{0} defined in (4.21). It follows that

Φ=v2𝒢0v1Φ+c0v2e¯1=v2(𝒢0v1Φ+c0e¯1)=v2Ψ.\Phi=-v_{2}\mathcal{G}_{0}v_{1}\Phi+c_{0}v_{2}\underline{e}_{1}=v_{2}(-\mathcal{G}_{0}v_{1}\Phi+c_{0}\underline{e}_{1})=v_{2}\Psi.

This proves (4.22). Next, we show (4.23). Denoting Φ=(Φ1,Φ2)\Phi=(\Phi_{1},\Phi_{2})^{\top} and using the definition of 𝒢0\mathcal{G}_{0} (c.f. (4.2)), we have

(0μI)𝒢0(v1Φ)=v1Φ,(\mathcal{H}_{0}-\mu I)\mathcal{G}_{0}(v_{1}\Phi)=v_{1}\Phi,

i.e.,

{(x2)|xy|2(a(y)Φ1(y)b(y)Φ2(y))dy=a(x)Φ1(x)b(x)Φ2(x),(x22μ)e2μ|xy|22μ(b(y)Φ1(y)+a(y)Φ2(y))dy=b(x)Φ1(x)+a(x)Φ2(x).\begin{split}\begin{cases}(-\partial_{x}^{2})\displaystyle\int_{\mathbb{R}}\frac{-|x-y|}{2}\big{(}-a(y)\Phi_{1}(y)-b(y)\Phi_{2}(y)\big{)}\,\mathrm{d}y=-a(x)\Phi_{1}(x)-b(x)\Phi_{2}(x),\\ (\partial_{x}^{2}-2\mu)\displaystyle\int_{\mathbb{R}}\frac{-e^{-\sqrt{2\mu}|x-y|}}{2\sqrt{2\mu}}\big{(}b(y)\Phi_{1}(y)+a(y)\Phi_{2}(y)\big{)}\,\mathrm{d}y=b(x)\Phi_{1}(x)+a(x)\Phi_{2}(x).\end{cases}\end{split}

This equation is well-defined, since v1Φx1L1()×x1L1()v_{1}\Phi\in\langle x\rangle^{-1-}L^{1}({\mathbb{R}})\times\langle x\rangle^{-1-}L^{1}({\mathbb{R}}). Using (4.20), (4.22), and (H0μI)(c0e¯1)=0(H_{0}-\mu I)(c_{0}\underline{e}_{1})=0, we have

(0μI)Ψ=(H0μI)[𝒢0(v1Φ)+c0e¯1]=v1Φ=v1v2Ψ=𝒱Ψ,(\mathcal{H}_{0}-\mu I)\Psi=(H_{0}-\mu I)[-\mathcal{G}_{0}(v_{1}\Phi)+c_{0}\underline{e}_{1}]=-v_{1}\Phi=-v_{1}v_{2}\Psi=-\mathcal{V}\Psi,

which implies (4.23). We now show that Ψ=(Ψ1,Ψ2)\Psi=(\Psi_{1},\Psi_{2})^{\top} is in L()×L()L^{\infty}({\mathbb{R}})\times L^{\infty}({\mathbb{R}}). Noting that

Ψ1(x)=c0+12|xy|(a(y)Φ1(y)+b(y)Φ2(y))dy,\Psi_{1}(x)=c_{0}+\frac{1}{2}\int_{\mathbb{R}}|x-y|\big{(}a(y)\Phi_{1}(y)+b(y)\Phi_{2}(y)\big{)}\,\mathrm{d}y,

by employing the orthogonality condition (a,b),Φ=0\langle(a,b)^{\top},\Phi\rangle=0, we have

Ψ1(x)=c0+12(|xy||x|)(a(y)Φ1(y)+b(y)Φ2(y))dy.\Psi_{1}(x)=c_{0}+\frac{1}{2}\int_{\mathbb{R}}(|x-y|-|x|)\big{(}a(y)\Phi_{1}(y)+b(y)\Phi_{2}(y)\big{)}\,\mathrm{d}y.

Using ||xy||x|||y|\big{|}|x-y|-|x|\big{|}\leq|y| and |a(y)|+|b(y)|y2|a(y)|+|b(y)|\lesssim\langle y\rangle^{-2}, we have the uniform bound

supx|Ψ1(x)||c0|+12|y||a(y)Φ1(y)+b(y)Φ2(y)|dyΦL2()×L2()1.\sup_{x\in{\mathbb{R}}}|\Psi_{1}(x)|\leq|c_{0}|+\frac{1}{2}\int|y|\left|a(y)\Phi_{1}(y)+b(y)\Phi_{2}(y)\right|\,\mathrm{d}y\lesssim\|\Phi\|_{L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})}\lesssim 1.

Since (a,b)(a,b)^{\top} and Φ\Phi are in L2()×L2()L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}}), we have the uniform bound on Ψ2\Psi_{2} by the Cauchy-Schwarz inequality

supx|Ψ2(x)||b(y)Φ1(y)+a(y)Φ2(y)|dybL2()Φ1L2()+aL2()Φ2L2()1.\sup_{x\in{\mathbb{R}}}|\Psi_{2}(x)|\lesssim\int_{\mathbb{R}}|b(y)\Phi_{1}(y)+a(y)\Phi_{2}(y)|\,\mathrm{d}y\leq\|b\|_{L^{2}({\mathbb{R}})}\|\Phi_{1}\|_{L^{2}({\mathbb{R}})}+\|a\|_{L^{2}({\mathbb{R}})}\|\Phi_{2}\|_{L^{2}({\mathbb{R}})}\lesssim 1.

Thus, we have shown that Ψ=(Ψ1,Ψ2)L()×L()\Psi=(\Psi_{1},\Psi_{2})^{\top}\in L^{\infty}({\mathbb{R}})\times L^{\infty}({\mathbb{R}}). Finally, we now assume c2,±=0c_{2,\pm}=0 and show that Ψ1\Psi_{1} cannot be in L2(){0}L^{2}({\mathbb{R}})\setminus\{0\} by a Volterra argument. Using (a,b),Φ=0\langle(a,b)^{\top},\Phi\rangle=0, for x0x\geq 0 large, we write

Ψ1(x)=c0c1+x(yx)(a(y)Φ1(y)+b(y)Φ2(y))dy.\begin{split}\Psi_{1}(x)&=c_{0}-c_{1}+\int_{x}^{\infty}(y-x)\big{(}a(y)\Phi_{1}(y)+b(y)\Phi_{2}(y)\big{)}\,\mathrm{d}y.\end{split} (4.34)

Using c2,±=0c_{2,\pm}=0, we insert e2μxc2,+=0-e^{-\sqrt{2\mu}x}c_{2,+}=0 to write

Ψ2(x)=122μx(e2μ(yx)e2μ(xy))(V2(y)Ψ1(y)+V1(y)Ψ2(y))dy.\begin{split}\Psi_{2}(x)&=\frac{1}{2\sqrt{2\mu}}\int_{x}^{\infty}\big{(}e^{-\sqrt{2\mu}(y-x)}-e^{-\sqrt{2\mu}(x-y)}\big{)}\big{(}V_{2}(y)\Psi_{1}(y)+V_{1}(y)\Psi_{2}(y)\big{)}\,\mathrm{d}y.\end{split} (4.35)

Similarly, for x<0x<0, using e2μxc2,=0e^{\sqrt{2\mu x}}c_{2,-}=0, we have

Ψ1(x)=c0+c1+x(xy)(V1(y)Ψ1(y)+V2(y)Ψ2(y))dy,\displaystyle\Psi_{1}(x)=c_{0}+c_{1}+\int_{-\infty}^{x}(x-y)(V_{1}(y)\Psi_{1}(y)+V_{2}(y)\Psi_{2}(y))\,\mathrm{d}y, (4.36)
Ψ2(x)=122μx(e2μ(xy)e2μ(yx))(V2(y)Ψ1(y)+V1(y)Ψ2(y))dy.\displaystyle\Psi_{2}(x)=\frac{1}{2\sqrt{2\mu}}\int_{-\infty}^{x}\big{(}e^{-\sqrt{2\mu}(x-y)}-e^{-\sqrt{2\mu}(y-x)}\big{)}\big{(}V_{2}(y)\Psi_{1}(y)+V_{1}(y)\Psi_{2}(y)\big{)}\,\mathrm{d}y. (4.37)

Suppose now that c0=c1=0c_{0}=c_{1}=0. Owing to the exponential decay of V1V_{1}, V2V_{2} by assumption (A3), we obtain from (4.34) and (4.34) a homogeneous Volterra equation for Ψ=(Ψ1,Ψ2)\Psi=(\Psi_{1},\Psi_{2})^{\top} satisfying

Ψ(x)=K(x,y)Ψ(y)dy,x0,\Psi(x)=\int_{{\mathbb{R}}}K(x,y)\Psi(y)\,\mathrm{d}y,\quad x\geq 0,

where |K(x,y)|eγ|y|𝟙y>x|K(x,y)|\lesssim e^{-\gamma|y|}\mathbbm{1}_{y>x} for some 0<γ<β0<\gamma<\beta, which is a quasi-nilpotent operator. By performing a standard contraction on L(M,)L^{\infty}(M,\infty), with M>0M>0 sufficiently large, one arrives at a solution Ψ(x)0\Psi(x)\equiv 0 for all xMx\geq M. By the uniqueness theorem for ODEs, this implies that Ψ0\Psi\equiv 0 on {\mathbb{R}}. Then, by the relation Φ=v2Ψ\Phi=v_{2}\Psi and the fact that v2v_{2} is a positive matrix, one finds that Φ0\Phi\equiv 0, which contradicts the hypothesis Φ0\Phi\neq 0. Thus, the conclusion is that c0c_{0} and c1c_{1} cannot be both zero. In particular, it follows from (4.34) and (4.36) that

limx±Ψ1(x)=c0c1.\lim_{x\rightarrow\pm\infty}\Psi_{1}(x)=c_{0}\mp c_{1}.

Since either c0+c10c_{0}+c_{1}\neq 0 or c0c10c_{0}-c_{1}\neq 0, we conclude that Ψ1L2()\Psi_{1}\not\in L^{2}({\mathbb{R}}).

Proof of (2). Define Φ=v2Ψ\Phi=v_{2}\Psi. Since Ψ\Psi is a distributional solution to (4.23), using 𝒱=v1v2\mathcal{V}=v_{1}v_{2}, we have

(0μI)Ψ=v1Φ{Ψ1′′=aΦ1+bΦ2,Ψ2′′2μΨ2=bΦ1+aΦ2.(\mathcal{H}_{0}-\mu I)\Psi=v_{1}\Phi\Longleftrightarrow\begin{cases}\Psi_{1}^{\prime\prime}=a\Phi_{1}+b\Phi_{2},\\ \Psi_{2}^{\prime\prime}-2\mu\Psi_{2}=b\Phi_{1}+a\Phi_{2}.\end{cases}

Let ηC0()\eta\in C_{0}^{\infty}({\mathbb{R}}) be a non-negative function satisfying η(x)=1\eta(x)=1 for |x|1|x|\leq 1 and η(x)=0\eta(x)=0 for |x|2|x|\geq 2. Using the first equation from above and integrating by parts, we have for any ε>0{\varepsilon}>0,

|(a(y)Φ1(y)+b(y)Φ2(y))η(εy)dy|=|Ψ1′′(y)η(εy)dy|=|Ψ1(y)ε2η′′(εy)dy|εΨ1L()|η′′(x)|dx.\begin{split}&\left|\int_{\mathbb{R}}\big{(}a(y)\Phi_{1}(y)+b(y)\Phi_{2}(y)\big{)}\eta({\varepsilon}y)\,\mathrm{d}y\right|=\left|\int_{\mathbb{R}}\Psi_{1}^{\prime\prime}(y)\eta({\varepsilon}y)\,\mathrm{d}y\right|\\ &=\left|\int_{\mathbb{R}}\Psi_{1}(y){\varepsilon}^{2}\eta^{\prime\prime}({\varepsilon}y)\,\mathrm{d}y\right|\leq{\varepsilon}\|\Psi_{1}\|_{L^{\infty}({\mathbb{R}})}\int_{\mathbb{R}}\left|\eta^{\prime\prime}(x)\right|\,\mathrm{d}x.\end{split}

By taking the limit ε0{\varepsilon}\to 0 and using the Lebesgue dominated convergence theorem, we find that (a,b),Φ=0\langle(a,b)^{\top},\Phi\rangle=0. Thus, PΦ=0P\Phi=0, i.e. ΦQ(L2()×L2())\Phi\in Q(L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})). Following this fact and using Φ=v2Ψ\Phi=v_{2}\Psi, we have

QTQΦ=QTΦ=Q(I+v2𝒢0v1)Φ=Qv2(Ψ+𝒢0(𝒱Ψ)).QTQ\Phi=QT\Phi=Q(I+v_{2}\mathcal{G}_{0}v_{1})\Phi=Qv_{2}\big{(}\Psi+\mathcal{G}_{0}(\mathcal{V}\Psi)\big{)}. (4.38)

Now set u:=Ψ+𝒢0(𝒱Ψ)u:=\Psi+\mathcal{G}_{0}(\mathcal{V}\Psi). Since u=(u1,u2)u=(u_{1},u_{2})^{\top} is a distributional solution of (0μI)u=0(\mathcal{H}_{0}-\mu I)u=0, i.e.

u1′′=0,u2′′2μu2=0,\begin{split}-u_{1}^{\prime\prime}&=0,\\ u_{2}^{\prime\prime}-2\mu u_{2}&=0,\end{split}

we find that

u1(x)=κ1+κ2x,u2(x)=κ3e2μx+κ4e2μx,\begin{split}&u_{1}(x)=\kappa_{1}+\kappa_{2}x,\\ &u_{2}(x)=\kappa_{3}e^{-\sqrt{2\mu}x}+\kappa_{4}e^{\sqrt{2\mu}x},\end{split}

for some κi\kappa_{i}\in{\mathbb{C}}, i{1,,4}i\in\{1,\ldots,4\}. By similar arguments from Item (1), we obtain that 𝒢0(𝒱Ψ)L()×L()\mathcal{G}_{0}(\mathcal{V}\Psi)\in L^{\infty}({\mathbb{R}})\times L^{\infty}({\mathbb{R}}). Since ΨL()×L()\Psi\in L^{\infty}({\mathbb{R}})\times L^{\infty}({\mathbb{R}}), it follows that uL()×L()u\in L^{\infty}({\mathbb{R}})\times L^{\infty}({\mathbb{R}}), which implies that κ2=κ3=κ4=0\kappa_{2}=\kappa_{3}=\kappa_{4}=0. Thus, we have u(x)(κ1,0)=κ1e¯1u(x)\equiv(\kappa_{1},0)^{\top}=\kappa_{1}\underline{e}_{1}. Since Qv2e¯1=0Qv_{2}\underline{e}_{1}=0, we conclude from (4.38) using the definition of u(x)u(x) that QTQΦ=0QTQ\Phi=0, whence ΦS1(L2()×L2())\Phi\in S_{1}(L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})).

Proof of (3). Suppose there are two linearly independent Φ,Φ~S1(L2()×L2())\Phi,{\widetilde{\Phi}}\in S_{1}(L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})). As in the proof of Item (1), for x0x\geq 0, we have

Ψ1(x)=c0c1+x(yx)(V1(y)Ψ1(y)+V2(y)Ψ2(y))dy,Ψ2(x)=122μx(e2μ(yx)e2μ(xy))(V2(y)Ψ1(y)+V1(y)Ψ2(y))dy,\begin{split}&\Psi_{1}(x)=c_{0}-c_{1}+\int_{x}^{\infty}(y-x)\big{(}V_{1}(y)\Psi_{1}(y)+V_{2}(y)\Psi_{2}(y)\big{)}\,\mathrm{d}y,\\ &\Psi_{2}(x)=\frac{1}{2\sqrt{2\mu}}\int_{x}^{\infty}\big{(}e^{-\sqrt{2\mu}(y-x)}-e^{-\sqrt{2\mu}(x-y)}\big{)}\big{(}V_{2}(y)\Psi_{1}(y)+V_{1}(y)\Psi_{2}(y)\big{)}\,\mathrm{d}y,\end{split}

and

Ψ~1(x)=d0d1+x(yx)(V1(y)Ψ~1(y)+V2(y)Ψ~2(y))dy,Ψ~2(x)=122μx(e2μ(yx)e2μ(xy))(V2(y)Ψ~1(y)+V1(y)Ψ~2(y))dy,\begin{split}&\widetilde{\Psi}_{1}(x)=d_{0}-d_{1}+\int_{x}^{\infty}(y-x)\big{(}V_{1}(y)\widetilde{\Psi}_{1}(y)+V_{2}(y)\widetilde{\Psi}_{2}(y)\big{)}\,\mathrm{d}y,\\ &\widetilde{\Psi}_{2}(x)=\frac{1}{2\sqrt{2\mu}}\int_{x}^{\infty}\big{(}e^{-\sqrt{2\mu}(y-x)}-e^{-\sqrt{2\mu}(x-y)}\big{)}\big{(}V_{2}(y)\widetilde{\Psi}_{1}(y)+V_{1}(y)\widetilde{\Psi}_{2}(y)\big{)}\,\mathrm{d}y,\end{split}

where d0d_{0} and d1d_{1} are constants defined from Φ~{\widetilde{\Phi}} which are analogous to c0c_{0} and c1c_{1}. There is some constant θ\theta\in{\mathbb{C}} such that

c0c1=θ(d0d1),c_{0}-c_{1}=-\theta(d_{0}-d_{1}),

which imply the Volterra integral equation

[Ψ1+θΨ~1Ψ2+θΨ~2](x)=x[yx00e2μ(yx)e2μ(xy)22μ]𝒱(y)[Ψ1(y)+θΨ~1(y)Ψ2(y)+θΨ~2(y)]𝑑y,\begin{bmatrix}\Psi_{1}+\theta\widetilde{\Psi}_{1}\\ \Psi_{2}+\theta\widetilde{\Psi}_{2}\end{bmatrix}(x)=\int_{x}^{\infty}\begin{bmatrix}y-x&0\\ 0&\frac{e^{-\sqrt{2\mu}(y-x)}-e^{-\sqrt{2\mu}(x-y)}}{2\sqrt{2\mu}}\end{bmatrix}\mathcal{V}(y)\begin{bmatrix}\Psi_{1}(y)+\theta\widetilde{\Psi}_{1}(y)\\ \Psi_{2}(y)+\theta\widetilde{\Psi}_{2}(y)\end{bmatrix}dy,

for any x0x\geq 0. By the same Volterra equation argument used in Item (1), we obtain Ψ+θΨ~0\Psi+\theta\widetilde{\Psi}\equiv 0, which implies that Φ+θΦ~0\Phi+\theta\widetilde{\Phi}\equiv 0, but this contradicts that Φ\Phi and Φ~\widetilde{\Phi} are linearly independent. Thus, we have shown that dimS1(L2()×L2())1\dim S_{1}(L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}}))\leq 1. Next, we prove (4.29)–(4.31). Write S1=ΦL2×L22Φ,ΦS_{1}=\|\Phi\|_{L^{2}\times L^{2}}^{-2}\langle\Phi,\cdot\rangle\Phi. By (4.33) and the fact that PP, S1S_{1}, and TT are self-adjoint, we compute for any uL2()×L2()u\in L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}}) that

S1TPTS1u=ΦL2×L22Φ,uS1TPTΦ=ΦL2×L22c0Φ,uS1T[ab]=|c0|2ΦL2×L22V1L1()S1u.S_{1}TPTS_{1}u=\|\Phi\|_{L^{2}\times L^{2}}^{-2}\langle\Phi,u\rangle S_{1}TPT\Phi=\|\Phi\|_{L^{2}\times L^{2}}^{-2}c_{0}\langle\Phi,u\rangle S_{1}T\begin{bmatrix}a\\ b\end{bmatrix}=|c_{0}|^{2}\|\Phi\|_{L^{2}\times L^{2}}^{-2}\|V_{1}\|_{L^{1}({\mathbb{R}})}S_{1}u.

A similar computation reveals

PTS1TPu=|c0|2ΦL2×L22V1L1()Pu.PTS_{1}TPu=|c_{0}|^{2}\|\Phi\|_{L^{2}\times L^{2}}^{-2}\|V_{1}\|_{L^{1}({\mathbb{R}})}Pu.

For the third identity (4.31), in view of (4.11) and (4.17), we write

M1(x,y)=v2(x)G1(x,y)v1(y)=i|xy|24[a(x)b(x)][a(y)b(y)].M_{1}(x,y)=v_{2}(x)G_{1}(x,y)v_{1}(y)=\frac{i|x-y|^{2}}{4}\begin{bmatrix}a(x)\\ b(x)\end{bmatrix}\begin{bmatrix}a(y)&b(y)\end{bmatrix}.

By using the orthogonality

Φ,(a,b)=(Φ1(x)a(x)+Φ2(x)b(x))dx=0,\langle\Phi,(a,b)^{\top}\rangle=\int_{\mathbb{R}}\big{(}\Phi_{1}(x)a(x)+\Phi_{2}(x)b(x)\big{)}\,\mathrm{d}x=0,

and the identity

|xy|2=x2+y22xy,|x-y|^{2}=x^{2}+y^{2}-2xy,

we have

[S1M1S1](x,y)=2S1(x,x1)M1(x1,y1)S1(y1,y)dx1dy1=i4Φ(x)ΦL2×L222(|x1y1|2Φ(x1)[a(x1)b(x1)][a(y1)b(y1)]Φ(y1))dx1dy1Φ(y)ΦL2×L22=2i(x12Φ(x1)[a(x1)b(x1)]dx1)(y12[a(y1)b(y1)]Φ(y1)dy1)ΦL2×L22S1(x,y)=2i|c1|2ΦL2×L22S1(x,y).\begin{split}&[S_{1}M_{1}S_{1}](x,y)=\int_{{\mathbb{R}}^{2}}S_{1}(x,x_{1})M_{1}(x_{1},y_{1})S_{1}(y_{1},y)\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\\ &\quad=\frac{i}{4}\frac{\Phi(x)}{\|\Phi\|_{L^{2}\times L^{2}}^{2}}\int_{{\mathbb{R}}^{2}}\left(|x_{1}-y_{1}|^{2}\Phi^{*}(x_{1})\begin{bmatrix}a(x_{1})\\ b(x_{1})\end{bmatrix}\begin{bmatrix}a(y_{1})&b(y_{1})\end{bmatrix}\Phi(y_{1})\right)\,\mathrm{d}x_{1}\mathrm{d}y_{1}\frac{\Phi^{*}(y)}{\|\Phi\|_{L^{2}\times L^{2}}^{2}}\\ &\quad=-2i\left(\int_{{\mathbb{R}}}\frac{x_{1}}{2}\Phi^{*}(x_{1})\begin{bmatrix}a(x_{1})\\ b(x_{1})\end{bmatrix}\,\mathrm{d}x_{1}\right)\left(\int_{{\mathbb{R}}}\tfrac{y_{1}}{2}\begin{bmatrix}a(y_{1})&b(y_{1})\end{bmatrix}\Phi(y_{1})\,\mathrm{d}y_{1}\right)\|\Phi\|_{L^{2}\times L^{2}}^{-2}S_{1}(x,y)\\ &\quad=-2i|c_{1}|^{2}\|\Phi\|_{L^{2}\times L^{2}}^{-2}S_{1}(x,y).\end{split}

This proves (4.31) and we are done. ∎

Remark 4.6.

By direct computation, the conjugation identity σ3=σ3\sigma_{3}\mathcal{H}=\mathcal{H}^{*}\sigma_{3} and the identity v1=σ3v2v_{1}=-\sigma_{3}v_{2} imply that the vector Ψ~:=σ3Ψ\widetilde{\Psi}:=\sigma_{3}\Psi solves

Ψ~=μΨ~,\mathcal{H}^{*}\widetilde{\Psi}=\mu\widetilde{\Psi}, (4.39)

where Ψ\Psi is the distribution solution to (4.23). Moreover, one has the identities

σ3Ψ=𝒢0(v2Φ)+(c0,0),Φ=v2Ψ=v1Ψ~\sigma_{3}\Psi=\mathcal{G}_{0}(v_{2}\Phi)+(c_{0},0)^{\top},\quad\Phi=v_{2}\Psi=-v_{1}^{\top}\widetilde{\Psi} (4.40)

Similarly, using the conjugation identity σ1=σ1\sigma_{1}\mathcal{H}=-\mathcal{H}\sigma_{1}, we note that the vector Ψ=σ1Ψ\Psi_{-}=\sigma_{1}\Psi solves the system

Ψ=μΨ.\mathcal{H}\Psi_{-}=-\mu\Psi_{-}. (4.41)

Following the preceding discussion, we assume the threshold μ\mu is irregular and we derive an expansion for the inverse operator M(z)1M(z)^{-1} on a small punctured disk near the origin. We employ the inversion lemma due to Jensen and Nenciu [JN01, Lemma 2.1].

Lemma 4.7.

Let HH be a Hilbert space, let AA be a closed operator and SS a projection. Suppose A+SA+S has a bounded inverse. Then AA has a bounded inverse if and only if

B=SS(A+S)1SB=S-S(A+S)^{-1}S

has a bounded inverse in SHSH, and in this case,

A1=(A+S)1+(A+S)1SB1S(A+S)1,on H.A^{-1}=(A+S)^{-1}+(A+S)^{-1}SB^{-1}S(A+S)^{-1},\quad\text{on $H$}.

We will now state the inverse operator of M(z)M(z) away from z=0z=0.

Proposition 4.8.

Suppose assumptions (A1) – (A6) hold. Let S1(L2()×L2())=span({Φ})S_{1}(L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}}))=\operatorname*{span}(\{\Phi\}) for some Φ=(Φ1,Φ2)0\Phi=(\Phi_{1},\Phi_{2})^{\top}\neq\vec{0}. Let κ:=(2i)1V1L1()\kappa:=(2i)^{-1}\|V_{1}\|_{L^{1}({\mathbb{R}})}, and let dd be the constant defined by

d:=2i(|c0|2+|c1|2)ΦL2×L220,d:=-2i(|c_{0}|^{2}+|c_{1}|^{2})\|\Phi\|_{L^{2}\times L^{2}}^{-2}\neq 0, (4.42)

with c0c_{0} and c1c_{1} defined by (4.21) and (4.26) respectively for this Φ\Phi. Then, there exists a positive radius z0>0z_{0}>0 such that for all 0<|z|<z00<|z|<z_{0}, M(z)M(z) is invertible on L2()×L2()L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}}) and

M(z)1=1d(1zS11κPTS11κS1TP)+(1κ+|c0|2ΦL2×L22V1L1()dκ2)zP+QΛ0(z)Q+zQΛ1(z)+zΛ2(z)Q+z2Λ3(z),\begin{split}M(z)^{-1}&=\frac{1}{d}\left(\frac{1}{z}S_{1}-\frac{1}{\kappa}PTS_{1}-\frac{1}{\kappa}S_{1}TP\right)+\left(\frac{1}{\kappa}+\frac{|c_{0}|^{2}\|\Phi\|_{L^{2}\times L^{2}}^{-2}\|V_{1}\|_{L^{1}({\mathbb{R}})}}{d\kappa^{2}}\right)zP\\ &\qquad+Q\Lambda_{0}(z)Q+zQ\Lambda_{1}(z)+z\Lambda_{2}(z)Q+z^{2}\Lambda_{3}(z),\end{split} (4.43)

where Λj(z)\Lambda_{j}(z) are absolutely bounded operators on L2()×L2()L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}}) satisfying the improved bounds

|zkΛj(z)|L2()×L2()L2()×L2()1,k=0,1,2,j=0,1,2,3,\||\partial_{z}^{k}\Lambda_{j}(z)|\|_{L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})\to L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})}\lesssim 1,\quad k=0,1,2,\quad j=0,1,2,3, (4.44)

uniformly in zz for |z|<z0|z|<z_{0}.

Proof.

Throughout the proof, we will denote by j(z)\mathcal{E}_{j}(z), for 0j30\leq j\leq 3, as error terms that satisfy the absolute bound

|z|k|zkj(z)|L2()×L2()L2()×L2()|z|j,k=0,1,2,|z|<z0,|z|^{k}\left\||\partial_{z}^{k}\mathcal{E}_{j}(z)|\right\|_{L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})\rightarrow L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})}\lesssim|z|^{j},\quad\forall\,k=0,1,2,\quad\forall\,|z|<z_{0},

for some z0>0z_{0}>0 small. This convenient notation will be useful in invoking Neumann series inversion for small values of zz. Since we only need the expansion of M(z)1M(z)^{-1} up to a few powers of zz, the exact expressions of j(z)\mathcal{E}_{j}(z) are insignificant and we allow it to vary from line to line. By Proposition 4.3, we rewrite M(z)M(z) by setting

M~(z):=zκM(z)=P+zκ(T+zM1+2(z)),{\widetilde{M}}(z):=\frac{z}{\kappa}M(z)=P+\frac{z}{\kappa}\big{(}T+zM_{1}+\mathcal{M}_{2}(z)\big{)}, (4.45)

where 2(z)\mathcal{M}_{2}(z) is the error term in Proposition 4.3. Using I=P+QI=P+Q, we write

M~(z)+Q=I+zκ(T+zM1+2(z)),{\widetilde{M}}(z)+Q=I+\frac{z}{\kappa}\big{(}T+zM_{1}+\mathcal{M}_{2}(z)\big{)}, (4.46)

and by choosing zz small enough, a Neumann series expansion yields the inverse operator

[M~(z)+Q]1=n0(1)n(zκ(T+zM1+2(z)))non L2()×L2().[{\widetilde{M}}(z)+Q]^{-1}=\sum_{n\geq 0}(-1)^{n}\left(\frac{z}{\kappa}\big{(}T+zM_{1}+\mathcal{M}_{2}(z)\big{)}\right)^{n}\quad\text{on $L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})$}. (4.47)

We collect the terms of power order up to 22 to obtain

[M~(z)+Q]1=IzκTz2(1κM11κ2T2)+3(z).[{\widetilde{M}}(z)+Q]^{-1}=I-\frac{z}{\kappa}T-z^{2}\left(\frac{1}{\kappa}M_{1}-\frac{1}{\kappa^{2}}T^{2}\right)+\mathcal{E}_{3}(z). (4.48)

Note that z2(z)z\mathcal{M}_{2}(z) is of the form 3(z)\mathcal{E}_{3}(z). Recall by Lemma 4.7 that the operator M~(z){\widetilde{M}}(z) is invertible on L2()×L2()L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}}) if and only if the operator

B1(z):=QQ[M~(z)+Q]1QB_{1}(z):=Q-Q\big{[}{\widetilde{M}}(z)+Q\big{]}^{-1}Q (4.49)

is invertible on the subspace QL2Q(L2()×L2())QL^{2}\equiv Q(L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})). Using (4.48), we find that

B1(z)=zκQTQ+z2(1κQM1Q1κ2QT2Q)+Q3(z)Q.B_{1}(z)=\frac{z}{\kappa}QTQ+z^{2}\left(\frac{1}{\kappa}QM_{1}Q-\frac{1}{\kappa^{2}}QT^{2}Q\right)+Q\mathcal{E}_{3}(z)Q.

We rewrite B1(z)B_{1}(z) by setting

B~1(z):=κzB1(z)=QTQ+z(QM1Q1κQT2Q)+Q2(z)Q.{\widetilde{B}}_{1}(z):=\frac{\kappa}{z}B_{1}(z)=QTQ+z\left(QM_{1}Q-\frac{1}{\kappa}QT^{2}Q\right)+Q\mathcal{E}_{2}(z)Q. (4.50)

Since the threshold μ\mu is not regular, the operator QTQQTQ is not invertible on QL2QL^{2} according to Definition 4.4. By considering the operator

B~1(z)+S1=(QTQ+S1)+z(QM1Q1κQT2Q)+Q2(z)Q,{\widetilde{B}}_{1}(z)+S_{1}=(QTQ+S_{1})+z\left(QM_{1}Q-\frac{1}{\kappa}QT^{2}Q\right)+Q\mathcal{E}_{2}(z)Q,

and the fact that we have QD0Q=D0=(QTQ+S1)1QD_{0}Q=D_{0}=(QTQ+S_{1})^{-1} on QL2QL^{2}, we can pick zz small enough such that

z(QM1Q1κQT2Q)+Q2(z)QL2×L2L2×L2<QD0QL2×L2L2×L21.\left\|z\left(QM_{1}Q-\frac{1}{\kappa}QT^{2}Q\right)+Q\mathcal{E}_{2}(z)Q\right\|_{L^{2}\times L^{2}\to L^{2}\times L^{2}}<\|QD_{0}Q\|_{L^{2}\times L^{2}\to L^{2}\times L^{2}}^{-1}.

This allows for the more complicated Neumann series expansion (c.f. Lemma A.1) on QL2QL^{2}:

(B~1(z)+S1)1=D0n0(1)n((z(QM1Qκ1QT2Q)+Q2(z)Q)D0)non QL2.({\widetilde{B}}_{1}(z)+S_{1})^{-1}=D_{0}\sum_{n\geq 0}(-1)^{n}\Big{(}\big{(}z(QM_{1}Q-\kappa^{-1}QT^{2}Q)+Q\mathcal{E}_{2}(z)Q\big{)}D_{0}\Big{)}^{n}\quad\text{on $QL^{2}$}. (4.51)

We collect the leading order terms in this expansion and write

(B~1(z)+S1)1=D0zD0(QM1Qκ1QT2Q)D0+Q2(z)Q.\begin{split}({\widetilde{B}}_{1}(z)+S_{1})^{-1}=D_{0}-zD_{0}\left(QM_{1}Q-\kappa^{-1}QT^{2}Q\right)D_{0}+Q\mathcal{E}_{2}(z)Q.\end{split} (4.52)

At this step, it is crucial that the operator D0D_{0} is absolutely bounded to ensure that the remainder term Q2(z)QQ\mathcal{E}_{2}(z)Q and its derivatives are absolutely bounded. Next, we set

B2(z):=S1S1(B~1(z)+S1)1S1,on S1L2S1(L2()×L2()).B_{2}(z):=S_{1}-S_{1}({\widetilde{B}}_{1}(z)+S_{1})^{-1}S_{1},\quad\text{on $S_{1}L^{2}\equiv S_{1}(L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}}))$}. (4.53)

Using the orthogonality conditions

S1D0=D0S1=S1,S1Q=QS1=S1,QTS1=S1TQ=0,\begin{split}&S_{1}D_{0}=D_{0}S_{1}=S_{1},\\ &S_{1}Q=QS_{1}=S_{1},\\ &QTS_{1}=S_{1}TQ=0,\end{split} (4.54)

we obtain

B2(z)=zS1(M1κ1T2)S1+S12(z)S1.B_{2}(z)=zS_{1}(M_{1}-\kappa^{-1}T^{2})S_{1}+S_{1}\mathcal{E}_{2}(z)S_{1}.

By Lemma 4.5, we note that S1L2S_{1}L^{2} is spanned by Φ(x)\Phi(x) and that PTΦ=TΦPT\Phi=T\Phi holds (c.f. (4.32)), whence S1T2S1=S1TPTS1S_{1}T^{2}S_{1}=S_{1}TPTS_{1}. Using Lemma 4.5 (c.f. (4.29), (4.31)), we obtain that

d:=Tr(S1(M1κ1T2)S1)=Tr(S1M1S1)κ1Tr(S1TPTS1)=2i(|c0|2+|c1|2)ΦL2×L220.d:=\operatorname{Tr}(S_{1}(M_{1}-\kappa^{-1}T^{2})S_{1})=\operatorname{Tr}(S_{1}M_{1}S_{1})-\kappa^{-1}\operatorname{Tr}(S_{1}TPTS_{1})=-2i(|c_{0}|^{2}+|c_{1}|^{2})\|\Phi\|_{L^{2}\times L^{2}}^{-2}\neq 0.

Hence, we apply another Neumann series expansion to invert the operator B2(z)B_{2}(z) on S1L2S_{1}L^{2} for small zz and write

B2(z)1=1dzS1+S10(z)S1on S1L2.B_{2}(z)^{-1}=\frac{1}{dz}S_{1}+S_{1}\mathcal{E}_{0}(z)S_{1}\quad\text{on $S_{1}L^{2}$}. (4.55)

Moreover, by Lemma 4.7, we have

B~1(z)1=(B~1(z)+S1)1+(B~1(z)+S1)1S1B2(z)1S1(B~1(z)+S1)1on QL2.{\widetilde{B}}_{1}(z)^{-1}=\big{(}{\widetilde{B}}_{1}(z)+S_{1}\big{)}^{-1}+\big{(}{\widetilde{B}}_{1}(z)+S_{1}\big{)}^{-1}S_{1}B_{2}(z)^{-1}S_{1}\big{(}{\widetilde{B}}_{1}(z)+S_{1}\big{)}^{-1}\quad\text{on $QL^{2}$}.

Using (4.52), (4.54), and (4.55), we find that

B~1(z)1=1dzS1+Q0(z)Qon QL2.\begin{split}{\widetilde{B}}_{1}(z)^{-1}&=\frac{1}{dz}S_{1}+Q\mathcal{E}_{0}(z)Q\quad\text{on $QL^{2}$}.\end{split}

Hence,

B1(z)1=κzB~1(z)1=κdz2S1+κzQ0(z)Qon QL2.B_{1}(z)^{-1}=\frac{\kappa}{z}{\widetilde{B}}_{1}(z)^{-1}=\frac{\kappa}{dz^{2}}S_{1}+\frac{\kappa}{z}Q\mathcal{E}_{0}(z)Q\quad\text{on $QL^{2}$}.

We return to the expansion of M~(z)1{\widetilde{M}}(z)^{-1} by using Lemma 4.7 with (4.48) to obtain that

M~(z)1=(M~(z)+Q)1+(M~(z)+Q)1QB1(z)1Q(M~(z)+Q)1=(IzκT)+κdz2S11dzTS11dzS1T+1dκTS1T+κz(Q0(z)Q+1(z)Q+Q1(z)+2(z)).\begin{split}{\widetilde{M}}(z)^{-1}&=\big{(}{\widetilde{M}}(z)+Q\big{)}^{-1}+\big{(}{\widetilde{M}}(z)+Q\big{)}^{-1}QB_{1}(z)^{-1}Q\big{(}{\widetilde{M}}(z)+Q\big{)}^{-1}\\ &=\big{(}I-\frac{z}{\kappa}T\big{)}+\frac{\kappa}{dz^{2}}S_{1}-\frac{1}{dz}TS_{1}-\frac{1}{dz}S_{1}T+\frac{1}{d\kappa}TS_{1}T\\ &\quad+\frac{\kappa}{z}\left(Q\mathcal{E}_{0}(z)Q+\mathcal{E}_{1}(z)Q+Q\mathcal{E}_{1}(z)+\mathcal{E}_{2}(z)\right).\end{split}

Here, we used the identity Q=IQ=QIQ=IQ=QI. By reverting back to M(z)=κzM~(z)M(z)=\frac{\kappa}{z}{\widetilde{M}}(z), we have

M(z)1=zκM~(z)1=zκI+1dzS11dκTS11dκS1T+zdκ2TS1T+Q0(z)Q+1(z)Q+Q1(z)+2(z).\begin{split}M(z)^{-1}=\frac{z}{\kappa}{\widetilde{M}}(z)^{-1}&=\frac{z}{\kappa}I+\frac{1}{dz}S_{1}-\frac{1}{d\kappa}TS_{1}-\frac{1}{d\kappa}S_{1}T+\frac{z}{d\kappa^{2}}TS_{1}T\\ &\qquad+Q\mathcal{E}_{0}(z)Q+\mathcal{E}_{1}(z)Q+Q\mathcal{E}_{1}(z)+\mathcal{E}_{2}(z).\end{split}

Note that we absorb the z2κ2T\frac{z^{2}}{\kappa^{2}}T term into the error 2(z)\mathcal{E}_{2}(z) above. By using the identities I=Q+PI=Q+P, QTS1=S1TQ=0QTS_{1}=S_{1}TQ=0, and by factoring the powers of zz from the error terms j(z)\mathcal{E}_{j}(z), we obtain the expansion of M(z)1M(z)^{-1} on L2L^{2}: for 0<|z|<z00<|z|<z_{0},

M(z)1=zκP+1d(1zS11κPTS11κS1TP+1κ2PTS1TP)+QΛ0(z)Q+zQΛ1(z)+zΛ2(z)Q+z2Λ3(z),\begin{split}M(z)^{-1}&=\frac{z}{\kappa}P+\frac{1}{d}\left(\frac{1}{z}S_{1}-\frac{1}{\kappa}PTS_{1}-\frac{1}{\kappa}S_{1}TP+\frac{1}{\kappa^{2}}PTS_{1}TP\right)\\ &\qquad+Q\Lambda_{0}(z)Q+zQ\Lambda_{1}(z)+z\Lambda_{2}(z)Q+z^{2}\Lambda_{3}(z),\end{split}

where the operators Λj(z)\Lambda_{j}(z), j=0,,3j=0,\ldots,3, satisfy (4.44). Here, we choose z0>0z_{0}>0 sufficiently small such that the expansion (4.45)\eqref{eqn: def wtilM} and the Neumann series inversions (4.47), (4.51), (4.55) are valid for all 0<|z|<z00<|z|<z_{0}. Finally, by Lemma 4.5 (c.f. (4.30)), the term PTS1TPPTS_{1}TP can be simplified to |c0|2ΦL2×L22V1L1()P|c_{0}|^{2}\|\Phi\|_{L^{2}\times L^{2}}^{-2}\|V_{1}\|_{L^{1}({\mathbb{R}})}P, which finishes the proof. ∎

Remark 4.9.

We appeal to the reader that each leading term in the expansion (4.43) plays an important role in revealing the cancellations among the finite rank operators that arise in the local decay estimate (1.14). Such a precise expression was also obtained for the one-dimensional Dirac operators in [EG21], even though the proof we give here is different. See Remark 3.7 in that paper. For the low-energy unweighted dispersive estimates, it is sufficient to work with the simpler expression

M(z)1=1zQΛ~0(z)Q+QΛ~1(z)+Λ~2(z)Q+zΛ~3(z),M(z)^{-1}=\frac{1}{z}Q\widetilde{\Lambda}_{0}(z)Q+Q\widetilde{\Lambda}_{1}(z)+\widetilde{\Lambda}_{2}(z)Q+z\widetilde{\Lambda}_{3}(z), (4.56)

where we absorb the operators S1,S1TP,PTS1,PS_{1},S_{1}TP,PTS_{1},P in (4.43) into the operators QΛ~0(z)QQ\widetilde{\Lambda}_{0}(z)Q, QΛ~1(z)Q\widetilde{\Lambda}_{1}(z), Λ~2(z)Q\widetilde{\Lambda}_{2}(z)Q, Λ~3(z)\widetilde{\Lambda}_{3}(z) respectively. The operators Λ~j(z)\widetilde{\Lambda}_{j}(z), for j=0,,3j=0,\ldots,3, satisfy the same estimates as (4.44).

5. Low energy estimates

In this section, we prove the low energy bounds for the perturbed evolution, following the ideas in Section 4 of [EG21]. We will frequently exploit the crucial orthogonality condition

e¯11v1(x)Q(x,y)dx=Q(x,y)v2(y)e¯11dy=𝟎2×2.\int_{\mathbb{R}}\underline{e}_{11}v_{1}(x)Q(x,y)\,\mathrm{d}x=\int_{\mathbb{R}}Q(x,y)v_{2}(y)\underline{e}_{11}\,\mathrm{d}y=\mathbf{0}_{2\times 2}. (5.1)

The following calculus lemma will be helpful for dealing with the lower entry of the free resolvent kernel.

Lemma 5.1.

For any m>0m>0 and r0r\geq 0, we define

gm(x):=erx2+m2x2+m2.g_{m}(x):=\frac{e^{-r\sqrt{x^{2}+m^{2}}}}{\sqrt{x^{2}+m^{2}}}. (5.2)

Then, there exists Cm>0C_{m}>0 (independent of rr) such that

xkgmL()Cm1,k=0,1,2.\|\partial_{x}^{k}\,g_{m}\|_{L^{\infty}({\mathbb{R}})}\leq C_{m}\lesssim 1,\quad\forall\ k=0,1,2. (5.3)
Proof.

First, by rescaling, we set gm(x)=1mg~(x/m)g_{m}(x)=\frac{1}{m}{\widetilde{g}}(x/m) where

g~(x):=ermx2+1x2+1=1er~xx,r~:=rm.{\widetilde{g}}(x):=\frac{e^{-rm\sqrt{x^{2}+1}}}{\sqrt{x^{2}+1}}=\frac{1}{e^{{\tilde{r}}\langle x\rangle}\langle x\rangle},\quad{\tilde{r}}:=rm. (5.4)

Hence, it sufficient to prove the same estimate (5.3) for g~(x){\widetilde{g}}(x). For k=0k=0, it is clear that |g~(x)|1|{\widetilde{g}}(x)|\leq 1 for all xx\in{\mathbb{R}}. For k=1,2k=1,2, direct computation shows that

xg~(x)=x(1+r~x)er~xx3,\partial_{x}\,{\widetilde{g}}(x)=-\frac{x(1+{\tilde{r}}\langle x\rangle)}{e^{{\tilde{r}}\langle x\rangle}\langle x\rangle^{3}}, (5.5)

and

x2g~(x)=3x2+3r~x2xx2+r~2x2x2r~x4er~xx5.\partial_{x}^{2}\,{\widetilde{g}}(x)=\frac{3x^{2}+3{\tilde{r}}x^{2}\langle x\rangle-\langle x\rangle^{2}+{\tilde{r}}^{2}x^{2}\langle x\rangle^{2}-{\tilde{r}}\langle x\rangle^{4}}{e^{{\tilde{r}}\langle x\rangle}\langle x\rangle^{5}}. (5.6)

Since er~xmax{1,r~,r~2}1e^{-{\tilde{r}}\langle x\rangle}\max\{1,{\tilde{r}},{\tilde{r}}^{2}\}\leq 1, it follows from (5.5), (5.6) that the estimate (5.3) holds for g~{\widetilde{g}} and thus for g(x)g(x) too. ∎

The next proposition establishes the dispersive estimates for the evolution semigroup eitPs+e^{it\mathcal{H}}P_{\mathrm{s}}^{+} for small energies close to the threshold μ\mu.

Proposition 5.2.

Let the assumptions of Theorem 1.4 hold. Let χ0(z)\chi_{0}(z) be a smooth, even, non-negative cut-off function satisfying χ0(z)=1\chi_{0}(z)=1 for |z|z02|z|\leq\frac{z_{0}}{2} and χ0(z)=0\chi_{0}(z)=0 for |z|z0|z|\geq z_{0}, where z0>0z_{0}>0 is given by Proposition 4.8. Then, for any |t|1|t|\geq 1, and u=(u1,u2)𝒮()×𝒮()\vec{u}=(u_{1},u_{2})\in\mathcal{S}({\mathbb{R}})\times\mathcal{S}({\mathbb{R}}), we have

eitχ0(μI)Ps+uL()×L()|t|12uL1()×L1(),\|e^{it\mathcal{H}}\chi_{0}(\mathcal{H}-\mu I)P_{\mathrm{s}}^{+}\vec{u}\|_{L^{\infty}({\mathbb{R}})\times L^{\infty}({\mathbb{R}})}\lesssim|t|^{-\frac{1}{2}}\|\vec{u}\|_{L^{1}({\mathbb{R}})\times L^{1}({\mathbb{R}})}, (5.7)

and

x2(eitχ0(μI)Ps+Ft+)uL()×L()|t|32x2uL1()×L1(),\|\langle x\rangle^{-2}(e^{it\mathcal{H}}\chi_{0}(\mathcal{H}-\mu I)P_{\mathrm{s}}^{+}-F_{t}^{+})\vec{u}\|_{L^{\infty}({\mathbb{R}})\times L^{\infty}({\mathbb{R}})}\lesssim|t|^{-\frac{3}{2}}\|\langle x\rangle^{2}\vec{u}\|_{L^{1}({\mathbb{R}})\times L^{1}({\mathbb{R}})}, (5.8)

where Ft+F_{t}^{+} is defined by

Ft+(x,y)=eitμ4πitΨ(x)[σ3Ψ(y)].F_{t}^{+}(x,y)=\frac{e^{it\mu}}{\sqrt{-4\pi it}}\vec{\Psi}(x)[\sigma_{3}\vec{\Psi}(y)]^{\top}. (5.9)

We begin with the proof of the dispersive decay estimate (5.7).

Proof of (5.7).

We recall the spectral representation from (3.15):

eitPs+=eitμπieitz2z0(z)dzeitμπieitz2z0(z)v1(M(z))1v20(z)dz.e^{it\mathcal{H}}P_{\mathrm{s}}^{+}=\frac{e^{it\mu}}{\pi i}\int_{{\mathbb{R}}}e^{itz^{2}}z\mathcal{R}_{0}(z)\,\mathrm{d}z-\frac{e^{it\mu}}{\pi i}\int_{{\mathbb{R}}}e^{itz^{2}}z\mathcal{R}_{0}(z)v_{1}(M(z))^{-1}v_{2}\mathcal{R}_{0}(z)\,\mathrm{d}z.

Note that the first term on the right is the spectral representation for the free evolution eit0Ps+e^{it\mathcal{H}_{0}}P_{\mathrm{s}}^{+} and it satisfies the same estimate as (5.7) thanks to Proposition 2.3. We insert the weaker expansion (4.56) for M(z)1M(z)^{-1} following Remark 4.9, and write

eitz2zχ0(z2)0(z)v1(M(z))1v20(z)dz=eitz2χ0(z2)0(z)v1QΛ~0(z)Qv20(z)dz+eitz2zχ0(z2)0(z)v1QΛ~1(z)v20(z)dz+eitz2zχ0(z2)0(z)v1Λ~2(z)Qv20(z)dz+eitz2z2χ0(z2)0(z)v1Λ~3(z)v20(z)dz=:J1+J2+J3+J4.\begin{split}&\int_{{\mathbb{R}}}e^{itz^{2}}z\chi_{0}(z^{2})\mathcal{R}_{0}(z)v_{1}(M(z))^{-1}v_{2}\mathcal{R}_{0}(z)\,\mathrm{d}z\\ &=\int_{{\mathbb{R}}}e^{itz^{2}}\chi_{0}(z^{2})\mathcal{R}_{0}(z)v_{1}Q\widetilde{\Lambda}_{0}(z)Qv_{2}\mathcal{R}_{0}(z)\,\mathrm{d}z+\int_{{\mathbb{R}}}e^{itz^{2}}z\chi_{0}(z^{2})\mathcal{R}_{0}(z)v_{1}Q\widetilde{\Lambda}_{1}(z)v_{2}\mathcal{R}_{0}(z)\,\mathrm{d}z\\ &\quad+\int_{{\mathbb{R}}}e^{itz^{2}}z\chi_{0}(z^{2})\mathcal{R}_{0}(z)v_{1}\widetilde{\Lambda}_{2}(z)Qv_{2}\mathcal{R}_{0}(z)\,\mathrm{d}z+\int_{{\mathbb{R}}}e^{itz^{2}}z^{2}\chi_{0}(z^{2})\mathcal{R}_{0}(z)v_{1}\widetilde{\Lambda}_{3}(z)v_{2}\mathcal{R}_{0}(z)\,\mathrm{d}z\\ &=:J_{1}+J_{2}+J_{3}+J_{4}.\end{split}

It remains to show that

JkL1LC|t|12,k=1,,4.\left\|J_{k}\right\|_{L^{1}\to L^{\infty}}\leq C|t|^{-\frac{1}{2}},\quad\forall\ k=1,\ldots,4. (5.10)

We treat the case for J1J_{1} since the other cases follow similarly. First, we recall the kernel of 0(z)\mathcal{R}_{0}(z) from (3.12) and write

0(z)(x,y):=1(z)(x,y)+2(z)(x,y):=ieiz|xy|2ze¯11+ez2+2μ|xy|2z2+2μe¯22,\mathcal{R}_{0}(z)(x,y):=\mathcal{R}_{1}(z)(x,y)+\mathcal{R}_{2}(z)(x,y):=\frac{ie^{iz|x-y|}}{2z}\underline{e}_{11}+\frac{-e^{-\sqrt{z^{2}+2\mu}|x-y|}}{2\sqrt{z^{2}+2\mu}}\underline{e}_{22}, (5.11)

and we further decompose the integral J1J_{1} as

J1=J1(1,1)+J1(1,2)+J1(2,1)+J1(2,2),J_{1}=J_{1}^{(1,1)}+J_{1}^{(1,2)}+J_{1}^{(2,1)}+J_{1}^{(2,2)},

where

J1(i,j)(x,y):=eitz2χ0(z2)[i(z)v1QΛ~0(z)Qv2j(z)](x,y)dz,i,j{1,2}.J_{1}^{(i,j)}(x,y):=\int_{\mathbb{R}}e^{itz^{2}}\chi_{0}(z^{2})[\mathcal{R}_{i}(z)v_{1}Q\widetilde{\Lambda}_{0}(z)Qv_{2}\mathcal{R}_{j}(z)](x,y)\,\mathrm{d}z,\quad i,j\in\{1,2\}.

We begin with the most singular term

J1(1,1)(x,y)=3eitz2+iz(|xx1|+|yy1|)χ0(z2)(2iz)2[e¯11v1QΛ~0(z)Qv2e¯11](x1,y1)dzdx1dy1.J_{1}^{(1,1)}(x,y)=\int_{{\mathbb{R}}^{3}}e^{itz^{2}+iz(|x-x_{1}|+|y-y_{1}|)}\frac{\chi_{0}(z^{2})}{(2iz)^{2}}[\underline{e}_{11}v_{1}Q\widetilde{\Lambda}_{0}(z)Qv_{2}\underline{e}_{11}](x_{1},y_{1})\,\mathrm{d}z\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}. (5.12)

The orthogonality conditions (5.1) imply that

eiz|x|e¯11v1(x1)Q(x1,x2)dx1=eiz|y|Q(y2,y1)v2(y1)e¯11dy1=𝟎.\int_{\mathbb{R}}e^{iz|x|}\underline{e}_{11}v_{1}(x_{1})Q(x_{1},x_{2})\,\mathrm{d}x_{1}=\int_{\mathbb{R}}e^{iz|y|}Q(y_{2},y_{1})v_{2}(y_{1})\underline{e}_{11}\,\mathrm{d}y_{1}=\mathbf{0}. (5.13)

Hence, writing

eiz|xx1|eiz|x|=iz|x||xx1|eizs1ds1andeiz|yy1|eiz|y|=iz|y||yy1|eizs2ds2,e^{iz|x-x_{1}|}-e^{iz|x|}=iz\int_{|x|}^{|x-x_{1}|}e^{izs_{1}}\,\mathrm{d}s_{1}\quad\text{and}\quad e^{iz|y-y_{1}|}-e^{iz|y|}=iz\int_{|y|}^{|y-y_{1}|}e^{izs_{2}}\,\mathrm{d}s_{2}, (5.14)

we obtain

J1(1,1)(x,y)=142|x||xx1||y||yy1|eitz2+iz(s1+s2)A(z,x1,y1)ds1ds2dx1dy1dz,J_{1}^{(1,1)}(x,y)=\frac{1}{4}\int_{{\mathbb{R}}^{2}}\int_{|x|}^{|x-x_{1}|}\int_{|y|}^{|y-y_{1}|}\int_{{\mathbb{R}}}e^{itz^{2}+iz(s_{1}+s_{2})}A(z,x_{1},y_{1})\,\mathrm{d}s_{1}\,\mathrm{d}s_{2}\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\,\mathrm{d}z,

where A(z,x1,y1)=χ0(z2)[e¯11v1QΛ~0(z)Qv2e¯11](x1,y1)A(z,x_{1},y_{1})=\chi_{0}(z^{2})[\underline{e}_{11}v_{1}Q\widetilde{\Lambda}_{0}(z)Qv_{2}\underline{e}_{11}](x_{1},y_{1}), and note that AA is differentiable and compactly supported in zz due to Proposition 4.8 and the compact support of χ0(z2)\chi_{0}(z^{2}). We obtain by Lemma 2.1 and the Fubini theorem that

|J1(1,1)(x,y)|C|t|122|x||xx1||y||yy1||zA(z,x1,x2)|dzds1ds2dx1dy1.\left|J_{1}^{(1,1)}(x,y)\right|\leq C|t|^{-\frac{1}{2}}\int_{{\mathbb{R}}^{2}}\int_{|x|}^{|x-x_{1}|}\int_{|y|}^{|y-y_{1}|}\int_{{\mathbb{R}}}\left|\partial_{z}A(z,x_{1},x_{2})\right|\,\mathrm{d}z\,\mathrm{d}s_{1}\,\mathrm{d}s_{2}\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}.

Using

|x||xx1||y||yy1|1ds1ds2||xx1||x||||yy1||y||x1y1,\int_{|x|}^{|x-x_{1}|}\int_{|y|}^{|y-y_{1}|}1\,\mathrm{d}s_{1}\,\mathrm{d}s_{2}\leq||x-x_{1}|-|x||\cdot||y-y_{1}|-|y||\lesssim\langle x_{1}\rangle\langle y_{1}\rangle, (5.15)

as well as

zA(z,x1,y1)=[e¯11v1Qz(χ0(z2)Λ0~(z))Qv2e¯11](x1,y1),\partial_{z}A(z,x_{1},y_{1})=[\underline{e}_{11}v_{1}Q\partial_{z}(\chi_{0}(z^{2})\widetilde{\Lambda_{0}}(z))Qv_{2}\underline{e}_{11}](x_{1},y_{1}), (5.16)

along with the bound (4.44) on Λ~0\widetilde{\Lambda}_{0}, we deduce that

2|x||xx1||y||yy1||zA(z,x1,x2)|dzds1ds2dx1dy1CQL2L22x1v1(x1)L2()y1v2(y1)L2()[z0,z0](|Λ~0(z)|L2×L2L2×L2+|zΛ~0(z)|L2×L2L2×L2)dz1.\begin{split}&\int_{{\mathbb{R}}^{2}}\int_{|x|}^{|x-x_{1}|}\int_{|y|}^{|y-y_{1}|}\int_{{\mathbb{R}}}\left|\partial_{z}A(z,x_{1},x_{2})\right|\,\mathrm{d}z\,\mathrm{d}s_{1}\,\mathrm{d}s_{2}\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\\ &\leq C\|Q\|_{L^{2}\to L^{2}}^{2}\|\langle x_{1}\rangle v_{1}(x_{1})\|_{L^{2}({\mathbb{R}})}\|\langle y_{1}\rangle v_{2}(y_{1})\|_{L^{2}({\mathbb{R}})}\\ &\qquad\cdot\int_{[-z_{0},z_{0}]}(\||\widetilde{\Lambda}_{0}(z)|\|_{L^{2}\times L^{2}\to L^{2}\times L^{2}}+\||\partial_{z}\widetilde{\Lambda}_{0}(z)|\|_{L^{2}\times L^{2}\to L^{2}\times L^{2}})\,\mathrm{d}z\\ &\lesssim 1.\end{split} (5.17)

Hence,

J1(1,1)L1×L1L×LC|t|12.\|J_{1}^{(1,1)}\|_{L^{1}\times L^{1}\rightarrow L^{\infty}\times L^{\infty}}\leq C|t|^{-\frac{1}{2}}.

Next, we consider the least singular term

J1(2,2)(x,y)=3eitz2B(z,x,y,x1,y1)dx1dy1dz,J_{1}^{(2,2)}(x,y)=\int_{{\mathbb{R}}^{3}}e^{itz^{2}}B(z,x,y,x_{1},y_{1})\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\,\mathrm{d}z, (5.18)

where

B(z,x,y,x1,y1):=ez2+2μ(|xx1|+|yy1|)χ0(z2)4(z2+2μ)[e¯22v1QΛ~0(z)Qv2e¯22](x1,y1).B(z,x,y,x_{1},y_{1}):=e^{-\sqrt{z^{2}+2\mu}(|x-x_{1}|+|y-y_{1}|)}\frac{\chi_{0}(z^{2})}{4(z^{2}+2\mu)}[\underline{e}_{22}v_{1}Q\widetilde{\Lambda}_{0}(z)Qv_{2}\underline{e}_{22}](x_{1},y_{1}). (5.19)

By Lemma 2.1, we have

|J1(2,2)(x,y)|C|t|12,|J_{1}^{(2,2)}(x,y)|\leq C|t|^{-\frac{1}{2}}, (5.20)

if we can show the uniform estimate

supx,y3|zB(z,x,y,x1,y1)|dzdx1dy11.\sup_{x,y\in{\mathbb{R}}}\int_{{\mathbb{R}}^{3}}|\partial_{z}B(z,x,y,x_{1},y_{1})|\,\mathrm{d}z\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\lesssim 1.

By Lemma 5.1, we have

supz|zk(ez2+2μ(|xx1|+|yy1|)4(z2+2μ))|Cμ1,k=0,1,\sup_{z\in{\mathbb{R}}}\left|\partial_{z}^{k}\left(\frac{e^{-\sqrt{z^{2}+2\mu}(|x-x_{1}|+|y-y_{1}|)}}{4(z^{2}+2\mu)}\right)\right|\leq C_{\mu}\lesssim 1,\quad k=0,1,

uniformly in the x,y,x1,y1x,y,x_{1},y_{1} variables. Hence, using the Cauchy-Schwarz inequality in the x1,y1x_{1},y_{1} variables and the bound (4.44) on Λ~0\widetilde{\Lambda}_{0}, we have

3|zB(z,x,y,x1,y1)|dzdx1dy1Cμ3|(1+z)χ0(z2)[e¯22v1QΛ~0(z)Qv2e¯22](x1,y1)|dzdx1dy1QL2×L2L2×L22v1L2()v2L2()[z0,z0](|Λ~0(z)|L2×L2L2×L2+|zΛ~0(z)|L2×L2L2×L2)dz1.\begin{split}&\int_{{\mathbb{R}}^{3}}|\partial_{z}B(z,x,y,x_{1},y_{1})|\,\mathrm{d}z\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\\ &\leq C_{\mu}\int_{{\mathbb{R}}^{3}}\left|(1+\partial_{z})\chi_{0}(z^{2})[\underline{e}_{22}v_{1}Q\widetilde{\Lambda}_{0}(z)Qv_{2}\underline{e}_{22}](x_{1},y_{1})\right|\,\mathrm{d}z\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\\ &\lesssim\|Q\|_{L^{2}\times L^{2}\to L^{2}\times L^{2}}^{2}\|v_{1}\|_{L^{2}({\mathbb{R}})}\|v_{2}\|_{L^{2}({\mathbb{R}})}\\ &\qquad\int_{[-z_{0},z_{0}]}\big{(}\||\widetilde{\Lambda}_{0}(z)|\|_{L^{2}\times L^{2}\to L^{2}\times L^{2}}+\||\partial_{z}\widetilde{\Lambda}_{0}(z)|\|_{L^{2}\times L^{2}\to L^{2}\times L^{2}}\big{)}\,\mathrm{d}z\\ &\lesssim 1.\end{split}

Hence, the bound (5.20) is proven. The remaining terms J1(1,2)J_{1}^{(1,2)} and J1(2,1)J_{1}^{(2,1)} can be treated similarly with the same techniques, while for the remaining cases J2,J3J_{2},J_{3}, and J4J_{4}, we use the additional powers of zz in place of the missing QQ operators to obtain the same bounds (5.10) as the term J1J_{1}. This finishes the proof of (5.7). ∎

Next, we turn to the proof of the low-energy weighted estimate (5.8).

Proof of (5.8).

Recall that the threshold resonance function Ψ=(Ψ1,Ψ2)\Psi=(\Psi_{1},\Psi_{2})^{\top} has been normalized in Theorem 1.4, which means that we need to carefully treat the constants relating to the function Φ\Phi where Φ:=v2Ψ\Phi:=v_{2}\Psi. By Lemma 4.5, note that Φ\Phi spans the subspace S1(L2()×L2())S_{1}(L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})). We define

η:=ΦL2()×L2()20,\eta:=\|\Phi\|_{L^{2}({\mathbb{R}})\times L^{2}({\mathbb{R}})}^{-2}\neq 0, (5.21)

so that S1(x,y)=ηΦ(x)Φ(y)S_{1}(x,y)=\eta\Phi(x)\Phi^{*}(y), and we fix the constants c0c_{0} and c1c_{1} defined by (4.21) and (4.26) respectively for this Φ\Phi. By Lemma 4.5, one finds the relation

2=limx(|Ψ1(x)|2+|Ψ1(x)|2)=2(|c0|2+|c1|2),2=\lim_{x\to\infty}(|\Psi_{1}(x)|^{2}+|\Psi_{1}(-x)|^{2})=2(|c_{0}|^{2}+|c_{1}|^{2}), (5.22)

by the polarization identity (c.f. (4.25)). Thus, the precise expansion (4.43) of M(z)1M(z)^{-1} from Proposition 4.8 simplifies to

M(z)1=i2ηzS1+1ηV1L1()PTS1+1ηV1L1()S1TP+(2iV1L1()+2|c0|2iV1L1())zP+QΛ0(z)Q+zQΛ1(z)+zΛ2(z)Q+z2Λ3(z),0<|z|<z0.\begin{split}M(z)^{-1}&=\frac{i}{2\eta z}S_{1}+\frac{1}{\eta\|V_{1}\|_{L^{1}({\mathbb{R}})}}PTS_{1}+\frac{1}{\eta\|V_{1}\|_{L^{1}({\mathbb{R}})}}S_{1}TP+\left(\frac{2i}{\|V_{1}\|_{L^{1}({\mathbb{R}})}}+\frac{2|c_{0}|^{2}}{i\|V_{1}\|_{L^{1}({\mathbb{R}})}}\right)zP\\ &\quad+Q\Lambda_{0}(z)Q+zQ\Lambda_{1}(z)+z\Lambda_{2}(z)Q+z^{2}\Lambda_{3}(z),\qquad 0<|z|<z_{0}.\end{split} (5.23)

We insert the above expression into the spectral representation of eitχ0(μI)Ps+e^{it\mathcal{H}}\chi_{0}(\mathcal{H}-\mu I)P_{\mathrm{s}}^{+}, and obtain that

eitχ0(μI)Ps+=eitμπieitz2zχ0(z2)0(z)dzeitμπieitz2zχ0(z2)0(z)v1(M(z))1v20(z)dz=eitμπiI1eitμπi(i2ηI2,1+1ηV1L1()I2,2+1ηV1L1()I2,3+(2iV1L1()+2|c0|2iV1L1())I2,4)eitμπi(I3,1+I3,2+I3,3+I3,4),\begin{split}&e^{it\mathcal{H}}\chi_{0}(\mathcal{H}-\mu I)P_{\mathrm{s}}^{+}\\ &=\frac{e^{it\mu}}{\pi i}\int_{{\mathbb{R}}}e^{itz^{2}}z\chi_{0}(z^{2})\mathcal{R}_{0}(z)\,\mathrm{d}z-\frac{e^{it\mu}}{\pi i}\int_{{\mathbb{R}}}e^{itz^{2}}z\chi_{0}(z^{2})\mathcal{R}_{0}(z)v_{1}(M(z))^{-1}v_{2}\mathcal{R}_{0}(z)\,\mathrm{d}z\\ &=\frac{e^{it\mu}}{\pi i}I_{1}\\ &\quad-\frac{e^{it\mu}}{\pi i}\left(\frac{i}{2\eta}I_{2,1}+\frac{1}{\eta\|V_{1}\|_{L^{1}({\mathbb{R}})}}I_{2,2}+\frac{1}{\eta\|V_{1}\|_{L^{1}({\mathbb{R}})}}I_{2,3}+\left(\frac{2i}{\|V_{1}\|_{L^{1}({\mathbb{R}})}}+\frac{2|c_{0}|^{2}}{i\|V_{1}\|_{L^{1}({\mathbb{R}})}}\right)I_{2,4}\right)\\ &\quad-\frac{e^{it\mu}}{\pi i}\left(I_{3,1}+I_{3,2}+I_{3,3}+I_{3,4}\right),\end{split} (5.24)

where

I1:=eitz2zχ0(z2)0(z)dz,\displaystyle I_{1}:=\int_{\mathbb{R}}e^{itz^{2}}z\chi_{0}(z^{2})\mathcal{R}_{0}(z)\,\mathrm{d}z, (5.25)
I2,1:=eitz2χ0(z2)[0(z)v1S1v20(z)]dz,\displaystyle I_{2,1}:=\int_{\mathbb{R}}e^{itz^{2}}\chi_{0}(z^{2})[\mathcal{R}_{0}(z)v_{1}S_{1}v_{2}\mathcal{R}_{0}(z)]\,\mathrm{d}z, (5.26)
I2,2:=eitz2zχ0(z2)[0(z)v1S1TPv20(z)]dz,\displaystyle I_{2,2}:=\int_{\mathbb{R}}e^{itz^{2}}z\chi_{0}(z^{2})[\mathcal{R}_{0}(z)v_{1}S_{1}TPv_{2}\mathcal{R}_{0}(z)]\,\mathrm{d}z, (5.27)
I2,3:=eitz2zχ0(z2)[0(z)v1PTS1v20(z)]dz,\displaystyle I_{2,3}:=\int_{\mathbb{R}}e^{itz^{2}}z\chi_{0}(z^{2})[\mathcal{R}_{0}(z)v_{1}PTS_{1}v_{2}\mathcal{R}_{0}(z)]\,\mathrm{d}z, (5.28)
I2,4:=eitz2z2χ0(z2)[0(z)v1Pv20(z)]dz,\displaystyle I_{2,4}:=\int_{\mathbb{R}}e^{itz^{2}}z^{2}\chi_{0}(z^{2})[\mathcal{R}_{0}(z)v_{1}Pv_{2}\mathcal{R}_{0}(z)]\,\mathrm{d}z, (5.29)

and

I3,1:=eitz2zχ0(z2)[0(z)v1QΛ0(z)Qv20(z)]dz,\displaystyle I_{3,1}:=\int_{\mathbb{R}}e^{itz^{2}}z\chi_{0}(z^{2})[\mathcal{R}_{0}(z)v_{1}Q\Lambda_{0}(z)Qv_{2}\mathcal{R}_{0}(z)]\,\mathrm{d}z, (5.30)
I3,2:=eitz2z2χ0(z2)[0(z)v1QΛ1(z)v20(z)]dz,\displaystyle I_{3,2}:=\int_{\mathbb{R}}e^{itz^{2}}z^{2}\chi_{0}(z^{2})[\mathcal{R}_{0}(z)v_{1}Q\Lambda_{1}(z)v_{2}\mathcal{R}_{0}(z)]\,\mathrm{d}z, (5.31)
I3,3:=eitz2z2χ0(z2)[0(z)v1Λ2(z)Qv20(z)]dz,\displaystyle I_{3,3}:=\int_{\mathbb{R}}e^{itz^{2}}z^{2}\chi_{0}(z^{2})[\mathcal{R}_{0}(z)v_{1}\Lambda_{2}(z)Qv_{2}\mathcal{R}_{0}(z)]\,\mathrm{d}z, (5.32)
I3,4:=eitz2z3χ0(z2)[0(z)v1Λ3(z)v20(z)]dz.\displaystyle I_{3,4}:=\int_{\mathbb{R}}e^{itz^{2}}z^{3}\chi_{0}(z^{2})[\mathcal{R}_{0}(z)v_{1}\Lambda_{3}(z)v_{2}\mathcal{R}_{0}(z)]\,\mathrm{d}z. (5.33)

Now we study the local decay of the terms I1I_{1}, I2,jI_{2,j}, I3,I_{3,\ell}, for j,{1,,4}j,\ell\in\{1,\ldots,4\} and we will observe in the following propositions that the terms I1,I2,1,,I2,4I_{1},I_{2,1},\ldots,I_{2,4} contribute to the leading order for the local decay estimate while the remainder terms I3,1,,I3,4I_{3,1},\ldots,I_{3,4} satisfy the stronger local decay estimate 𝒪(|t|32xy)\mathcal{O}(|t|^{-\frac{3}{2}}\langle x\rangle\langle y\rangle). We first handle these remainder terms by Lemma 2.1 in a similar spirit to the proof for the (unweighted) dispersive bound (5.7), exploiting the additional power of zz.

Proposition 5.3.

For i{1,2,,4}i\in\{1,2,\ldots,4\} and |t|1|t|\geq 1, we have

|I3,i(x,y)|C|t|32xy.|I_{3,i}(x,y)|\leq C|t|^{-\frac{3}{2}}\langle x\rangle\langle y\rangle. (5.34)
Proof.

We treat the case for I3,1I_{3,1} as the other cases follow similarly by using the additional powers of zz in place of the missing operators QQ. As before, we consider the decomposition

I3,1=I3,1(1,1)+I3,1(1,2)+I3,1(2,1)+I3,1(2,2),I_{3,1}=I_{3,1}^{(1,1)}+I_{3,1}^{(1,2)}+I_{3,1}^{(2,1)}+I_{3,1}^{(2,2)},

where

I3,1(i,j):=eitz2zχ0(z2)[i(z)v1QΛ0(z)Qv2j(z)]dz,i,j{1,2},I_{3,1}^{(i,j)}:=\int_{{\mathbb{R}}}e^{itz^{2}}z\chi_{0}(z^{2})[\mathcal{R}_{i}(z)v_{1}Q\Lambda_{0}(z)Qv_{2}\mathcal{R}_{j}(z)]\,\mathrm{d}z,\quad i,j\in\{1,2\},

with 1\mathcal{R}_{1} and 2\mathcal{R}_{2} defined in (5.11). We begin with the term

I3,1(1,1)(x,y)=3eitz2+iz(|xx1|+|yy1|)zχ0(z2)(2iz)2[e¯11v1QΛ0(z)Qv2e¯11](x1,y1)dzdx1dy1.I_{3,1}^{(1,1)}(x,y)=\int_{{\mathbb{R}}^{3}}e^{itz^{2}+iz(|x-x_{1}|+|y-y_{1}|)}\frac{z\chi_{0}(z^{2})}{(2iz)^{2}}[\underline{e}_{11}v_{1}Q{\Lambda}_{0}(z)Qv_{2}\underline{e}_{11}](x_{1},y_{1})\,\mathrm{d}z\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}.

Using the orthogonality condition (5.1) like in (5.14), we obtain

I3,1(1,1)(x,y)=142|x||xx1||y||yy1|eitz2+iz(s1+s2)zA(z,x1,y1)dzds1ds2dx1dy1,\begin{split}I_{3,1}^{(1,1)}(x,y)&=\frac{1}{4}\int_{{\mathbb{R}}^{2}}\int_{|x|}^{|x-x_{1}|}\int_{|y|}^{|y-y_{1}|}\int_{{\mathbb{R}}}e^{itz^{2}+iz(s_{1}+s_{2})}zA(z,x_{1},y_{1})\,\mathrm{d}z\,\mathrm{d}s_{1}\,\mathrm{d}s_{2}\,\mathrm{d}x_{1}\,\mathrm{d}y_{1},\end{split}

where A(z,x1,y1):=χ0(z2)[e¯11v1QΛ0(z)v2Qe¯11](x1,y1)A(z,x_{1},y_{1}):=\chi_{0}(z^{2})[\underline{e}_{11}v_{1}Q\Lambda_{0}(z)v_{2}Q\underline{e}_{11}](x_{1},y_{1}). By Lemma 2.1, we obtain that

|I3,1(1,1)(x,y)||t|322|x||xx1||y||yy1|[z0,z0](|z2A|+(s1+s2)|zA|+|A|)dzds1ds2dx1dy1.\begin{split}&\left|I_{3,1}^{(1,1)}(x,y)\right|\\ &\lesssim|t|^{-\frac{3}{2}}\int_{{\mathbb{R}}^{2}}\int_{|x|}^{|x-x_{1}|}\int_{|y|}^{|y-y_{1}|}\int_{[-z_{0},z_{0}]}\big{(}|\partial_{z}^{2}A|+(s_{1}+s_{2})|\partial_{z}A|+|A|\big{)}\,\mathrm{d}z\,\mathrm{d}s_{1}\,\mathrm{d}s_{2}\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}.\end{split} (5.35)

Using the bounds

|x||xx1||y||yy1|1ds1ds2x1y1,|x||xx1||y||yy1|(s1+s2)ds1ds2x12y12xy,\begin{split}&\int_{|x|}^{|x-x_{1}|}\int_{|y|}^{|y-y_{1}|}1\,\mathrm{d}s_{1}\,\mathrm{d}s_{2}\lesssim\langle x_{1}\rangle\langle y_{1}\rangle,\\ &\int_{|x|}^{|x-x_{1}|}\int_{|y|}^{|y-y_{1}|}(s_{1}+s_{2})\,\mathrm{d}s_{1}\,\mathrm{d}s_{2}\lesssim\langle x_{1}\rangle^{2}\langle y_{1}\rangle^{2}\langle x\rangle\langle y\rangle,\end{split} (5.36)

we have

|I3,1(1,1)(x,y)||t|322[|z|z0]x1y1(|z2A|+x1y1xy|zA|+|A|)dzdx1dy1.\left|I_{3,1}^{(1,1)}(x,y)\right|\lesssim|t|^{-\frac{3}{2}}\int_{{\mathbb{R}}^{2}}\int_{[|z|\leq z_{0}]}\langle x_{1}\rangle\langle y_{1}\rangle(|\partial_{z}^{2}A|+\langle x_{1}\rangle\langle y_{1}\rangle\langle x\rangle\langle y\rangle|\partial_{z}A|+|A|)\,\mathrm{d}z\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}.

Noting that xv1(x1)\langle x\rangle v_{1}(x_{1}) and y1v2(y1)\langle y_{1}\rangle v_{2}(y_{1}) are in L2L^{2} and that Λ0\Lambda_{0} satisfies the bound (4.44), we apply Cauchy-Schwarz inequality in x1x_{1} and y1y_{1} variables to obtain the bound

|I3,1(1,1)(x,y)||t|32QL2L22x1v1Lx12()y1v2Ly12()[|z|z0](|z2Λ0(z)|L2×L2L2×L2+|Λ0(z)|L2×L2L2×L2)dz+|t|32xyQL2L22x1v1Lx12()y1v2Ly12()[|z|z0]|zΛ0(z)|L2×L2L2×L2dz|t|32xy.\begin{split}|I_{3,1}^{(1,1)}(x,y)|&\lesssim|t|^{-\frac{3}{2}}\|Q\|_{L^{2}\to L^{2}}^{2}\|\langle x_{1}\rangle v_{1}\|_{L_{x_{1}}^{2}({\mathbb{R}})}\|\langle y_{1}\rangle v_{2}\|_{L_{y_{1}}^{2}({\mathbb{R}})}\\ &\qquad\cdot\int_{[|z|\leq z_{0}]}(\||\partial_{z}^{2}\Lambda_{0}(z)|\|_{L^{2}\times L^{2}\to L^{2}\times L^{2}}+\||\Lambda_{0}(z)|\|_{L^{2}\times L^{2}\to L^{2}\times L^{2}})\,\mathrm{d}z\\ &\quad+|t|^{-\frac{3}{2}}\langle x\rangle\langle y\rangle\|Q\|_{L^{2}\to L^{2}}^{2}\|\langle x_{1}\rangle v_{1}\|_{L_{x_{1}}^{2}({\mathbb{R}})}\|\langle y_{1}\rangle v_{2}\|_{L_{y_{1}}^{2}({\mathbb{R}})}\\ &\qquad\cdot\int_{[|z|\leq z_{0}]}\||\partial_{z}\Lambda_{0}(z)|\|_{L^{2}\times L^{2}\to L^{2}\times L^{2}}\,\mathrm{d}z\\ &\lesssim|t|^{-\frac{3}{2}}\langle x\rangle\langle y\rangle.\end{split} (5.37)

Next, we consider the term

I3,1(1,2)(x,y)=3eitz2+iz|xx1|z2+2μ|yy1|χ0(z2)4iz2+2μ[e¯11v1QΛ0(z)Qv2e¯22](x1,y1)dzdx1dy1.I_{3,1}^{(1,2)}(x,y)=\int_{{\mathbb{R}}^{3}}e^{itz^{2}+iz|x-x_{1}|-\sqrt{z^{2}+2\mu}|y-y_{1}|}\frac{\chi_{0}(z^{2})}{4i\sqrt{z^{2}+2\mu}}[\underline{e}_{11}v_{1}Q{\Lambda}_{0}(z)Qv_{2}\underline{e}_{22}](x_{1},y_{1})\,\mathrm{d}z\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}.

By using the QQ orthogonality (c.f. (5.1)) condition, we write

I3,1(1,2)(x,y)=3|x||xx1|eitz2+izs1zB(z,x1,y1,x,y)ds1dzdx1dy1,\begin{split}I_{3,1}^{(1,2)}(x,y)=\int_{{\mathbb{R}}^{3}}\int_{|x|}^{|x-x_{1}|}e^{itz^{2}+izs_{1}}zB(z,x_{1},y_{1},x,y)\,\mathrm{d}s_{1}\,\mathrm{d}z\,\mathrm{d}x_{1}\,\mathrm{d}y_{1},\end{split} (5.38)

where

B(z,x1,y1,x,y):=ez2+2μ|yy1|4iz2+2μχ0(z2)[e¯11v1QΛ0(z)Qv2e¯22](x1,y1).B(z,x_{1},y_{1},x,y):=\frac{e^{-\sqrt{z^{2}+2\mu}|y-y_{1}|}}{4i\sqrt{z^{2}+2\mu}}\chi_{0}(z^{2})[\underline{e}_{11}v_{1}Q{\Lambda}_{0}(z)Qv_{2}\underline{e}_{22}](x_{1},y_{1}). (5.39)

Since BB is compactly supported in zz, we can exchange the order of integration and we use Lemma 2.1 to obtain

|I3,1(1,2)(x,y)|C|t|322|x||xx1||[z2+is1z]B(z,x1,y1,x,y)|dzds1dx1dy1.\left|I_{3,1}^{(1,2)}(x,y)\right|\leq C|t|^{-\frac{3}{2}}\int_{{\mathbb{R}}^{2}}\int_{|x|}^{|x-x_{1}|}\int_{{\mathbb{R}}}|[\partial_{z}^{2}+is_{1}\partial_{z}]B(z,x_{1},y_{1},x,y)|\,\mathrm{d}z\,\mathrm{d}s_{1}\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}.

By Lemma 5.1, we have

supz|zk(ez2+2μ|yy1|4iz2+2μ)|Cμ1,k=0,1,2,\sup_{z\in{\mathbb{R}}}\left|\partial_{z}^{k}\left(\tfrac{e^{-\sqrt{z^{2}+2\mu}|y-y_{1}|}}{4i\sqrt{z^{2}+2\mu}}\right)\right|\leq C_{\mu}\lesssim 1,\quad\forall k=0,1,2, (5.40)

which implies by Hölder’s inequality and Leibniz rule that

|[z2+is1z]B(z,x1,y1,x,y)|dzCs1|e¯11v1Q[1+z+z2](χ0(z2)Λ0(z))Qv2e¯22|dz.\begin{split}&\int_{{\mathbb{R}}}\left|[\partial_{z}^{2}+is_{1}\partial_{z}]B(z,x_{1},y_{1},x,y)\right|\,\mathrm{d}z\\ &\quad\leq C\langle s_{1}\rangle\int_{{\mathbb{R}}}\left|\underline{e}_{11}v_{1}Q[1+\partial_{z}+\partial_{z}^{2}](\chi_{0}(z^{2}){\Lambda}_{0}(z))Qv_{2}\underline{e}_{22}\right|\,\mathrm{d}z.\end{split}

Repeating the arguments from (5.35)–(5.37), we obtain

|I3,1(1,2)(x,y)|C|t|32x.\left|I_{3,1}^{(1,2)}(x,y)\right|\leq C|t|^{-\frac{3}{2}}\langle x\rangle.

Similarly, one has the bounds

|I3,1(2,1)(x,y)|C|t|32y,|I3,1(2,2)(x,y)|C|t|32,\left|I_{3,1}^{(2,1)}(x,y)\right|\leq C|t|^{-\frac{3}{2}}\langle y\rangle,\qquad\left|I_{3,1}^{(2,2)}(x,y)\right|\leq C|t|^{-\frac{3}{2}},

and we are done. ∎

Proposition 5.4.

For all |t|1|t|\geq 1, we have

|I2,1(x,y)Ft1(x,y)|C|t|32x2y2,\left|I_{2,1}(x,y)-F_{t}^{1}(x,y)\right|\leq C|t|^{-\frac{3}{2}}\langle x\rangle^{2}\langle y\rangle^{2}, (5.41)

where

Ft1(x,y):=ηπit[c0e¯1Ψ(x)][σ3Ψ(y)c0e¯1].F_{t}^{1}(x,y):=\frac{\eta\sqrt{\pi}}{{\sqrt{-it}}}[c_{0}\underline{e}_{1}-\Psi(x)][\sigma_{3}\Psi(y)-{c_{0}}\underline{e}_{1}]^{*}. (5.42)
Proof.

As in the previous propositions, we decompose I2,1I_{2,1} into the sum

I2,1=I2,1(1,1)+I2,1(1,2)+I2,1(2,1)+I2,1(2,2),I_{2,1}=I_{2,1}^{(1,1)}+I_{2,1}^{(1,2)}+I_{2,1}^{(2,1)}+I_{2,1}^{(2,2)},

with

I2,1(i,j):=eitz2χ0(z2)[i(z)v1S1v2j(z)]dz,i,j{1,2}.I_{2,1}^{(i,j)}:=\int_{\mathbb{R}}e^{itz^{2}}\chi_{0}(z^{2})[\mathcal{R}_{i}(z)v_{1}S_{1}v_{2}\mathcal{R}_{j}(z)]\,\mathrm{d}z,\quad i,j\in\{1,2\}.

We start with the most singular term

I2,1(1,1)(x,y)=3eitz2+iz(|xx1|+|yy1|)χ0(z2)(2iz)2[e¯11v1S1v2e¯11](x1,y1)dx1dy1dz.I_{2,1}^{(1,1)}(x,y)=\int_{{\mathbb{R}}^{3}}e^{itz^{2}+iz(|x-x_{1}|+|y-y_{1}|)}\frac{\chi_{0}(z^{2})}{(2iz)^{2}}[\underline{e}_{11}v_{1}S_{1}v_{2}\underline{e}_{11}](x_{1},y_{1})\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\,\mathrm{d}z.

Noting that S1L2QL2S_{1}L^{2}\subset QL^{2}, the orthogonality conditions (5.1) imply that

eiz|x|e¯11v1(x1)S1(x1,x2)dx1=eiz|y|S1(y2,y1)v2(y1)e¯11dy1=𝟎2×2,x,y.\int_{\mathbb{R}}e^{iz|x|}\underline{e}_{11}v_{1}(x_{1})S_{1}(x_{1},x_{2})\,\mathrm{d}x_{1}=\int_{\mathbb{R}}e^{iz|y|}S_{1}(y_{2},y_{1})v_{2}(y_{1})\underline{e}_{11}\,\mathrm{d}y_{1}=\mathbf{0}_{2\times 2},\quad\forall x,y\in{\mathbb{R}}. (5.43)

Hence, by the Fubini theorem,

I2,1(1,1)(x,y)=142|x||xx1||y||yy1|eitz2+iz(s1+s2)χ0(z2)[e¯11v1S1v2e¯11](x1,y1)dzds1ds2dx1dy1=14|x||xx1||y||yy1|Gt(s1+s2)ds1ds22[e¯11v1S1v2e¯11](x1,y1)dx1dy1,\begin{split}I_{2,1}^{(1,1)}(x,y)&=\frac{1}{4}\int_{{\mathbb{R}}^{2}}\int_{|x|}^{|x-x_{1}|}\int_{|y|}^{|y-y_{1}|}\int_{{\mathbb{R}}}e^{itz^{2}+iz(s_{1}+s_{2})}\chi_{0}(z^{2})[\underline{e}_{11}v_{1}S_{1}v_{2}\underline{e}_{11}](x_{1},y_{1})\,\mathrm{d}z\,\mathrm{d}s_{1}\,\mathrm{d}s_{2}\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\\ &=\frac{1}{4}\int_{|x|}^{|x-x_{1}|}\int_{|y|}^{|y-y_{1}|}G_{t}(s_{1}+s_{2})\,\mathrm{d}s_{1}\,\mathrm{d}s_{2}\int_{{\mathbb{R}}^{2}}[\underline{e}_{11}v_{1}S_{1}v_{2}\underline{e}_{11}](x_{1},y_{1})\,\mathrm{d}x_{1}\,\mathrm{d}y_{1},\end{split}

where Gt()G_{t}(\cdot) is the function defined in Lemma 2.2, which satisfies the estimate

|Gt(s1+s2)πiteis124teis224t|C|t|32s1s2.\left|G_{t}(s_{1}+s_{2})-\frac{\sqrt{\pi}}{{\sqrt{-it}}}e^{-i\frac{s_{1}^{2}}{4t}}e^{-i\frac{s_{2}^{2}}{4t}}\right|\leq C|t|^{-\frac{3}{2}}\langle s_{1}\rangle\langle s_{2}\rangle. (5.44)

Using the bound

|x||xx1||y||yy1|s1s2ds1ds2x12y12xy,\int_{|x|}^{|x-x_{1}|}\int_{|y|}^{|y-y_{1}|}\langle s_{1}\rangle\langle s_{2}\rangle\,\mathrm{d}s_{1}\,\mathrm{d}s_{2}\lesssim\langle x_{1}\rangle^{2}\langle y_{1}\rangle^{2}\langle x\rangle\langle y\rangle, (5.45)

the decay assumptions on v1,v2v_{1},v_{2}, and the estimate (5.44), we have

|I2,1(1,1)(x,y)π4iteiπ42Ht(x1,x)[e¯11v1S1v2e¯11](x1,y1)Ht(y1,y)dx1dy1|C|t|32xyS1L2×L2L2×L2x12v1(x1)L2y12v2(y2)L2C|t|32xy,\begin{split}&\left|I_{2,1}^{(1,1)}(x,y)-\frac{\sqrt{\pi}}{{4\sqrt{-it}}}e^{i\frac{\pi}{4}}\int_{{\mathbb{R}}^{2}}H_{t}(x_{1},x)[\underline{e}_{11}v_{1}S_{1}v_{2}\underline{e}_{11}](x_{1},y_{1})H_{t}(y_{1},y)\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\right|\\ &\quad\leq C|t|^{-\frac{3}{2}}\langle x\rangle\langle y\rangle\|S_{1}\|_{L^{2}\times L^{2}\to L^{2}\times L^{2}}\|\langle x_{1}\rangle^{2}v_{1}(x_{1})\|_{L^{2}}\|\langle y_{1}\rangle^{2}v_{2}(y_{2})\|_{L^{2}}\leq C|t|^{-\frac{3}{2}}\langle x\rangle\langle y\rangle,\end{split}

where we set

Ht(x1,x):=|x||x1x|eis24tds.H_{t}(x_{1},x):=\int_{|x|}^{|x_{1}-x|}e^{-i\frac{s^{2}}{4t}}\,\mathrm{d}s. (5.46)

Since S1(x,y)=ηΦ(x)Φ(y)S_{1}(x,y)=\eta\Phi(x)\Phi^{*}(y), the orthogonality conditions (5.1) imply that

|x|e¯11v1(x1)S1(x1,y1)dx1=η|x|e¯11v1(x1)Φ(x1)dx1Φ(y1)=𝟎2×2,y,|y|S1(x1,y1)v2(y1)e¯11dy1=ηΦ(x1)|y|Φ(y1)v2(y1)e¯11dy1=𝟎2×2,x.\begin{split}\int_{{\mathbb{R}}}|x|\underline{e}_{11}v_{1}(x_{1})S_{1}(x_{1},y_{1})\,\mathrm{d}x_{1}&=\eta\int_{\mathbb{R}}|x|\underline{e}_{11}v_{1}(x_{1})\Phi(x_{1})\,\mathrm{d}x_{1}\Phi^{*}(y_{1})=\bm{0}_{2\times 2},\quad\forall y\in{\mathbb{R}},\\ \int_{{\mathbb{R}}}|y|S_{1}(x_{1},y_{1})v_{2}(y_{1})\underline{e}_{11}\,\mathrm{d}y_{1}&=\eta\Phi(x_{1})\int_{\mathbb{R}}|y|\Phi^{*}(y_{1})v_{2}(y_{1})\underline{e}_{11}\,\mathrm{d}y_{1}=\bm{0}_{2\times 2},\quad\forall x\in{\mathbb{R}}.\end{split}

Hence, using the bound

|Ht(x1,x)(|xx1||x|)|C|t|1x2x13,\begin{split}|H_{t}(x_{1},x)-(|x-x_{1}|-|x|)|\leq C|t|^{-1}\langle x\rangle^{2}\langle x_{1}\rangle^{3},\end{split} (5.47)

and the exponential decay of v1,v2v_{1},v_{2}, we conclude the estimate

|I2,1(1,1)(x,y)ηπit[G0(e¯11v1Φ)(x)][G0(e¯11v2Φ)(y)]|C|t|32x2y2,\begin{split}\left|I_{2,1}^{(1,1)}(x,y)-\frac{\eta\sqrt{\pi}}{{\sqrt{-it}}}[G_{0}(\underline{e}_{11}v_{1}\Phi)(x)][G_{0}(\underline{e}_{11}v_{2}\Phi)(y)]^{*}\right|\leq C|t|^{-\frac{3}{2}}\langle x\rangle^{2}\langle y\rangle^{2},\end{split} (5.48)

where

G0(x,y):=12|xy|,G_{0}(x,y):=-\frac{1}{2}|x-y|, (5.49)

and

[G0(e¯11v1Φ)(x)]:=12|xx1|e¯11v1(x1)Φ(x1)dx1,[G0(e¯11v2Φ)(y)]:=12|yy1|Φ(y1)v2(y1)e¯11dy1.\begin{split}&[G_{0}(\underline{e}_{11}v_{1}\Phi)(x)]:=-\frac{1}{2}\int_{{\mathbb{R}}}|x-x_{1}|\underline{e}_{11}v_{1}(x_{1})\Phi(x_{1})\,\mathrm{d}x_{1},\\ &[G_{0}(\underline{e}_{11}v_{2}\Phi)(y)]^{*}:=-\frac{1}{2}\int_{{\mathbb{R}}}|y-y_{1}|\Phi^{*}(y_{1})v_{2}(y_{1})\underline{e}_{11}\,\mathrm{d}y_{1}.\end{split}

In the preceding definition, we used the identity v2=v2v_{2}^{*}=v_{2}. Next, we treat the term

I2,1(2,2)(x,y)=3eitz2χ0(z2)ez2+2μ|xx1|2z2+2μ[e¯22v1S1v2e¯22](x1,y1)ez2+2μ|yy1|2z2+2μdx1dy1dz.I_{2,1}^{(2,2)}(x,y)=\int_{{\mathbb{R}}^{3}}e^{itz^{2}}\chi_{0}(z^{2})\frac{e^{-\sqrt{z^{2}+2\mu}|x-x_{1}|}}{-2\sqrt{z^{2}+2\mu}}[\underline{e}_{22}v_{1}S_{1}v_{2}\underline{e}_{22}](x_{1},y_{1})\frac{e^{-\sqrt{z^{2}+2\mu}|y-y_{1}|}}{-2\sqrt{z^{2}+2\mu}}\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\,\mathrm{d}z.

By Taylor expansion, we have

I2,1(2,2)(x,y)=3eitz2χ0(z2)e2μ|xx1|22μ[e¯22v1S1v2e¯22](x1,y1)e2μ|yy1|22μdx1dy1dz+3eitz2z2χ0(z2)[e¯22v1S1v2e¯22](x1,y1)κ(x,x1)κ(y,y1)dx1dy1dz=ηeitz2χ0(z2)dz[G2(e¯22v1Φ)(x)][G2(e¯22v2Φ)(y)]+3eitz2z2χ0(z2)[e¯22v1S1v2e¯22](x1,y1)κ(x,x1)κ(y,y1)dx1dy1dz,\begin{split}I_{2,1}^{(2,2)}(x,y)&=\int_{{\mathbb{R}}^{3}}e^{itz^{2}}\chi_{0}(z^{2})\frac{e^{-\sqrt{2\mu}|x-x_{1}|}}{-2\sqrt{2\mu}}[\underline{e}_{22}v_{1}S_{1}v_{2}\underline{e}_{22}](x_{1},y_{1})\frac{e^{-\sqrt{2\mu}|y-y_{1}|}}{-2\sqrt{2\mu}}\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\,\mathrm{d}z\\ &\quad+\int_{{\mathbb{R}}^{3}}e^{itz^{2}}z^{2}\chi_{0}(z^{2})[\underline{e}_{22}v_{1}S_{1}v_{2}\underline{e}_{22}](x_{1},y_{1})\kappa(x,x_{1})\kappa(y,y_{1})\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\,\mathrm{d}z\\ &=\eta\int_{\mathbb{R}}e^{itz^{2}}\chi_{0}(z^{2})\,\mathrm{d}z[G_{2}(\underline{e}_{22}v_{1}\Phi)(x)][G_{2}(\underline{e}_{22}v_{2}\Phi)(y)]^{*}\\ &\quad+\int_{{\mathbb{R}}^{3}}e^{itz^{2}}z^{2}\chi_{0}(z^{2})[\underline{e}_{22}v_{1}S_{1}v_{2}\underline{e}_{22}](x_{1},y_{1})\kappa(x,x_{1})\kappa(y,y_{1})\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\,\mathrm{d}z,\end{split} (5.50)

where we set

G2(x,y):=e2μ|xy|22μ,G_{2}(x,y):=\frac{e^{-\sqrt{2\mu}|x-y|}}{-2\sqrt{2\mu}}, (5.51)

and where κ(x,x1)κ(y,y1)\kappa(x,x_{1})\kappa(y,y_{1}) is an error term bounded by Cxx1yy1ec(|xx1|+|yy1|)C\langle x\rangle\langle x_{1}\rangle\langle y\rangle\langle y_{1}\rangle e^{-c(|x-x_{1}|+|y-y_{1}|)}, for some C,c>0C,c>0, (c.f. (4.7)). The definitions for G2(e¯22v1Φ)(x)G_{2}(\underline{e}_{22}v_{1}\Phi)(x) and G2(e¯22v2Φ)(y)G_{2}(\underline{e}_{22}v_{2}\Phi)(y) are defined analogously to the ones for G0(e¯11v1Φ)(x)G_{0}(\underline{e}_{11}v_{1}\Phi)(x) and G0(e¯11v2Φ)(y)G_{0}(\underline{e}_{11}v_{2}\Phi)(y). By non-stationary phase, one has the uniform estimate

|eitz2z2χ0(z2)dz|C|t|32.\left|\int_{\mathbb{R}}e^{itz^{2}}z^{2}\chi_{0}(z^{2})\,\mathrm{d}z\right|\leq C|t|^{-\frac{3}{2}}. (5.52)

Hence, we can control the remainder term in I2,1(2,2)I_{2,1}^{(2,2)} by

|3eitz2z2χ0(z2)[e¯22v1S1v2e¯22](x1,y1)κ(x,x1)κ(y,y1)dx1dy1dz|C|t|32xy.\left|\int_{{\mathbb{R}}^{3}}e^{itz^{2}}z^{2}\chi_{0}(z^{2})[\underline{e}_{22}v_{1}S_{1}v_{2}\underline{e}_{22}](x_{1},y_{1})\kappa(x,x_{1})\kappa(y,y_{1})\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\,\mathrm{d}z\right|\leq C|t|^{-\frac{3}{2}}\langle x\rangle\langle y\rangle. (5.53)

On the other hand, by Lemma 2.2, one has

eitz2χ0(z2)dz=πit+Rt,|Rt|C|t|32.\int_{{\mathbb{R}}}e^{itz^{2}}\chi_{0}(z^{2})\,\mathrm{d}z=\frac{\sqrt{\pi}}{{\sqrt{-it}}}+R_{t},\quad|R_{t}|\leq C|t|^{-\frac{3}{2}}.

Hence, the leading contribution of I2,1(2,2)I_{2,1}^{(2,2)} can be written as

|eitz2χ0(z2)dz[G2(e¯22v1Φ)(x)][G2(e¯22v2Φ)(y)]ηπit[G2(e¯22v1Φ)(x)][G2(e¯22v2Φ)(y)]|C|t|32.\left|\int_{\mathbb{R}}e^{itz^{2}}\chi_{0}(z^{2})\,\mathrm{d}z[G_{2}(\underline{e}_{22}v_{1}\Phi)(x)][G_{2}(\underline{e}_{22}v_{2}\Phi)(y)]^{*}-\frac{\eta\sqrt{\pi}}{{\sqrt{-it}}}[G_{2}(\underline{e}_{22}v_{1}\Phi)(x)][G_{2}(\underline{e}_{22}v_{2}\Phi)(y)]^{*}\right|\leq C|t|^{-\frac{3}{2}}.

Thus, one concludes the estimate for I2,1(2,2)I_{2,1}^{(2,2)}:

|I2,1(2,2)ηπit[G2(e¯22v1Φ)(x)][G2(e¯22v2Φ)(y)]|C|t|32xy.\left|I_{2,1}^{(2,2)}-\frac{\eta\sqrt{\pi}}{{\sqrt{-it}}}[G_{2}(\underline{e}_{22}v_{1}\Phi)(x)][G_{2}(\underline{e}_{22}v_{2}\Phi)(y)]^{*}\right|\leq C|t|^{-\frac{3}{2}}\langle x\rangle\langle y\rangle. (5.54)

Finally, we note that a similar analysis holds for the terms I2,1(1,2)I_{2,1}^{(1,2)} and I2,1(2,1)I_{2,1}^{(2,1)} yielding the contributions

|I2,1(1,2)ηπit[G0(e¯11v1Φ)(x)][G2(e¯22v2Φ)(y)]|C|t|32x2y,|I2,1(2,1)ηπit[G2(e¯22v1Φ)(x)][G0(e¯11v2Φ)(y)]|C|t|32xy2.\begin{split}&\left|I_{2,1}^{(1,2)}-\frac{\eta\sqrt{\pi}}{{\sqrt{-it}}}[G_{0}(\underline{e}_{11}v_{1}\Phi)(x)][G_{2}(\underline{e}_{22}v_{2}\Phi)(y)]^{*}\right|\leq C|t|^{-\frac{3}{2}}\langle x\rangle^{2}\langle y\rangle,\\ &\left|I_{2,1}^{(2,1)}-\frac{\eta\sqrt{\pi}}{{\sqrt{-it}}}[G_{2}(\underline{e}_{22}v_{1}\Phi)(x)][G_{0}(\underline{e}_{11}v_{2}\Phi)(y)]^{*}\right|\leq C|t|^{-\frac{3}{2}}\langle x\rangle\langle y\rangle^{2}.\end{split} (5.55)

By adding all leading order contributions, we obtain

Ft1(x,y)=ηπit[(G0e¯11+G2e¯22)v1Φ](x)[(G0e¯11+G2e¯22)v2Φ](y).F_{t}^{1}(x,y)=\frac{\eta\sqrt{\pi}}{{\sqrt{-it}}}[(G_{0}\underline{e}_{11}+G_{2}\underline{e}_{22})v_{1}\Phi](x)[(G_{0}\underline{e}_{11}+G_{2}\underline{e}_{22})v_{2}\Phi]^{*}(y).

Recalling that 𝒢0=G0e¯11+G2e¯22\mathcal{G}_{0}=G_{0}\underline{e}_{11}+G_{2}\underline{e}_{22} from Lemma 4.1, that 𝒢0(v1Φ)=c0e¯1Ψ\mathcal{G}_{0}(v_{1}\Phi)=c_{0}\underline{e}_{1}-\Psi from Lemma 4.5, and that 𝒢0(v2Φ)=σ3Ψc0e¯1\mathcal{G}_{0}(v_{2}\Phi)=\sigma_{3}\Psi-c_{0}\underline{e}_{1} from Remark 4.6 (c.f. (4.40)), we arrive at

Ft1(x,y)=ηπit[c0e¯1Ψ(x)][σ3Ψ(y)c0e¯1],F_{t}^{1}(x,y)=\frac{\eta\sqrt{\pi}}{{\sqrt{-it}}}[c_{0}\underline{e}_{1}-\Psi(x)][\sigma_{3}\Psi(y)-{c_{0}}\underline{e}_{1}]^{*},

as claimed ∎

We continue the analysis for the terms involving the operators S1TPS_{1}TP and PTS1PTS_{1}.

Proposition 5.5.

For all |t|1|t|\geq 1, we have

|I2,2(x,y)Ft2(x,y)|C|t|32x2y2,|I_{2,2}(x,y)-F_{t}^{2}(x,y)|\leq C|t|^{-\frac{3}{2}}\langle x\rangle^{2}\langle y\rangle^{2}, (5.56)
|I2,3(x,y)Ft3(x,y)|C|t|32x2y2,|I_{2,3}(x,y)-F_{t}^{3}(x,y)|\leq C|t|^{-\frac{3}{2}}\langle x\rangle^{2}\langle y\rangle^{2}, (5.57)

where

Ft2(x,y):=iηV1L1()2πit[c0e¯1Ψ(x)][eiy24tc0e¯1],F_{t}^{2}(x,y):=\frac{i\eta\|V_{1}\|_{L^{1}({\mathbb{R}})}}{2}\frac{\sqrt{\pi}}{{\sqrt{-it}}}[c_{0}\underline{e}_{1}-\Psi(x)][e^{i\frac{y^{2}}{4t}}{c_{0}}\underline{e}_{1}]^{*}, (5.58)
Ft3(x,y):=iηV1L1()2πit[eix24tc0e¯1][σ3Ψ(y)c0e¯1].F_{t}^{3}(x,y):=-\frac{i\eta\|V_{1}\|_{L^{1}({\mathbb{R}})}}{2}\frac{\sqrt{\pi}}{{\sqrt{-it}}}[e^{-i\frac{x^{2}}{4t}}c_{0}\underline{e}_{1}][\sigma_{3}\Psi(y)-{c_{0}}\underline{e}_{1}]^{*}. (5.59)
Proof.

As in the proof of Proposition 5.4, we decompose I2,2I_{2,2} into

I2,2=I2,2(1,1)+I2,2(1,2)+I2,2(2,1)+I2,2(2,2),I_{2,2}=I_{2,2}^{(1,1)}+I_{2,2}^{(1,2)}+I_{2,2}^{(2,1)}+I_{2,2}^{(2,2)},

with

I2,2(i,j):=eitz2zχ0(z2)[i(z)v1S1TPv2j(z)]dz,i,j{1,2},I_{2,2}^{(i,j)}:=\int_{{\mathbb{R}}}e^{itz^{2}}z\chi_{0}(z^{2})[\mathcal{R}_{i}(z)v_{1}S_{1}TPv_{2}\mathcal{R}_{j}(z)]\,\mathrm{d}z,\quad i,j\in\{1,2\},

where 1\mathcal{R}_{1} and 2\mathcal{R}_{2} were defined in (5.11). We start with

I2,2(1,1)(x,y)=3eitz2zχ0(z2)eiz|xx1|2iz[e¯11v1S1TPv2e¯11](x1,y1)eiz|yy1|2izdx1dy1dz.I_{2,2}^{(1,1)}(x,y)=\int_{{\mathbb{R}}^{3}}e^{itz^{2}}z\chi_{0}(z^{2})\frac{e^{iz|x-x_{1}|}}{-2iz}[\underline{e}_{11}v_{1}S_{1}TPv_{2}\underline{e}_{11}](x_{1},y_{1})\frac{e^{iz|y-y_{1}|}}{-2iz}\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\,\mathrm{d}z.

Using the orthogonality (5.43), we have

I2,2(1,1)(x,y)=143|x||xx1||y||yy1|eitz2+iz(s1+s2)zχ0(z2)[e¯11v1S1TPv2e¯11](x1,y1)ds1ds2dx1dy1dz+14i3|x||xx1|eitz2+izs1χ0(z2)[e¯11v1S1TPv2e¯11](x1,y1)eiz|y|ds1dx1dy1dz=:I2,2;1(1,1)+I2,2;2(1,1).\begin{split}I_{2,2}^{(1,1)}(x,y)&=\frac{1}{4}\int_{{\mathbb{R}}^{3}}\int_{|x|}^{|x-x_{1}|}\int_{|y|}^{|y-y_{1}|}e^{itz^{2}+iz(s_{1}+s_{2})}z\chi_{0}(z^{2})[\underline{e}_{11}v_{1}S_{1}TPv_{2}\underline{e}_{11}](x_{1},y_{1})\,\mathrm{d}s_{1}\,\mathrm{d}s_{2}\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\,\mathrm{d}z\\ &\quad+\frac{1}{4i}\int_{{\mathbb{R}}^{3}}\int_{|x|}^{|x-x_{1}|}e^{itz^{2}+izs_{1}}\chi_{0}(z^{2})[\underline{e}_{11}v_{1}S_{1}TPv_{2}\underline{e}_{11}](x_{1},y_{1})e^{iz|y|}\,\mathrm{d}s_{1}\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\,\mathrm{d}z\\ &=:I_{2,2;1}^{(1,1)}+I_{2,2;2}^{(1,1)}.\end{split}

By Lemma 2.1, we have

|eitz2+iz(s1+s2)zχ0(z2)dz|C|t|32s1s2.\left|\int_{{\mathbb{R}}}e^{itz^{2}+iz(s_{1}+s_{2})}z\chi_{0}(z^{2})\,\mathrm{d}z\right|\leq C|t|^{-\frac{3}{2}}\langle s_{1}\rangle\langle s_{2}\rangle.

Using this estimate, the bound

|x||xx1||y||yy1|s1s2ds1ds2x12y22xy,\int_{|x|}^{|x-x_{1}|}\int_{|y|}^{|y-y_{1}|}\langle s_{1}\rangle\langle s_{2}\rangle\,\mathrm{d}s_{1}\,\mathrm{d}s_{2}\lesssim\langle x_{1}\rangle^{2}\langle y_{2}\rangle^{2}\langle x\rangle\langle y\rangle,

the absolute boundedness of S1TPS_{1}TP, and the exponential decay of v1,v2v_{1},v_{2}, we deduce that

|I2,2;1(1,1)(x,y)||t|32xy2|x12y22[e¯11v1S1TPv2e¯11](x1,y1)|dx1dy1|t|32xy.\begin{split}\left|I_{2,2;1}^{(1,1)}(x,y)\right|\lesssim|t|^{-\frac{3}{2}}\langle x\rangle\langle y\rangle\int_{{\mathbb{R}}^{2}}|\langle x_{1}\rangle^{2}\langle y_{2}\rangle^{2}[\underline{e}_{11}v_{1}S_{1}TPv_{2}\underline{e}_{11}](x_{1},y_{1})|\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\lesssim|t|^{-\frac{3}{2}}\langle x\rangle\langle y\rangle.\end{split} (5.60)

By Lemma 4.5 and direct computation,

S1TP(x1,y1)v2(y1)e¯11dy1=ηV1L1()Φ(x1)[c0e¯1].\int_{{\mathbb{R}}}S_{1}TP(x_{1},y_{1})v_{2}(y_{1})\underline{e}_{11}\,\mathrm{d}y_{1}=\eta\|V_{1}\|_{L^{1}({\mathbb{R}})}\Phi(x_{1})[c_{0}\underline{e}_{1}]^{*}. (5.61)

Hence, integrating in y1y_{1}, we have

I2,2;2(1,1)(x,y)=ηV1L1()4i(|x||xx1|eitz2+iz(s1+|y|)χ0(z2)e¯11v1(x1)Φ(x1)dzds1dx1)[c0e¯1]=ηV1L1()4i(|x||xx1|Gt(s1+|y|)ds1e¯11v1(x1)Φ(x1)dx1)[c0e¯1],\begin{split}I_{2,2;2}^{(1,1)}(x,y)&=\frac{\eta\|V_{1}\|_{L^{1}({\mathbb{R}})}}{4i}\left(\int_{{\mathbb{R}}}\int_{|x|}^{|x-x_{1}|}\int_{{\mathbb{R}}}e^{itz^{2}+iz(s_{1}+|y|)}\chi_{0}(z^{2})\underline{e}_{11}v_{1}(x_{1})\Phi(x_{1})\,\mathrm{d}z\,\mathrm{d}s_{1}\,\mathrm{d}x_{1}\right)[c_{0}\underline{e}_{1}]^{*}\\ &=\frac{\eta\|V_{1}\|_{L^{1}({\mathbb{R}})}}{4i}\left(\int_{{\mathbb{R}}}\int_{|x|}^{|x-x_{1}|}G_{t}(s_{1}+|y|)\,\mathrm{d}s_{1}\underline{e}_{11}v_{1}(x_{1})\Phi(x_{1})\,\mathrm{d}x_{1}\right)[c_{0}\underline{e}_{1}]^{*},\end{split}

where GtG_{t} is the function defined in Lemma 2.2. By Lemma 2.2 (c.f. (5.44)–(5.48) for similar computations), we have

|I2,2;2(1,1)(x,y)iηV1L1()2[G0(e¯11v1Φ)(x)][eiy24tc0e¯1]|C|t|32x2y2,\left|I_{2,2;2}^{(1,1)}(x,y)-\frac{i\eta\|V_{1}\|_{L^{1}({\mathbb{R}})}}{2}[G_{0}(\underline{e}_{11}v_{1}\Phi)(x)][e^{i\frac{y^{2}}{4t}}{c_{0}}\underline{e}_{1}]^{*}\right|\leq C|t|^{-\frac{3}{2}}\langle x\rangle^{2}\langle y\rangle^{2},

where G0G_{0} is the operator defined in (5.49). This completes the analysis of the term I2,2(1,1)I_{2,2}^{(1,1)}. Next, we treat the term

I2,2(2,1)(x,y)=3eitz2zχ0(z2)ez2+2μ|xx1|2z2+2μ[e¯22v1S1TPv2e¯11](x1,y1)eiz|yy1|2izdx1dy1dz.\begin{split}I_{2,2}^{(2,1)}(x,y)=\int_{{\mathbb{R}}^{3}}e^{itz^{2}}z\chi_{0}(z^{2})\frac{e^{-\sqrt{z^{2}+2\mu}|x-x_{1}|}}{-2\sqrt{z^{2}+2\mu}}[\underline{e}_{22}v_{1}S_{1}TPv_{2}\underline{e}_{11}](x_{1},y_{1})\frac{e^{iz|y-y_{1}|}}{-2iz}\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\,\mathrm{d}z.\end{split} (5.62)

By inserting eiz|y|e^{iz|y|}, we write

I2,2(2,1)(x,y)=123eitz2zχ0(z2)ez2+2μ|xx1|2z2+2μ[e¯22v1S1TPv2e¯11](x1,y1)|y||yy1|eizs2ds2dx1dy1dz+3eitz2zχ0(z2)ez2+2μ|xx1|2z2+2μ[e¯22v1S1TPv2e¯11](x1,y1)eiz|y|2izdx1dy1dz=:I2,2;1(2,1)(x,y)+I2,2;2(2,1)(x,y),\begin{split}I_{2,2}^{(2,1)}(x,y)&=-\frac{1}{2}\int_{{\mathbb{R}}^{3}}e^{itz^{2}}z\chi_{0}(z^{2})\frac{e^{-\sqrt{z^{2}+2\mu}|x-x_{1}|}}{-2\sqrt{z^{2}+2\mu}}[\underline{e}_{22}v_{1}S_{1}TPv_{2}\underline{e}_{11}](x_{1},y_{1})\int_{|y|}^{|y-y_{1}|}e^{izs_{2}}\,\mathrm{d}s_{2}\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\,\mathrm{d}z\\ &\quad+\int_{{\mathbb{R}}^{3}}e^{itz^{2}}z\chi_{0}(z^{2})\frac{e^{-\sqrt{z^{2}+2\mu}|x-x_{1}|}}{-2\sqrt{z^{2}+2\mu}}[\underline{e}_{22}v_{1}S_{1}TPv_{2}\underline{e}_{11}](x_{1},y_{1})\frac{e^{iz|y|}}{-2iz}\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\,\mathrm{d}z\\ &=:I_{2,2;1}^{(2,1)}(x,y)+I_{2,2;2}^{(2,1)}(x,y),\end{split}

where I2,2;2(2,1)I_{2,2;2}^{(2,1)} is the leading term. By Lemma 2.1 and Lemma 5.1,

|eitz2+izs2zχ0(z2)ez2+2μ|xx1|2z2+2μdz|C|t|32s2.\left|\int_{{\mathbb{R}}}e^{itz^{2}+izs_{2}}z\chi_{0}(z^{2})\frac{e^{-\sqrt{z^{2}+2\mu}|x-x_{1}|}}{-2\sqrt{z^{2}+2\mu}}\,\mathrm{d}z\right|\leq C|t|^{-\frac{3}{2}}\langle s_{2}\rangle. (5.63)

Hence, using the absolute boundedness of S1TPS_{1}TP and the bound (5.45), we have

|I2,2;1(2,1)(x,y)||t|322y12y[e¯22v1S1TPv2e¯11](x1,y1)dx1dy1|t|32y.\begin{split}\left|I_{2,2;1}^{(2,1)}(x,y)\right|\lesssim|t|^{-\frac{3}{2}}\int_{{\mathbb{R}}^{2}}\langle y_{1}\rangle^{2}\langle y\rangle[\underline{e}_{22}v_{1}S_{1}TPv_{2}\underline{e}_{11}](x_{1},y_{1})\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\lesssim|t|^{-\frac{3}{2}}\langle y\rangle.\end{split}

On the other hand, we treat I2,2;1(2,1)I_{2,2;1}^{(2,1)} similarly as in (5.50) - (5.53) and find that

|I2,2;2(2,1)(x,y)i23eitz2+iz|y|χ0(z2)G2(x,x1)[e¯22v1S1TPv2e¯11](x1,y1)dx1dy1dz|C|t|32xy,\begin{split}\left|I_{2,2;2}^{(2,1)}(x,y)-\frac{i}{2}\int_{{\mathbb{R}}^{3}}e^{itz^{2}+iz|y|}\chi_{0}(z^{2})G_{2}(x,x_{1})[\underline{e}_{22}v_{1}S_{1}TPv_{2}\underline{e}_{11}](x_{1},y_{1})\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\,\mathrm{d}z\right|\leq C|t|^{-\frac{3}{2}}\langle x\rangle\langle y\rangle,\end{split}

where G2G_{2} is defined in (5.51). Hence, by Lemma 2.2 and (5.61), we conclude that

|I2,2(2,1)(x,y)iηV1L1()2πit[G2(e¯22v1Φ)(x)][eiy24tc0e¯1]|C|t|32xy.\left|I_{2,2}^{(2,1)}(x,y)-\frac{i\eta\|V_{1}\|_{L^{1}({\mathbb{R}})}}{2}\frac{\sqrt{\pi}}{{\sqrt{-it}}}[G_{2}(\underline{e}_{22}v_{1}\Phi)(x)][e^{i\frac{y^{2}}{4t}}{c_{0}}\underline{e}_{1}]^{*}\right|\leq C|t|^{-\frac{3}{2}}\langle x\rangle\langle y\rangle. (5.64)

Finally, we show that the terms I2,2(1,2)I_{2,2}^{(1,2)} and I2,2(2,2)I_{2,2}^{(2,2)} satisfy the better decay rates of 𝒪(|t|32xy)\mathcal{O}(|t|^{-\frac{3}{2}}\langle x\rangle\langle y\rangle). By orthogonality (c.f. (5.43)),

I2,2(1,2)(x,y)=123eitz2zχ0(z2)|x||xx1|eizs1ds1[e¯11v1S1TPv2e¯22](x1,y1)ez2+2μ|yy1|2z2+2μdx1dy1dz.\begin{split}&I_{2,2}^{(1,2)}(x,y)\\ &=\frac{1}{-2}\int_{{\mathbb{R}}^{3}}e^{itz^{2}}z\chi_{0}(z^{2})\int_{|x|}^{|x-x_{1}|}e^{izs_{1}}\,\mathrm{d}s_{1}[\underline{e}_{11}v_{1}S_{1}TPv_{2}\underline{e}_{22}](x_{1},y_{1})\frac{e^{-\sqrt{z^{2}+2\mu}|y-y_{1}|}}{-2\sqrt{z^{2}+2\mu}}\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\,\mathrm{d}z.\end{split}

By Lemma 2.1 and Lemma 5.1, we note that the zz-integral satisfy the bound

|eitz2+izs1zχ0(z2)ez2+2μ|yy1|2z2+2μdz|C|t|32s1.\left|\int_{{\mathbb{R}}}e^{itz^{2}+izs_{1}}z\chi_{0}(z^{2})\frac{e^{-\sqrt{z^{2}+2\mu}|y-y_{1}|}}{-2\sqrt{z^{2}+2\mu}}\,\mathrm{d}z\right|\leq C|t|^{-\frac{3}{2}}\langle s_{1}\rangle.

Hence, by the absolute boundedness of S1TPS_{1}TP and decay of v1v_{1}, v2v_{2}, we conclude that

|I2,2(1,2)(x,y)|C|t|32x.\left|I_{2,2}^{(1,2)}(x,y)\right|\leq C|t|^{-\frac{3}{2}}\langle x\rangle.

The analysis of I2,2(2,2)I_{2,2}^{(2,2)} is analogous to the preceeding one, yielding the bound

|I2,2(2,2)(x,y)|C|t|32y.\left|I_{2,2}^{(2,2)}(x,y)\right|\leq C|t|^{-\frac{3}{2}}\langle y\rangle.

Thus, using 𝒢0=G0e¯11+G2e¯22\mathcal{G}_{0}=G_{0}\underline{e}_{11}+G_{2}\underline{e}_{22}, and 𝒢0(v1Φ)=c0e¯1Ψ\mathcal{G}_{0}(v_{1}\Phi)=c_{0}\underline{e}_{1}-\Psi from Lemma 4.5, we conclude (5.56) and (5.58). For the estimate (5.57) involving I2,3I_{2,3}, one should instead use the identity

e¯11v1(x1)PTS1(x1,y1)dx1=ηV1L1()c0e¯1Φ(y1),\int_{{\mathbb{R}}}\underline{e}_{11}v_{1}(x_{1})PTS_{1}(x_{1},y_{1})\,\mathrm{d}x_{1}=-\eta\|V_{1}\|_{L^{1}({\mathbb{R}})}c_{0}\underline{e}_{1}\Phi(y_{1})^{*}, (5.65)

and we leave the remaining details to the reader. ∎

Next, we remark that the analysis for I2,4I_{2,4} involving the operator PP leads to a similar estimate as the free evolution in Proposition 2.3.

Proposition 5.6.

For all |t|1|t|\geq 1, we have

|I2,4(x,y)Ft4(x,y)|C|t|32x2y2,\left|I_{2,4}(x,y)-F_{t}^{4}(x,y)\right|\leq C|t|^{-\frac{3}{2}}\langle x\rangle^{2}\langle y\rangle^{2}, (5.66)

where

Ft4(x,y):=V1L1()4πiteix24te¯1eiy24te¯1.F_{t}^{4}(x,y):=\frac{\|V_{1}\|_{L^{1}({\mathbb{R}})}}{4}\frac{\sqrt{\pi}}{{\sqrt{-it}}}e^{-i\frac{x^{2}}{4t}}\underline{e}_{1}e^{-i\frac{y^{2}}{4t}}\underline{e}_{1}^{\top}. (5.67)
Proof.

As before, we write

I2,4=I2,4(1,1)+I2,4(1,2)+I2,4(2,1)+I2,4(2,2),I_{2,4}=I_{2,4}^{(1,1)}+I_{2,4}^{(1,2)}+I_{2,4}^{(2,1)}+I_{2,4}^{(2,2)},

with

I2,4(i,j):=eitz2z2χ0(z2)[i(z)v1Pv2j(z)]dz,i,j{1,2},I_{2,4}^{(i,j)}:=\int_{{\mathbb{R}}}e^{itz^{2}}z^{2}\chi_{0}(z^{2})[\mathcal{R}_{i}(z)v_{1}Pv_{2}\mathcal{R}_{j}(z)]\,\mathrm{d}z,\quad i,j\in\{1,2\},

where 1\mathcal{R}_{1} and 2\mathcal{R}_{2} were defined in (5.11). We first treat the leading term

I2,4(1,1)(x,y)=eitz2z2χ0(z2)eiz|xx1|2iz[e¯11v1Pv2e¯11](x1,y1)eiz|yy1|2izdx1dy1dz.I_{2,4}^{(1,1)}(x,y)=\int_{{\mathbb{R}}}e^{itz^{2}}z^{2}\chi_{0}(z^{2})\frac{e^{iz|x-x_{1}|}}{2iz}[\underline{e}_{11}v_{1}Pv_{2}\underline{e}_{11}](x_{1},y_{1})\frac{e^{iz|y-y_{1}|}}{2iz}\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\,\mathrm{d}z. (5.68)

By adding and subtracting eiz|x|e^{iz|x|} and eiz|y|e^{iz|y|} twice, we further consider

I2,4(1,1)(x,y)=3eitz2z2χ0(z2)eiz|x|2iz[e¯11v1Pv2e¯11](x1,y1)eiz|y|2izdx1dy1dz+123eitz2z2χ0(z2)eiz|x|2iz[e¯11v1Pv2e¯11](x1,y1)|y||yy1|eizs2ds2dx1dy1dz+123eitz2z2χ0(z2)|x||xx1|eizs1𝑑s1[e¯11v1Pv2e¯11](x1,y1)eiz|y|2izdx1dy1dz+143eitz2z2χ0(z2)|x||xx1|eizs1𝑑s1[e¯11v1Pv2e¯11](x1,y1)|y||yy1|eizs2ds2dx1dy1dz=:I2,4;1(1,1)(x,y)+I2,4;2(1,1)(x,y)+I2,4;3(1,1)(x,y)+I2,4;4(1,1)(x,y).\begin{split}I_{2,4}^{(1,1)}(x,y)&=\int_{{\mathbb{R}}^{3}}e^{itz^{2}}z^{2}\chi_{0}(z^{2})\frac{e^{iz|x|}}{2iz}[\underline{e}_{11}v_{1}Pv_{2}\underline{e}_{11}](x_{1},y_{1})\frac{e^{iz|y|}}{2iz}\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\,\mathrm{d}z\\ &\quad+\frac{1}{2}\int_{{\mathbb{R}}^{3}}e^{itz^{2}}z^{2}\chi_{0}(z^{2})\frac{e^{iz|x|}}{2iz}[\underline{e}_{11}v_{1}Pv_{2}\underline{e}_{11}](x_{1},y_{1})\int_{|y|}^{|y-y_{1}|}e^{izs_{2}}\,\mathrm{d}s_{2}\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\,\mathrm{d}z\\ &\quad+\frac{1}{2}\int_{{\mathbb{R}}^{3}}e^{itz^{2}}z^{2}\chi_{0}(z^{2})\int_{|x|}^{|x-x_{1}|}e^{izs_{1}}ds_{1}[\underline{e}_{11}v_{1}Pv_{2}\underline{e}_{11}](x_{1},y_{1})\frac{e^{iz|y|}}{2iz}\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\,\mathrm{d}z\\ &\quad+\frac{1}{4}\int_{{\mathbb{R}}^{3}}e^{itz^{2}}z^{2}\chi_{0}(z^{2})\int_{|x|}^{|x-x_{1}|}e^{izs_{1}}ds_{1}[\underline{e}_{11}v_{1}Pv_{2}\underline{e}_{11}](x_{1},y_{1})\int_{|y|}^{|y-y_{1}|}e^{izs_{2}}\,\mathrm{d}s_{2}\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\,\mathrm{d}z\\ &=:I_{2,4;1}^{(1,1)}(x,y)+I_{2,4;2}^{(1,1)}(x,y)+I_{2,4;3}^{(1,1)}(x,y)+I_{2,4;4}^{(1,1)}(x,y).\end{split}

By direct computation,

2[e¯11v1Pv2e¯11](x1,y1)dx1dy1=V1L1()e¯1e¯1.\int_{{\mathbb{R}}^{2}}[\underline{e}_{11}v_{1}Pv_{2}\underline{e}_{11}](x_{1},y_{1})\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}=-\|V_{1}\|_{L^{1}({\mathbb{R}})}\underline{e}_{1}\underline{e}_{1}^{\top}.

Hence, by Lemma 2.2,

|I2,4;1(1,1)(x,y)V1L1()4πiteix24te¯1eiy24te¯1|C|t|32xy.\left|I_{2,4;1}^{(1,1)}(x,y)-\frac{\|V_{1}\|_{L^{1}({\mathbb{R}})}}{4}\frac{\sqrt{\pi}}{{\sqrt{-it}}}e^{-i\frac{x^{2}}{4t}}\underline{e}_{1}e^{-i\frac{y^{2}}{4t}}\underline{e}_{1}^{\top}\right|\leq C|t|^{-\frac{3}{2}}\langle x\rangle\langle y\rangle.

For the terms I2,4;2(1,1)I_{2,4;2}^{(1,1)}, I2,4;3(1,1)I_{2,4;3}^{(1,1)}, the additional factor of zz allows to invoke Lemma 2.1,

|eitz2+iz(|x|+s2)zχ0(z2)dz|C|t|32xs2,|eitz2+iz(s1+|y|)zχ0(z2)dz|C|t|32ys1.\begin{split}\left|\int_{{\mathbb{R}}}e^{itz^{2}+iz(|x|+s_{2})}z\chi_{0}(z^{2})\,\mathrm{d}z\right|\leq C|t|^{-\frac{3}{2}}\langle x\rangle\langle s_{2}\rangle,\\ \left|\int_{{\mathbb{R}}}e^{itz^{2}+iz(s_{1}+|y|)}z\chi_{0}(z^{2})\,\mathrm{d}z\right|\leq C|t|^{-\frac{3}{2}}\langle y\rangle\langle s_{1}\rangle.\end{split}

Thus, we infer from the exponential decay of v1v_{1} and v2v_{2} that

|I2,4;2(1,1)(x,y)|+|I2,4;3(1,1)(x,y)|C|t|32xy.\left|I_{2,4;2}^{(1,1)}(x,y)\right|+\left|I_{2,4;3}^{(1,1)}(x,y)\right|\leq C|t|^{-\frac{3}{2}}\langle x\rangle\langle y\rangle. (5.69)

For the term I2,4;4(1,1)I_{2,4;4}^{(1,1)}, we can use non-stationary phase to conclude the same bound. Hence, we have

|I2,4(1,1)(x,y)V1L1()4πiteix24te¯1eiy24te¯1|C|t|32xy.\left|I_{2,4}^{(1,1)}(x,y)-\frac{\|V_{1}\|_{L^{1}({\mathbb{R}})}}{4}\frac{\sqrt{\pi}}{{\sqrt{-it}}}e^{-i\frac{x^{2}}{4t}}\underline{e}_{1}e^{-i\frac{y^{2}}{4t}}\underline{e}_{1}^{\top}\right|\leq C|t|^{-\frac{3}{2}}\langle x\rangle\langle y\rangle.

Thus, it remains to prove that the other terms I2,4(1,2)I_{2,4}^{(1,2)}, I2,4(2,1)I_{2,4}^{(2,1)}, I2,4(2,2)I_{2,4}^{(2,2)} have the better 𝒪(|t|32xy)\mathcal{O}(|t|^{-\frac{3}{2}}\langle x\rangle\langle y\rangle) weighted decay estimate to finish the proposition. We first treat the term

I2,4(1,2)(x,y)=12i3eitz2zχ0(z2)eiz|xx1|[e¯11v1Pv2e¯22](x1,y1)ez2+2μ|yy1|2z2+2μdx1dy1dz.\begin{split}I_{2,4}^{(1,2)}(x,y)&=\frac{1}{2i}\int_{{\mathbb{R}}^{3}}e^{itz^{2}}z\chi_{0}(z^{2})e^{iz|x-x_{1}|}[\underline{e}_{11}v_{1}Pv_{2}\underline{e}_{22}](x_{1},y_{1})\frac{e^{-\sqrt{z^{2}+2\mu}|y-y_{1}|}}{-2\sqrt{z^{2}+2\mu}}\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\,\mathrm{d}z.\\ \end{split} (5.70)

By Lemma 2.1 and Lemma 5.1,

|eitz2+iz(|xx1|)zχ0(z2)ez2+2μ|yy1|2z2+2μdz|C|t|32xx1.\left|\int_{{\mathbb{R}}}e^{itz^{2}+iz(|x-x_{1}|)}z\chi_{0}(z^{2})\frac{e^{-\sqrt{z^{2}+2\mu}|y-y_{1}|}}{-2\sqrt{z^{2}+2\mu}}\,\mathrm{d}z\right|\leq C|t|^{-\frac{3}{2}}\langle x\rangle\langle x_{1}\rangle.

Hence, using the decay assumptions on v1v_{1} and v2v_{2}, we conclude that

|I2,4(1,2)(x,y)|C|t|32xy.\left|I_{2,4}^{(1,2)}(x,y)\right|\leq C|t|^{-\frac{3}{2}}\langle x\rangle\langle y\rangle.

The same bound holds for the term I2,4(2,1)I_{2,4}^{(2,1)} and we will skip the details. Finally, we are left with

I2,4(2,2)(x,y)=3eitz2z2χ0(z2)ez2+2μ|xx1|2z2+2μ[e¯22v1Pv2e¯22](x1,y1)ez2+2μ|yy1|2z2+2μdx1dy1dz.I_{2,4}^{(2,2)}(x,y)=\int_{{\mathbb{R}}^{3}}e^{itz^{2}}z^{2}\chi_{0}(z^{2})\frac{e^{-\sqrt{z^{2}+2\mu}|x-x_{1}|}}{-2\sqrt{z^{2}+2\mu}}[\underline{e}_{22}v_{1}Pv_{2}\underline{e}_{22}](x_{1},y_{1})\frac{e^{-\sqrt{z^{2}+2\mu}|y-y_{1}|}}{-2\sqrt{z^{2}+2\mu}}\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\,\mathrm{d}z.

By direct computation using (3.2),

[e¯22v1Pv2e¯22](x1,y1)=1V1L1()[V2e¯2](x1)[V2e¯2](y1),[\underline{e}_{22}v_{1}Pv_{2}\underline{e}_{22}](x_{1},y_{1})=\frac{1}{\|V_{1}\|_{L^{1}({\mathbb{R}})}}[V_{2}\underline{e}_{2}](x_{1})[V_{2}\underline{e}_{2}]^{\top}(y_{1}),

and by Lemma 2.1 and Lemma 5.1, we have the uniform estimate

|eitz2z2χ0(z2)ez2+2μ|xx1|2z2+2μez2+2μ|yy1|2z2+2μ𝑑z|Cμ|t|32.\left|\int_{{\mathbb{R}}}e^{itz^{2}}z^{2}\chi_{0}(z^{2})\frac{e^{-\sqrt{z^{2}+2\mu}|x-x_{1}|}}{-2\sqrt{z^{2}+2\mu}}\frac{e^{-\sqrt{z^{2}+2\mu}|y-y_{1}|}}{-2\sqrt{z^{2}+2\mu}}dz\right|\leq C_{\mu}|t|^{-\frac{3}{2}}.

Hence, by exchanging the order of integration, we conclude that

|I2,4(2,2)(x,y)|C|t|32.\left|I_{2,4}^{(2,2)}(x,y)\right|\leq C|t|^{-\frac{3}{2}}. (5.71)

Thus, we conclude (5.66) by summing over the four terms. ∎

Finally, we are ready to complete the proof of the local decay estimate (5.8). We sum the leading contributions of the spectral representation of eitχ0(μI)Ps+e^{it\mathcal{H}}\chi_{0}(\mathcal{H}-\mu I)P_{\mathrm{s}}^{+} in (5.24) by invoking Proposition 2.3, Proposition 5.4, Proposition 5.5, and Proposition 5.6 to obtain

Ft0eitμπi(i2ηFt1+1ηV1L1()Ft2+1ηV1L1()Ft3+(2iV1L1()+2|c0|2iV1L1())Ft4)=eitμ4πit([c0e¯1Ψ(x)][σ3Ψ(y)c0e¯1][c0e¯1Ψ(x)][eiy24tc0e¯1]+[eix24tc0e¯1][σ3Ψ(y)c0e¯1]+|c0|2eix24teiy24te¯1e¯1)=eitμ4πit(Ψ(x)[σ3Ψ(y)]+(eix24t1)c0[σ3Ψ(y)]+(eiy24t1)Ψ(x)[c0e¯1]+(1eix24teiy24t+eix24teiy24t)|c0|2e¯1e¯1),\begin{split}&F_{t}^{0}-\frac{e^{it\mu}}{\pi i}\left(\frac{i}{2\eta}F_{t}^{1}+\frac{1}{\eta\|V_{1}\|_{L^{1}({\mathbb{R}})}}F_{t}^{2}+\frac{1}{\eta\|V_{1}\|_{L^{1}({\mathbb{R}})}}F_{t}^{3}+\left(\frac{2i}{\|V_{1}\|_{L^{1}({\mathbb{R}})}}+\frac{2|c_{0}|^{2}}{i\|V_{1}\|_{L^{1}({\mathbb{R}})}}\right)F_{t}^{4}\right)\\ &=\frac{e^{it\mu}}{\sqrt{-4\pi it}}\left(-[c_{0}\underline{e}_{1}-\Psi(x)][\sigma_{3}\Psi(y)-{c_{0}}\underline{e}_{1}]^{*}-[c_{0}\underline{e}_{1}-\Psi(x)][e^{i\frac{y^{2}}{4t}}{c_{0}}\underline{e}_{1}]^{*}\right.\\ &\left.\hskip 142.26378pt+[e^{-i\frac{x^{2}}{4t}}c_{0}\underline{e}_{1}][\sigma_{3}\Psi(y)-{c_{0}}\underline{e}_{1}]^{*}+|c_{0}|^{2}e^{-i\frac{x^{2}}{4t}}e^{-i\frac{y^{2}}{4t}}\underline{e}_{1}\underline{e}_{1}^{\top}\right)\\ &=\frac{e^{it\mu}}{\sqrt{-4\pi it}}\left(\Psi(x)[\sigma_{3}\Psi(y)]^{*}+(e^{-i\frac{x^{2}}{4t}}-1)c_{0}[\sigma_{3}\Psi(y)]^{*}+(e^{-i\frac{y^{2}}{4t}}-1)\Psi(x)[{c_{0}}\underline{e}_{1}]^{*}\right.\\ &\left.\hskip 142.26378pt+(1-e^{-i\frac{x^{2}}{4t}}-e^{-i\frac{y^{2}}{4t}}+e^{-i\frac{x^{2}}{4t}}e^{-i\frac{y^{2}}{4t}})|c_{0}|^{2}\underline{e}_{1}\underline{e}_{1}^{\top}\right),\end{split}

where we use the cancellation Ft0eitμπi2iV1L1()Ft4=0F_{t}^{0}-\frac{e^{it\mu}}{\pi i}\frac{2i}{\|V_{1}\|_{L^{1}({\mathbb{R}})}}F_{t}^{4}=0 in the first equality. We note that the first term gives us the finite rank operator

Ft+(x,y)=eitμ4πitΨ(x)[σ3Ψ(y)],F_{t}^{+}(x,y)=\frac{e^{it\mu}}{\sqrt{-4\pi it}}\Psi(x)[\sigma_{3}\Psi(y)]^{*}, (5.72)

and we show that the last three terms satisfy the better decay rate. Using,

|1eix24t|x24|t|,|1-e^{-i\frac{x^{2}}{4t}}|\leq\frac{x^{2}}{4|t|}, (5.73)

and the fact that ΨL()×L()\Psi\in L^{\infty}({\mathbb{R}})\times L^{\infty}({\mathbb{R}}), we have

|eitμeiπ42πt(eix24t1)c0e¯1[σ3Ψ(y)]||t|32x2,\left|\frac{e^{it\mu}e^{i\frac{\pi}{4}}}{2\sqrt{\pi}\sqrt{t}}(e^{-i\frac{x^{2}}{4t}}-1)c_{0}\underline{e}_{1}[\sigma_{3}\Psi(y)]^{*}\right|\lesssim|t|^{-\frac{3}{2}}\langle x\rangle^{2},

and similarly

|eitμeiπ42πt(eiy24t1)c0¯Ψ(x)e¯1||t|32y2.\left|\frac{e^{it\mu}e^{i\frac{\pi}{4}}}{2\sqrt{\pi}\sqrt{t}}(e^{-i\frac{y^{2}}{4t}}-1)\overline{c_{0}}\Psi(x)\underline{e}_{1}^{\top}\right|\lesssim|t|^{-\frac{3}{2}}\langle y\rangle^{2}.

For the last term, we have

|1eix24teiy24t+eix24teiy24t|=|1eix24t||1eiy24t||t|2x2y2.\left|1-e^{-i\frac{x^{2}}{4t}}-e^{-i\frac{y^{2}}{4t}}+e^{-i\frac{x^{2}}{4t}}e^{-i\frac{y^{2}}{4t}}\right|=\left|1-e^{-i\frac{x^{2}}{4t}}\right|\left|1-e^{-i\frac{y^{2}}{4t}}\right|\lesssim|t|^{-2}\langle x\rangle^{2}\langle y\rangle^{2}. (5.74)

Thus, the leading contribution to eitχ0(μI)Ps+e^{it\mathcal{H}}\chi_{0}(\mathcal{H}-\mu I)P_{\mathrm{s}}^{+} is Ft+F_{t}^{+}. ∎

6. Intermediate and high energy estimates

In order to complete the proof of Theorem 1.4, we also need to prove the dispersive estimates when the spectral variable is bounded away from the thresholds ±μ\pm\mu. As usual, we focus on the positive semi-axis [μ,)[\mu,\infty) of the essential spectrum and prove the dispersive estimates for energies λ>μ\lambda>\mu. The negative semi-axis (,μ](-\infty,-\mu] can be treated by symmetry of \mathcal{H}. We recall from Section 2 that the kernel of the limiting resolvent operator for 0\mathcal{H}_{0} has the formula

0±(z)(x,y):=(0(z2+μ±i0))1=[±ie±iz|xy|2z00ez2+2μ|xy|2z2+2μ], 0<z<.\mathcal{R}_{0}^{\pm}(z)(x,y):=(\mathcal{H}_{0}-(z^{2}+\mu\pm i0))^{-1}=\begin{bmatrix}\pm\frac{ie^{\pm iz|x-y|}}{2z}&0\\ 0&-\frac{e^{-\sqrt{z^{2}+2\mu}|x-y|}}{2\sqrt{z^{2}+2\mu}}\end{bmatrix},\quad\forall\ 0<z<\infty. (6.1)

From this, we have the following bound

0±(z)L1×L1L×LC|z|1.\|\mathcal{R}_{0}^{\pm}(z)\|_{L^{1}\times L^{1}\to L^{\infty}\times L^{\infty}}\leq C|z|^{-1}.

Hence, for sufficiently large zz, the perturbed resolvent ±(z)\mathcal{R}^{\pm}(z) can be expanded into the infinite Born series

±(z)=n=00±(z)(𝒱0±(z))n.\mathcal{R}^{\pm}(z)=\sum_{n=0}^{\infty}\mathcal{R}_{0}^{\pm}(z)(-\mathcal{V}\mathcal{R}_{0}^{\pm}(z))^{n}. (6.2)

More precisely, since the operator norm L1×L1L×LL^{1}\times L^{1}\to L^{\infty}\times L^{\infty} in the nn-th summand above is bounded by C|z|1(C𝒱1|z|1)nC|z|^{-1}(C\|\mathcal{V}\|_{1}|z|^{-1})^{n}, the Born series converges in the operator norm whenever |z|>z1:=2C𝒱L1×L1|z|>z_{1}:=2C\|\mathcal{V}\|_{L^{1}\times L^{1}}. We define the high-energy cut-off by

χh(z):=1χ(z),\chi_{\mathrm{h}}(z):=1-\chi(z), (6.3)

where χ(z)\chi(z) is a standard smooth even cut-off supported on [z1,z1][-z_{1},z_{1}] satisfying χ(z)=1\chi(z)=1 for |z|z12|z|\leq\frac{z_{1}}{2} and χ(z)=0\chi(z)=0 for |z|z1|z|\geq z_{1}. We insert the cut-off and the Born series expansion into the spectral representation eitχh(μI)Ps+e^{it\mathcal{H}}\chi_{\mathrm{h}}(\mathcal{H}-\mu I)P_{\mathrm{s}}^{+} and look to bound the following

|eitχh(μI)Ps+u,v|=|0eitz2zχh(z2)[+(z)(z)]u,vdz|C±n=0|0eitz2zχh(z2)0±(z)(𝒱0±(z))nu,vdz|,\begin{split}|\langle e^{it\mathcal{H}}\chi_{\mathrm{h}}(\mathcal{H}-\mu I)P_{\mathrm{s}}^{+}\vec{u},\vec{v}\rangle|&=\left|\int_{0}^{\infty}e^{itz^{2}}z\chi_{\mathrm{h}}(z^{2})\langle[\mathcal{R}^{+}(z)-\mathcal{R}^{-}(z)]\vec{u},\vec{v}\rangle\,\mathrm{d}z\right|\\ &\leq C\sum_{\pm}\sum_{n=0}^{\infty}\left|\int_{0}^{\infty}e^{itz^{2}}z\chi_{\mathrm{h}}(z^{2})\langle\mathcal{R}_{0}^{\pm}(z)(\mathcal{V}\mathcal{R}_{0}^{\pm}(z))^{n}\vec{u},\vec{v}\rangle\,\mathrm{d}z\right|,\end{split} (6.4)

where u,v𝒮()×𝒮()\vec{u},\vec{v}\in\mathcal{S}({\mathbb{R}})\times\mathcal{S}({\mathbb{R}}). From [KS06], we have the following dispersive estimates:

Proposition 6.1.

Under the same hypothesis as Theorem 1.4, we have

eitχh(μI)Ps+uL()×L()|t|12uL1()×L1(),\left\|e^{it\mathcal{H}}\chi_{\mathrm{h}}(\mathcal{H}-\mu I)P_{\mathrm{s}}^{+}\vec{u}\,\right\|_{L^{\infty}({\mathbb{R}})\times L^{\infty}({\mathbb{R}})}\lesssim|t|^{-\frac{1}{2}}\left\|\vec{u}\,\right\|_{L^{1}({\mathbb{R}})\times L^{1}({\mathbb{R}})}, (6.5)

and

x1eitχh(μI)Ps+uL()×L()|t|32xuL1()×L1(),\left\|\langle x\rangle^{-1}e^{it\mathcal{H}}\chi_{\mathrm{h}}(\mathcal{H}-\mu I)P_{\mathrm{s}}^{+}\vec{u}\,\right\|_{L^{\infty}({\mathbb{R}})\times L^{\infty}({\mathbb{R}})}\lesssim|t|^{-\frac{3}{2}}\left\|\langle x\rangle\vec{u}\,\right\|_{L^{1}({\mathbb{R}})\times L^{1}({\mathbb{R}})}, (6.6)

for any |t|1|t|\geq 1.

Proof.

For (6.5), see the proof of [KS06, Proposition 7.1], and for (6.6), see the proof of [KS06, Proposition 8.1]. Note that the high-energy dispersive estimate holds irrespective of the regularity of the thresholds ±μ\pm\mu. ∎

Let z0>0z_{0}>0 be the constant from Proposition 4.8. It may happen that z1z_{1} is strictly larger than z0z_{0}. In this case, we need to derive estimates analogous to the above proposition in the remaining intermediate energy regime [z1,z0][z0,z1][-z_{1},-z_{0}]\cup[z_{0},z_{1}]. To this end, we set χm(z)\chi_{\mathrm{m}}(z) to be the intermediate energy cut-off given by

χm(z):=1χ0(z)χh(z),\chi_{\mathrm{m}}(z):=1-\chi_{0}(z)-\chi_{\mathrm{h}}(z), (6.7)

where χ0(z)\chi_{0}(z) was the cut-off defined in the previous section in Proposition 5.2.

Proposition 6.2.

For any |t|1|t|\geq 1, we have

eitχm(μI)Ps+uLx()×Lx()|t|12uLx1()×Lx1(),\left\|e^{it\mathcal{H}}\chi_{\mathrm{m}}(\mathcal{H}-\mu I)P_{\mathrm{s}}^{+}\vec{u}\,\right\|_{L_{x}^{\infty}({\mathbb{R}})\times L_{x}^{\infty}({\mathbb{R}})}\lesssim|t|^{-\frac{1}{2}}\left\|\vec{u}\,\right\|_{L_{x}^{1}({\mathbb{R}})\times L_{x}^{1}({\mathbb{R}})}, (6.8)

and

x1eitχm(μI)Ps+uLx()×Lx()|t|32xuLx1()×Lx1().\left\|\langle x\rangle^{-1}e^{it\mathcal{H}}\chi_{\mathrm{m}}(\mathcal{H}-\mu I)P_{\mathrm{s}}^{+}\vec{u}\,\right\|_{L_{x}^{\infty}({\mathbb{R}})\times L_{x}^{\infty}({\mathbb{R}})}\lesssim|t|^{-\frac{3}{2}}\left\|\langle x\rangle\vec{u}\,\right\|_{L_{x}^{1}({\mathbb{R}})\times L_{x}^{1}({\mathbb{R}})}. (6.9)

Before proving the above proposition, we need the following lemmas for pointwise bounds and operator norm bounds on the resolvent operators and its derivatives. The first lemma follows immediately from the expression (6.1) and the triangle inequality ||xx1||x|||x1|||x-x_{1}|-|x||\leq|x_{1}|.

Lemma 6.3.

Let γ0>0\gamma_{0}>0. For every z>γ0z>\gamma_{0}, and k{0,1,2}k\in\{0,1,2\}, we have

|zk0±(z)(x,y)|Cγ01kxyk,\left|\partial_{z}^{k}\mathcal{R}_{0}^{\pm}(z)(x,y)\right|\leq C\gamma_{0}^{-1-k}\langle x-y\rangle^{k}, (6.10)

and hence

zk0±(z)(x,)X(12+k)Cγ01kxk.\left\|\partial_{z}^{k}\mathcal{R}_{0}^{\pm}(z)(x,\cdot)\right\|_{X_{-(\frac{1}{2}+k)-}}\leq C\gamma_{0}^{-1-k}\langle x\rangle^{k}. (6.11)

Moreover, define

𝒢±(z)(x,x1)=[eiz|x|001]0±(z)(x,x1)=[±ie±iz(|xx1||x|)2z00ez2+2μ|xx1|2z2+2μ].\mathcal{G}_{\pm}(z)(x,x_{1})=\begin{bmatrix}e^{\mp iz|x|}&0\\ 0&1\end{bmatrix}\mathcal{R}_{0}^{\pm}(z)(x,x_{1})=\begin{bmatrix}\pm\frac{ie^{\pm iz(|x-x_{1}|-|x|)}}{2z}&0\\ 0&-\frac{e^{-\sqrt{z^{2}+2\mu}|x-x_{1}|}}{2\sqrt{z^{2}+2\mu}}\end{bmatrix}. (6.12)

Then, for any k0k\geq 0,

supx|zk𝒢±(z)(x,x1)|Cγ01k|x1|.\sup_{x\in{\mathbb{R}}}\left|\partial_{z}^{k}\mathcal{G}^{\pm}(z)(x,x_{1})\right|\leq C\gamma_{0}^{-1-k}|x_{1}|. (6.13)

With these bounds, we are able to give operator norm bounds on the perturbed resolvent via the resolvent identity.

Lemma 6.4.

Let γ0>0\gamma_{0}>0. We have

sup|z|>γ0z±(z)X32+X321,\sup_{|z|>\gamma_{0}}\left\|\partial_{z}\mathcal{R}^{\pm}(z)\right\|_{X_{\frac{3}{2}+}\to X_{-\frac{3}{2}-}}\lesssim 1, (6.14)
sup|z|>γ0z2±(z)X52+X521.\sup_{|z|>\gamma_{0}}\left\|\partial_{z}^{2}\mathcal{R}^{\pm}(z)\right\|_{X_{\frac{5}{2}+}\to X_{-\frac{5}{2}-}}\lesssim 1. (6.15)
Proof.

By Lemma 3.1, for any |z|>γ0|z|>\gamma_{0}, we have

±(z)=(I+0±(z)𝒱)10±(z)=:S±(z)10±(z),\mathcal{R}^{\pm}(z)=(I+\mathcal{R}_{0}^{\pm}(z)\mathcal{V})^{-1}\mathcal{R}_{0}^{\pm}(z)=:S^{\pm}(z)^{-1}\mathcal{R}_{0}^{\pm}(z), (6.16)

as a bounded operator from X12+X_{\frac{1}{2}+} to X12X_{-\frac{1}{2}-}. Note that S±(z)S^{\pm}(z) is boundedly invertible on XσX_{-\sigma} for any σ>0\sigma>0. By differentiation, we have

z±(z)=S±(z)1z0±(z)𝒱S±(z)10±(z)+S±(z)1z0±(z).\partial_{z}\mathcal{R}^{\pm}(z)=-S^{\pm}(z)^{-1}\partial_{z}\mathcal{R}_{0}^{\pm}(z)\mathcal{V}S^{\pm}(z)^{-1}\mathcal{R}_{0}^{\pm}(z)+S^{\pm}(z)^{-1}\partial_{z}\mathcal{R}_{0}^{\pm}(z). (6.17)

Moreover, as a multiplication operator, 𝒱:XσXσ\mathcal{V}:X_{-\sigma}\to X_{\sigma} is bounded for any σ>0\sigma>0 due to the exponential decay of 𝒱\mathcal{V}. By Lemma 6.3, zR0±(z):X32+X32\partial_{z}R_{0}^{\pm}(z):X_{\frac{3}{2}+}\to X_{-\frac{3}{2}-} is bounded and since the embedding X12X32X_{-\frac{1}{2}-}\subset X_{-\frac{3}{2}-} is continuous, we infer the bound (6.14) by taking composition. By a similar argument,

z2±(z)X52+X521.\|\partial_{z}^{2}\mathcal{R}^{\pm}(z)\|_{X_{\frac{5}{2}+}\to X_{-\frac{5}{2}-}}\lesssim 1. (6.18)

Proof of Proposition 6.2.

By iterating the second resolvent identity, we write the perturbed resolvent as a finite sum

±(z)=0±(z)0±(z)𝒱0±(z)+0±(z)𝒱±(z)𝒱0±(z),\mathcal{R}^{\pm}(z)=\mathcal{R}_{0}^{\pm}(z)-\mathcal{R}_{0}^{\pm}(z)\mathcal{V}\mathcal{R}_{0}^{\pm}(z)+\mathcal{R}_{0}^{\pm}(z)\mathcal{V}\mathcal{R}^{\pm}(z)\mathcal{V}\mathcal{R}_{0}^{\pm}(z), (6.19)

and we write

eitχm(μI)Ps+(x,y)=j=130eitz2zχm(z2)(1)j+1(j+(z)j(z))(x,y)𝑑z,e^{it\mathcal{H}}\chi_{\mathrm{m}}(\mathcal{H}-\mu I)P_{\mathrm{s}}^{+}(x,y)=\sum_{j=1}^{3}\int_{0}^{\infty}e^{itz^{2}}z\chi_{\mathrm{m}}(z^{2})(-1)^{j+1}(\mathcal{E}_{j}^{+}(z)-\mathcal{E}_{j}^{-}(z))(x,y)dz, (6.20)

with

1±(z)=0±(z),2±(z)=0±(z)𝒱0±(z),3±(z)=0±(z)𝒱±(z)𝒱0±(z).\mathcal{E}_{1}^{\pm}(z)=\mathcal{R}_{0}^{\pm}(z),\quad\mathcal{E}_{2}^{\pm}(z)=\mathcal{R}_{0}^{\pm}(z)\mathcal{V}\mathcal{R}_{0}^{\pm}(z),\quad\mathcal{E}_{3}^{\pm}(z)=\mathcal{R}_{0}^{\pm}(z)\mathcal{V}\mathcal{R}^{\pm}(z)\mathcal{V}\mathcal{R}_{0}^{\pm}(z). (6.21)

Hence, to prove (6.8) and (6.9), it is sufficient to establish the estimates

sup±supj=1,2,3|0eitz2zχm(z2)j±(z)(x,y)dz|min{|t|12,|t|32xy}.\sup_{\pm}\sup_{j=1,2,3}\left|\int_{0}^{\infty}e^{itz^{2}}z\chi_{\mathrm{m}}(z^{2})\mathcal{E}_{j}^{\pm}(z)(x,y)\,\mathrm{d}z\right|\lesssim\min\{|t|^{-\frac{1}{2}},|t|^{-\frac{3}{2}}\langle x\rangle\langle y\rangle\}. (6.22)

The term involving 1±\mathcal{E}_{1}^{\pm} is handled by the earlier Proposition 2.3, while the second term involving 2±\mathcal{E}_{2}^{\pm} can be treated analogously as in Proposition 6.1. We refer the reader to [GS04, Lemma 3] and [Gol07, Proposition 3] for similar computations. For the term involving 3±\mathcal{E}_{3}^{\pm}, we first write

0±(z)(s1,s2)=[e±iz|s1|001]𝒢±(z)(s1,s2),\mathcal{R}_{0}^{\pm}(z)(s_{1},s_{2})=\begin{bmatrix}e^{\pm iz|s_{1}|}&0\\ 0&1\end{bmatrix}\mathcal{G}_{\pm}(z)(s_{1},s_{2}),

where the operator 𝒢±(z)\mathcal{G}_{\pm}(z) was defined in (6.12). Then, using that the kernel 0±(z)(x,y)\mathcal{R}_{0}^{\pm}(z)(x,y) is symmetric in xx and yy variables, and using the matrix identity

e¯jj[a11a12a21a22]e¯kk=ajke¯je¯k,j,k{1,2},\underline{e}_{jj}\begin{bmatrix}a_{11}&a_{12}\\ a_{21}&a_{22}\end{bmatrix}\underline{e}_{kk}=a_{jk}\underline{e}_{j}\underline{e}_{k}^{\top},\quad j,k\in\{1,2\}, (6.23)

we compute the following kernel identity

3±(z)(x,y)=20±(x,x1)[𝒱±(z)𝒱](x1,y1)0±(y,y1)dx1dy1=[e±iz|x|001]2𝒢±(x,x1)[𝒱±(z)𝒱](x1,y1)𝒢±(y,y1)dx1dy1[e±iz|y|001]=e±iz(|x|+|y|)(𝒢±)(z)(x,)e¯1,𝒱±(z)𝒱𝒢±(z)(y,)e¯1e¯1e¯1+e±iz|x|(𝒢±)(z)(x,)e¯2,𝒱±(z)𝒱𝒢±(z)(y,)e¯1e¯1e¯2+e±iz|y|(𝒢±)(z)(x,)e¯1,𝒱±(z)𝒱𝒢±(z)(y,)e¯2e¯2e¯1+(𝒢±)(z)(x,)e¯2,𝒱±(z)𝒱𝒢±(z)(y,)e¯2e¯2e¯2=:e±iz(|x|+|y|)A1±(z,x,y)+e±iz|x|A2±(z,x,y)+e±iz|y|A3±(z,x,y)+A4±(z,x,y).\begin{split}\mathcal{E}_{3}^{\pm}(z)(x,y)&=\int_{{\mathbb{R}}^{2}}\mathcal{R}_{0}^{\pm}(x,x_{1})[\mathcal{V}\mathcal{R}^{\pm}(z)\mathcal{V}](x_{1},y_{1})\mathcal{R}_{0}^{\pm}(y,y_{1})\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\\ &=\begin{bmatrix}e^{\pm iz|x|}&0\\ 0&1\end{bmatrix}\int_{{\mathbb{R}}^{2}}\mathcal{G}^{\pm}(x,x_{1})[\mathcal{V}\mathcal{R}^{\pm}(z)\mathcal{V}](x_{1},y_{1})\mathcal{G}^{\pm}(y,y_{1})\,\mathrm{d}x_{1}\,\mathrm{d}y_{1}\begin{bmatrix}e^{\pm iz|y|}&0\\ 0&1\end{bmatrix}\\ &=e^{\pm iz(|x|+|y|)}\langle(\mathcal{G}^{\pm})^{*}(z)(x,\cdot)\underline{e}_{1},\mathcal{V}\mathcal{R}^{\pm}(z)\mathcal{V}\mathcal{G}^{\pm}(z)(y,\cdot)\underline{e}_{1}\rangle\,\underline{e}_{1}\underline{e}_{1}^{\top}\\ &\quad+e^{\pm iz|x|}\langle(\mathcal{G}^{\pm})^{*}(z)(x,\cdot)\underline{e}_{2},\mathcal{V}\mathcal{R}^{\pm}(z)\mathcal{V}\mathcal{G}^{\pm}(z)(y,\cdot)\underline{e}_{1}\rangle\,\underline{e}_{1}\underline{e}_{2}^{\top}\\ &\quad+e^{\pm iz|y|}\langle(\mathcal{G}^{\pm})^{*}(z)(x,\cdot)\underline{e}_{1},\mathcal{V}\mathcal{R}^{\pm}(z)\mathcal{V}\mathcal{G}^{\pm}(z)(y,\cdot)\underline{e}_{2}\rangle\,\underline{e}_{2}\underline{e}_{1}^{\top}\\ &\quad+\langle(\mathcal{G}^{\pm})^{*}(z)(x,\cdot)\underline{e}_{2},\mathcal{V}\mathcal{R}^{\pm}(z)\mathcal{V}\mathcal{G}^{\pm}(z)(y,\cdot)\underline{e}_{2}\rangle\,\underline{e}_{2}\underline{e}_{2}^{\top}\\ &=:e^{\pm iz(|x|+|y|)}A_{1}^{\pm}(z,x,y)+e^{\pm iz|x|}A_{2}^{\pm}(z,x,y)+e^{\pm iz|y|}A_{3}^{\pm}(z,x,y)+A_{4}^{\pm}(z,x,y).\end{split}

We plug this identity into the left hand side of (6.22), and hence it will be sufficient to provide the bounds

|0eitz2±izrzχm(z2)Ak±(z,x,y)dz|min{|t|12,|t|32r},k{1,,4},\left|\int_{0}^{\infty}e^{itz^{2}\pm izr}z\chi_{\mathrm{m}}(z^{2})A_{k}^{\pm}(z,x,y)\,\mathrm{d}z\right|\lesssim\min\{|t|^{-\frac{1}{2}},|t|^{-\frac{3}{2}}\langle r\rangle\},\quad k\in\{1,\ldots,4\}, (6.24)

where rr can represent 0 or |x||x|, |y||y|, or the sum of both variables. For the case k=1k=1, by Lemma 2.1, we have that

|0eitz2±iz(|x|+|y|)zχm(z2)A1±(z,x,y)dz|C|t|12z(zχm(z2)A1±(z,x,y))Lz1().\left|\int_{0}^{\infty}e^{itz^{2}\pm iz(|x|+|y|)}z\chi_{\mathrm{m}}(z^{2})A_{1}^{\pm}(z,x,y)\,\mathrm{d}z\right|\leq C|t|^{-\frac{1}{2}}\|\partial_{z}\left(z\chi_{\mathrm{m}}(z^{2})A_{1}^{\pm}(z,x,y)\right)\|_{L_{z}^{1}({\mathbb{R}})}.

Since the term zχm(z2)z\chi_{\mathrm{m}}(z^{2}) is smooth and has compact support, we only need to track the derivatives when they fall onto either 𝒢±(z)\mathcal{G}^{\pm}(z) or ±(z)\mathcal{R}^{\pm}(z). In any case, thanks to the exponential decay of 𝒱\mathcal{V}, and the bounds (6.13), (6.14) from the previous lemmas, we have the following uniform bound

sup±supzsupp(χm)supj,k=1,2|z(𝒢±)(z)(y,)e¯j,𝒱±(z)𝒱𝒢±(z)(x,)e¯k|sup±supzsupp(χm)supj,k=1,2|𝒱|(x1)(|±(z)(x1,x2)|+|z±(z)(x1,x2)|)|𝒱|(x2)Lx22Lx12|𝒱|(x1)(|𝒢±(z)(x,x1)|+|z𝒢±(z)(x,x1)|)e¯jLx12|𝒱|(x2)(|𝒢±(z)(x2,y)|+|z𝒢±(z)(x2,y)|)e¯kLx221,\begin{split}&\sup_{\pm}\sup_{z\in\operatorname*{supp}(\chi_{\mathrm{m}})}\sup_{j,k=1,2}|\partial_{z}\langle(\mathcal{G}^{\pm})^{*}(z)(y,\cdot)\underline{e}_{j},\mathcal{V}\mathcal{R}^{\pm}(z)\mathcal{V}\mathcal{G}^{\pm}(z)(x,\cdot)\underline{e}_{k}\rangle|\\ &\lesssim\sup_{\pm}\sup_{z\in\operatorname*{supp}(\chi_{\mathrm{m}})}\sup_{j,k=1,2}\left\|\sqrt{|\mathcal{V}|}(x_{1})\left(|\mathcal{R}^{\pm}(z)(x_{1},x_{2})|+|\partial_{z}\mathcal{R}^{\pm}(z)(x_{1},x_{2})|\right)\sqrt{|\mathcal{V}|}(x_{2})\right\|_{L_{x_{2}}^{2}\to L_{x_{1}}^{2}}\\ &\quad\cdot\|\sqrt{|\mathcal{V}|}(x_{1})\left(|\mathcal{G}^{\pm}(z)(x,x_{1})|+|\partial_{z}\mathcal{G}^{\pm}(z)(x,x_{1})|\right)\underline{e}_{j}\|_{L_{x_{1}}^{2}}\\ &\quad\cdot\|\sqrt{|\mathcal{V}|}(x_{2})\left(|\mathcal{G}^{\pm}(z)(x_{2},y)|+|\partial_{z}\mathcal{G}^{\pm}(z)(x_{2},y)|\right)\underline{e}_{k}\|_{L_{x_{2}}^{2}}\\ &\lesssim 1,\end{split} (6.25)

for all x,yx,y\in{\mathbb{R}}.

To prove the weighted dispersive estimate, we invoke the stronger estimate in Lemma 2.1:

|0eitz2±iz(|x|+|y|)zχm(z2)A1±(z,x,y)dz|C|t|32[z2±i(|x|+|y|)z](χm(z2)A1±(z,x,y))Lz1()\left|\int_{0}^{\infty}e^{itz^{2}\pm iz(|x|+|y|)}z\chi_{\mathrm{m}}(z^{2})A_{1}^{\pm}(z,x,y)\,\mathrm{d}z\right|\leq C|t|^{-\frac{3}{2}}\left\|[\partial_{z}^{2}\pm i(|x|+|y|)\partial_{z}]\left(\chi_{\mathrm{m}}(z^{2})A_{1}^{\pm}(z,x,y)\right)\right\|_{L_{z}^{1}({\mathbb{R}})}

Here, we can apply the same argument as in (6.25) for the two derivatives bound on A1±A_{1}^{\pm} using the estimates (6.13) and (6.15), whereas the bound on one derivative for A1±A_{1}^{\pm} leads to the weights xy\langle x\rangle\langle y\rangle. Thus, we prove (6.24) for k=1k=1. The other cases follow by the same argument and we are done. ∎

Finally, we conclude with the proof of Theorem 1.4.

Proof of Theorem 1.4.

By combining the estimates from Proposition 5.2, Proposition 6.1, and Proposition 6.2, we have established the bounds

eitPs+uLx()×Lx()|t|12uLx1()×Lx1(),\left\|e^{it\mathcal{H}}P_{\mathrm{s}}^{+}{\vec{u}}\,\right\|_{L_{x}^{\infty}({\mathbb{R}})\times L_{x}^{\infty}({\mathbb{R}})}\lesssim|t|^{-\frac{1}{2}}\|{\vec{u}}\,\|_{L_{x}^{1}({\mathbb{R}})\times L_{x}^{1}({\mathbb{R}})},

as well as

x2(eitPs+Ft+)uLx()×Lx()|t|32uLx1()×Lx1(),\left\|\langle x\rangle^{-2}(e^{it\mathcal{H}}P_{\mathrm{s}}^{+}-F_{t}^{+}){\vec{u}}\,\right\|_{L_{x}^{\infty}({\mathbb{R}})\times L_{x}^{\infty}({\mathbb{R}})}\lesssim|t|^{-\frac{3}{2}}\|{\vec{u}}\,\|_{L_{x}^{1}({\mathbb{R}})\times L_{x}^{1}({\mathbb{R}})},

for any u:=(u1,u2)𝒮()×𝒮(){\vec{u}}:=(u_{1},u_{2})^{\top}\in\mathcal{S}({\mathbb{R}})\times\mathcal{S}({\mathbb{R}}) and |t|1|t|\geq 1, with Ft+F_{t}^{+} given by (5.9). By Remark 3.3, we can similarly deduce that the unweighted dispersive estimate for the evolution eitPse^{it\mathcal{H}}P_{\mathrm{s}}^{-} using the identity (3.17). On the other hand, for the weighted estimate, we find that the leading contribution to eitPse^{it\mathcal{H}}P_{\mathrm{s}}^{-} is given by

Ft(x,y)=σ1Ft+(x,y)σ1=eitμ4πit[σ1Ψ(x)][σ3σ1Ψ(y)],F_{t}^{-}(x,y)=\sigma_{1}F_{-t}^{+}(x,y)\sigma_{1}=-\frac{e^{-it\mu}}{\sqrt{4\pi it}}[\sigma_{1}\Psi(x)][\sigma_{3}\sigma_{1}\Psi(y)]^{*}, (6.26)

where we used the anti-commutation identity σ3σ1=σ1σ3\sigma_{3}\sigma_{1}=-\sigma_{1}\sigma_{3}. Thus, we conclude the local decay estimate (1.14) and the formula (1.15) by setting Ft:=Ft++FtF_{t}:=F_{t}^{+}+F_{t}^{-}. ∎

Appendix A Neumann series

Lemma A.1.

Let AA be an invertible operator and BB be a bounded operator satisfying B<A11\|B\|<\|A^{-1}\|^{-1}. Then, ABA-B is invertible with

(AB)1=A1n=0(BA1)n=A1+A1BA1+A1BA1BA1+,(A-B)^{-1}=A^{-1}\sum_{n=0}^{\infty}(BA^{-1})^{n}=A^{-1}+A^{-1}BA^{-1}+A^{-1}BA^{-1}BA^{-1}+\cdots, (A.1)

and

(AB)1(A11B)1.\|(A-B)^{-1}\|\leq(\|A^{-1}\|^{-1}-\|B\|)^{-1}. (A.2)
Proof.

By the hypothesis B<A11\|B\|<\|A^{-1}\|^{-1}, we have A1B<1\|A^{-1}B\|<1. Consider the identity

(AB)1=(IA1B)1A1.(A-B)^{-1}=(I-A^{-1}B)^{-1}A^{-1}.

The term on the right hand side can be written in the usual Neumann series

(IA1B)1=n=0(A1B)n.(I-A^{-1}B)^{-1}=\sum_{n=0}^{\infty}(A^{-1}B)^{n}.

Thus, by multiplying A1A^{-1}, we deduce (A.1). Note that the argument also holds true for (AB)1=A1(IBA1)1(A-B)^{-1}=A^{-1}(I-BA^{-1})^{-1}. Now, since we have the estimate

(IA1B)1(1A1B)1,\|(I-A^{-1}B)^{-1}\|\leq(1-\|A^{-1}B\|)^{-1},

we deduce (A.2) by the sub-multiplicative property for operator norms. ∎

References

  • [Agm75] S. Agmon. Spectral properties of Schrödinger operators and scattering theory. Annali della Scuola Normale Superiore di Pisa-Classe di Scienze, 2(2):151–218, 1975.
  • [BC81] H. Berestycki and T. Cazenave. Instabilité des états stationnaires dans les équations de Schrödinger et de Klein-Gordon non linéaires. C. R. Acad. Sci. Paris Sér. I Math., 293(9):489–492, 1981.
  • [Bec08] M. Beceanu. A centre-stable manifold for the focussing cubic NLS in R1+3R^{1+3}. Comm. Math. Phys., 280(1):145–205, 2008.
  • [Bec16] M. Beceanu. Dispersive estimates in R3R^{3} with threshold eigenstates and resonances. Analysis & PDE, 9(4):813–858, 2016.
  • [BGW85] D. Bollé, F. Gesztesy, and S.F.J. Wilk. A complete treatment of low-energy scattering in one dimension. Journal of Operator Theory, pages 3–32, 1985.
  • [BL83] H. Berestycki and P.-L. Lions. Nonlinear scalar field equations, I Existence of a Ground State. Archive for Rational Mechanics and Analysis, 82:313–345, 1983.
  • [BP92] V. S. Buslaev and G. S. Perelman. Scattering for the nonlinear Schrödinger equation: states that are close to a soliton. Algebra i Analiz, 4(6):63–102, 1992.
  • [BP95] V. S. Buslaev and G. S. Perelman. On the stability of solitary waves for nonlinear Schrödinger equations. Nonlinear evolution equations, 164:75–98, 1995.
  • [Caz03] T. Cazenave. Semilinear Schrödinger Equations. Courant lecture notes in mathematics. American Mathematical Society, 2003.
  • [CG23] C. Collot and P. Germain. Asymptotic stability of solitary waves for one dimensional nonlinear Schrödinger equations. arXiv preprint arXiv:2306.03668, 2023.
  • [CGNT08] S.-M. Chang, S. Gustafson, K. Nakanishi, and T.-P. Tsai. Spectra of linearized operators for NLS solitary waves. SIAM J. Math. Anal., 39(4):1070–1111, 2008.
  • [CHS11] O. Costin, M. Huang, and W. Schlag. On the spectral properties of L±L_{\pm} in three dimensions. Nonlinearity, 25(1):125, 2011.
  • [CL82] T. Cazenave and P.-L. Lions. Orbital stability of standing waves for some nonlinear Schrödinger equations. Communications in Mathematical Physics, 85:549–561, 1982.
  • [CM21] S. Cuccagna and M. Maeda. A survey on asymptotic stability of ground states of nonlinear Schrödinger equations II. Discrete Contin. Dyn. Syst. Ser. S, 14(5):1693–1716, 2021.
  • [CP03] A. Comech and D. Pelinovsky. Purely nonlinear instability of standing waves with minimal energy. Comm. Pure Appl. Math., 56(11):1565–1607, 2003.
  • [CP14] S. Cuccagna and D. E. Pelinovsky. The asymptotic stability of solitons in the cubic NLS equation on the line. Appl. Anal., 93(4):791–822, 2014.
  • [Cuc14] S. Cuccagna. On asymptotic stability of moving ground states of the nonlinear Schrödinger equation. Trans. Amer. Math. Soc., 366(6):2827–2888, 2014.
  • [EG13] M. B. Erdoǧan and W. R. Green. Dispersive estimates for matrix Schrödinger operators in dimension two. Discrete and Continuous Dynamical Systems, Apr 2013.
  • [EG21] M. B. Erdoğan and W. R. Green. On the one dimensional Dirac equation with potential. Journal de Mathématiques Pures et Appliquées, 151:132–170, 2021.
  • [EGG14] M. B. Erdoğan, M. Goldberg, and W. R. Green. Dispersive estimates for four dimensional Schrödinger and wave equations with obstructions at zero energy. Communications in Partial Differential Equations, 39(10):1936–1964, 2014.
  • [ES04] M. B. Erdoğan and W. Schlag. Dispersive estimates for Schrödinger operators in the presence of a resonance and/or an eigenvalue at zero energy in dimension three: I. Dynamics of Partial Differential Equations, 1(4):359–379, 2004.
  • [ES06] M. B. Erdoğan and W. Schlag. Dispersive estimates for Schrödinger operators in the presence of a resonance and/or an eigenvalue at zero energy in dimension three. II. J. Anal. Math., 99:199–248, 2006.
  • [GG15] M. Goldberg and W. R. Green. Dispersive estimates for higher dimensional Schrödinger operators with threshold eigenvalues I: The odd dimensional case. Journal of Functional Analysis, 269(3):633–682, 2015.
  • [GG17] M. Goldberg and W. R. Green. Dispersive estimates for higher dimensional Schrödinger operators with threshold eigenvalues. II: The even dimensional case. Journal of Spectral Theory, 7(1):33–86, 2017.
  • [Gol07] M. Goldberg. Transport in the one-dimensional Schrödinger equation. Proc. Amer. Math. Soc., 135(10):3171–3179, 2007.
  • [Gol10] M. Goldberg. A dispersive bound for three-dimensional Schrödinger operators with zero energy eigenvalues. Communications in Partial Differential Equations, 35(9):1610–1634, 2010.
  • [Gre12] W. R. Green. Dispersive estimates for matrix and scalar Schrödinger operators in dimension five. Illinois Journal of Mathematics, 56(2):307–341, 2012.
  • [Gri88] M. Grillakis. Linearized instability for nonlinear Schrödinger and Klein-Gordon equations. Comm. Pure Appl. Math., 41(6):747–774, 1988.
  • [Gri90] M. Grillakis. Analysis of the linearization around a critical point of an infinite-dimensional Hamiltonian system. Comm. Pure Appl. Math., 43(3):299–333, 1990.
  • [GS04] M. Goldberg and W. Schlag. Dispersive estimates for Schrödinger operators in dimensions one and three. Comm. Math. Phys., 251(1):157–178, 2004.
  • [GS05] Z. Gang and I.M. Sigal. Asymptotic stability of nonlinear Schrödinger equations with potential. Reviews in Mathematical Physics, 17(10):1143–1207, 2005.
  • [GSS87] M. Grillakis, J. Shatah, and W. Strauss. Stability theory of solitary waves in the presence of symmetry, I. Journal of functional analysis, 74(1):160–197, 1987.
  • [GV79] J. Ginibre and G. Velo. On a class of nonlinear Schrödinger equations. I. The Cauchy problem, general case. J. Functional Analysis, 32(1):1–32, 1979.
  • [HL07] D. Hundertmark and Y.-R. Lee. Exponential decay of eigenfunctions and generalized eigenfunctions of a non-self-adjoint matrix Schrödinger operator related to NLS. Bulletin of the London Mathematical Society, 39(5):709–720, 2007.
  • [JN01] A. Jensen and G. Nenciu. A unified approach to resolvent expansions at thresholds. Rev. Math. Phys., 13(6):717–754, 2001.
  • [JSS91] J.-L. Journé, A. Soffer, and C. D. Sogge. Decay estimates for Schrödinger operators. Communications on Pure and Applied mathematics, 44(5):573–604, 1991.
  • [Kat87] T. Kato. On nonlinear Schrödinger equations. Ann. Inst. H. Poincaré Phys. Théor., 46(1):113–129, 1987.
  • [KMM17] M. Kowalczyk, Y. Martel, and C. Muñoz. On asymptotic stability of nonlinear waves. In Séminaire Laurent Schwartz—Équations aux dérivées partielles et applications. Année 2016–2017, pages Exp. No. XVIII, 27. Ed. Éc. Polytech., Palaiseau, 2017.
  • [KS06] J. Krieger and W. Schlag. Stable manifolds for all monic supercritical focusing nonlinear Schrödinger equations in one dimension. J. Amer. Math. Soc., 19(4):815–920, 2006.
  • [Kwo89] M.K. Kwong. Uniqueness of positive solutions of Δuu+up=0\Delta u-u+u^{p}=0 in RnR^{n}. Archive for Rational Mechanics and Analysis, 105:243–266, 1989.
  • [LLSS23] H. Lindblad, J. Lührmann, W. Schlag, and A. Soffer. On modified scattering for 1D quadratic Klein-Gordon equations with non-generic potentials. International Mathematics Research Notices, 2023(6):5118–5208, 2023.
  • [LS21] J. Lührmann and W. Schlag. Asymptotic stability of the sine-Gordon kink under odd perturbations. arXiv preprint arXiv:2106.09605, 2021.
  • [LS23] J. Lührmann and W. Schlag. On codimension one stability of the soliton for the 1D focusing cubic Klein-Gordon equation. arXiv preprint arXiv:2302.05273, 2023.
  • [Mae08] M. Maeda. Stability and instability of standing waves for 1-dimensional nonlinear Schrödinger equation with multiple-power nonlinearity. Kodai Math. J., 31(2):263–271, 2008.
  • [Mar11] Jeremy L. Marzuola. Dispersive estimates using scattering theory for matrix Hamiltonian equations. Discrete and Continuous Dynamical Systems, 30(4):995–1035, 2011.
  • [Mar23] Y. Martel. Asymptotic stability of solitary waves for the 1D cubic-quintic Schrödinger equation with no internal mode. Probability and Mathematical Physics, 3(4):839–867, 2023.
  • [Miz11] H. Mizutani. Dispersive estimates and asymptotic expansions for Schrödinger equations in dimension one. Journal of the Mathematical Society of Japan, 63(1):239–261, 2011.
  • [MMS20] S. Masaki, J. Murphy, and J.-I. Segata. Asymptotic stability of solitary waves for the 1d1d NLS with an attractive delta potential. arXiv preprint arXiv:2008.11645, 2020.
  • [Mur82] M. Murata. Asymptotic expansions in time for solutions of Schrödinger-type equations. J. Funct. Anal., 49(1):10–56, 1982.
  • [Oht95] M. Ohta. Stability and instability of standing waves for one-dimensional nonlinear Schrödinger equations with double power nonlinearity. Kodai Math. J., 18(1):68–74, 1995.
  • [Sch05] W. Schlag. Dispersive estimates for Schrödinger operators in dimension two. Comm. Math. Phys., 257(1):87–117, 2005.
  • [Sch06] W. Schlag. Spectral theory and nonlinear partial differential equations: a survey. Discrete Contin. Dyn. Syst., 15(3):703–723, 2006.
  • [Sch09] W. Schlag. Stable manifolds for an orbitally unstable nonlinear Schrödinger equation. Ann. of Math. (2), 169(1):139–227, 2009.
  • [SS85] J. Shatah and W. Strauss. Instability of nonlinear bound states. Communications in Mathematical Physics, 100(2):173–190, 1985.
  • [SS07] C. Sulem and P.-L. Sulem. The nonlinear Schrödinger equation: self-focusing and wave collapse, volume 139. Springer Science & Business Media, 2007.
  • [Ste93] E. M. Stein. Harmonic analysis: real-variable methods, orthogonality, and oscillatory integrals, volume 3. Princeton University Press, 1993.
  • [Top17] E. Toprak. A weighted estimate for two dimensional Schrödinger, matrix Schrödinger, and wave equations with resonance of the first kind at zero energy. Journal of Spectral Theory, 7(4):1235–1284, 2017.
  • [Tsu87] Y. Tsutsumi. Global strong solutions for nonlinear Schrödinger equations. Nonlinear Anal., 11(10):1143–1154, 1987.
  • [Vou10] V. Vougalter. On threshold eigenvalues and resonances for the linearized NLS equation. Math. Model. Nat. Phenom., 5(4):448–469, 2010.
  • [Wed00] R. Weder. LpL^{p}-Lp˙L^{\dot{p}} estimates for the Schrödinger equation on the line and inverse scattering for the nonlinear Schrödinger equation with a potential. J. Funct. Anal., 170(1):37–68, 2000.
  • [Wei85] M. I. Weinstein. Modulational stability of ground states of nonlinear Schrödinger equations. SIAM journal on mathematical analysis, 16(3):472–491, 1985.
  • [Wei86] M. I. Weinstein. Lyapunov stability of ground states of nonlinear dispersive evolution equations. Comm. Pure Appl. Math., 39(1):51–67, 1986.